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 ability of proteins and their complexes to withstand or respond to mechanical stimuli is vital for cells to maintain their structural organisation, to relay external signals and to facilitate unfolding and remodelling. Force spectroscopy using the atomic force microscope allows the behaviour of single protein molecules under an applied extension to be investigated and their mechanical strength to be quantified. protein L, a simple model protein, displays moderate mechanical strength and is thought to unfold by the shearing of two mechanical sub-domains. Here, we investigate the importance of side-chain packing for the mechanical strength of protein L by measuring the mechanical strength of a series of protein L variants containing single conservative hydrophobic volume deletion mutants. Of the five thermodynamically destabilised variants characterised, only one residue (I60V) close to the interface between two mechanical sub-domains was found to differ in mechanical properties to wild type (ΔFI60V-WT = − 36 pN at 447 nm s− 1, ΔxuI60V-WT = 0.2 nm). Φ-value analysis of the unfolding data revealed a highly native transition state. To test whether the number of hydrophobic contacts across the mechanical interface does affect the mechanical strength of protein L, we measured the mechanical properties of two further variants. protein L L10F, which increases core packing but does not enhance interfacial contacts, increased mechanical strength by 13 ± 11 pN at 447 nm s− 1. By contrast, protein L I60F, which increases both core and cross-interface contacts, increased mechanical strength by 72 ± 13 pN at 447 nm s− 1. These data suggest a method by which nature can evolve a varied mechanical response from a limited number of topologies and demonstrate a generic but facile method by which the mechanical strength of proteins can be rationally modified.  相似文献   

2.
Yeast enolase is stabilized by its natural cofactor Mg2+. This stabilization is ascribed to the reduced subunit dissociation of the holoprotein. Nevertheless, how Mg2+ alters the unfolding mechanism has yet to be fully characterized. Here, we investigate the role of Mg2+ in the denaturation mechanism and unfolding kinetics of yeast enolase. Apo-enolase unfolds through a three-state process (N2 ↔ 2I → 2D). The intermediate species is described as a monomeric molten globule-like conformation that becomes noticeable in the presence of phosphate and is able to recover its native secondary structure when cooled down. Kinetic studies confirmed the presence of the intermediate species, even though it was not noticeable in the thermal scans. The cofactor increases the cooperativity of the unfolding transitions, while the intermediate species becomes less noticeable or nonexistent. Thus, holo-enolase follows a simple two-state mechanism (N2 → 2D). Our results indicate smaller unfolding rate-constants in the presence of Mg2+, thus favoring the native state. The temperature dependence of the unfolding rates allowed us to calculate the activation enthalpies of denaturation. Interestingly, despite the different unfolding mechanisms of the apo and holo forms of enolase, they both have similar activation barriers of denaturation (185-190 kJ mol−1).  相似文献   

3.
The factors controlling the stability, folding, and dynamics of integral membrane proteins are not fully understood. The high stability of the membrane protein bacteriorhodopsin (bR), an archetypal member of the rhodopsin photoreceptor family, has been ascribed to its covalently bound retinal cofactor. We investigate here the role of this cofactor in the thermodynamic stability and folding kinetics of bR. Multiple spectroscopic probes were used to determine the kinetics and energetics of protein folding in mixed lipid/detergent micelles in the presence and absence of retinal. The presence of retinal increases extrapolated values for the overall unfolding free energy from 6.3 ± 0.4 kcal mol− 1 to 23.4 ± 1.5 kcal mol− 1 at zero denaturant, suggesting that the cofactor contributes 17.1 kcal mol− 1 towards the overall stability of bR. In addition, the cooperativity of equilibrium unfolding curves is markedly reduced in the absence of retinal with overall m-values decreasing from 31.0 ± 2.0 kcal mol− 1 to 10.9 ± 1.0 kcal mol− 1, indicating that the folded state of the apoprotein is less compact than the equivalent for the holoprotein. This change in the denaturant response means that the difference in the unfolding free energy at a denaturant concentration midway between the two unfolding curves is only ca 3-6 kcal mol− 1. Kinetic data show that the decrease in stability upon removal of retinal is associated with an increase in the apparent intrinsic rate constant of unfolding, kuH2O, from ~1 × 10− 16 s− 1 to ~1 × 10− 4 s− 1 at 25 °C. This correlates with a decrease in the unfolding activation energy by 16.3 kcal mol− 1 in the apoprotein, extrapolated to zero SDS. These results suggest that changes in bR stability induced by retinal binding are mediated solely by changes in the activation barrier for unfolding. The results are consistent with a model in which bR is kinetically stabilized via a very slow rate of unfolding arising from protein-retinal interactions that increase the rigidity and compactness of the polypeptide chain.  相似文献   

4.
Folding mechanisms and stability of membrane proteins are poorly understood because of the known difficulties in finding experimental conditions under which reversible denaturation could be possible. In this work, we describe the equilibrium unfolding of Archaeoglobus fulgidus CopA, an 804-residue α-helical membrane protein that is involved in transporting Cu+ throughout biological membranes. The incubation of CopA reconstituted in phospholipid/detergent mixed micelles with high concentrations of guanidinium hydrochloride induced a reversible decrease in fluorescence quantum yield, far-UV ellipticity, and loss of ATPase and phosphatase activities. Refolding of CopA from this unfolded state led to recovery of full biological activity and all the structural features of the native enzyme. CopA unfolding showed typical characteristics of a two-state process, with ΔGw° = 12.9 kJ mol 1, = 4.1 kJ mol− 1 M− 1, Cm = 3 M, and ΔCpw° = 0.93 kJ mol− 1 K− 1. These results point out to a fine-tuning mechanism for improving protein stability. Circular dichroism spectroscopic analysis of the unfolded state shows that most of the secondary and tertiary structures were disrupted. The fraction of Trp fluorescence accessible to soluble quenchers shifted from 0.52 in the native state to 0.96 in the unfolded state, with a significant spectral redshift. Also, hydrophobic patches in CopA, mainly located in the transmembrane region, were disrupted as indicated by 1-anilino-naphtalene-8-sulfonate fluorescence. Nevertheless, the unfolded state had a small but detectable amount of residual structure, which might play a key role in both CopA folding and adaptation for working at high temperatures.  相似文献   

5.
The stabilization of RNA tertiary structures by ions is well known, but the neutral osmolyte trimethylamine oxide (TMAO) can also effectively stabilize RNA tertiary structure. To begin to understand the physical basis for the effects of TMAO on RNA, we have quantitated the TMAO-induced stabilization of five RNAs with known structures. So-called m values, the increment in unfolding free energy per molal of osmolyte at constant KCl activity, are ∼ 0 for a hairpin secondary structure and between 0.70 and 1.85 kcal mol− 1m− 1 for four RNA tertiary structures (30-86 nt). Further analysis of two RNAs by small-angle X-ray scattering and hydroxyl radical probing shows that TMAO reduces the radius of gyration of the unfolded ensemble to the same endpoint as seen in titration with Mg2+ and that the structures stabilized by TMAO and Mg2+ are indistinguishable. Remarkably, TMAO induces the native conformation of a Mg2+ ion chelation site formed in part by a buried phosphate, even though Mg2+ is absent. TMAO interacts weakly, if at all, with KCl, ruling out the possibility that TMAO stabilizes RNA indirectly by increasing salt activity. TMAO is, however, strongly excluded from the vicinity of dimethylphosphate (unfavorable interaction free energy, + 211 cal mol− 1m− 1 for the potassium salt), an ion that mimics the RNA backbone phosphate. We suggest that formation of RNA tertiary structure is accompanied by substantial phosphate dehydration (loss of 66-173 water molecules in the RNA structures studied) and that TMAO works principally by reducing the energetic penalty associated with this dehydration. The strong parallels we find between the effects of TMAO and Mg2+ suggest that RNA sequence is more important than specific ion interactions in specifying the native structure.  相似文献   

6.
An early step in the morphogenesis of the double-stranded DNA (dsDNA) bacteriophage HK97 is the assembly of a precursor shell (prohead I) from 420 copies of a 384-residue subunit (gp5). Although formation of prohead I requires direct participation of gp5 residues 2-103 (Δ-domain), this domain is eliminated by viral protease prior to subsequent shell maturation and DNA packaging. The prohead I Δ-domain is thought to resemble a phage scaffolding protein, by virtue of its highly α-helical secondary structure and a tertiary fold that projects inward from the interior surface of the shell. Here, we employ factor analysis of temperature-dependent Raman spectra to characterize the thermostability of the Δ-domain secondary structure and to quantify the thermodynamic parameters of Δ-domain unfolding. The results are compared for the Δ-domain within the prohead I architecture (in situ) and for a recombinantly expressed 111-residue peptide (in vitro). We find that the α-helicity (∼ 70%), median melting temperature (Tm = 58 °C), enthalpy (ΔHm = 50 ± 5 kcal mol− 1), entropy (ΔSm = 150 ± 10 cal mol− 1 K− 1), and average cooperative melting unit (〈nc〉 ∼ 3.5) of the in situ Δ-domain are altered in vitro, indicating specific interdomain interactions within prohead I. Thus, the in vitro Δ-domain, despite an enhanced helical secondary structure (∼ 90% α-helix), exhibits diminished thermostability (Tm = 40 °C; ΔHm = 27 ± 2 kcal mol− 1; ΔSm = 86 ± 6 cal mol− 1 K− 1) and noncooperative unfolding (〈nc〉 ∼ 1) vis-à-vis the in situ Δ-domain. Temperature-dependent Raman markers of subunit side chains, particularly those of Phe and Trp residues, also confirm different local interactions for the in situ and in vitro Δ-domains. The present results clarify the key role of the gp5 Δ-domain in prohead I architecture by providing direct evidence of domain structure stabilization and interdomain interactions within the assembled shell.  相似文献   

7.
The bindings of biogenic polyamines spermine (spm), spermidine (spmd) and synthetic polyamines 3,7,11,15-tetrazaheptadecane·4HCl (BE-333) and 3,7,11,15,19-pentazahenicosane·5HCl (BE-3333) with β-lactoglobulin (β-LG) were determined in aqueous solution. FTIR, UV-vis, CD and fluorescence spectroscopic methods as well as molecular modeling were used to determine the polyamine binding sites and the effect of polyamine complexation on protein stability and secondary structure. Structural analysis showed that polyamines bind β-LG via both hydrophilic and hydrophobic contacts. Stronger polyamine-protein complexes formed with synthetic polyamines than biogenic polyamines, with overall binding constants of Kspm-β-LG = 3.2(±0.6) × 104 M−1, Kspmd-β-LG = 1.8(±0.5) × 104 M−1, KBE-333-β-LG = 5.8(±0.3) × 104 M−1 and KBE-3333-β-LG = 6.2(±0.05) × 104 M−1. Molecular modeling showed the participation of several amino acids in the polyamine complexes with the following order of polyamine-protein binding affinity: BE-3333 > BE-333 > spermine > spermidine, which correlates with their positively charged amino group content. Alteration of protein conformation was observed with a reduction of β-sheet from 57% (free protein) to 55-51%, and a major increase of turn structure from 13% (free protein) to ∼21% in the polyamine-β-LG complexes, indicating a partial protein unfolding.  相似文献   

8.
The X-ray crystal structures of two related trans-N2S2 copper macrocycles are reported. One was isolated with the copper in the divalent form and the other with copper in its univalent form affording a valuable insight into the changes of geometry and metrical parameters that occur during redox processes in macrocyclic copper complexes. A variable temperature NMR study of the copper(I) complex is reported, indicative of a chair-boat conformational change within the alkyl chain backbone of the macrocycle. It was possible to extract the relevant kinetic and thermodynamic parameters (ΔG, 57.8 kJ mol−1; ΔH, 52.1 kJ mol−1; ΔS, −19.2 J K−1 mol−1) for this process at 298 K. DFT molecular orbital calculations were used to confirm these observations and to calculate the energy difference (26.2 kJmol−1) between the copper(I) macrocycle in a planar and a distorted tetrahedral disposition.  相似文献   

9.
Escherichia coli RecBCD is a highly processive DNA helicase involved in double-strand break repair and recombination that possesses two helicase/translocase subunits with opposite translocation directionality (RecB (3′ to 5′) and RecD (5′ to 3′)). RecBCD has been shown to melt out ∼ 5-6 bp upon binding to a blunt-ended duplex DNA in a Mg2+-dependent, but ATP-independent reaction. Here, we examine the binding of E. coli RecBC helicase (minus RecD), also a processive helicase, to duplex DNA ends in the presence and in the absence of Mg2+ in order to determine if RecBC can also melt a duplex DNA end in the absence of ATP. Equilibrium binding of RecBC to DNA substrates with ends possessing pre-formed 3′ and/or 5′ single-stranded (ss)-(dT)n flanking regions (tails) (n ranging from zero to 20 nt) was examined by competition with a fluorescently labeled reference DNA and by isothermal titration calorimetry. The presence of Mg2+ enhances the affinity of RecBC for DNA ends possessing 3′ or 5′-(dT)n ssDNA tails with n < 6 nt, with the relative enhancement decreasing as n increases from zero to six nt. No effect of Mg2+ was observed for either the binding constant or the enthalpy of binding (ΔHobs) for RecBC binding to DNA with ssDNA tail lengths, n ≥ 6 nucleotides. Upon RecBC binding to a blunt duplex DNA end in the presence of Mg2+, at least 4 bp at the duplex end become accessible to KMnO4 attack, consistent with melting of the duplex end. Since Mg2+ has no effect on the affinity or binding enthalpy of RecBC for a DNA end that is fully pre-melted, this suggests that the role of Mg2+ is to overcome a kinetic barrier to melting of the DNA by RecBC and presumably also by RecBCD. These data also provide an accurate estimate (ΔHobs = 8 ± 1 kcal/mol) for the average enthalpy change associated with the melting of a DNA base-pair by RecBC.  相似文献   

10.
The effect of ions on the thermostability and unfolding of Na,K-ATPase from shark salt gland was studied and compared with that of Na,K-ATPase from pig kidney by using differential scanning calorimetry (DSC) and activity assays. In 1 mM histidine at pH 7, the shark enzyme inactivates rapidly at 20 °C, as does the kidney enzyme at 42 °C (but not at 20 °C). Increasing ionic strength by addition of 20 mM histidine, or of 1 mM NaCl or KCl, protects both enzymes against this rapid inactivation. As detected by DSC, the shark enzyme undergoes thermal unfolding at lower temperature (Tm ≈ 45 °C) than does the kidney enzyme (Tm ≈ 55 °C). Both calorimetric endotherms indicate multi-step unfolding, probably associated with different cooperative domains. Whereas the overall heat of unfolding is similar for the kidney enzyme in either 1 mM or 20 mM histidine, components with high mid-point temperatures are lost from the unfolding transition of the shark enzyme in 1 mM histidine, relative to that in 20 mM histidine. This is attributed to partial unfolding of the enzyme due to a high hydrostatic pressure during centrifugation of DSC samples at low ionic strength, which correlates with inactivation measurements. Addition of 10 mM NaCl to shark enzyme in 1 mM histidine protects against inactivation during centrifugation of the DSC sample, but incubation for 1 h at 20 °C prior to addition of NaCl results in loss of components with lower mid-point temperatures within the unfolding transition. Cations at millimolar concentration therefore afford at least two distinct modes of stabilization, likely affecting separate cooperative domains. The different thermal stabilities and denaturation temperatures of the two Na,K-ATPases correlate with the respective physiological temperatures, and may be attributed to the different lipid environments.  相似文献   

11.
Escherichiacoli RecBCD is a bipolar DNA helicase possessing two motor subunits (RecB, a 3′-to-5′ translocase, and RecD, a 5′-to-3′ translocase) that is involved in the major pathway of recombinational repair. Previous studies indicated that the minimal kinetic mechanism needed to describe the ATP-dependent unwinding of blunt-ended DNA by RecBCD in vitro is a sequential n-step mechanism with two to three additional kinetic steps prior to initiating DNA unwinding. Since RecBCD can “melt out” ∼ 6 bp upon binding to the end of a blunt-ended DNA duplex in a Mg2+-dependent but ATP-independent reaction, we investigated the effects of noncomplementary single-stranded (ss) DNA tails [3′-(dT)6 and 5′-(dT)6 or 5′-(dT)10] on the mechanism of RecBCD and RecBC unwinding of duplex DNA using rapid kinetic methods. As with blunt-ended DNA, RecBCD unwinding of DNA possessing 3′-(dT)6 and 5′-(dT)6 noncomplementary ssDNA tails is well described by a sequential n-step mechanism with the same unwinding rate (mkU = 774 ± 16 bp s− 1) and kinetic step size (m = 3.3 ± 1.3 bp), yet two to three additional kinetic steps are still required prior to initiation of DNA unwinding (kC = 45 ± 2 s− 1). However, when the noncomplementary 5′ ssDNA tail is extended to 10 nt [5′-(dT)10 and 3′-(dT)6], the DNA end structure for which RecBCD displays optimal binding affinity, the additional kinetic steps are no longer needed, although a slightly slower unwinding rate (mkU = 538 ± 24 bp s− 1) is observed with a similar kinetic step size (m = 3.9 ± 0.5 bp). The RecBC DNA helicase (without the RecD subunit) does not initiate unwinding efficiently from a blunt DNA end. However, RecBC does initiate well from a DNA end possessing noncomplementary twin 5′-(dT)6 and 3′-(dT)6 tails, and unwinding can be described by a simple uniform n-step sequential scheme, without the need for the additional kC initiation steps, with a similar kinetic step size (m = 4.4 ± 1.7 bp) and unwinding rate (mkobs = 396 ± 15 bp s− 1). These results suggest that the additional kinetic steps with rate constant kC required for RecBCD to initiate unwinding of blunt-ended and twin (dT)6-tailed DNA reflect processes needed to engage the RecD motor with the 5′ ssDNA.  相似文献   

12.
Insulin-degrading enzyme (IDE) can degrade insulin and amyloid-β, peptides involved in diabetes and Alzheimer's disease, respectively. IDE selects its substrates based on size, charge, and flexibility. From these criteria, we predict that IDE can cleave and inactivate ubiquitin (Ub). Here, we show that IDE cleaves Ub in a biphasic manner, first, by rapidly removing the two C-terminal glycines (kcat = 2 s− 1) followed by a slow cleavage between residues 72 and 73 (kcat = 0.07 s−  1), thereby producing the inactive 1-74 fragment of Ub (Ub1-74) and 1-72 fragment of Ub (Ub1-72). IDE is a ubiquitously expressed cytosolic protein, where monomeric Ub is also present. Thus, Ub degradation by IDE should be regulated. IDE is known to bind the cytoplasmic intermediate filament protein nestin with high affinity. We found that nestin potently inhibits the cleavage of Ub by IDE. In addition, Ub1-72 has a markedly increased affinity for IDE (∼ 90-fold). Thus, the association of IDE with cellular regulators and product inhibition by Ub1-72 can prevent inadvertent proteolysis of cellular Ub by IDE. Ub is a highly stable protein. However, IDE instead prefers to degrade peptides with high intrinsic flexibility. Indeed, we demonstrate that IDE is exquisitely sensitive to Ub stability. Mutations that only mildly destabilize Ub (ΔΔG <  0.6 kcal/mol) render IDE hypersensitive to Ub with rate enhancements greater than 12-fold. The Ub-bound IDE structure and IDE mutants reveal that the interaction of the exosite with the N-terminus of Ub guides the unfolding of Ub, allowing its sequential cleavages. Together, our studies link the control of Ub clearance with IDE.  相似文献   

13.
The crystal structure of the yeast SLN1 response regulator (RR) domain bound to both a phosphoryl analog [beryllium fluoride (BeF3 −)] and Mg2 +, in complex with its downstream phosphorelay signaling partner YPD1, has been determined at a resolution of 1.70 Å. Comparisons between the BeF3 −-activated complex and the unliganded (or apo) complex determined previously reveal modest but important differences. The SLN1-R1·Mg2 +·BeF3 − structure from the complex provides evidence for the first time that the mechanism of phosphorylation-induced activation is highly conserved between bacterial RR domains and this example from a eukaryotic organism. Residues in and around the active site undergo slight rearrangements in order to form bonds with the essential divalent cation and fluorine atoms of BeF3 −. Two conserved switch-like residues (Thr1173 and Phe1192) occupy distinctly different positions in the apo versus BeF3 −-bound structures, consistent with the “Y-T” coupling mechanism proposed for the activation of CheY and other bacterial RRs. Several loop regions and the α4-β5-α5 surface of the SLN1-R1 domain undergo subtle conformational changes (∼ 1-3 Å displacements relative to the apo structure) that lead to significant changes in terms of contacts that are formed with YPD1. Detailed structural comparisons of protein-protein interactions in the apo and BeF3 −-bound complexes suggest at least a two-state equilibrium model for the formation of a transient encounter complex, in which phosphorylation of the RR promotes the formation of a phosphotransfer-competent complex. In the BeF3 −-activated complex, the position of His64 from YPD1 needs to be within ideal distance of and in near-linear geometry with Asp1144 from the SLN1-R1 domain for phosphotransfer to occur. The ground-state structure presented here suggests that phosphoryl transfer will likely proceed through an associative mechanism involving the formation of a pentacoordinate phosphorus intermediate.  相似文献   

14.
Alexander Wiedenmann 《BBA》2008,1777(10):1301-1310
The membrane-embedded F0 part of ATP synthases is responsible for ion translocation during ATP synthesis and hydrolysis. Here, we describe an in vitro system for measuring proton fluxes through F0 complexes by fluorescence changes of the entrapped fluorophore pyranine. Starting from purified enzyme, the F0 part was incorporated unidirectionally into phospholipid vesicles. This allowed analysis of proton transport in either synthesis or hydrolysis direction with Δψ or ΔpH as driving forces. The system displayed a high signal-to-noise ratio and can be accurately quantified. In contrast to ATP synthesis in the Escherichia coli F1F0 holoenzyme, no significant difference was observed in the efficiency of ΔpH or Δψ as driving forces for H+-transport through F0. Transport rates showed linear dependency on the driving force. Proton transport in hydrolysis direction was about 2400 H+/(s × F0) at Δψ of 120 mV, which is approximately twice as fast as in synthesis direction. The chloroplast enzyme was faster and catalyzed H+-transport at initial rates of 6300 H+/(s × F0) under similar conditions. The new method is an ideal tool for detailed kinetic investigations of the ion transport mechanism of ATP synthases from various organisms.  相似文献   

15.
Much of our understanding of protein folding mechanisms is derived from experiments using intrinsic fluorescence of natural or genetically inserted tryptophan (Trp) residues to monitor protein refolding and site-directed mutagenesis to determine the energetic role of amino acids in the native (N), intermediate (I) or transition (T) states. However, this strategy has limited use to study complex folding reactions because a single fluorescence probe may not detect all low-energy folding intermediates. To overcome this limitation, we suggest that protein refolding should be monitored with different solvent-exposed Trp probes. Here, we demonstrate the utility of this approach by investigating the controversial folding mechanism of ubiquitin (Ub) using Trp probes located at residue positions 1, 28, 45, 57, and 66. We first show that these Trp are structurally sensitive and minimally perturbing fluorescent probes for monitoring folding/unfolding of the protein. Using a conventional stopped-flow instrument, we show that ANS and Trp fluorescence detect two distinct transitions during the refolding of all five Trp mutants at low concentrations of denaturant: T1, a denaturant-dependent transition and T2, a slower transition, largely denaturant-independent. Surprisingly, some Trp mutants (UbM1W, UbS57W) display Trp fluorescence changes during T1 that are distinct from the expected U → N transition suggesting that the denaturant-dependent refolding transition of Ub is not a U → N transition but represents the formation of a structurally distinct I-state (U → I). Alternatively, this U → I transition could be also clearly distinguished by using a combination of two Trp mutations UbF45W-T66W for which the two Trp probes that display fluorescence changes of opposite sign during T1 and T2 (UbF45W-T66W). Global fitting of the folding/unfolding kinetic parameters and additional folding-unfolding double-jump experiments performed on UbM1W, a mutant with enhanced fluorescence in the I-state, demonstrate that the I-state is stable, compact, misfolded, and on-pathway. These results illustrate how transient low-energy I-states can be characterized efficiently in complex refolding reactions using multiple Trp probes.  相似文献   

16.
The cellular isoform of the prion protein PrPC is a Cu2+-binding cell surface glycoprotein that, when misfolded, is responsible for a range of transmissible spongiform encephalopathies. As changes in PrPC conformation are intimately linked with disease pathogenesis, the effect of Cu2+ ions on the structure and stability of the protein has been investigated. Urea unfolding studies indicate that Cu2+ ions destabilise the native fold of PrPC. The midpoint of the unfolding transition is reduced by 0.73 ± 0.07 M urea in the presence of 1 mol equiv of Cu2+. This equates to an appreciable difference in free energy of unfolding (2.02 ± 0.05 kJ mol− 1 at the midpoint of unfolding). We relate Cu2+-induced changes in secondary structure for full-length PrP(23-231) to smaller Cu2+ binding fragments. In particular, Cu2+-induced structural changes can directly be attributed to Cu2+ binding to the octarepeat region of PrPC. Furthermore, a β-sheet-like transition that is observed when Cu ions are bound to the amyloidogenic fragment of PrP (residues 90-126) is due only to local Cu2+ coordination to the individual binding sites centred at His95 and His110. Cu2+ binding does not directly generate a β-sheet conformation within PrPC; however, Cu2+ ions do destabilise the native fold of PrPC and may make the transition to a misfolded state more favourable.  相似文献   

17.
Retroviruses selectively package two copies of their RNA genomes via mechanisms that have yet to be fully deciphered. Recent studies with small fragments of the Moloney murine leukemia virus (MoMuLV) genome suggested that selection may be mediated by an RNA switch mechanism, in which conserved UCUG elements that are sequestered by base-pairing in the monomeric RNA become exposed upon dimerization to allow binding to the cognate nucleocapsid (NC) domains of the viral Gag proteins. Here we show that a large fragment of the MoMuLV 5′ untranslated region that contains all residues necessary for efficient RNA packaging (ΨWT; residues 147-623) also exhibits a dimerization-dependent affinity for NC, with the native dimer ([ΨWT]2) binding 12 ± 2 NC molecules with high affinity (Kd = 17 ± 7 nM) and with the monomer, stabilized by substitution of dimer-promoting loop residues with hairpin-stabilizing sequences (ΨM), binding 1-2 NC molecules. Identical dimer-inhibiting mutations in MoMuLV-based vectors significantly inhibit genome packaging in vivo (∼ 100-fold decrease), whereas a large deletion of nearly 200 nucleotides just upstream of the gag start codon has minimal effects. Our findings support the proposed RNA switch mechanism and further suggest that virus assembly may be initiated by a complex comprising as few as 12 Gag molecules bound to a dimeric packaging signal.  相似文献   

18.
Kinetic studies of X exchange on [AuX4] square-planar complexes (where X=Cl and CN) were performed at acidic pH in the case of chloride system and as a function of pH for the cyanide one. Chloride NMR study (330-365 K) gives a second-order rate law on [AuCl4] with the kinetic parameters: (k2Au,Cl)298=0.56±0.03 s−1 mol−1 kg; ΔH2‡ Au,Cl=65.1±1 kJ mol−1; ΔS2‡ Au,Cl=−31.3±3 J mol−1 K−1 and ΔV2 Au,Cl=−14±2 cm3 mol−1. The variable pressure data clearly indicate the operation of an Ia or A mechanism for this exchange pathway. The proton exchange on HCN was determined by 13C NMR as a function of pH and the rate constant of the three reaction pathways involving H2O, OH and CN were determined: k0HCN,H=113±17 s−1, k1HCN,H=(2.9±0.7)×109 s−1 mol−1 kg and k2HCN,H=(0.6±0.2)×106 s−1 mol−1 kg at 298.1 K. The rate law of the cyanide exchange on [Au(CN)4] was found to be second order with the following kinetic parameters: (k2Au,CN)298=6240±85 s−1 mol−1 kg, ΔH2 Au,CN=40.0±0.8 kJ mol−1, ΔS2 Au,CN=−37.8±3 J mol−1 K−1 and ΔV2 Au,CN=+2±1 cm3 mol−1. The rate constant observed varies about nine orders of magnitude depending on the pH and HCN does not act as a nucleophile. The observed rate constant of X exchange on [AuX4] are two or three orders of magnitude faster than the Pt(II) analogue.  相似文献   

19.
A series of [R3EP7W(CO)3]2− complexes where (E = Si, Ge, Sn, Pb; R = alkyl, phenyl) were prepared from [P7W(CO)3]3− and R3EX reagents (X = Cl, Br) in dmf or CH3CN solutions. The Pb derivatives were prepared at −50 °C and are not thermally stable. The compounds were characterized by 31P NMR spectroscopy and selected ESI-MS studies. All compounds undergo rapid inversion at the ER3-bound phosphorus atom. The barriers to inversion were measured by way of 2D 31P EXSY experiments at various temperatures. The analysis showed very low barriers to pyramidal inversion (ΔG 10.3-13.5 kcal/mol) that were essentially enthalpic in origin. The activation barriers generally increased with increasing electronegativity of the E atom and the steric bulk of the ER3 substituents. The latter was interpreted by way of a non-linear transition state.  相似文献   

20.
A subclass of proteins with the SEA (sea urchin sperm protein, enterokinase, and agrin) domain fold exists as heterodimers generated by autoproteolytic cleavage within a characteristic G− 1S+ 1VVV sequence. Autoproteolysis occurs by a nucleophilic attack of the serine hydroxyl on the vicinal glycine carbonyl followed by an N → O acyl shift and hydrolysis of the resulting ester. The reaction has been suggested to be accelerated by the straining of the scissile peptide bond upon protein folding. In an accompanying article, we report the mechanism; in this article, we provide further key evidence and account for the energetics of coupled protein folding and autoproteolysis. Cleavage of the GPR116 domain and that of the MUC1 SEA domain occur with half-life (t½) values of 12 and 18 min, respectively, with lowering of the free energy of the activation barrier by ∼ 10 kcal mol− 1 compared with uncatalyzed hydrolysis. The free energies of unfolding of the GPR116 and MUC1 SEA domains were measured to ∼ 11 and ∼ 15 kcal mol− 1, respectively, but ∼ 7 kcal mol− 1 of conformational energy is partitioned as strain over the scissile peptide bond in the precursor to catalyze autoproteolysis by substrate destabilization. A straining energy of ∼ 7 kcal mol− 1 was measured by using both a pre-equilibrium model to analyze stability and cleavage kinetics data obtained with the GPR116 SEA domain destabilized by core mutations or urea addition, as well as the difference in thermodynamic stabilities of the MUC1 SEA precursor mutant S1098A (with a G− 1A+ 1VVV motif) and the wild-type protein. The results imply that cleavage by N → O acyl shift alone would proceed with a t½ of ∼ 2.3 years, which is too slow to be biochemically effective. A subsequent review of structural data on other self-cleaving proteins suggests that conformational strain of the scissile peptide bond may be a common mechanism of autoproteolysis.  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