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.
The helix–coil transition and conformational structure of poly(8-bromoadenylic acid) [poly(8BrA)] have been investigated using 1H- and 13C-nmr, CD, and ir spectroscopy. The results have been compared with the structure of the related 5′-mono- and polynucleotides. The chemical shifts of H(2′), H(3′), C(2′), and C(3′) nmr signals show an interesting correlation with both the puckering of ribose ring and glycosidic bond torsion angle. Poly(8BrA) shows an upfield shift of the C(3′) signal and a downfield shift of the H(3′) signal compared to the chemical shifts in poly(A). These shifts are consistent with a C(3′) endo-syn conformation for poly(8BrA). A similar effect has been reported previously and is also observed here on the C(2′) and H(2′) signals when the preferred conformation is C(2′)endo-syn (e.g., in 5′-8BrAMP). The chemical-shift parameters thus act as a probe for studying syn ? anti and N ? S equilibria in solutions. The three-bond 1H-′13C coupling constants between H(1′) and C(8) and C(4) have been measured in poly(8BrA) and 5′-8BrAMP and their structural implications have been discussed. The observed preference of a C(3′)endo-syn conformation for poly(8BrA), coupled with other evidence, throws doubt on the validity of a correlation previously reported whereby a syn conformation is associated with a C(2′)endo ribose pucker. The backbone conformation of randomly coiled poly(8BrA) is very similar to the structures found in polyribonucleotides: poly(A) and poly(U). All three polymers show strong preferences for the backbone angles found in RNA helices. The CD spectrum of poly(8BrA) has a striking relationship to that of poly(A). The signs of all extrema are inverted, and the magnitudes are related by a constant factor. We suggest that these differences result from a change in the angle between coupled transition moment vectors in the two polymers. Infrared spectra of poly(8BrA) in H2O and D2O solution are reported for the frequency range below 1400 cm?1. The antisymmetric >PO stretching vibration is observed at an unusually low frequency in the helix (1214 cm?1). The symmetric >PO stretch occurs at ~1095 cm?1 but is not resolved from a ring vibration near this frequency. A conformationally sensitive band, characteristic of helical RNA structures, is observed at 817 cm?1 and disappears when the helix is melted. This observation confirms the conclusion that ordered poly(8BrA) has a regular helical structure with an RNA backbone conformation. A stereochemical explanation is provided for the failure of poly(8BrA) (or other syn polymers) to form double helices with anti-polyribonucleotides.  相似文献   

2.
Abstract

The interaction between glycated human serum albumin (gHSA) and folic acid (FA) was investigated by various spectroscopic techniques, such as fluorescence, circular dichroism, UV–vis absorption spectroscopy and electrophoretic light scattering technique. These methods characterize the binding properties of an albumin–folic acid system. The binding constants values (Ka) at 300 and 310 K are about 104 M?1. The standard enthalpy change (ΔH) and the standard entropy change (ΔS) were calculated to be ~?20?kJ mol?1 and ~16 J mol?1 K?1, respectively, which indicate characteristic electrostatic interactions between gHSA and folic acid. The CD studies showed that there are no significant conformational changes in the secondary structure of the protein. Moreover, the zeta potential measurements proved that under physiological conditions the gHSA–folic acid complex shows instability. No significant changes in the secondary structure of the protein and reversible drug binding are the desirable effect from pharmacological point of view.

Communicated by Ramaswamy H. Sarma  相似文献   

3.
Complexes of poly(BU) with various adenine derivatives were investigated by circular dichroism (CD) and absorption spectroscopy. A 1:2 stoichiometry was indicated on CD mixing curves for typical complexes of 9-substituted adenine and 2-aminoadenine derivatives with poly(BU). The CD spectrum of adenosine·2poly(BU) is characterized by well-resolved bands in the range of 210–350 nm. Other adenine derivative–poly(BU) complexes also afford similar CD spectra, while 2-aminoadenine derivative–poly(BU) complexes give quite different spectra. Attempts to assign representative CD spectra were made using the transition of helical poly(BU) and the respective purine polynucleotides. The similarity of the CD spectra suggests that poly(A)·2poly(BU) and adenine derivative–poly(BU) complexes are nearly identical in structure except for the ribose–phosphate linkage. The fact that the uv isosbestic point of adenosine·2poly(BU) falls in close proximity to that of the corresponding polymer complex also supports this conclusion. In the formation of stable helices, the ribose moiety is dispensable in the “strand” of purine. The Tm of 9-methyladenine·2poly(BU) is somewhat higher than that of adenosine·2poly(BU) under equivalent conditions. The Tm difference with the monomer–poly(U) system was found to be about 20°C in 0.4M NaCl–0.02M Na–cacodylate–5 × 10?4M EDTA (pH 7.0). Further, it was noted that the monomer–poly(BU) complexes are formed even when the Tm is lower than that of self-folded poly(BU).  相似文献   

4.
Background: In contrast to wild type, interleukin‐10‐deficient (IL‐10?/–) mice are able to clear Helicobacter infection. In this study, we investigated the immune response of IL‐10?/– mice leading to the reduction of Helicobacter infection. Materials and Methods: We characterized the immune responses of Helicobacter felis‐infected IL‐10?/– mice by studying the systemic antibody and cellular responses toward Helicobacter. We investigated the role of CD4+ T cells in the Helicobacter clearance by injecting H. felis‐infected IL‐10?/– mice with anti‐CD4 depleting antibodies. To examine the role of mast cells in Helicobacter clearance, we constructed and infected mast cells and IL‐10 double‐deficient mice. Results: Reduction of Helicobacter infection in IL‐10?/– mice is associated with strong humoral (fivefold higher serum antiurease antibody titers were measured in IL‐10?/– in comparison to wild‐type mice, p < .008) and cellular (urease‐stimulated splenic CD4+ T cells isolated from infected IL‐10?/– mice produce 150‐fold more interferon‐γ in comparison to wild‐type counterparts, p < .008) immune responses directed toward Helicobacter. Depletion of CD4+ cells from Helicobacter‐infected IL‐10?/– mice lead to the loss of bacterial clearance (rapid urease tests are threefold higher in CD4+ depleted IL‐10?/– in comparison to nondepleted IL‐10?/– mice, p < .02). Mast cell IL‐10?/– double‐deficient mice clear H. felis infection, indicating that mast cells are unnecessary for the bacterial eradication in IL‐10?/– mice. Conclusion: Taken together, these results suggest that CD4+ cells are required for Helicobacter clearance in IL‐10?/– mice. This reduction of Helicobacter infection is, however, not dependent on the mast cell population.  相似文献   

5.
In this work, the interaction of chlortetracycline with bovine serum albumin (BSA) was investigated by fluorescence spectroscopy, circular dichroism (CD) spectroscopy, and molecular docking. Results indicated that chlortetracycline quenches BSA fluorescence mainly by a static quenching mechanism. The quenching constants (KSV) were obtained as 5.64 × 104, 4.49 × 104/, and 3.44 × 104/ M?1 at 283, 295, and 307 K, respectively. The thermodynamic parameters of enthalpy change Δ H°, entropy change Δ S°, and free energy change Δ G° were ?5.12 × 104/ J mol?1, ?97.6 J mol?1 K?1, and ?2.24 × 104/ J mol?1 (295 K), respectively. The association constant (KA) and the number of binding sites (n) were 9.41 × 103/ M?1 and 0.86, respectively. The analysis results suggested that the interaction was spontaneous, and van der Waals force and hydrogen‐bonding interactions played key roles in the reaction process. In addition, CD spectra proved secondary structure alteration of BSA in the presence of chlortetracycline. © 2012 Wiley Periodicals, Inc. J Biochem Mol Toxicol 26:331–336, 2012; View this article online at wileyonlinelibrary.com . DOI 10:1002/jbt.21424  相似文献   

6.
Antimicrobial peptides (AMPs) are components of the innate immune system and may be potential alternatives to conventional antibiotics because they exhibit broad‐spectrum antimicrobial activity. The AMP cecropin P1 (CP1), isolated from nematodes found in the stomachs of pigs, is known to exhibit antimicrobial activity against Gram‐negative bacteria. In this study, we investigated the interaction between CP1 and lipopolysaccharide (LPS), which is the main component of the outer membrane of Gram‐negative bacteria, using circular dichroism (CD) and nuclear magnetic resonance (NMR). CD results showed that CP1 formed an α‐helical structure in a solution containing LPS. For NMR experiments, we expressed 15N‐labeled and 13C‐labeled CP1 in bacterial cells and successfully assigned almost all backbone and side‐chain proton resonance peaks of CP1 in water for transferred nuclear Overhauser effect (Tr‐NOE) experiments in LPS. We performed 15N‐edited and 13C‐edited Tr‐NOE spectroscopy for CP1 bound to LPS. Tr‐NOE peaks were observed at the only C‐terminal region of CP1 in LPS. The results of structure calculation indicated that the C‐terminal region (Lys15–Gly29) formed the well‐defined α‐helical structure in LPS. Finally, the docking study revealed that Lys15/Lys16 interacted with phosphate at glucosamine I via an electrostatic interaction and that Ile22/Ile26 was in close proximity with the acyl chain of lipid A. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

7.
The conformation of segments corresponding to the four α-helical stretches found in human granulocyte-macrophage colony-stimulating factor was studied in water solution in the presence of different amounts of 2,2,2-trifluoroethanol (TFE). The CD spectra reveal the onset of secondary structure upon addition of TFE. The final amount of helical conformation varies among the four peptides. In all cases, the conformational transition is complete before 50% TFE (v/v). 1H-NMR studies were conducted at this solvent composition, leading to the assignment of all the resonances and to the definition of the secondary structure for all four fragments. © 1997 European Peptide Society and John Wiley & Sons, Ltd. J. Pep. Sci.3: 336–346 No. of Figures: 12. No. of Tables: 5. No. of References: 25  相似文献   

8.
Alan E. Tonelli 《Biopolymers》1978,17(5):1175-1179
Solution conformations of the cyclic pentapeptide plant-hormone malformin A, whose conformational freedom is constrained by an intramolecular disulfide bridge, are derived and presented here. The nmr and CD data of Ptak are used to place restrictions on the search for possible malformin A solution conformers of low energy. Only two distinct conformers were found to be consistent with Ptak's data. Both structures are characterized by an internally buried (solvent-shielded) D -Cys2 amide proton, a seven-membered (1–3)hydrogen bond between (N–H) and (O?C), and a disulfide bridge conformation with a P chirality as manifested in the nmr study by the temperature independence of the amide proton chemical shifts for the D -Cys2 and D -Leu4 residues and the negative sign of the long wavelength maximum in the CD spectrum, respectively. Inspection of space-filling molecular models of both structures indicates severe steric barriers to their rapid interconversion. Thus, it appears that only one of the two conformers may be present in solution. The difference in their calculated dipole moments (4.6 and 6.9D) suggests an experimental method for distinguishing between the two proposed solution structures.  相似文献   

9.
The ability of lethally irradiated C57BL/6 mice to acutely reject H-2d bone marrow is due to a lymphocyte population that is NK1+, ASGM1+, CD4, CD8, CD3+. Transfer of spleen cells from C57BL/6 mice expressing these antigens into nonresponder 129 mice adoptively transfers the ability to reject H-2d marrow grafts. The specificity of this rejection maps to the H-2D major histocompatibility complex (MHC) region. Transplantation of high doses of H-2d marrow into C57BL/6 overrides the acute rejection mechanism leading to graft survival. During growth of the graft, a cytolytic activity develops that is due to ASGM1+, CD8+ cytolytic T lymphocytes (CTLs) with H-2Ld specificity. The possibility that the ASGM1+, CD8+ CTLs are descendents of the CD3+, NK1+, ASGM1+, CD8 cells responsible for acute rejection is investigated by adoptive cell transfer experiments. We show that beige mice that lack NK1+ cells as well as the ability to acutely reject H-2d marrow fail to generate specific CTLs after transplantation with a high dose of H-2d marrow. Transfer of highly purified NK1+ cells from B6.PL-Ly-2 a /Ly-3 a (Lyt-2.1) into beige mice together with H-2d marrow leads to generation of Lyt-2.1 CTLs from donor NK1+ cells. These results show that specific CTLs are generated from NK1+ cells during acute marrow graft rejection. Offprint requests to: G. Dennert.  相似文献   

10.
Growth factor-dependent gut intraepithelial lymphocyte (IEL) cell lines were established from a long-term in vitro culture of BALB/c IEL with syngeneic irradiated spleen cells in the presence of concanavalin A-stimulated spleen supernatant fluids. The cell lines were preferentially consisted of very limited thymoindependent subsets of IEL; i.e., Thy-1+ CD5TCRαβ+ CD4+CD8 α+β (double-positive; DP) IEL and Thy-1+ CD5 TCRαβ+ CD4CD8α+β (CD8 single-positive; CD8 SP) IEL. The CD8 SP IEL cell line had cytotoxic activities and was triggered to proliferate by T-cell receptor (TCR)-directed stimuli. The DP IEL cell line expressed high levels of the CD3-TCRαβ, exhibited cytotoxic activity in redirected lysis assays, and had perforin in the cytoplasm, indicating the functional maturity of this cell line. However, the DP IEL cell line did not proliferate in response to TCRαβ-directed stimuli, which indicated that TCRαβ-mediated signalling was able to initiate cytotoxic function but not to induce proliferation of the DP IEL cell line. Although both cell lines were shown to have functional competence, they expressed J11d antigen which marks immaturity in thymocyte differentiation pathways. These results indicate that the established thymoindependent DP and CD8 SP IEL cell lines have unique properties distinct from DP thymocytes and CD8 SP peripheral T cells. Together with a recent report on freshly isolated DP IEL (10), the unique properties of the DP IEL cell line seems to support the notion that DP IEL may undergo a unique maturation process in the gut microenvironment.  相似文献   

11.
The N‐terminal 1–34 segments of both parathyroid hormone (PTH) and parathyroid hormone‐related protein (PTHrP) bind and activate the same membrane receptor in spite of major differences in their amino acid sequence. The hypothesis was made that they share the same bioactive conformation when bound to the receptor. A common structural motif in all bioactive fragments of the hormone in water/trifluoroethanol mixtures or in aqueous solution containing detergent micelles is the presence of two helical segments at the N‐ and C‐termini of the sequence. In order to stabilize the helical structures, we have recently synthesized and studied the PTHrP(1–34) analog [(Lys13–As p17, Lys26–As p30)]PTHrP(1–34)NH2, which contains lactam‐constrained Lys‐Asp side chains at positions i, i+4. This very potent agonist exhibits enhanced helix stability with respect to the corresponding linear peptide and also two flexible sites at positions 12 and 19 in 1:1 trifluoroethanol/water. These structural elements have been suggested to play a critical role in bioactivity. In the present work we have extended our conformational studies on the bicyclic lactam‐constrained analog to aqueous solution. By CD, 2D‐NMR and structure calculations we have shown that in water two helical segments are present in the region of the lactam bridges (13–18, and 26–31) with high flexibility around Gly12 and Arg19. Thus, the essential structural features observed in the aqueous‐organic medium are maintained in water even if, in this solvent, the overall structure is more flexible. Our findings confirm the stabilizing effect of side‐chain lactam constraints on the α‐helical structure. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
The high-resolution (300 MHz) proton nmr spectrum of E. coli tRNAfMet has been examined in 0.17M NaCl, with and without Mg2+, and at various temperatures. In light of recent studies of other E. coli tRNA and fragments of tRNAfMet, some low field (11–15 ppm) resonances previously assigned to secondary structure base pairs are reassigned to a tertiary structure A14–S4U8 base pair and a protected uridine residue in the anticodon loop. These two resonances and other low field resonances which are assigned to secondary structure base pairs are used to monitor the thermal unfolding of the molecule. In the absence of Mg2+ the tertiary structure base pair is present only to ~45°C, but in the presence of Mg2+ it remains until at least 70°C. Analysis of the temperature dependence of other low field resonances indicates that the melting of the dihydrouridine stem occurs more or less simultaneously with the loss of tertiary structure. The observation of the resonance from the A14–S4U8 base pair proves that tertiary structure is present in this molecule below 40°C, even in the absence of Mg2+.  相似文献   

13.
Human plasma protein S is a nonenzymatic cofactor for activated protein C (APC) in the inactivation of coagulation factors Va and VIIIa, and helps to provide an essential negative feedback on blood coagulation. Previous indirect evidence suggested that the thrombin-sensitive region (TSR:residues 47–75, 1 disulfide) and the first epidermal growth factorlike region (EGF1: residues 76–116, 3 disulfides) of protein S may be functionally important for expression of its APC cofactor activity. To study the functional importance of these modules directly, access to the isolated TSR and EGF1 modules would be preferred. Recombinant expression of protein S intact TSR and correctly folded EGF1 has not been possible. Here we describe the synthesis of both TSR and EGF1 modules by stepwise solid phase peptide synthesis using the in situ neutralization/2-(1H-benzotriazol-1-yl)-1,1,3,3-tetramethyluronium hexafluorophosphate activation procedure for tert-butoxycarbonyl chemistry. For the TSR, correct intramodular disulfide bonding was confirmed. To overcome folding difficulties with the EGF1, a two-step oxidation procedure was used in which the cysteines involved in the middle, crossing, disulfide bond (Cys85-Cys102) remained protected with acetamidomethyl (Acm) groups after hydrogen fluoride treatment of the peptide resin. Selective formation of the first two disulfide bonds (Cys80-Cys93 and Cys104-Cys113) was followed by release of the Acm groups and subsequent formation of the third disulfide bond (Cys85-Cys102). CD studies revealed 54% of β-sheet/turn in the EGF1 that is characteristic for EGF modules. Deuterium exchange studies suggested a very tightly packed core in EGF1 that is not accessible to the bulk solvent, likely a result from the compact structure caused by its three disulfide bonds. The 30% β-sheet structure observed in the TSR involved amide protons that could be readily exchanged by deuterons, likely reflecting a more flexible structure of the TSR loop in contrast to the rigid structure of EGF1. The establishment of synthetic access to the TSR and EGF1 of protein S provides a versatile tool to study interactions of these modules with the blood coagulation components of the anticoagulant plasma protein C pathway. © 1998 John Wiley & Sons, Inc. Biopoly 46: 53–63, 1998  相似文献   

14.
The three‐dimensional solution structure of harzianin HC IX, a peptaibol antibiotic isolated from the fungus Trichoderma harzianum, was determined using CD, homonuclear, and heteronuclear two‐dimensional nmr spectroscopy combined with molecular modeling. This 14‐residue peptide, Ac Aib1 Asn2 Leu3 Aib4 Pro5 Ala6 Ile7 Aib8 Pro9 Iva10 Leu11 Aib12 Pro13 Leuol14 (Aib, α‐aminoisobutyric acid; Iva, isovaline; Leuol, leucinol), is a main representative of a short‐sequence peptaibol class characterized by an acetylated N‐terminus, a C‐terminal amino alcohol, and the presence of three Aib‐L ‐Pro motifs at positions 4–5, 8–9, and 12–13, separated by two dipeptide units. In spite of a lower number of residues, compared to the 18/20‐residue peptaibols such as alamethicin, harzianin HC IX exhibits remarkable membrane‐perturbing properties. It interacts with phospholipid bilayers, increasing their permeability and forming voltage‐gated ion channels through a mechanism slightly differing from that proposed for alamethicin. Sequence‐specific 1H‐ and 13C‐nmr assignments and conformational nmr parameters (3JNHCαH coupling constants, quantitative nuclear Overhauser enhancement data, temperature coefficients of amide and carbonyl groups, NH–ND exchange rates) were obtained in methanol solution. Sixty structures were calculated based on 98 interproton distance restraints and 6 Φ dihedral angle restraints, using high temperature restrained molecular dynamics and energy minimization. Thirty‐seven out of the sixty generated structures were consistent with the nmr data and were convergent. The peptide backbone consists in a ribbon of overlapping β‐turns twisted into a continuous spiral from Asn2 to Leuol14 and forming a 26 Å long helix‐like structure. This structure is slightly amphipathic, with the three Aib–Pro motifs aligned on the less hydrophobic face of the spiral where the Asn2 side chain is also present, while the more hydrophobic bulky side chains of leucines, isoleucine, isovaline, and leucinol are located on the concave side. The repetitive (Xaa–Yaa–Aib–Pro) tetrapeptide subunit, making up the peptide sequence, is characterized by four sets of (Φ,Ψ) torsional angles, with the following mean values: Φi = −90°, Ψi = −27°; Φi+1 = −98°, Ψi+1 = −17°; Φi+2 = −49°, Ψi+2 = −50°; Φi+3 = −78°, Ψi+3 = +3°. We term this particular structure, specifically occurring in the case of (Xaa–Yaa–Aib–Pro)n sequences, the (Xaa–Yaa–Aib–Pro)‐β‐bend ribbon spiral. It is stabilized by 4 → 1 intramolecular hydrogen bonds and differs from both the canonical 310‐helix made of a succession of type III β‐turns and from the β‐bend ribbon spiral that has been described in the case of (Aib–Pro)n peptide segments. © 1999 John Wiley & Sons, Inc. Biopoly 50: 71–85, 1999  相似文献   

15.
Z-Dehydrophenylalanine (ΔzPhe) possessing four oligopeptides, Boc-(L -Ala-ΔzPhe-Aib)n-OCH3 (n = 1–4: Boc, t-butoxycarbonyl; Aib, α-aminoisobutyric acid), were synthesized, and their solution conformations were investigated by 1H-nmr, ir, uv, and CD spectroscopy and theoretical CD calculation. 1H-nmr (the solvent accessibility of NH groups) and ir studies indicated that all the NH groups except for those belonging to the N-terminal L -Ala-ΔzPhe moiety participate in intramolecular hydrogen bonding in chloroform. This suggests that the peptides n = 2–4 have a 4 → 1 hydrogen-bonding pattern characteristic of 310-helical structures. The uv spectra of all these peptides recorded in chloroform and in trimethyl phosphate showed an intense maximum around 276 nm assigned to the ΔzPhe chromophores. The corresponding CD spectra of the peptides n = 2–4 showed exciton couplets with a negative peak at longer wavelengths, whereas that of the peptide n = 1 showed only weak signals. Theoretical CD spectra were calculated for the peptides n = 2–4 of several helical conformations, on the basis of exciton chirality method. This calculation indicated that the three peptides form a helical conformation deviating from the perfect 310-helix that contains three residues per turn, and that their side chains of Δz Phe residues are arranged regularly along the helix. The center-to-center distance between the nearest phenyl pair(s) was estimated to be ~ 5.5 Å. The chemical shifts of the ΔzPhe side-chain protons (Hβ and aromatic H) for the peptides n = 2–4 indicated anisotropic shielding effect of neighboring phenyl group(s); the effect also supports a regular arrangement of the Δz Phe side chains along the helical axis. © 1993 John Wiley & Sons, Inc.  相似文献   

16.
《Chirality》2017,29(8):451-457
A major challenge in pharmaceuticals for clinical applications is to alter the solubility, stability, and toxicity of drug molecules in living systems. Cyclodextrins (CDs) have the ability to form host–guest inclusion complexes with pharmaceuticals for further development of new drug formulations. The inclusion complex of clomiphene citrate (CL), a poorly water‐soluble drug, with native β‐cyclodextrin (β‐CD) was characterized by a one and two‐dimensional nuclear magnetic resonance (NMR) spectroscopic approach and also by molecular docking techniques. Here we report NMR and a computational approach in preferential isomeric selection of CL, which exists in two stereochemical isomers, enclomiphene citrate (ENC; E isomer) and zuclomiphene citrate (ZNC; Z isomer) with β‐CD. β‐CD cavity protons, namely, H‐3′ and H‐5′, experienced shielding in the presence of CL. The aromatic ring protons of the CL molecule were observed to be deshielded in the presence of β‐CD. The stoichiometric ratio of the β‐CD:CL inclusion complex was observed by NMR and found to be 1:1. The overall binding constant of β‐CD:CL inclusion complexes was based on NMR chemical shifts and was calculated to be 50.21 M−1. The change in Gibb's free energy (∆G) was calculated to be −9.80 KJ mol−1. The orientation and structure of the β‐CD:CL inclusion complexes are proposed on the basis of NMR and molecular docking studies. 2D 1H‐1H ROESY confirmed the involvement of all three aromatic rings of CL in the inclusion complexation with β‐CD in the solution, confirming the multiple equilibria between β‐CD and CL. Molecular docking and 2D 1H‐1H ROESY provide insight into the inclusion complexation of two isomers of CL into the β‐CD cavity. A molecular docking technique further provided the different binding affinities of the E and Z isomers of CL with β‐CD and confirmed the preference of the Z isomer binding for β‐CD:CL inclusion complexes. The study indicates that the formation of a hydrogen bond between –O– of CL and the hydrogen atom of the hydroxyl group of β‐CD was the main factor for noncovalent β‐CD:CL inclusion complex formation and stabilization in the aqueous phase.  相似文献   

17.
T‐cell population consists of two major subsets, CD4+ T cells and CD8+ T cells, which can be distinguished by the expression of CD4 or CD8 molecules, respectively. Although they play quite different roles in the immune system, many of their basic cellular processes such as proliferation following stimulation are presumably common. In this study, we have carefully analyzed time–course of G0/1 transition as well as cell cycle progression in the two subsets of quiescent T‐cell population following in vitro growth stimulation. We found that CD8+ T cells promote G0/1 transition more rapidly and drive their cell cycle progression faster compared to CD4+ T cells. In addition, expression of CD25 and effects of its blockade revealed that IL‐2 is implicated in the rapid progression, but not the earlier G0/1 transition, of CD8+ T cells.  相似文献   

18.
Purpose: Sulfur mustard (SM) lung is a heterogeneous disease associated with abnormal inflammatory immune responses. The Th17/Treg axis imbalance is associated with the pathogenesis of chronic inflammatory pulmonary disease. We aimed to determine the distribution of different Th17 and Treg cells in patients with SM lung and chronic obstructive pulmonary disease (COPD) and evaluate the clinical implications in this homeostasis. Methods: In this analytical cross-sectional study, CD4?+?Foxp3+?Treg and CD4+?IL-17+?Th17 cells were measured in peripheral blood mononuclear cells (PBMCs) and transbronchial biopsy (TBB) samples of 15 SM-exposed patients, 12 COPD and 13 healthy controls (HCs). The potential correlation between the ratio of Th17/Tregs and lung function was evaluated with multivariate logistic regression (MLR) analysis. Results: The frequency of CD4?+?FoxP3+?Tregs and CD4?+?IL-17+?Th17 was increased ~1.7-fold (8.71/4.95) and ~2.7-fold (1.028/0.371) respectively, in the PBMC of SM patients compared with the health controls (p?<?0.001). The results indicated that there were increases in the frequency of Th17 and Tregs cells in the patients with COPD versus the HC, that is, ~2.6-fold (0.987/0.371) and ~1.4-fold (7.12/4.95), respectively; but they did not reach to SM level (p?≥?0.05). Moreover, in the TBB samples, the CD4?+?IL-17+?Th17 and CD4+?FoxP3+?Tregs numbers were significantly higher in SM and COPD patients than HC (p?<?0.05). The Th17 and Treg cells were inversely correlated with forced expiratory volume in 1s (FEV1%) (r?=??0.351, p =?0.001; r?=??0.344, p?=?0.021) and FEV1/FVC (r?=??0.44, p?=?0.001; r?=??0.302, p?=?0.011), respectively. Instead, positive correlations were found between Treg/Th17 ratios and forced FEV1%pred (r?=?0.156, p?=?0.007), as well as FEV1/FVC ratio (r?=?0.334, p?=?0.006). Conclusions: The imbalance of Th17/Treg has a key role in immunopathogenesis of chronic phase of mustard lung disease.  相似文献   

19.
Soybean calmodulin isoform 4 (sCaM4) is a plant calcium‐binding protein, regulating cellular responses to the second messenger Ca2+. We have found that the metal ion free (apo‐) form of sCaM4 possesses a half unfolded structure, with the N‐terminal domain unfolded and the C‐terminal domain folded. This result was unexpected as the apo‐forms of both soybean calmodulin isoform 1 (sCaM1) and mammalian CaM (mCaM) are fully folded. Because of the fact that free Mg2+ ions are always present at high concentrations in cells (0.5–2 mM), we suggest that Mg2+ should be bound to sCaM4 in nonactivated cells. CD studies revealed that in the presence of Mg2+ the initially unfolded N‐terminal domain of sCaM4 folds into an α‐helix‐rich structure, similar to the Ca2+ form. We have used the NMR backbone residual dipolar coupling restraints 1DNH, 1DCαHα, and 1DC′Cα to determine the solution structure of the N‐terminal domain of Mg2+‐sCaM4 (Mg2+‐sCaM4‐NT). Compared with the known structure of Ca2+‐sCaM4, the structure of the Mg2+‐sCaM4‐NT does not fully open the hydrophobic pocket, which was further confirmed by the use of the fluorescent probe ANS. Tryptophan fluorescence experiments were used to study the interactions between Mg2+‐sCaM4 and CaM‐binding peptides derived from smooth muscle myosin light chain kinase and plant glutamate decarboxylase. These results suggest that Mg2+‐sCaM4 does not bind to Ca2+‐CaM target peptides and therefore is functionally similar to apo‐mCaM. The Mg2+‐ and apo‐structures of the sCaM4‐NT provide unique insights into the structure and function of some plant calmodulins in resting cells.  相似文献   

20.
The host–guest interaction between sulfobutylether‐β‐cyclodextrin (SBE‐β‐CD) and reserpine (RSP) is described using flow injection‐chemiluminescence (FI‐CL) and site‐directed molecular docking methods. It was found that RSP could inhibit the CL intensity produced by a luminol/SBE‐β‐CD system. The decrease in CL intensity was logarithmic over an RSP concentration range of 0.03 to 700.0 nM, giving a regression equation of ?I = 107.1lgCRES + 186.1 with a detection limit of 10 pM (3σ). The CL assay was successfully applied in the determination of RSP in injection, saliva and urine samples with recoveries in the range 93.5–106.1%. Using the proposed CL model, the binding constant (KCD‐R) and the stoichiometric ratio of SBE‐β‐CD/RSP were calculated to be 7.4 × 106 M‐1 and 1 : 1, respectively. Using molecular docking, it was confirmed that luminol binds to the small cavity of SBE‐β‐CD with a nonpolar interaction, while RSP targeted the larger cavity of SBE‐β‐CD and formed a 1 : 1 complex with hydrogen bonds. The proposed new CL method has the potential to become a powerful tool for revealing the host–guest interaction between CDs and drugs, as well as monitoring drugs with high sensitivity. Copyright © 2013 John Wiley & Sons, Ltd.  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