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.
Size exclusion chromatographic analyses showed that Ca2+‐free VILIP‐1 contained both monomeric and dimeric forms, while no appreciable dimerization was noted with Ca2+‐free VILIP‐3. Swapping of EF‐hands 3 and 4 of VILIP‐1 with those of VILIP‐3 caused the inability of the resulting chimeric protein to form dimeric protein. Nonreducing SDS‐PAGE analyses revealed that most of the dimeric VILIP‐1 was noncovalently bound together. Reduced glutathione (GSH)/oxidized glutathione (GSSG) treatment notably enhanced the formation of disulfide‐linked VILIP‐1 dimer, while Ca2+ and Mg2+ enhanced disulfide dimerization of VILIP‐1 marginally in the presence of thiol compounds. Cys‐187 at the C‐terminus of VILIP‐1 contributed greatly to form S‐S‐crosslinked dimer as revealed by mutagenesis studies. The ability of GSH/GSSG‐treated VILIP‐1 to activate guanylyl cyclase B was reduced by substituting Cys‐187 with Ala. Together with disulfide dimer of VILIP‐1 detected in rat brain extracts, our data may imply the functional contribution of disulfide dimer to the interaction of VILIP‐1 with its physiological target(s). Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

2.
Studying the dimeric RNA structural organization is a step toward the understanding of retroviral genomic RNA dimerization. A kissing loop dimer is rearranged into an extended dimer during maturation of the virus particle. The extended dimer formation may be inhibited by ligands interacting with the RNA kissing loop dimer. A study was made of the interaction of dimeric RNA with paromomycin and magnesium ions. RNA dimers were formed from two hairpin RNAs having complementary sequences in the loop. The structural features of RNA dimers and the influence of the ligands were inferred from the fluorescence of 2-aminopurine (2-AP) incorporated in one of the two RNA hairpin sequences. As dimeric RNA interacted with paromomycin, 2-AP fluorescence increased. The increase was explained by a flipping of the fluorescent base out of the RNA structure. The binding constants and stoichiometry were estimated for dimeric RNA binding with paromomycin. An RNA dimer was found to interact with two paromomycin molecules; the binding constant was approximately the same (about 3 × 105 M−1) for both types of dimers. It was observed that the antibiotic and Mg2+ ions compete for binding to the hairpin RNA dimer and that one paromomycin molecule is displaced by one Mg2+ ion.  相似文献   

3.
In chloroform solution, the D ,L -alternating stereo-co-oligopeptide HCO-L -Phe-(D -Phe-L -Phe)3-OMe (I) forms three major species, two of which are dimeric and one tetrameric. One of the two dimeric species gives a specific set of 1H-nmr signals at 25°C; the other, together with the tetrameric species, gives another set of resonance signals. In a carbon tetrachloride or cyclohexane solution at 25°C, I forms virtually only the tetrameric species. From the nmr data, it can be shown that the dimeric and tetrameric species, that are in rapid equilibrium with each other in chloroform solutions, are a right-handed ↑↑β5.6 helical dimer and the head-to-head (formyl-ends-to-formyl-ends) dimerization product of this dimer. It is suggested that the linear gramicidins may also form head-to-head dimers of parallel β helices, as observed for the model oligopeptide I.  相似文献   

4.
Protein disulfide isomerases (PDIs) are responsible for catalyzing the proper oxidation and isomerization of disulfide bonds of newly synthesized proteins in the endoplasmic reticulum (ER). Here, it is shown that human PDI (PDIA1) dimerizes in vivo and proposed that the dimerization of PDI has physiological relevance by autoregulating its activity. The crystal structure of the dimeric form of noncatalytic bb′ domains of human PDIA1 determined to 2.3 Å resolution revealed that the formation of dimers occludes the substrate binding site and may function as a mechanism to regulate PDI activity in the ER.  相似文献   

5.
Hisatoshi Mimura 《BBA》2005,1708(3):393-403
The H+-pyrophosphatase (H+-PPase) consists of a single polypeptide, containing 16 or 17 transmembrane domains. To determine the higher order oligomeric state of Streptomyces coelicolor H+-PPase, we constructed a series of cysteine substitution mutants and expressed them in Escherichia coli. Firstly, we analyzed the formation of disulfide bonds, promoted by copper, in mutants with single cysteine substitutions. 28 of 39 mutants formed disulfide bonds, including S545C, a substitution at the periplasmic side. The formation of intermolecular disulfide bonds suppressed the enzyme activity of several, where the substituted residues were located in the cytosol. Creating disulfide links in the cytosol may interfere with the enzyme's catalytic function. Secondly, we prepared double mutants by introducing second cysteine substitutions into the S545C mutant. These double-cysteine mutants produced cross-linked complexes, estimated to be at least tetramers and possibly hexamers. Thirdly, we co-expressed epitope-tagged, wild type, and inactive mutant H+-PPases in E. coli and confirmed the formation of oligomers by co-purifying one subunit using the epitope tag used to label the other. The enzyme activity of these oligomers was markedly suppressed. We propose that H+-PPase is present as an oligomer made up of at least two or three sets of dimers.  相似文献   

6.
Hepatitis B virus “e-antigen” (HBeAg) is thought to be a soluble dimeric protein that is associated with chronic infection. It shares 149 residues with the viral capsid protein “core-antigen” (HBcAg), but has an additional 10-residue, hydrophobic, cysteine-containing amino-terminal propeptide whose presence correlates with physical, serological, and immunological differences between the two proteins. In HBcAg dimers, the subunits pair by forming a four-helix bundle stabilized by an intermolecular disulfide bond. The structure of HBeAg is probably similar but, instead, has two intramolecular disulfide bonds involving the propeptide. To compare the proteins directly and thereby clarify the role of the propeptide, we identified mutations and solution conditions that render both proteins as either soluble dimers or assembled capsids. Thermally induced unfolding monitored by circular dichroism, and electrophoresis of oxidized and reduced dimers, showed that the propeptide has a destabilizing effect and that the intramolecular disulfide bond forms preferentially and blocks the formation of the intermolecular disulfide bond that otherwise stabilizes the dimer. The HBeAg capsids are less regular than the HBcAg capsids; nevertheless, cryo-electron microscopy reconstructions confirm that they are constructed of dimers resembling those of HBcAg capsids. In them, a portion of the propeptide is visible near the dimer interface, suggesting that it intercalates there, consistent with the known formation of a disulfide bond between C(− 7) in the propeptide and C61 in the dimer interface. However, this intercalation distorts the dimer into an assembly-reluctant conformation.  相似文献   

7.
The hexameric complex of globular domains of type IV collagen was isolated after collagenase digestion of human placenta and the different monomers and dimers present were chromatographically separated. The ratio of alpha 1(IV)NC1 to alpha 2(IV)NC1 was 2:1. About 50% of the NC1 domains were connected to dimers. Predominantly alpha 1-alpha 1 dimers were found. Only 12% were alpha 2-alpha 2 dimers and no alpha 1-alpha 2 dimers could be detected. The majority (88%) of the intermolecular bonds was found to be disulfide bridges. The remainder could not be cleaved by reduction. To elucidate the arrangement of the disulfide bonds, the unreduced alpha 1(IV)NC1 monomers were treated with cyanogen bromide, the disulfide-bridged peptides isolated and characterized by Edman degradation. Each of the two homologous subdomains within a monomer is stabilized by an identical set of three disulfide bonds. In subdomain I, cysteines at positions 20 and 53 are connected with the C-terminal cysteine pair 108 and 111. Thus formed, the disulfide knot stabilizes two interconnected loops of 32 and 54 residues, respectively. A smaller loop of five residues occurs due to a disulfide bond between the cysteines 65 and 71. A similar disulfide arrangement is indicated for subdomain II which is separated from subdomain I by a segment of 20 amino acid residues. The same arrangement of disulfide bonds has been strongly suggested for the alpha 2(IV)NC1 monomer by the isolation and characterization of its disulfide-bridged tryptic fragments. Similar investigations on the dimeric alpha 1(IV)NC1 domain established the arrangement of the intermolecular disulfide bonds. They are formed by a complete disulfide exchange between corresponding disulfide knots of two monomeric NC1 domains.  相似文献   

8.
Purified enolase from Bacillus subtilis has a native mass of approximately 370 kDa. Since B. subtilis enolase was found to have a subunit mass of 46.58 kDa, the quaternary structure of B. subtilis is octameric. The pl for B. subtilis enolase is 6.1, the pH optimum (pHo) for activity is 8.1–8.2, and the K m for 2-PGA is approximately 0.67 mM. Using the dimeric Cα structure of yeast dimeric enolase as a guide, these dimers were arranged as a tetramer of dimers to simulate the electron microscopy image processing obtained for the octameric enolase purified from Thermotoga maritima. This arrangement allowed identification of helix J of one dimer (residues 86–96) and the loop between helix L and strand 1 (HL–S1 loop) of another dimer as possible subunit interaction regions. Alignment of available enolase amino acid sequences revealed that in 16 there are two tandem glycines at the C-terminal end of helix L and the HL–S1 loop is truncated by 4–6 residues relative to the yeast polypeptide, two structural features absent in enolases known to be dimers. From these arrangements and alignments it is proposed that the GG tandem at the C-terminal end of helix L and truncation of the HL–S1 loop may play a critical role in octamer formation of enolases. Interestingly, the sequence features associated with dimeric quaternary structure are found in three phylogenetically disparate groups, suggesting that the ancestral enolase was an octamer and that the dimeric structure has arisen independently multiple times through evolutionary history.  相似文献   

9.
Aspartokinase I - homoserine dehydrogenase I from Escherichia coli K-12, a homotetrameric enzyme, dissociates into dimers upon alkaline treatment. Both aspartokinase and homoserine dehydrogenase inactivation, as well as desensitazion towards L-threonine, occur in a multi-step process. Dithiothreitol stabilizes a dimeric form retaining full activity and sensitivity; L-homoserine stabilizing another dimeric form devoid of aspartokinase activity and retaining a substantial dehydrogenase activity insensitive toward L-threonine. A model is proposed showing that dissociation into dimers occurs in a first step, the resulting dimer losing both aspartokinase and homoserine dehydrogenase sensitivity in two subsequent steps involving the formation of intrachain disulfide bonds.  相似文献   

10.
S100 proteins constitute a large subfamily of the EF-hand superfamily of calcium binding proteins. They possess one classical EF-hand Ca2+-binding domain and an atypical EF-hand domain. Most of the S100 proteins form stable symmetric homodimers. An analysis of literature data on S100 proteins showed that their physiological concentrations could be much lower than dissociation constants of their dimeric forms. It means that just monomeric forms of these proteins are important for their functioning. In the present work, thermal denaturation of apo-S100P protein monitored by intrinsic tyrosine fluorescence has been studied at various protein concentrations within the region from 0.04–10 μM. A transition from the dimeric to monomeric form results in a decrease in protein thermal stability shifting the mid-transition temperature from 85 to 75 °C. Monomeric S100P immobilized on the surface of a sensor chip of a surface plasmon resonance instrument forms calcium dependent 1 to 1 complexes with human interleukin-11 (equilibrium dissociation constant 1.2 nM). In contrast, immobilized interleukin-11 binds two molecules of dimeric S100P with dissociation constants of 32 nM and 288 nM. Since effective dissociation constant of dimeric S100P protein is very low (0.5 μM as evaluated from our data) the sensitivity of the existing physical methods does not allow carrying out a detailed study of S100P monomer properties. For this reason, we have used molecular dynamics methods to evaluate structural changes in S100P upon its transition from the dimeric to monomeric state. 80-ns molecular dynamics simulations of kinetics of formation of S100P, S100B and S100A11 monomers from the corresponding dimers have been carried out. It was found that during the transition from the homo-dimer to monomer form, the three S100 monomer structures undergo the following changes: (1) the helices in the four-helix bundles within each monomer rotate in order to shield the exposed non-polar residues; (2) almost all lost contacts at the dimer interface are substituted with equivalent and newly formed interactions inside each monomer, and new stabilizing interactions are formed; and (3) all monomers recreate functional hydrophobic cores. The results of the present study show that both dimeric and monomeric forms of S100 proteins can be functional.  相似文献   

11.
Bovine β-casein (β-CN) is a highly amphiphilic micellising phospho-protein showing chaperone-like activity in vitro. Recently, existence of multiple sequential epitopes on β-CN polypeptide chain in both hydrophilic–polar (ψ) and hydrophobic–apolar domains (φ) has been evidenced. In order to clarify specific contribution of polar and apolar domains in micellisation process and in shaping immunoreactivity of β-CN, its dimeric/bi-amphiphilic “quasi palindromic” forms covalently connected by a disulfide bond linking either N-terminal (C4 β-CND) or C-terminal domain (C208 β-CND) were produced and studied. Depending on the C- or N-terminal position of inserted cysteine, each dimeric β-CN contains one polar/apolar region at the centre and two external hydrophobic/hydrophilic ends. Consequently, such casein dimers have radically different polarities/hydrophobicities on their outside surfaces. Dynamic light scattering (DLS) measurements indicate that these dimeric casein molecules form micelles of different sizes depending on arrangement of polar fragments of the β-CN mutants in their constrained dimers. Non-aggregated dimers have different hydrodynamic diameters that could be explained by their different geometries. Measurements of fluorescence showed more hydrophobic environment of Trp residues of C208 β-CND, while in similar experimental conditions Trp residues of C4 β-CND and native β-CN were more exposed to the polar medium. Both fluorescence and DLS studies showed greater propensity for micellisation of the dimeric β-CNs, suggesting that the factors inducing the formation of micelles are stronger in the bi-amphiphilic dimers. 1-Anilino-naphthalene-8-sulfonate (ANS) binding studies showed different binding of ANS by these dimers as well as different exposition of ANS binding (hydrophobic) regions in the micellar states. The differences in fluorescence resonance energy transfer (FRET) profiles of C4 β-CND and C208 β-CND can be explained by differences of distances and/or by differences of relative orientations of the donor (Trp) and acceptor (ANS), as well as by differences in quenching properties of the disulfide bridges and intra-molecular hydrophobic interactions. The immunoreactivity assays showed somewhat lower IgE response to C208 β-CND than to C4 β-CND. Thus, dimerization of C208 β-CN, connecting two C-terminal hydrophobic domains of two monomers doubling long-range hydrophobic interactions, possibly may hide a part of epitopes in the hydrophobic interface/core of C208 β-CND that is consistent with the results of DLS and fluorescence studies. The obtained results indicate structural differences of dimers – possibly the formation of Y- and U-shaped structures for C208 β-CND and C4 β-CND, respectively. This study not only demonstrated the importance of the organization of polar and hydrophobic regions during micellisation of the constrained and oriented β-CN dimers but also confirmed a possible role of C-terminal hydrophobic domain in the immunoreactivity profile of native β-CN.  相似文献   

12.
The d-/l-peptide gramicidin A (gA) is well known as a pivotal ion channel model and shows a broad spectrum of bioactivities such as antibiosis, antimalarial activity, as well as hemolysis. We applied inter-chain disulfide bonds to constrain the conformational freedom of gA into parallel and antiparallel dimeric topologies. Albeit the constructs were not found to be monoconformational, CD- and IR-spectroscopic studies suggested that this strategy indeed restricted the conformational space of the d-/l-peptide construct, and that β-helical secondary structures prevail. Correlative testing of gA dimers in antimicrobial, antimalarial, and ion conduction assays suggested that the tail-to-tail antiparallel single stranded β6.3 helix dominantly mediates the bioactivity of gA. Other conformers are unlikely to contribute to these activities. From these investigations, only weakly ion conducting gA dimers were identified that retained nM antimalarial activity.  相似文献   

13.
Phospholipases A2 may exist in solution both as monomers and dimers, but enzymes that form strong dimers (K D approximately 10–9 M) have been found, thus far, only in venoms of the snake family Crotilidae. The complete amino acid sequences of a basic monomeric and an acidic dimeric phospholipase A2 fromAgkistrodon piscivorus piscivorus (American cotton-mouth water moccasin) venom have been determined by protein sequencing methods as part of a search for aspects of structure contributing to formation of stable dimers. Both the monomeric and dimeric phospholipases A2 are highly homologous to the dimeric phospholipases A2 fromCrotalus atrox andCrotalus adamanteus venoms, and both have the seven residue carboxy-terminal extension characteristic of the crotalid and viperid enzymes. Thus, it is clear that the extension is not a prerequisite for dimerization. Studies to date have revealed two characteristic features of phosphilipases A2 that exist in solution as strong dimers. One is the presence in the dimers of a Pro-Pro sequence at position 112 and 113 which just precedes the seven residue carboxy-terminal extension (residues 116–122). The other is a low isoelectric point; only the acidic phospholipases A2 have been observed, thus far, to form stable dimers. These, alone or together, may be necessary, though not sufficient conditions for phospholipase A2 dimer formation. Ideas regarding subunit interactions based upon crystallographic data are evaluated relative to the new sequence information on the monomeric and dimeric phospholipases A2 fromA. p. piscivorus venom.  相似文献   

14.
Phospholipases A2 may exist in solution both as monomers and dimers, but enzymes that form strong dimers (K D approximately 10?9 M) have been found, thus far, only in venoms of the snake family Crotilidae. The complete amino acid sequences of a basic monomeric and an acidic dimeric phospholipase A2 fromAgkistrodon piscivorus piscivorus (American cotton-mouth water moccasin) venom have been determined by protein sequencing methods as part of a search for aspects of structure contributing to formation of stable dimers. Both the monomeric and dimeric phospholipases A2 are highly homologous to the dimeric phospholipases A2 fromCrotalus atrox andCrotalus adamanteus venoms, and both have the seven residue carboxy-terminal extension characteristic of the crotalid and viperid enzymes. Thus, it is clear that the extension is not a prerequisite for dimerization. Studies to date have revealed two characteristic features of phosphilipases A2 that exist in solution as strong dimers. One is the presence in the dimers of a Pro-Pro sequence at position 112 and 113 which just precedes the seven residue carboxy-terminal extension (residues 116–122). The other is a low isoelectric point; only the acidic phospholipases A2 have been observed, thus far, to form stable dimers. These, alone or together, may be necessary, though not sufficient conditions for phospholipase A2 dimer formation. Ideas regarding subunit interactions based upon crystallographic data are evaluated relative to the new sequence information on the monomeric and dimeric phospholipases A2 fromA. p. piscivorus venom.  相似文献   

15.
Purified enolase from Bacillus subtilis has a native mass of approximately 370 kDa. Since B. subtilis enolase was found to have a subunit mass of 46.58 kDa, the quaternary structure of B. subtilis is octameric. The pl for B. subtilis enolase is 6.1, the pH optimum (pHo) for activity is 8.1–8.2, and the K m for 2-PGA is approximately 0.67 mM. Using the dimeric C structure of yeast dimeric enolase as a guide, these dimers were arranged as a tetramer of dimers to simulate the electron microscopy image processing obtained for the octameric enolase purified from Thermotoga maritima. This arrangement allowed identification of helix J of one dimer (residues 86–96) and the loop between helix L and strand 1 (HL–S1 loop) of another dimer as possible subunit interaction regions. Alignment of available enolase amino acid sequences revealed that in 16 there are two tandem glycines at the C-terminal end of helix L and the HL–S1 loop is truncated by 4–6 residues relative to the yeast polypeptide, two structural features absent in enolases known to be dimers. From these arrangements and alignments it is proposed that the GG tandem at the C-terminal end of helix L and truncation of the HL–S1 loop may play a critical role in octamer formation of enolases. Interestingly, the sequence features associated with dimeric quaternary structure are found in three phylogenetically disparate groups, suggesting that the ancestral enolase was an octamer and that the dimeric structure has arisen independently multiple times through evolutionary history.  相似文献   

16.
Isolated oxoglutarate carrier (OGC) can be cross-linked to dimers by disulfide-forming reagents such as Cu2+-phenanthroline and diamide. Acetone and other solvents increase the extent of Cu2+-phenanthroline-induced cross-linking of OGC. Cross-linked OGC re-incorporated in photeoliposomes fully retains the oxoglutarate transport activity. The amount of cross-linked OGC calculated by densitometry of scanned gels depends on the method of staining, since cross-linked OGC exhibits a higher sensitivity to Coomassie brilliant blue as compared to silver nitrate. Under optimal conditions the formation of cross-linked OGC dimer (stained with Coomassie brilliant blue) amounts to 75% of the total protein. Approximately the same cross-linking efficiency was evaluated from Western blots. Cross-linking of OGC is prevented by SH reagents and reversed by SH-reducing reagents, which shows that it is mediated by disulfide bridge(s). The formation of SS bridge(s) requires the native state of the protein, since it is suppressed by SDS and by heating. Furthermore, the extent of cross-linking is independent of OGC concentration indicating that disulfide bridge(s) must be formed between the two subunits of native dimers. The number and localization of disulfide bridge(s) in the cross-linked OGC were examined by peptide fragmentation and subsequent cleavage of disulfide bond(s) by β-mercaptoethanol. Our experimental results show that cross-linking of OGC is accomplished by a single disulfide bond between the cysteines 184 of the two subunits and suggest that these residues in the putative transmembrane helix four are fairly close to the twofold axis of the native dimer structure.  相似文献   

17.
Human fibrinogen is a dimer composed of two identical halves. Each dimeric half contains three peptide chains (alpha, beta, and gamma) linked by disulfide bonds. The two half-molecules are joined by three disulfide bonds, one between the two alpha-chains (residue alpha-28) and two between the two gamma-chains (residues gamma-8 and gamma-9). In the absence of any difinitive experimental evidence, it has been presumed that the joined halves were aligned in a parallel orientation similar to the situation found in immunoglobulins. We have now determined that the two gamma-chains--hence, the dimeric halves--are connected in an antiparallel manner. A tryptic peptide containing gamma-chain residues 6-14 was isolated as a disulfide-linked dimer from CNBr-treated fragment E. Synthetic peptides corresponding to this sequence were prepared, from which parallel and antiparallel dimers were constructed. During the syntheses, cysteine thiol groups were protected as p-methoxybenzyl and acetamidomethyl sulfides; the peptides were dimerized by selective deprotection and disulfide bond formation. First, the p-methoxybenzyl groups were removed by liquid hydrogen fluoride and the newly exposed thiols oxidized in the presence of potassium ferricyanide. Then the monocystine compound was converted to the double-cystine product by iodolytic cleavage of the acetamidomethyl group with concomitant disulfide bond formation. This selectivity was used to prepare peptide dimers which modeled both parallel and antiparallel arrangements. The antiparallel-oriented synthetic peptide was indistinguishable from the native tryptic peptide as judged by elution from reverse-phase high-performance liquid chromatography and circular dichroism spectroscopy. The parallel-oriented synthetic peptide differed from the native material by both criteria.  相似文献   

18.
The investigation of Encelia canescens afforded, in addition to several known compounds, four new dimeric p-hydroxyacetophenone derivatives, two epimeric chromene dimers and two epimeric mixed dimers of euparin and encecalin. Furthermore, derivatives of tremetone and of encecalin were present. The structures were elucidated hy high field 1H NMR spectroscopy.  相似文献   

19.
The secondary structure of xanthan in solutions of relatively low salt concentration and at room temperature has been investigated using static light scattering experiments. Additional evidence has been found for a dimeric structure at 25°C in 0.01M NaCl. From the experimental z-average mean square (ms) radius of gyration, a value for the persistence length p has been estimated, taking explicitly into account the polydispersity of the three samples used, which has been established by gel permeation chromatography (GPC) measurements. The experimental particle scattering functions of the three samples are consistent with theoretical estimates for polydisperse systems with the same value of p = 65 ± 10 nm and the molar mass per unit length for a dimeric structure. This secondary structure remains unaffected by the ionic strength in the 0.005–0.0lM range. Partial aggregation seems to occur at higher NaCl concentrations. Light scattering and GPC data show that heating the xanthan 0.01M NaCl solutions to about 70°C considerably reduces the Mw of the low molar mass sample (2.3 × 105-g·mol?1), contrary to what is observed for the high molar mass sample (1.8 × 106-g·mol?1). These experimental findings can be accounted for by a partial temperature-induced dissociation of the xanthan dimers according to an all-or-none mechanism. © 1994 John Wiley & Sons, Inc.  相似文献   

20.
Covalent single-chain dimers of the chemokine interleukin-8 (IL-8) have been designed to mimic the dimeric form of IL-8 in solution and facilitate the production of heterodimer variants of IL-8. Physical studies indicated that use of a simple peptide linker to join two subunits, while allowing receptor binding and activation, led to self-association of the tethered dimers. However, addition of a single disulfide crosslink between the tethered subunits prevented this multimer from forming, yielding a species of dimer molecular weight. Crosslinked single-chain dimers bind to both IL-8 neutrophil receptors CXCR1 and CXCR2 as well as to DARC, as does a double disulfide-linked dimer with no peptide linker. In addition, neutrophil response to these dimers as measured by chemotaxis or beta-glucuronidase release is similar to that elicited by wild-type IL-8, providing evidence that the dissociation of the dimeric species is not required for these biologically relevant activities. Finally, through construction of single-chain heterodimer mutants, we show that only the first subunit's ELR motif is the single-chain variants.  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