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 31 毫秒
1.
Tissue transglutaminase (tTG) has been implicated in the pathogenesis of Parkinson disease (PD). However, exactly how tTG modulates the structural and functional properties of α-synuclein (α-syn) and contributes to the pathogenesis of PD remains unknown. Using site-directed mutagenesis combined with detailed biophysical and mass spectrometry analyses, we sought to identify the exact residues involved in tTG-catalyzed cross-linking of wild-type α-syn and α-syn mutants associated with PD. To better understand the structural consequences of each cross-linking reaction, we determined the effect of tTG-catalyzed cross-linking on the oligomerization, fibrillization, and membrane binding of α-syn in vitro. Our findings show that tTG-catalyzed cross-linking of monomeric α-syn involves multiple cross-links (specifically 2-3). We subjected tTG-catalyzed cross-linked monomeric α-syn composed of either wild-type or Gln → Asn mutants to sequential proteolysis by multiple enzymes and peptide mapping by mass spectrometry. Using this approach, we identified the glutamine and lysine residues involved in tTG-catalyzed intramolecular cross-linking of α-syn. These studies demonstrate for the first time that Gln79 and Gln109 serve as the primary tTG reactive sites. Mutating both residues to asparagine abolishes tTG-catalyzed cross-linking of α-syn and tTG-induced inhibition of α-syn fibrillization in vitro. To further elucidate the sequence and structural basis underlying these effects, we identified the lysine residues that form isopeptide bonds with Gln79 and Gln109. This study provides mechanistic insight into the sequence and structural basis of the inhibitory effects of tTG on α-syn fibrillogenesis in vivo, and it sheds light on the potential role of tTG cross-linking on modulating the physiological and pathogenic properties of α-syn.Parkinson disease (PD)2 is a progressive movement disorder that is caused by the loss of dopaminergic neurons in the substantia nigra, the part of the brain responsible for controlling movement. Clinically, PD is manifested in symptoms that include tremors, rigidity, and difficulty in initiating movement (bradykinesia). Pathologically, PD is characterized by the presence of intraneuronal, cytoplasmic inclusions known as Lewy bodies (LB), which are composed primarily of the protein “α-synuclein” (α-syn) (1) and are seen in the post-mortem brains of PD patients with the sporadic or familial forms of the disease (2). α-Syn is a presynaptic protein of 140 residues with a “natively” unfolded structure (3). Three missense point mutations in α-syn (A30P, E46K, and A53T) are associated with the early-onset, dominant, inherited form of PD (4, 5). Moreover, duplication or triplication of the α-syn gene has been linked to the familial form of PD, suggesting that an increase in α-syn expression is sufficient to cause PD. Together, these findings suggest that α-syn plays a central role in the pathogenesis of PD.The molecular and cellular determinants that govern α-syn oligomerization and fibrillogenesis in vivo remain poorly understood. In vitro aggregation studies have shown that the mutations associated with PD (A30P, E46K, and A53T) accelerate α-syn oligomerization, but only E46K and A53T α-syn show higher propensity to fibrillize than wild-type (WT) α-syn (6-8). This suggests that oligomerization, rather than fibrillization, is linked to early-onset familial PD (9). Our understanding of the molecular composition and biochemical state of α-syn in LBs has provided important clues about protein-protein interactions and post-translational modifications that may play a role in modulating oligomerization, fibrillogenesis, and LB formation of the protein. In addition to ubiquitination (10), phosphorylation (11, 12), nitration (13, 14), and C-terminal truncation (15, 16), analysis of post-mortem brain tissues from PD and Lewy bodies in dementia patients has confirmed the colocalization of tissue transglutaminase (tTG)-catalyzed cross-linked α-syn monomers and higher molecular aggregates in LBs within dopaminergic neurons (17, 18). Tissue transglutaminase catalyzes a calcium-dependent transamidating reaction involving glutamine and lysine residues, which results in the formation of a covalent cross-link via ε-(γ-glutamyl) lysine bonds (Fig. 2F). To date, seven different isoforms of tTGs have been reported, of which only tTG2 seems to be expressed in the human brain (19), whereas tTG1 and tTG3 are more abundantly found in stratified squamous epithelia (20). Subsequent immuno-histochemical, colocalization, and immunoprecipitation studies have shown that the levels of tTG and cross-linked α-syn species are increased in the substantia nigra of PD brains (17). These findings, combined with the known role of tTG in cross-linking and stabilizing bimolecular assemblies, led to the hypothesis that tTG plays an important role in the initiation and propagation of α-syn fibril formation and that it contributes to fibril stability in LBs. This hypothesis was initially supported by in vitro studies demonstrating that tTG catalyzes the polymerization of the α-syn-derived non-amyloid component (NAC) peptide via intermolecular covalent cross-linking of residues Gln79 and Lys80 (21) and by other studies suggesting that tTG promotes the fibrillization of amyloidogenic proteins implicated in the pathogenesis of other neurodegenerative diseases such as Alzheimer disease, supranuclear palsy, Huntington disease, and other polyglutamine diseases (22-24). However, recent in vitro studies with full-length α-syn have shown that tTG catalyzes intramolecular cross-linking of monomeric α-syn and inhibits, rather than promotes, its fibrillization in vitro (25, 26). The structural basis of this inhibitory effect and the exact residues involved in tTG-mediated cross-linking of α-syn, as well as structural and functional consequences of these modifications, remain poorly understood.Open in a separate windowFIGURE 2.tTG-catalyzed cross-linking of α-syn involves one to three intramolecular cross-links. A-C, MALDI-TOF/TOF analysis of native (—) and cross-linked (- - -) α-syn, showing that most tTG-catalyzed cross-linking products of WT or disease-associated mutant forms of α-syn are intramolecularly linked (predominant peak with two cross-links), and up to three intramolecular cross-links can occur (left shoulder). The abbreviations M and m/cl are used to designate native and cross-linked α-synuclein, respectively. D and E, kinetic analysis of α-syn (A30P) cross-linking monitored by MALDI-TOF and SDS-PAGE. F, schematic depiction of the tTG-catalyzed chemical reaction (isodipeptide formation) between glutamine and lysine residues.In this study, we have identified the primary glutamine and lysine residues involved in tTG-catalyzed, intramolecularly cross-linked monomeric α-syn and investigated how cross-linking these residues affects the oligomerization, fibrillization, and membrane binding of α-syn in vitro. Using single-site mutagenesis and mass spectrometry applied to exhaustive proteolytic digests of native and cross-linked monomeric α-syn, we identified Gln109 and Gln79 as the major tTG substrates. We demonstrate that the altered electrophoretic mobility of the intramolecularly cross-linked α-syn in SDS-PAGE occurs as a result of tTG-catalyzed cross-linking of Gln109 to lysine residues in the N terminus of α-syn, which leads to the formation of more compact monomers. Consistent with previous studies, we show that intramolecularly cross-linked α-syn forms off-pathway oligomers that are distinct from those formed by the wild-type protein and that do not convert to fibrils within the time scale of our experiments (3-5 days). We also show that membrane-bound α-syn is a substrate of tTG and that intramolecular cross-linking does not interfere with the ability of monomeric α-syn to adopt an α-helical conformation upon binding to synthetic membranes. These studies provide novel mechanistic insight into the sequence and structural basis of events that allow tTG to inhibit α-syn fibrillogenesis, and they shed light on the potential role of tTG-catalyzed cross-linking in modulating the physiological and pathogenic properties of α-syn.  相似文献   

2.
3.
Carbonic anhydrase (CA) (EC 4.2.1.1) enzymes catalyze the reversible hydration of CO2, a reaction that is important in many physiological processes. We have cloned and sequenced a full-length cDNA encoding an intracellular β-CA from the unicellular green alga Coccomyxa. Nucleotide sequence data show that the isolated cDNA contains an open reading frame encoding a polypeptide of 227 amino acids. The predicted polypeptide is similar to β-type CAs from Escherichia coli and higher plants, with an identity of 26% to 30%. The Coccomyxa cDNA was overexpressed in E. coli, and the enzyme was purified and biochemically characterized. The mature protein is a homotetramer with an estimated molecular mass of 100 kD. The CO2-hydration activity of the Coccomyxa enzyme is comparable with that of the pea homolog. However, the activity of Coccomyxa CA is largely insensitive to oxidative conditions, in contrast to similar enzymes from most higher plants. Fractionation studies further showed that Coccomyxa CA is extrachloroplastic.  相似文献   

4.
Six species of Gastrocopta have been identified from the Pilbara region, Western Australia, by means of comparative analyses of shell and mtDNA variation. Three of these species, Gastrocopta hedleyi, Gastrocopta larapinta and Gastrocopta servilis, have been recorded in the Pilbara for the first time. Gastrocopta sp. CW1 is probably new to science and might be endemic to the region. By contrast, Gastrocopta hedleyi, Gastrocopta larapinta and Gastrocopta mussoni are shown to be widespread.  相似文献   

5.
Jaskuła R 《ZooKeys》2011,(100):487-502
The tiger beetle fauna of the Balkan Peninsula is one of the richest in Europe and includes 19 species or 41% of the European tiger beetle fauna. Assembled by their biogeographical origins, the Balkan tiger beetle species fall into 14 different groups that include, Mediterranean, Middle Oriental, Central Asiatic, Euro-Siberian, South and East European, Pannonian-Sarmatian, West Palaearctic, Turano-European and Afrotropico Indo-Mediterranean species. The Mediterranean Sclerophyl and the Pontian Steppe are the Balkan biogeographical provinces with the highest species richness, while the Balkan Highlands has the lowest Cicindelidae diversity. Most species are restricted to single habitat types in lowland areas of the Balkan Peninsula and only Calomera aulica aulica and Calomera littoralis nemoralis occur in respectively 3 and 4 different types of habitat. About 60% of all Balkan Cicindelidae species are found in habitats potentially endangered by human activity.  相似文献   

6.
7.
ATP synthase uses a unique rotational mechanism to convert chemical energy into mechanical energy and back into chemical energy. The helix-turn-helix motif, termed “DELSEED-loop,” in the C-terminal domain of the β subunit was suggested to be involved in coupling between catalysis and rotation. Here, the role of the DELSEED-loop was investigated by functional analysis of mutants of Bacillus PS3 ATP synthase that had 3–7 amino acids within the loop deleted. All mutants were able to catalyze ATP hydrolysis, some at rates several times higher than the wild-type enzyme. In most cases ATP hydrolysis in membrane vesicles generated a transmembrane proton gradient, indicating that hydrolysis occurred via the normal rotational mechanism. Except for two mutants that showed low activity and low abundance in the membrane preparations, the deletion mutants were able to catalyze ATP synthesis. In general, the mutants seemed less well coupled than the wild-type enzyme, to a varying degree. Arrhenius analysis demonstrated that in the mutants fewer bonds had to be rearranged during the rate-limiting catalytic step; the extent of this effect was dependent on the size of the deletion. The results support the idea of a significant involvement of the DELSEED-loop in mechanochemical coupling in ATP synthase. In addition, for two deletion mutants it was possible to prepare an α3β3γ subcomplex and measure nucleotide binding to the catalytic sites. Interestingly, both mutants showed a severely reduced affinity for MgATP at the high affinity site.F1F0-ATP synthase catalyzes the final step of oxidative phosphorylation and photophosphorylation, the synthesis of ATP from ADP and inorganic phosphate. F1F0-ATP synthase consists of the membrane-embedded F0 subcomplex, with, in most bacteria, a subunit composition of ab2c10, and the peripheral F1 subcomplex, with a subunit composition of α3β3γδε. The energy necessary for ATP synthesis is derived from an electrochemical transmembrane proton (or, in some organisms, a sodium ion) gradient. Proton flow down the gradient through F0 is coupled to ATP synthesis on F1 by a unique rotary mechanism. The protons flow through (half) channels at the interface of the a and c subunits, which drives rotation of the ring of c subunits. The c10 ring, together with F1 subunits γ and ε, forms the rotor. Rotation of γ leads to conformational changes in the catalytic nucleotide binding sites on the β subunits, where ADP and Pi are bound. The conformational changes result in the formation and release of ATP. Thus, ATP synthase converts electrochemical energy, the proton gradient, into mechanical energy in the form of subunit rotation and back into chemical energy as ATP. In bacteria, under certain physiological conditions, the process runs in reverse. ATP is hydrolyzed to generate a transmembrane proton gradient, which the bacterium requires for such functions as nutrient import and locomotion (for reviews, see Refs. 16).F1 (or F1-ATPase) has three catalytic nucleotide binding sites located on the β subunits at the interface to the adjacent α subunit. The catalytic sites have pronounced differences in their nucleotide binding affinity. During rotational catalysis, the sites switch their affinities in a synchronized manner; the position of γ determines which catalytic site is the high affinity site (Kd1 in the nanomolar range), which site is the medium affinity site (Kd2 ≈ 1 μm), and which site is the low affinity site (Kd3 ≈ 30–100 μm; see Refs. 7 and 8). In the original crystal structure of bovine mitochondrial F1 (9), one of the three catalytic sites, was filled with the ATP analog AMP-PNP,2 a second was filled with ADP (plus azide) (see Ref. 10), and the third site was empty. Hence, the β subunits are referred to as βTP, βDP, and βE. The occupied β subunits, βTP and βDP, were in a closed conformation, and the empty βE subunit was in an open conformation. The main difference between these two conformations is found in the C-terminal domain. Here, the “DELSEED-loop,” a helix-turn-helix structure containing the conserved DELSEED motif, is in an “up” position when the catalytic site on the respective β subunit is filled with nucleotide and in a “down” position when the site is empty (Fig. 1A). When all three catalytic sites are occupied by nucleotide, the previously open βE subunit assumes an intermediate, half-closed (βHC) conformation. It cannot close completely because of steric clashes with γ (11).Open in a separate windowFIGURE 1.The βDELSEED-loop. A, interaction of the βTP and βE subunits with theγ subunit.β subunits are shown in yellow andγ in blue. The DELSEED-loop (shown in orange, with the DELSEED motif itself in green)of βTP interacts with the C-terminal helixγ and the short helix that runs nearly perpendicular to the rotation axis. The DELSEED-loop of βE makes contact with the convex portion of γ, formed mainly by the N-terminal helix. A nucleotide molecule (shown in stick representation) occupies the catalytic site of βTP, and the subunit is in the closed conformation. The catalytic site on βE is empty, and the subunit is in the open conformation. This figure is based on Protein Data Bank file 1e79 (32). B, deletions in the βDELSEED-loop. The loop was “mutated” in silico to represent the PS3 ATP synthase. The 3–4-residue segments that are removed in the deletion mutants are color-coded as follows: 380LQDI383, pink; 384IAIL387, green; 388GMDE391, yellow; 392LSD394, cyan; 395EDKL398, orange; 399VVHR402, blue. Residues that are the most involved in contacts with γ are labeled. All figures were generated using the program PyMOL (DeLano Scientific, San Carlos, CA).The DELSEED-loop of each of the three β subunits makes contact with the γ subunit. In some cases, these contacts consist of hydrogen bonds or salt bridges between the negatively charged residues of the DELSEED motif and positively charged residues on γ. The interactions of the DELSEED-loop with γ, its movement during catalysis, the conservation of the DELSEED motif (see 1214). Thus, the finding that an AALSAAA mutant in the α3β3γ complex of ATP synthase from the thermophilic Bacillus PS3, where several hydrogen bonds/salt bridges to γ are removed simultaneously, could drive rotation of γ with the same torque as the wild-type enzyme (14) came as a surprise. On the other hand, it seems possible that it is the bulk of the DELSEED-loop, more so than individual interactions, that drives rotation of γ. According to a model favored by several authors (6, 15, 16) (see also Refs. 1719), binding of ATP (or, more precisely, MgATP) to the low affinity catalytic site on βE and the subsequent closure of this site, accompanied by its conversion into the high affinity site, are responsible for driving the large (80–90°) rotation substep during ATP hydrolysis, with the DELSEED-loop acting as a “pushrod.” A recent molecular dynamics (20) study supports this model and implicates mainly the region around several hydrophobic residues upstream of the DELSEED motif (specifically βI386 and βL387)3 as being responsible for making contact with γ during the large rotation substep.

TABLE 1

Conservation of residues in the DELSEED-loop Amino acids found in selected species in the turn region of the DELSEED-loop. Listed are all positions subjected to deletions in the present study. Residue numbers refer to the PS3 enzyme. Consensus annotation: p, polar residue; s, small residue; h, hydrophobic residue; –, negatively charged residue; +, positively charged residue.Open in a separate windowIn the present study, we investigated the function of the DELSEED-loop using an approach less focused on individual residues, by deleting stretches of 3–7 amino acids between positions β380 and β402 of ATP synthase from the thermophilic Bacillus PS3. We analyzed the functional properties of the deletion mutants after expression in Escherichia coli. The mutants showed ATPase activities, which were in some cases surprisingly high, severalfold higher than the activity of the wild-type control. On the other hand, in all cases where ATP synthesis could be measured, the rates where below or equal to those of the wild-type enzyme. In Arrhenius plots, the hydrolysis rates of the mutants were less temperature-dependent than those of wild-type ATP synthase. In those cases where nucleotide binding to the catalytic sites could be tested, the deletion mutants had a much reduced affinity for MgATP at high affinity site 1. The functional role of the DELSEED-loop will be discussed in light of the new information.  相似文献   

8.
9.
The three key players in the exocytotic release of neurotransmitters from synaptic vesicles are the SNARE (soluble N-ethylmaleimide-sensitive factor attachment protein receptor) proteins synaptobrevin 2, syntaxin 1a, and SNAP-25. Their assembly into a tight four-helix bundle complex is thought to pull the two membranes into close proximity. It is debated, however, whether the energy generated suffices for membrane fusion. Here, we have determined the thermodynamic properties of the individual SNARE assembly steps by isothermal titration calorimetry. We found extremely large favorable enthalpy changes counterbalanced by positive entropy changes, reflecting the major conformational changes upon assembly. To circumvent the fact that ternary complex formation is essentially irreversible, we used a stabilized syntaxin-SNAP-25 heterodimer to study synaptobrevin binding. This strategy revealed that the N-terminal synaptobrevin coil binds reversibly with nanomolar affinity. This suggests that individual, membrane-bridging SNARE complexes can provide much less pulling force than previously claimed.The molecular machinery that drives the Ca2+-dependent release of neurotransmitters from synaptic vesicles is studied intensively. Three key players in the underlying exocytotic fusion of the vesicle with the plasma membrane are the proteins synaptobrevin 2/VAMP2 (vesicle-associated membrane protein), syntaxin 1a, and SNAP-253 (for review, see Refs. 1-7). They belong to the so-called SNARE protein family, the members of which are involved in all vesicle fusion steps in the endocytic and secretory pathway. In general, SNARE proteins are relatively small, tail-anchored membrane proteins. Their key characteristic is the so-called SNARE motif, an extended stretch of heptad repeats that is usually connected to a single transmembrane domain by a short linker. Syntaxin and synaptobrevin each contain a single SNARE motif, whereas SNAP-25 contains two SNARE motifs connected by a palmitoylated linker region serving as a membrane anchor. The SNARE motifs of the three proteins assemble into a very tight four-helix bundle between opposing membranes; during this process the plasma membrane proteins syntaxin and SNAP-25 provide the binding site for the vesicular synaptobrevin. Formation of this complex is accompanied by extensive structural rearrangements (8-10). Based on these findings, it was put forward that the formation of the SNARE bundle provides the energy that drives membrane fusion. As the bundle is oriented in parallel, it is thought that formation of this complex starts from the membrane-proximal N termini and proceeds toward the C-terminal membrane anchors, effectively pulling the membranes together (the “zipper” model) (11). Although the zipper scenario is intuitive, it has been difficult to demonstrate directly.A decade ago it was shown that the three neuronal SNARE proteins are sufficient to fuse artificial vesicles (12). However, this reductionist approach yields rather slow fusion rates (12-14). Over the years various different end products of SNARE catalysis (complete fusion, hemifusion, and only tethered membranes) have been reported (15-19). These unsatisfactory results have fueled the debate over whether the assembly process indeed provides enough impetus to fuse bilayers. Not surprisingly, an alternative scenario has been put forward in which repulsive forces between membranes bring the SNARE assembly to a grinding halt. According to this idea, other factors like the Ca2+ sensor synaptotagmin or the small soluble protein complexin are needed to induce membrane merging (20-22).In simple terms, to find out whether the SNARE complex assembly is enough for membrane fusion, only the amount of energy released during complex formation and the amount of energy needed for membrane fusion need to be compared. However, the physics of membrane fusion are very complicated, and it is even more challenging to understand how proteins modulate the process. The free energy for bilayer fusion in an aqueous environment is not very high, but fusion is thought to require a large activation energy of about 40 kBT, as two charged membranes have to be brought into close apposition. According to a theoretical model, the apposing membranes then need to be modified into a stalk-like configuration. Before fusion occurs, the process is thought to pass through a hemifusion intermediate in which only the outer monolayers are merged (for review, see Refs. 23 and 24). The role of fusion proteins is thought to lower the energy barrier for membrane fusion, but understanding how they modulate the lipid membrane and how their conformational changes are translated into a mechanical force is still in its infancy. It is not clear, for instance, whether SNARE-catalyzed fusion indeed proceeds through a stalk-like structure or just locally alters the membranes, a mechanism that might need much less activation energy.As the folding and unfolding transitions of the ternary SNARE complex exhibit a marked hysteresis (25), the question of how much energy is released during complex formation has been difficult to answer as well. To avoid the quasi-irreversibility of the process, the problem has been elegantly tackled by atomic force microscopy by two different research groups (26-28). In these experiments individual complexes affixed to solid supports were ruptured, yielding energy values of 43 and 33 kBT. In another approach, which used a surface-force apparatus (SFA), a comparable energy of 35 kBT has recently been determined (29). Strikingly, these values appear to correspond closely with the activation energy needed to fuse two membranes, substantiating the view that SNAREs are nano-fusion machineries. However, one should be cautious about the conclusion that these sophisticated procedures in fact yield the genuine SNARE assembly energy. For example, with the SFA approach, the number of complexes had to be deduced rather indirectly to estimate the free energy. Moreover, these approaches offered only indirect information about the reaction pathway.In this study we set out to determine the SNARE assembly energy more directly by using isothermal titration calorimetry (ITC) complemented by kinetic measurements. ITC is a powerful technique for studying the thermodynamics of macromolecular interactions by directly measuring the heat changes associated with complex formation, which at constant pressure is equal to the enthalpy change (ΔH). The titration approach also yields the stoichiometry (n), the entropy change (ΔS), and the association constant (KA) of the reaction. We studied the consecutive reaction steps individually to gain deeper insights into the rugged energy landscape of complex formation. To study synaptobrevin binding in isolation, we used a stabilized syntaxin-SNAP-25 heterodimer, which has been shown to greatly accelerate liposome fusion rates (30). This strategy revealed that the N-terminal coil of synaptobrevin binds reversibly, making it feasible to access the free energy of SNARE assembly. Overall, our results suggest that individual SNARE complexes might provide much less pulling energy than previously claimed.  相似文献   

10.
An analysis of the dietary content of haematophagous insects can provide important information about the transmission networks of certain zoonoses. The present study evaluated the potential of polymerase chain reaction-restriction fragment length polymorphism (PCR-RFLP) analysis of the mitochondrial cytochrome B (cytb) gene to differentiate between vertebrate species that were identified as possible sources of sandfly meals. The complete cytb gene sequences of 11 vertebrate species available in the National Center for Biotechnology Information database were digested with Aci I, Alu I, Hae III and Rsa I restriction enzymes in silico using Restriction Mapper software. The cytb gene fragment (358 bp) was amplified from tissue samples of vertebrate species and the dietary contents of sandflies and digested with restriction enzymes. Vertebrate species presented a restriction fragment profile that differed from that of other species, with the exception of Canis familiaris and Cerdocyon thous. The 358 bp fragment was identified in 76 sandflies. Of these, 10 were evaluated using the restriction enzymes and the food sources were predicted for four: Homo sapiens (1), Bos taurus (1) and Equus caballus (2). Thus, the PCR-RFLP technique could be a potential method for identifying the food sources of arthropods. However, some points must be clarified regarding the applicability of the method, such as the extent of DNA degradation through intestinal digestion, the potential for multiple sources of blood meals and the need for greater knowledge regarding intraspecific variations in mtDNA.  相似文献   

11.
The DSM-IV major depression "bereavement exclusion" (BE), which recognizes that depressive symptoms are sometimes normal in recently bereaved individuals, is proposed for elimination in DSM-5. Evidence cited for the BE's invalidity comes from two 2007 reviews purporting to show that bereavement-related depression is similar to other depression across various validators, and a 2010 review of subsequent research. We examined whether the 2007 and 2010 reviews and subsequent relevant literature support the BE's invalidity. Findings were: a) studies included in the 2007 reviews sampled bereavement-related depression groups most of whom were not BE-excluded, making them irrelevant for evaluating BE validity; b) three subsequent studies cited by the 2010 review as supporting BE elimination did examine BE-excluded cases but were in fact inconclusive; and c) two more recent articles comparing recurrence of BE-excluded and other major depressive disorder cases both support the BE's validity. We conclude that the claimed evidence for the BE's invalidity does not exist. The evidence in fact supports the BE's validity and its retention in DSM-5 to prevent false positive diagnoses. We suggest some improvements to increase validity and mitigate risk of false negatives.  相似文献   

12.
The multifunctional protein encoded by gene 4 of bacteriophage T7 (gp4) provides both helicase and primase activity at the replication fork. T7 DNA helicase preferentially utilizes dTTP to unwind duplex DNA in vitro but also hydrolyzes other nucleotides, some of which do not support helicase activity. Very little is known regarding the architecture of the nucleotide binding site in determining nucleotide specificity. Crystal structures of the T7 helicase domain with bound dATP or dTTP identified Arg-363 and Arg-504 as potential determinants of the specificity for dATP and dTTP. Arg-363 is in close proximity to the sugar of the bound dATP, whereas Arg-504 makes a hydrogen bridge with the base of bound dTTP. T7 helicase has a serine at position 319, whereas bacterial helicases that use rATP have a threonine in the comparable position. Therefore, in the present study we have examined the role of these residues (Arg-363, Arg-504, and Ser-319) in determining nucleotide specificity. Our results show that Arg-363 is responsible for dATP, dCTP, and dGTP hydrolysis, whereas Arg-504 and Ser-319 confer dTTP specificity. Helicase-R504A hydrolyzes dCTP far better than wild-type helicase, and the hydrolysis of dCTP fuels unwinding of DNA. Substitution of threonine for serine 319 reduces the rate of hydrolysis of dTTP without affecting the rate of dATP hydrolysis. We propose that different nucleotides bind to the nucleotide binding site of T7 helicase by an induced fit mechanism. We also present evidence that T7 helicase uses the energy derived from the hydrolysis of dATP in addition to dTTP for mediating DNA unwinding.Helicases are molecular machines that translocate unidirectionally along single-stranded nucleic acids using the energy derived from nucleotide hydrolysis (13). The gene 4 protein encoded by bacteriophage T7 consists of a helicase domain and a primase domain, located in the C-terminal and N-terminal halves of the protein, respectively (4). The T7 helicase functions as a hexamer and has been used as a model to study ring-shaped replicative helicases. In the presence of dTTP, T7 helicase binds to single-stranded DNA (ssDNA)3 as a hexamer and translocates 5′ to 3′ along the DNA strand using the energy of hydrolysis of dTTP (57). T7 helicase hydrolyzes a variety of ribo and deoxyribonucleotides; however, dTTP hydrolysis is optimally coupled to DNA unwinding (5).Most hexameric helicases use rATP to fuel translocation and unwind DNA (3). T7 helicase does hydrolyze rATP but with a 20-fold higher Km as compared with dTTP (5, 8). It has been suggested that T7 helicase actually uses rATP in vivo where the concentration of rATP is 20-fold that of dTTP in the Escherichia coli cell (8). However, hydrolysis of rATP, even at optimal concentrations, is poorly coupled to translocation and unwinding of DNA (9). Other ribonucleotides (rCTP, rGTP, and rUTP) are either not hydrolyzed or the poor hydrolysis observed is not coupled to DNA unwinding (8). Furthermore, Patel et al. (10) found that the form of T7 helicase found in vivo, an equimolar mixture of the full-length gp4 and a truncated form lacking the zinc binding domain of the primase, prefers dTTP and dATP. Therefore, in the present study we have restricted our examination of nucleotides to the deoxyribonucleotides.The nucleotide binding site of the replicative DNA helicases, such as T7 gene 4 protein, bind nucleotides at the subunit interface (Fig. 1) located between two RecA-like subdomains that bind ATP (11, 12). The location of the nucleotide binding site at the subunit interface provides multiple interactions of residues with the bound NTP. A number of cis- and trans-acting amino acids stabilize the bound nucleotide in the nucleotide binding site and also provide for communication between subunits (1315). Earlier reports revealed that the arginine finger (Arg-522) in T7 helicase is positioned to interact with the γ-phosphate of the bound nucleotide in the adjacent subunit (12, 16). However, His-465 (phosphate sensor), Glu-343 (catalytic base), and Asp-424 (Walker motif B) interacts with the γ-phosphate of the bound nucleotide in the same subunit (12, 17, 18). The arginine finger and the phosphate sensor have been proposed to couple NTP hydrolysis to DNA unwinding. Substitution of Glu-343, the catalytic base, eliminates dTTP hydrolysis (19), and substitution of Asp-424 with Asn leads to a severe reduction in dTTP hydrolysis (20). The conserved Lys-318 in Walker motif A interacts with the β-phosphate of the bound nucleotide and plays an important role in dTTP hydrolysis (21).Open in a separate windowFIGURE 1.Crystal structure of T7 helicase. A, crystal structure of the hexameric helicase C-terminal domain of gp4 (17). The structure reveals a ring-shaped molecule with a central core through which ssDNA passes. The inset shows the interface between two subunits of the helicase with adenosine 5′-{β,γ-imidol}-triphosphate in the nucleotide binding site. B, the nucleotide binding site of a monomer of the gp4 with the crucial amino acid residues reported earlier and in the present study is shown in sticks. The crystal structures of the T7 gene 4 helicase domain (12) with bound dTTP (C) and dATP (D). The structures shown are the nucleotide binding site of T7 helicase as viewed in Pymol by analyzing the PDB files 1cr1 and 1cr2 (12). Arg-504 and Tyr-535 sandwiches the base of the bound dNTP. Additionally, Arg-504 forms a hydrogen bridge with dTTP. Arg-363 interacts specifically with the 3-OH group of bound dATP. AMPPNP, adenosine 5′-(β,γ-imino)triphosphate.Considering the wealth of information on the above residues that are involved in the hydrolysis of dTTP and the coupling of hydrolysis to unwinding, it is intriguing that little information is available on nucleotide specificity. Several crystal structures of T7 helicase in complex with a nucleotide triphosphate are available. However, most of structures were crystallized with a non-hydrolyzable analogue of dTTP or the nucleotide was diffused into the crystal. The crystal structure of the T7 helicase domain bound with dTTP or dATP was reported by Sawaya et al. (12). These structures assisted us in identifying two basic residues (Arg-363 and Arg-504) in close proximity to the sugar and base of the bound nucleotide whose orientation suggested that these residues could be involved in nucleotide selection. Arg-504 together with Tyr-535 sandwich the base of the bound nucleotide at the subunit interface of the hexameric helicase (Fig. 1). Arg-504 and Tyr-535 are structurally well conserved in various helicases (12). However, Arg-504 could make a hydrogen bridge with the OH group of thymidine, thus suggesting a role in dTTP specificity. On the other hand, Arg-363 is in close proximity (∼3.4 Å) to the sugar 3′-OH of bound dATP, whereas in the dTTP-bound structure this residue is displaced by 7.12 Å (Fig. 1) from the equivalent position. Consequently Arg-363 could play a role in dATP binding. The crystal structures do not provide any information on different interaction of residues with the phosphates of dATP and dTTP. However, alignment of the residues in the P-loops of different hexameric helicases reveals that the serine adjacent to the invariant lysine at position 319 (Ser-319) is conserved in bacteriophages, whereas bacterial helicases have a conserved threonine in the equivalent position (supplemental Fig. 1). Bacterial helicases use rATP in the DNA unwinding reactions. whereas T7 helicase preferentially uses dTTP, and bacteriophage T4 gene 41 uses rGTP or rATP (22).Although considerable information is available on the role of residues in nucleotide binding and dTTP hydrolysis, very little is known on the determinants of nucleotide specificity. In the present study we made an attempt to address the role of a few selected residues (Arg-363, Arg-504, and Ser-319) in determining nucleotide specificity, especially dTTP and dATP, both of which are hydrolyzed and mediate DNA unwinding. We show that under physiological conditions T7 helicase uses the energy derived from the hydrolysis of dATP in addition to dTTP for mediating DNA unwinding.  相似文献   

13.
14.
15.
The European bone-skippers (Diptera: Piophilidae: Thyreophorina), long considered extinct, have recently been the object of much interest by dipterists after their unexpected rediscovery. Considerable faunistic work has been done on these flies in recent years. However, some nomenclatural and taxonomic issues still require attention. A neotype is designated for Thyreophora anthropophaga Robineau-Desvoidy, 1830 (now in the genus Centrophlebomyia Hendel, 1903) to fix the identity of this nominal species. Centrophlebomyia anthropophaga is recognized as a valid species. It is described and illustrated in detail, and information on its preimaginal instars is provided for the first time. Four Palaearctic species of Centrophlebomyia are recognized and reviewed and a key is provided for their identification. Centrophlebomyia orientalis Hendel, 1907 from northern India, is removed from synonymy with Centrophlebomyia anthropophaga and recognized as a valid species of Centrophlebomyia, stat. r. The nominal genus Protothyreophora Ozerov, 1984 is considered a junior synonym of Centrophlebomyia, syn. n.  相似文献   

16.
We present here three expression plasmids for Trypanosoma cruzi adapted to the Gateway® recombination cloning system. Two of these plasmids were designed to express trypanosomal proteins fused to a double tag for tandem affinity purification (TAPtag). The TAPtag and Gateway® cassette were introduced into an episomal (pTEX) and an integrative (pTREX) plasmid. Both plasmids were assayed by introducing green fluorescent protein (GFP) by recombination and the integrity of the double-tagged protein was determined by western blotting and immunofluorescence microscopy. The third Gateway adapted vector assayed was the inducible pTcINDEX. When tested with GFP, pTcINDEX-GW showed a good response to tetracycline, being less leaky than its precursor (pTcINDEX).  相似文献   

17.
We have previously shown that the C-terminal cystathionine β-synthase (CBS) domains of the nucleotide-binding domains of the ABC transporter OpuA, in conjunction with an anionic membrane surface function, act as sensor of internal ionic strength (Iin). Here, we show that a surface-exposed cationic region in the CBS module domain is critical for ion sensing. The consecutive substitution of up to five cationic residues led to a gradual decrease of the ionic strength dependence of transport. In fact, a 5-fold mutant was essentially independent of salt in the range from 0 to 250 mm KCl (or NaCl), supplemented to medium of 30 mm potassium phosphate. Importantly, the threshold temperature for transport was lowered by 5–7 °C and the temperature coefficient Q10 was lowered from 8 to ∼1.5 in the 5-fold mutant, indicating that large conformational changes are accompanying the CBS-mediated regulation of transport. Furthermore, by replacing the anionic C-terminal tail residues that extend the CBS module with histidines, the transport of OpuA became pH-dependent, presumably by additional charge interactions of the histidine residues with the membrane. The pH dependence was not observed at high ionic strength. Altogether the analyses of the CBS mutants support the notion that the osmotic regulation of OpuA involves a simple biophysical switching mechanism, in which nonspecific electrostatic interactions of a protein module with the membrane are sufficient to lock the transporter in the inactive state.In their natural habitats microorganisms are often exposed to changes in the concentration of solutes in the environment (1). A sudden increase in the medium osmolality results in loss of water from the cell, loss of turgor, a decrease in cell volume, and an increase in intracellular osmolyte concentration. Osmoregulatory transporters such as OpuA in Lactococcus lactis, ProP in Escherichia coli, and BetP in Corynebacterium glutamicum diminish the consequences of the osmotic stress by mediating the uptake of compatible solutes upon an increase in extracellular osmolality (24). For the ATP-binding cassette (ABC)5 transporter OpuA, it has been shown that the system, reconstituted in proteoliposomes, is activated by increased concentrations of lumenal ions (increased internal ionic strength) (2, 5, 6). This activation is instantaneous both in vivo and in vitro and only requires threshold levels of ionic osmolytes. Moreover, the ionic threshold for activation is highly dependent of the ionic lipid content (charge density) of the membrane and requires the presence of so-called cystathionine β-synthase (CBS) domains, suggesting that the ionic signal is transduced to the transporter via critical interactions of the protein with membrane lipids.The ABC transporter OpuA consists of two identical nucleotide-binding domains (NBD) fused to CBS domains and two identical substrate-binding domains fused to transmembrane domains. The NBD-CBS and substrate-binding domain-transmembrane domain subunits are named OpuAA and OpuABC, respectively. Two tandem CBS domains are linked to the C-terminal end of the NBD; each domain (CBS1 and CBS2) has a β-α-β-β-α secondary structure (5) (Fig. 1A). The CBS domains are widely distributed in most if not all species of life but their function is largely unknown. Most of the CBS domains are found as tandem repeats but data base searches have also revealed tetra-repeat units (5). The crystal structures of several tandem CBS domains have been elucidated (79, 32), and in a number of cases it has been shown that two tandem CBS domains form dimeric structures with a total of four CBS domains per structural module (hereafter referred to as CBS module). The crystal structures of the full-length MgtE Mg2+ transporter confirm the dimeric configuration and show that the CBS domains undergo large conformational changes upon Mg2+ binding or release (10, 11). In general, ABC transporters are functional as dimers, which implies that two tandem CBS domains are present in the OpuA complex. Preliminary experiments with disulfides engineered at the interface of two tandem CBS domains in OpuA suggest that large structural rearrangements (association-dissociation of the interfaces) play a determining role in the ionic strength-regulated transport. Finally, a subset of CBS-containing proteins has a C-terminal extension, which in OpuA is highly anionic (sequence: ADIPDEDEVEEIEKEEENK) and modulates the ion sensing activity (6).Open in a separate windowFIGURE 1.Domain structure of CBS module of OpuA. A, sequence of tandem CBS domains. The predicted secondary structure is indicated above the sequence. The residues modified in this study are underlined. The amino acid sequence end-points of OpuAΔ61 and OpuAΔ119 are indicated by vertical arrows. B, homology model of tandem CBS domain of OpuA. The CBS domains were individually modeled on the crystal structure of the tandem CBS protein Ta0289 from T. acidophilum (PDB entry 1PVM), using Phyre. Ta0289 was used for the initial modeling, because its primary sequence was more similar to the CBS domains of OpuA than those of the other crystallized CBS proteins. The individual domain models were then assembled with reference to the atomic coordinates of the tandem CBS domains of IMPDH from Streptococcus pyogenes (PDB entry 1ZFJ) to form the tandem CBS pair, using PyMOL (DeLano). The positions of the (substituted) cationic residues are indicated.In this study, we have engineered the surface-exposed cationic residues of the CBS module and the C-terminal anionic tail of OpuA (Fig. 1B). The ionic strength and lipid dependence of the OpuA mutants were determined in vivo and in vitro. We show that substitution of five cationic residues for neutral amino acids is sufficient to inactivate the ionic strength sensor and convert OpuA into a constitutively active transporter. Moreover, by substituting six anionic plus four neutral residues of the C-terminal anionic tail for histidines, the transport reaction becomes strongly pH-dependent.  相似文献   

18.
Here we describe and illustrate a new parasitoid wasp species, Lathrolestes gauldi sp. n. from the lowland rainforest of eastern Ecuador and provide a key to the Neotropical species of the genus. This is the first record of the subfamily Ctenopelmatinae from Ecuador.  相似文献   

19.
Classic and novel protein kinase C (PKC) isozymes contain two zinc finger motifs, designated “C1a” and “C1b” domains, which constitute the recognition modules for the second messenger diacylglycerol (DAG) or the phorbol esters. However, the individual contributions of these tandem C1 domains to PKC function and, reciprocally, the influence of protein context on their function remain uncertain. In the present study, we prepared PKCδ constructs in which the individual C1a and C1b domains were deleted, swapped, or substituted for one another to explore these issues. As isolated fragments, both the δC1a and δC1b domains potently bound phorbol esters, but the binding of [3H]phorbol 12,13-dibutyrate ([3H]PDBu) by the δC1a domain depended much more on the presence of phosphatidylserine than did that of the δC1b domain. In intact PKCδ, the δC1b domain played the dominant role in [3H]PDBu binding, membrane translocation, and down-regulation. A contribution from the δC1a domain was nonetheless evident, as shown by retention of [3H]PDBu binding at reduced affinity, by increased [3H]PDBu affinity upon expression of a second δC1a domain substituting for the δC1b domain, and by loss of persistent plasma membrane translocation for PKCδ expressing only the δC1b domain, but its contribution was less than predicted from the activity of the isolated domain. Switching the position of the δC1b domain to the normal position of the δC1a domain (or vice versa) had no apparent effect on the response to phorbol esters, suggesting that the specific position of the C1 domain within PKCδ was not the primary determinant of its activity.One of the essential steps for protein kinase C (PKC)2 activation is its translocation from the cytosol to the membranes. For conventional (α, βI, βII, and γ) and novel (δ, ε, η, and θ) PKCs, this translocation is driven by interaction with the lipophilic second messenger sn-1,2-diacylglycerol (DAG), generated from phosphatidylinositol 4,5-bisphosphate upon the activation of receptor-coupled phospholipase C or indirectly from phosphatidylcholine via phospholipase D (1). A pair of zinc finger structures in the regulatory domain of the PKCs, the “C1” domains, are responsible for the recognition of the DAG signal. The DAG-C1 domain-membrane interaction is coupled to a conformational change in PKC, both causing the release of the pseudosubstrate domain from the catalytic site to activate the enzyme and triggering the translocation to the membrane (2). By regulating access to substrates, PKC translocation complements the intrinsic enzymatic specificity of PKC to determine its substrate profile.The C1 domain is a highly conserved cysteine-rich motif (∼50 amino acids), which was first identified in PKC as the interaction site for DAG or phorbol esters (3). It possesses a globular structure with a hydrophilic binding cleft at one end surrounded by hydrophobic residues. Binding of DAG or phorbol esters to the C1 domain caps the hydrophilic cleft and forms a continuous hydrophobic surface favoring the interaction or penetration of the C1 domain into the membrane (4). In addition to the novel and classic PKCs, six other families of proteins have also been identified, some of whose members possess DAG/phorbol ester-responsive C1 domains. These are the protein kinase D (5), the chimaerin (6), the munc-13 (7), the RasGRP (guanyl nucleotide exchange factors for Ras and Rap1) (8), the DAG kinase (9), and the recently characterized MRCK (myotonic dystrophy kinase-related Cdc42-binding kinase) families (10). Of these C1 domain-containing proteins, the PKCs have been studied most extensively and are important therapeutic targets (11). Among the drug candidates in clinical trials that target PKC, a number such as bryostatin 1 and PEP005 are directed at the C1 domains of PKC rather than at its catalytic site.Both the classic and novel PKCs contain in their N-terminal regulatory region tandem C1 domains, C1a and C1b, which bind DAG/phorbol ester (12). Multiple studies have sought to define the respective roles of these two C1 domains in PKC regulation, but the issue remains unclear. Initial in vitro binding measurements with conventional PKCs suggested that 1 mol of phorbol ester bound per mole of PKC (13-15). On the other hand, Stubbs et al., using a fluorescent phorbol ester analog, reported that PKCα bound two ligands per PKC (16). Further, site-directed mutagenesis of the C1a and C1b domains of intact PKCα indicated that the C1a and C1b domains played equivalent roles for membrane translocation in response to phorbol 12-myristate 13-acetate (PMA) and (-)octylindolactam V (17). Likewise, deletion studies indicated that the C1a and C1b domains of PKCγ bound PDBu equally with high potency (3, 18). Using a functional assay with PKCα expression in yeast, Shieh et al. (19) deleted individual C1 domains and reported that C1a and C1b were both functional and equivalent upon stimulation by PMA, with either deletion causing a similar reduction in potency of response, whereas for mezerein the response depended essentially on the C1a domain, with much weaker response if only the C1b domain was present. Using isolated C1 domains, Irie et al. (20) suggested that the C1a domain of PKCα but not those of PKCβ or PKCγ bound [3H]PDBu preferentially; different ligands showed a generally similar pattern but with different extents of selectivity. Using synthesized dimeric bisphorbols, Newton''s group reported (21) that, although both C1 domains of PKCβII are oriented for potential membrane interaction, only one C1 domain bound ligand in a physiological context.In the case of novel PKCs, many studies have been performed on PKCδ to study the equivalency of the twin C1 domains. The P11G point mutation of the C1a domain, which caused a 300-fold loss of binding potency in the isolated domain (22), had little effect on the phorbol ester-dependent translocation of PKCδ in NIH3T3 cells, whereas the same mutation of the C1b caused a 20-fold shift in phorbol ester potency for inducing translocation, suggesting a major role of the C1b domain for phorbol ester binding (23). A secondary role for the C1a domain was suggested, however, because mutation in the C1a domain as well as the C1b domain caused a further 7-fold shift in potency. Using the same mutations in the C1a and C1b domains, Bögi et al. (24) found that the binding selectivity for the C1a and C1b domains of PKCδ appeared to be ligand-dependent. Whereas PMA and the indole alkaloids indolactam and octylindolactam were selectively dependent on the C1b domain, selectivity was not observed for mezerein, the 12-deoxyphorbol 13-monoesters prostratin and 12-deoxyphorbol 13-phenylacetate, and the macrocyclic lactone bryostatin 1 (24). In in vitro studies using isolated C1a and C1b domains of PKCδ, Cho''s group (25) described that the two C1 domains had opposite affinities for DAG and phorbol ester; i.e. the C1a domain showed high affinity for DAG and the C1b domain showed high affinity for phorbol ester. No such difference in selectivity was observed by Irie et al. (20).PKC has emerged as a promising therapeutic target both for cancer and for other conditions, such as diabetic retinopathy or macular degeneration (26-30). Kinase inhibitors represent one promising approach for targeting PKC, and enzastaurin, an inhibitor with moderate selectivity for PKCβ relative to other PKC isoforms (but still with activity on some other non-PKC kinases) is currently in multiple clinical trials. An alternative strategy for drug development has been to target the regulatory C1 domains of PKC. Strong proof of principle for this approach is provided by multiple natural products, e.g. bryostatin 1 and PEP005, which are likewise in clinical trials and which are directed at the C1 domains. A potential advantage of this approach is the lesser number of homologous targets, <30 DAG-sensitive C1 domains compared with over 500 kinases, as well as further opportunities for specificity provided by the diversity of lipid environments, which form a half-site for ligand binding to the C1 domain. Because different PKC isoforms may induce antagonistic activities, inhibition of one isoform may be functionally equivalent to activation of an antagonistic isoform (31).Along with the benzolactams (20, 32), the DAG lactones have provided a powerful synthetic platform for manipulating ligand: C1 domain interactions (31). For example, the DAG lactone derivative 130C037 displayed marked selectivity among the recombinant C1a and C1b domains of PKCα and PKCδ as well as substantial selectivity for RasGRP relative to PKCα (33). Likewise, we have shown that a modified DAG lactone (dioxolanones) can afford an additional point of contact in ligand binding to the C1b domain of PKCδ (34). Such studies provide clear examples that ligand-C1 domain interactions can be manipulated to yield novel patterns of recognition. Further selectivity might be gained with bivalent compounds, exploiting the spacing and individual characteristics of the C1a and C1b domains (35). A better understanding of the differential roles of the two C1 domains in PKC regulation is critical for the rational development of such compounds. In this study, by molecularly manipulating the C1a or C1b domains in intact PKCδ, we find that both the C1a and C1b domains play important roles in PKCδ regulation. The C1b domain is predominant for ligand binding and for membrane translocation of the whole PKCδ molecule. The C1a domain of intact PKCδ plays only a secondary role in ligand binding but stabilizes the PKCδ molecule at the plasma membrane for downstream signaling. In addition, we show that the effect of the individual C1 domains of PKCδ does not critically depend on their position within the regulatory domain.  相似文献   

20.
Duvalius (sg. Neoduvalius) gejzadunayi sp. n. from Pećina u Dubokom potoku cave ( Donje Biševo village near Rožaje, Montenegro), the first known representative of this subgenus from the territory of Montenegro is described, illustrated and compared with the related species of the subgenus Neoduvalius Müller, 1913. This new species is characterised by depigmented, medium sized body, totally reduced eyes, deep and complete frontal furrows, 3–4 pairs of discal setae in third elytral stria, as well as by the shape of aedeagus. Data on the distribution and the ecology of this remarkable species, as well as a check-list of the subgenus Neoduvalius are also provided. Recently described genera Serboduvalius Ćurčić, S. B. Pavićević & Ćurčić, B.P.M., 2001, Rascioduvalius Ćurčić, S. B. Brajković, Mitić & Ćurčić, B.P.M., 2003, Javorella Ćurčić, S. B. Brajković, Ćurčić, B.P.M. & Mitić, 2003 and Curcicia Ćurčić, S. B. & Brajković, 2003 are regarded as junior synonyms of the genus Duvalius Delarouzée.  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