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.
Chellgren BW  Creamer TP 《Biochemistry》2004,43(19):5864-5869
The left-handed polyproline II (P(II)) helix is a structure that has been given a great deal of attention lately because of its role in a wide variety of physiologically important processes and potential significance in protein unfolded states. Recent work by several authors has shown that residues besides proline can adopt this structure. A scale of relative P(II)-helix-forming propensities has been generated but only for single guest residues in a proline-based host system. Here, we present multiple guest residues in a proline-based host system. Using circular dichroism spectroscopy, we have shown that not only single residues, but also short sequences of non-proline residues can adopt the P(II) conformation.  相似文献   

2.
There has been growing interest in polyproline type II (PPII) helices since PPII helices have been found in folded and unfolded proteins and involved in a variety of biological activities. Polyproline can also form type I helices (PPI) which are very different from PPII conformation and only exist in certain organic solvents. Recent studies have shown that stereoelectronic effects play a critical role in stabilizing a PPI or PPII helix. Here, we have synthesized a series of host–guest peptides with an electron‐withdrawing substituent at the 4R or 4S position of proline and used a kinetic approach to further explore stereoelectronic effects on the transition barrier of the interconversion between PPI and PPII conformations. Time‐dependent circular dichroism measurements revealed that the rates of PPII → PPI conversion were reduced upon incorporating the hydroxyl‐, fluoro‐, and methoxy‐groups at the 4R position while the rates would be increased if these substituents were at the 4S position. We quantified the changes in transition free energy by comparing their rate constants. (4R,2S)‐4‐Fluoroproline and (4S,2S)‐4‐fluoroproline have the largest effect on the transition energy barrier for PPII → PPI conversion. Our results provide important insights into the role of stereoelectronic effects on the PPII → PPI transition state barrier, which has not been reported in past thermodynamic studies.  相似文献   

3.
The effects of N-terminal amino acid stereochemistry on prolyl amide geometry and peptide turn conformation were investigated by coupling both L- and D-amino acids to (2S, 5R)-5-tert-butylproline and L-proline to generate, respectively, N-(acetyl)dipeptide N'-methylamides 1 and 2. Prolyl amide cis- and trans-isomers were, respectively, favored for peptides 1 and 2 as observed by proton NMR spectroscopy in water, DMSO and chloroform. The influence of solvent composition on amide proton chemical shift indicated an intramolecular hydrogen bond between the N'-methylamide proton and the acetamide carbonyl for the major conformer of dipeptides (S)-1, that became less favorable in (R)-1 and 2. The coupling constant (3J(NH,alpha)) values for the cis-isomer of (R)-1 indicated a phi2 dihedral angle value characteristic of a type VIb beta-turn conformation in solution. X-ray crystallographic analysis of N-acetyl-D-leucyl-5-tert-butylproline N'-methylamide (R)-lb showed the prolyl residue in a type VIb beta-turn geometry possessing an amide cis-isomer and psi3-dihedral angle having a value of 157 degrees, which precluded an intramolecular hydrogen bond. Intermolecular hydrogen bonding between the leucyl residues of two turn structures within the unit cell positioned the N-terminal residue in a geometry where their phi2 and psi2 dihedral angle values were not characteristic of an ideal type VIb turn. The circular dichroism spectra of tert-butylprolyl peptides (S)- and (R)-1b were found not to be influenced by changes in solvent composition from water to acetonitrile. The type B spectrum exhibited by (S)-1b has been previously assigned to a type VIa beta-turn conformation [Halab L, Lubell WD. J. Org. Chem. 1999; 64: 3312-3321]. The type C spectrum exhibited by the (R)-lb has previously been associated with type II' beta-turn and alpha-helical conformations in solution and appears now to be also characteristic for a type VIb geometry.  相似文献   

4.
Polyproline II (PPII) is reported to be a dominant conformation in the unfolded state of peptides, even when no prolines are present in the sequence. Here we use isothermal titration calorimetry (ITC) to investigate the PPII bias in the unfolded state by studying the binding of the SH3 domain of SEM-5 to variants of its putative PPII peptide ligand, Sos. The experimental system is unique in that it provides direct access to the conformational entropy change of the substituted amino acids. Results indicate that the denatured ensemble can be characterized by at least two thermodynamically distinct states, the PPII conformation and an unfolded state conforming to the previously held idea of the denatured state as a random collection of conformations determined largely by hard-sphere collision. The probability of the PPII conformation in the denatured states for Ala and Gly were found to be significant, approximately 30% and approximately 10%, respectively, resulting in a dramatic reduction in the conformational entropy of folding.  相似文献   

5.
6.
A library of peptides and glycopeptides containing (4R)-hydroxy-l-proline (Hyp) residues were designed with a view to providing stable polyproline II (PPII) helical molecules with antifreeze activity. A library of dodecapeptides containing contiguous Hyp residues or an Ala-Hyp-Ala tripeptide repeat sequence were synthesized with and without α-O-linked N-acetylgalactosamine and α-O-linked galactose-β-(1→3)-N-acetylgalactosamine appended to the peptide backbone. All (glyco)peptides possessed PPII helical secondary structure with some showing significant thermal stability. The majority of the (glyco)peptides did not exhibit thermal hysteresis (TH) activity and were not capable of modifying the morphology of ice crystals. However, an unglycosylated Ala-Hyp-Ala repeat peptide did show significant TH and ice crystal re-shaping activity suggesting that it was capable of binding to the surface of ice. All (glyco)peptides synthesized displayed some ice recrystallization inhibition (IRI) activity with unglycosylated peptides containing the Ala-Hyp-Ala motif exhibiting the most potent inhibitory activity. Interestingly, although glycosylation is critical to the activity of native antifreeze glycoproteins (AFGPs) that possess an Ala-Thr-Ala tripeptide repeat, this same structural modification is detrimental to the antifreeze activity of the Ala-Hyp-Ala repeat peptides studied here.  相似文献   

7.
A diamide, N-acetyl-L -proline-N,N-dimethylamide (AcProDMA), in water solution has optical rotatory dispersion (ORD) and circular dichroism (CD) spectra very similar to those of poly-L -proline II and the fibrous protein collagen. In contrast, AcProDMA in cyclohexane solution has optical activity resembling that of poly-L -proline I. Conformational analysis shows that AcProDMA is confined by steric constraints to either of two narrow regions of conformational space. The trans isomer of AcProDMA assumes conformations near those of polyproline II and collagen nearest neighbors, while cis-AcProDMA assumes conformations near that of polyproline I nearest neighbors. Nuclear magnetic resonance (NMR) experiments show that an equilibrium mixture of the cis and trans isomers of AcProDMA is present in solution. The trans isomer predominates in aqueous solution, but the equilibrium shifts to favor the cis isomer in nonpolar organic solvents such as cyclohexane. Analysis of the ORD spectra in terms of two basic spectra reveals a solvent dependent isomerization which parallels that observed by NMR. The optical activity of the pure isomers of AcProDMA can be derived from the ORD, CD and NMR data. A comparison of component cotton effects confirms the similarity in optical activity of trans-AcProDMA, polyproline II, and collagen on the one hand, and of cis-AcProDMA and Polyproline I on the other.  相似文献   

8.
Six [Pen(6)]oxytocin analogs were synthesized by substituting penicillamine for cysteine in oxytocin, [Mpa(1)]oxytocin, [dPen(1)]oxytocin, [5-t-BuPro(7)]oxytocin, [Mpa(1), 5-t-BuPro(7)]oxytocin and [dPen(1), 5-t-BuPro(7)]oxytocin. When tested in the uterotonic test in vitro [Pen(6)]oxytocin, [Pen(6), 5-t-BuPro(7)]oxytocin, [Mpa(1), Pen(6)]oxytocin and [Mpa(1), Pen(6), 5-t-BuPro(7)]oxytocin, all were found to possess both agonistic and antagonistic properties. Their agonistic potency ranged from negligible (0.08 IU/mg) to low (5.85 IU/mg) and their antagonistic potency (pA2) was estimated to range from 6.6 to 7.9. [dPen(1), Pen(6)]Oxytocin and [dPen(1), Pen(6), 5-t-BuPro(7)]oxytocin were found to be pure antagonists with similarly high pA2 values of approximately 8.2. Replacement of proline by 5-tert-butylproline increased binding affinity by a factor of two in [Pen(6)]oxytocin and had no influence on the binding affinity of [Mpa(1), Pen(6)]oxytocin and [dPen(1), Pen(6)]oxytocin. Assignment of the proton signals for prolyl amide cis- and trans-isomers by NMR experiments in water indicated that the Pen(6)-5-tert-BuPro(7) peptide bond cis-isomer population was augmented relative to the prolyl peptides and measured, respectively, at 20, 35 and 35% in the 5-tert-butylproline(7) analogs of [Pen(6)]oxytocin, [Mpa(1), Pen(6)]oxytocin and [dPen(1), Pen(6)]oxytocin. This augmentation in cis-isomer population was correlated with a 21-fold reduction in the agonistic potency and 2-fold augmentation in antagonistic potency for [Pen(6), 5-t-BuPro(7)]oxytocin relative to [Pen(6)]oxytocin. Augmentation of cis-isomer population was also correlated to reduced agonist potency without effect on antagonism on conversion of [Mpa(1), Pen(6)]oxytocin to [Mpa(1), Pen(6), 5-t-BuPro(7)]oxytocin. In the potent oxytocin antagonist, [dPen(1), Pen(6)]oxytocin, substitution of 5-tert-butylproline for proline augmented the cis-isomer population without affecting antagonistic potency. The synthesis and evaluation of [Pen(6)]oxytocin and [Pen(6), 5-t-BuPro(7)]oxytocin analogs 1-6 indicated that steric interactions influenced agonist and antagonist activity by modifying peptide conformation. Augmentations in the prolyl cis-isomer population caused by 5-tert-butylproline occurred concurrently with enhanced or maintained antagonistic potency and binding affinity and reduced agonistic potency.  相似文献   

9.
It has been widely accepted that cAMP activates the protein kinase A (PKA) holoenzyme by dissociating the regulatory and catalytic subunits, thus freeing the catalytic subunit to phosphorylate its targets. However, recent experiments suggest that cAMP does not fully dissociate the holoenzyme. Here, we investigate this mechanism further by using small-angle X-ray scattering to study, at physiological enzyme concentrations, the type Ialpha and type IIbeta holoenzyme structures under equilibrium solution conditions without any labeling of the protein subunits. We observe that while the addition of a molar excess of cAMP to the type Ialpha PKA holoenzyme causes partial dissociation, it is only upon addition of a PKA peptide substrate together with cAMP that full dissociation occurs. Similarly, addition of excess cAMP to the type IIbeta holoenzyme causes only a partial dissociation. However, while the addition of peptide substrate as well as excess cAMP causes somewhat more dissociation, a significant percentage of intact type IIbeta holoenzyme remains. These results confirm that both the type Ialpha and the type IIbeta holoenzymes are more stable in the presence of cAMP than previously thought. They also demonstrate that substrate plays a differential role in the activation of type I versus type II holoenzymes, which could explain some important functional differences between PKA isoforms. On the basis of these data and other recently published data, we propose a structural model of type I holoenzyme activation by cAMP.  相似文献   

10.
The 5'-deiodination of thyroxine is catalyzed by two enzymes which differ in their tissue distribution, substrate specificities, sensitivity to the inhibitor, propylthiouracil, and response to thyroid status. By using the affinity label, N-bromoacetyl-L-thyroxine, both isoenzymes have been found to have substrate binding subunits of approximately 27 kDa. In this study, we compared the substrate binding subunits and hydrodynamic properties of the type I and the type II isozymes using the affinity label, N-bromoacetyl-L-thyroxine, to identify the enzymes. High resolution sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed that the substrate binding subunit of the type I enzyme had an Mr of 27,000, while that of the type II enzyme had a slightly higher Mr of 29,000. This difference was not accounted for by glycosylation. Partial staphylococcal V8-protease digests of the substrate binding subunit of the type I enzyme yielded fragments of 14.6, 13.7, and 7.0 kDa, while V8-protease digests of the substrate binding subunit for the type II enzyme produced fragments of 28.0, 25.1, 19.0, 9.5, 7.2, and 5.8 kDa. Unique cyanogen bromide fragmentation patterns were also observed for the two substrate binding subunits. Sedimentation coefficients of the detergent-soluble type I and type II holoenzymes were 3.67 and 5.22 S, respectively, as determined by sucrose density centrifugation. The type I enzyme behaved as a globular protein, whereas the type II enzyme showed sedimentation properties typical of asymmetric integral membrane proteins. The Stokes radii were 3.78 and 4.97 nm, respectively. From these data, the calculated Mr for detergent-solubilized type I 5'-iodothyronine deiodinase was 55,400 and for the type II enzyme was 198,700. These data indicate that the two isozymes of iodothyronine 5'-deiodinase are multimeric, differ in holoenzyme size and subunit composition, and that their substrate binding subunits are distinct.  相似文献   

11.
12.
The relative effects of treatment with an anticonvulsant, phenytoin, on the production of interferons were determined for both the murine and human systems. Phenytoin treatment was found to have differential effects on the in vitro production of Type I and Type II interferons. Phenytoin had either no effect (HuIFN-alpha) or an enhancing effect (MuIFN-alpha/beta) on the in vitro production of Type I interferons. In contrast, phenytoin pretreatment had an inhibitory effect on the in vitro production of Type II interferons (IFN-gamma) for both the murine and human systems. Phenytoin appeared to exert its inhibitory effect directly on the IFN-gamma-producing cell and was active even when added as late as 6 h after IFN-gamma induction. This inhibition was not related to a toxic effect of the phenytoin and occurred at phenytoin concentrations which were pharmacologically relevant (10-20 micrograms/ml). The effects of phenytoin on the in vivo production of MuIFN-gamma were also examined. In parallel to the in vitro observations, phenytoin treatment of mice significantly reduced the in vivo induction of MuIFN-gamma. The results raise the possibility that phenytoin therapy in humans may significantly affect the production of HuIFN-gamma.  相似文献   

13.
Using data from the National Longitudinal Survey of Youth 1979, the current study examines the effect of type I and type II diabetes on employment status and wages. The results suggest that both the probability of employment and wages are negatively related to the number of years since the initial diagnosis of diabetes. Moreover, the effect of diabetes duration on the probability of employment appears to be nonlinear, peaking around 16 years for females and 10 years for males. A similar negative effect on wages is found only in male diabetics. Finally, the results suggest that failure to distinguish between type I and type II diabetics may lead to some counterintuitive results.  相似文献   

14.
The mechanistic basis underlying the striking cooperativity observed for the assembly of TGF-β family ligand/receptor complexes is not well understood. We report here an investigation in which we used a novel ligand sequestration assay, in combination with immunofluorescent light microscopy and flow cytometry analyses, to examine and quantify cooperative assembly of TGF-β ligand/receptor complexes on the cell surface, as well as ligand/receptor complex internalization. We analyzed the roles played by the ecto/transmembrane (ecto/TM) domains and endodomains of RI and RII TGF-β receptors in these processes by transfecting 293 or HeLa cells with different combinations of receptor mutants. We found that the ecto/TM domains of RII and RI cooperated together to promote the formation of cell surface receptor/ligand complexes. Furthermore, in agreement with the recently determined structure of the TGF-β3/RII ectodomain/RI ectodomain complex [J. Groppe, C.S. Hinck, P. Samavarchi-Tehrani, C. Zubieta, J.P. Schuermann, A.B. Taylor, P.M. Schwarz, J.L. Wrana, A.P. Hinck, Cooperative assembly of TGF-beta superfamily signaling complexes is mediated by two disparate mechanisms and distinct modes of receptor binding, Mol. Cell 29 (2008) 157–168], we observed that the N-terminus of the RII ectodomain was required for full assembly. With respect to endodomains, we found that the RI endodomain enhanced cooperative complex assembly at the cell surface, whereas both the RI and RII endodomains enhanced internalization. Finally, we observed that ligand/receptor internalization, but not complex assembly at the cell surface, was partly raft-dependent. In light of these results, currently proposed mechanisms of cooperative ligand/receptor assembly are discussed.  相似文献   

15.
Absorption and CD spectra of the pi-pi* transition near 200 nm are calculated for helical (Pro)10 forms I and II with a variable proline ring conformation characterized by torsion angle chi 2 in the range -60 degrees to 60 degrees. The spectra for poly(Pro) I are not sufficiently sensitive to chi 2 to suggest a preferred ring conformation. The spectra for poly(Pro) II are more sensitive to chi 2, and suggest preferred ring conformations near either or both of the chi 2 regions -50 +/- 10 degrees and 50 +/- 10 degrees.  相似文献   

16.
17.
The left-handed polyproline II helical structure (P(II)) is observed to be a dominant conformation in the disordered states of protein and small polypeptide chains, even when no prolines are present in the sequence. Recently, in work by Ferreon and Hilser, the energetics associated with Ala and Gly substitutions at a surface exposed proline site were determined calorimetrically by measuring the binding energetics of Sos peptide variants to the C-terminal Src Homology 3 domain of SEM-5. The results were interpreted as a significant conformational bias toward the bound conformation (i.e., P(II)), even when the ligand is unbound. That study was not able to determine, however, whether the conformational bias of the peptides could be explained in terms other than that of a P(II) preference. Here, we test, using a computer algorithm based on the hard sphere collision (HSC) model, the notion of whether a bias in the unbound states of the peptide ligands is specific for the P(II) conformation, or if a bias to any other region of (phi, psi) space can also result in the same observed binding energetics. The results of these computer simulations indicate that, of the regions of (phi, psi) modeled for bias in the small peptides, only the bias to the P(II) conformation, and at rates of bias similar to the experimentally observed rates, quantitatively reproduced the experimental binding energetics.  相似文献   

18.
Adams JJ  Webb BA  Spencer HL  Smith SP 《Biochemistry》2005,44(6):2173-2182
The assembly of a functional cellulose-degrading complex termed the cellulosome involves two specific calcium-dependent cohesin-dockerin interactions: type I and type II. Extensive structural and mutagenesis studies have been performed on the type I modules and their interaction in an attempt to identify the underlying molecular determinants responsible for this specificity. However, very little structural information exists for the type II interaction. We have performed a variety of biophysical studies on the type II dockerin-X-module modular pair (DocX), which comprises the C-terminal region of cellulosomal scaffoldin subunit from Clostridium thermocellum, to determine the effect of calcium on its structure and interaction with type II cohesin. Our results indicate that calcium binding to type II dockerin occurs with an apparent dissociation constant (K(d)) of 7 microM, induces stable secondary and tertiary structure, and leads to the exposure of a hydrophobic surface. Calcium binding also results in the homodimerization of DocX. Analytical ultracentrifugation experiments indicate that the DocX homodimer has an elongated shape and a K(d) of approximately 40 microM. However, addition of the SdbA type II cohesin binding partner led to the dissociation of the DocX homodimer and to the formation of a 1:1 heterodimer. We propose that the exposed hydrophobic surface forms, at least in part, the type II cohesin-binding site, which in the absence of cohesin results in the dimerization of DocX.  相似文献   

19.
A eucaryotic cell assay was established to identify novel, dual and selective inhibitors of human 5alpha-reductase. For this purpose the cDNAs encoding 5alpha-reductase type I and type II were inserted into a pRcCMV vector and expressed in human embryonic kidney (HEK293) cells. Single cell clones with substantially high enzymatic activity were selected and established as permanent cell lines. KM values were determined for both isozymes. The inhibitory potency of several steroidal and non-steroidal compounds synthesized in our group, as well as finasteride and 4MA as controls, were tested by measuring the conversion of [3H]androstenedione. Reaction products were quantified by a HPLC reversed phase technique. Using the new cell assays, selective as well as novel dual 5alpha-reductase inhibitors with IC50 values between 1.0 and 2.5 microM were identified.  相似文献   

20.
Prenylation primarily by geranylgeranylation is required for membrane attachment and function of type I Rho of Plants (ROPs) and Gγ proteins, while type II ROPs are attached to the plasma membrane by S-acylation. Yet, it is not known how prenylation affects ROP membrane interaction dynamics and what are the functional redundancy and specificity of type I and type II ROPs. Here, we have used the expression of ROPs in mammalian cells together with geranylgeranylation and CaaX prenylation-deficient mutants to answer these questions. Our results show that the mechanism of type II ROP S-acylation and membrane attachment is unique to plants and likely responsible for the viability of plants in the absence of CaaX prenylation activity. The prenylation of ROPs determines their steady-state distribution between the plasma membrane and the cytosol but has little effect on membrane interaction dynamics. In addition, the prenyl group type has only minor effects on ROP function. Phenotypic analysis of the CaaX prenylation-deficient pluripetala mutant epidermal cells revealed that type I ROPs affect cell structure primarily on the adaxial side, while type II ROPs are functional and induce a novel cell division phenotype in this genetic background. Taken together, our studies show how prenyl and S-acyl lipid modifications affect ROP subcellular distribution, membrane interaction dynamics, and function.  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