首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Ann L. Umbach  James N. Siedow 《BBA》2006,1757(2):135-142
Two Cys residues, CysI and CysII, are present in most plant alternative oxidases (AOXs). CysI inactivates AOX by forming a disulfide bond with the corresponding CysI residue on the adjacent subunit of the AOX homodimer. When reduced, CysI associates with α-keto acids, such as pyruvate, to activate AOX, an effect mimicked by charged amino acid substitutions at the CysI site. CysII may also be a site of AOX activity regulation, through interaction with the small α-keto acid, glyoxylate. Comparison of Arabidopsis AOX1a (AtAOX1a) mutants with single or double substitutions at CysI and CysII confirmed that glyoxylate interacted with either Cys, while the effect of pyruvate (or succinate for AtAOX1a substituted with Ala at CysI) was limited to CysI. A variety of CysII substitutions constitutively activated AtAOX1a, indicating that neither the catalytic site nor, unlike at CysI, charge repulsion is involved. Independent effects at each Cys were suggested by lack of CysII substitution interference with pyruvate stimulation at CysI, and close to additive activation at the two sites. However, results obtained using diamide treatment to covalently link the AtAOX1a subunits by the disulfide bond indicated that CysI must be in the reduced state for activation at CysII to occur.  相似文献   

2.
The homodimeric cyanide-resistant alternative oxidase of plant mitochondria reduces oxygen to water without forming ATP. Arabidopsis thaliana alternative oxidase AOX1a is stimulated by pyruvate or other alpha-keto acids associating with a regulatory cysteine at position 78, by succinate in a serine-78 mutant, and by site-directed mutation of position 78 to glutamate. The mechanism of activation was explored with additional amino acid substitutions made at Cys-78 in AOX1a, which was functionally expressed in Escherichia coli. Oxidases with positively charged substitutions (Lys and Arg) were insensitive to pyruvate or succinate but were more active than the wild type without pyruvate. Uncharged substitutions (Gln, Leu) produced an inactive enzyme. These results indicate that activation may be due to conformational changes caused by charge repulsion between the dimer subunits and not through a direct role of alpha-keto acids in catalysis. Oxygen isotope fractionation experiments suggest that the charge of the amino acid at position 78 also affects the entry of oxygen into the active site. Therefore, the N-terminal portion of the protein containing residue 78 can indirectly affect both catalysis at the diiron active site and the path of oxygen to that site. In addition, both positively and negatively substituted alternative oxidases were stimulated by glyoxylate, suggesting the presence of a second activation site, possibly Cys-128.  相似文献   

3.
GC Vanlerberghe  L McIntosh    JY Yip 《The Plant cell》1998,10(9):1551-1560
Using in organellar assays, we found that significant tobacco alternative oxidase (AOX) activity is dependent on both reduction of a putative regulatory disulfide bond and the presence of pyruvate, which may interact with a Cys sulfhydryl. This redox modulation and pyruvate activation thus may be important in determining the partitioning of electrons to AOX in vivo. To investigate these regulatory mechanisms, we generated tobacco plants expressing mutated AOX proteins. Mutation of the most N-terminal Cys residue (Cys-126) to an Ala residue produced an AOX that could not be converted to the disulfide-linked form, thus identifying this Cys residue as being responsible for redox modulation. Although this mutation might be expected to produce an AOX with constitutive high activity in the presence of pyruvate, we found it to have minimal in organellar activity in the presence of pyruvate. Nonetheless, the Cys-126 mutation did not appear to have compromised the catalytic function of AOX, given that cells expressing the protein displayed high rates of cyanide-resistant respiration in vivo. The striking difference between in vivo and in organellar results suggests that an additional mechanism(s), as yet unidentified by in organellar assays, may promote activity in vivo. Mutation of the Cys residue nearest the presumptive active site (Cys-176) to an Ala residue did not prevent disulfide bond formation or affect the ability of AOX to be stimulated by pyruvate, indicating that this Cys residue is involved in neither redox modulation nor pyruvate activation.  相似文献   

4.
Both plant and fungal mitochondria have cyanide-resistant alternative oxidases that use reductant from the mitochondrial ubiquinone pool to reduce oxygen to water in a reaction that conserves no energy for ATP synthesis. The dimeric plant alternative oxidase is relatively inactive when its subunits are linked by a disulfide bond. When this bond is reduced, the enzyme can then be stimulated by its activators, alpha-keto acids. A Cys in the N-terminal section of the protein is responsible for both of these features. We examined the alternative oxidases in mitochondria isolated from two fungi Neurospora crassa and Pichia stipitis for dimeric structure, ability to form an intermolecular disulfide, and sensitivity to alpha-keto acids. Neither of the two fungal alternative oxidases could be covalently linked by diamide, which induces disulfide bond formation between nearby Cys residues, nor could they be cross-linked by a Lys-specific reagent or glutaraldehyde at concentrations which cross-link the plant alternative oxidase dimer completely. Alternative oxidase activity in fungal mitochondria was not stimulated by the alpha-keto acids pyruvate and glyoxylate. Pyruvate did stimulate activity when succinate was the respiratory substrate, but this was not a direct effect on the alternative oxidase. In contrast, added GMP was a strong activator of fungal alternative oxidase activity. Analysis of plant and fungal alternative oxidase protein sequences revealed a unique domain of about 40 amino acids surrounding the regulatory Cys in the plant sequences that is not present in the fungal sequences. This domain may be where dimerization of the plant enzymes occurs. In contrast to plant enzymes, the fungal alternative oxidases studied here are monomeric and their activities are independent of alpha-keto acids.  相似文献   

5.
Alternative oxidase (AOX) is encoded in small multigene families in plants. Functional analysis of the Arabidopsis (Arabidopsis thaliana) alternative oxidase 1c (AtAOX1c) promoter, an AOX gene not induced by oxidative stress, indicated that regulation of expression was complex, with the upstream promoter region containing positive and negative response regions. Comparison to the promoter region of soybean (Glycine max) alternative oxidase 2b (GmAOX2b), another AOX gene not induced by oxidative stress, revealed that they contained seven sequence elements in common. All elements were active in the promoter region of AtAOX1c in suspension cells and in leaf tissue from Columbia and mutant plants, where a mitochondrial protein import receptor was inactivated. Analysis of coexpressed and putatively coregulated genes, the latter defined as containing five or more sequence elements functional in AtAOX1c, indicated that AtAOX1c was coregulated with components involved with cell division and growth. Consistent with this analysis, we demonstrated that site II elements, previously shown to regulate the proliferating cell nuclear antigen, are present in the upstream promoter region of AtAOX1c and were strong negative regulators of AtAOX1c expression. It was demonstrated that NDB4, a gene encoding an external NAD(P)H dehydrogenase, displayed strong coexpression with AtAOX1c. Overall, these results indicate that AtAOX1c is regulated by growth and developmental signals.  相似文献   

6.
Rat 3-mercaptopyruvate sulfurtransferase (MST) contains three exposed cysteines as follows: a catalytic site cysteine, Cys(247), in the active site and Cys(154) and Cys(263) on the surface of MST. The corresponding cysteine to Cys(263) is conserved in mammalian MSTs, and Cys(154) is a unique cysteine. MST has monomer-dimer equilibrium with the assistance of oxidants and reductants. The monomer to dimer ratio is maintained at approximately 92:8 in 0.2 m potassium phosphate buffer containing no reductants under air-saturated conditions; the dimer might be symmetrical via an intersubunit disulfide bond between Cys(154) and Cys(154) and between Cys(263) and Cys(263), or asymmetrical via an intersubunit disulfide bond between Cys(154) and Cys(263). Escherichia coli reduced thioredoxin (Trx) cleaved the intersubunit disulfide bond to activate MST to 2.3- and 4.9-fold the levels of activation of dithiothreitol (DTT)-treated and DTT-untreated MST, respectively. Rat Trx also activated MST. On the other hand, reduced glutathione did not affect MST activity. E. coli C35S Trx, in which Cys(35) was replaced with Ser, formed some adducts with MST and activated MST after treatment with DTT. Thus, Cys(32) of E. coli Trx reacted with the redox-active cysteines, Cys(154) and Cys(263), by forming an intersubunit disulfide bond and a sulfenyl Cys(247). A consecutively formed disulfide bond between Trx and MST must be cleaved for the activation. E. coli C32S Trx, however, did not activate MST. Reduced Trx turns on a redox switch for the enzymatic activation of MST, which contributes to the maintenance of cellular redox homeostasis.  相似文献   

7.
Delonix regia trypsin inhibitor (DrTI) consists of a single-polypeptide chain with a molecular mass of 22 kDa and containing two disulfide bonds (Cys44–Cys89 and Cys139–Cys149). Sequence comparison with other plant trypsin inhibitors of the Kunitz family reveals that DrTI contains a negatively charged residue (Glu68) at the reactive site rather than the conserved Arg or Lys found in other Kunitz-type trypsin inhibitors. Site-directed mutagenesis yielded five mutants containing substitutions at the reactive site and at one of the disulfide bonds. Assay of the recombinant proteins showed mutant Glu68Leu and Glu68Lys to have only 4–5% of the wild-type activity. These provide evidence that the Glu68 residue is the reactive site for DrTI and various other Kunitz-type trypsin inhibitors. The Cys139Gly mutant lost its inhibitory activity, whereas the Cys44Gly mutant did not, indicating that the second disulfide bond (Cys139–Cys149) is critical to DrTI inhibitory activity, while the first disulfide bond (Cys44–Cys89) is not required.  相似文献   

8.
After isolation of tobacco (Nicotiana tabacum) leaf mitochondria, alternative oxidase (AOX) is predominantly present as the disulfide-linked, less-active “oxidized” form. In an in organello assay, significant AOX activity was dependent upon both the reduction of the regulatory disulfide bond (such as occurs by dithiothreitol) and upon the presence of the activator pyruvate. However, AOX activity in these assays was substantially affected when mitochondria were isolated in the presence of pyruvate. First, pyruvate protects against the oxidation of the regulatory sulfhydryl during isolation, such that subsequent in organello AOX activity is not dependent upon dithiothreitol. Second, pyruvate stabilizes AOX activity, such that mitochondria kept in the presence of pyruvate have higher maximum rates of AOX activity than mitochondria kept for some time in the absence of pyruvate. The ability of pyruvate to protect against AOX oxidation was exploited to assess the in vivo status of the regulatory sulfhydryl/disulfide system. In both tobacco suspension cells and tobacco leaves with high levels of AOX protein, the protein is predominantly present as the “reduced” active form in vivo under a range of respiratory conditions. Experiments also indicate that, while the presence of reduced protein may be a necessary prerequisite for significant AOX activity, it is not sufficient for activity and other factors must also be critical.  相似文献   

9.
The primary structure of human IGF-I, except for the disulfide bond system, has been reported by Rinderknecht and Humbel. IGF-I afforded the corresponding characteristic peptide fragment on V8 protease digestion, which contained Cys6, Cys47, Cys48, and Cys52. Two possible fragments, Type I with Cys6-Cys47 and Cys48-Cys52, and Type II with Cys6-Cys48 and Cys47-Cys52, were synthesized. The disulfide bond system of IGF-I was unequivocally determined to be the Type II form along with Cys18-Cys61. Interestingly, the Type I system was included in the disulfide bond isomer produced as the main by-product in the refolding step on IGF-I synthesis by the recombinant DNA method.  相似文献   

10.
TF (tissue factor) is a transmembrane cofactor that initiates blood coagulation in mammals by binding Factor VIIa to activate Factors X and IX. The cofactor can reside in a cryptic configuration on primary cells and de-encryption may involve a redox change in the C-terminal domain Cys(186)-Cys(209) disulfide bond. The redox potential of the bond, the spacing of the reduced cysteine thiols and their oxidation by TF activators was investigated to test the involvement of the dithiol/disulfide in TF activation. A standard redox potential of -278 mV was determined for the Cys(186)-Cys(209) disulfide of recombinant soluble TF. Notably, ablating the N-terminal domain Cys(49)-Cys(57) disulfide markedly increased the redox potential of the Cys(186)-Cys(209) bond, suggesting that the N-terminal bond may be involved in the regulation of redox activity at the C-terminal bond. Using As(III) and dibromobimane as molecular rulers for closely spaced sulfur atoms, the reduced Cys(186) and Cys(209) sulfurs were found to be within 3-6 ? (1 ?=0.1 nm) of each other, which is close enough to reform the disulfide bond. HgCl2 is a very efficient activator of cellular TF and activating concentrations of HgCl2-mediated oxidation of the reduced Cys(186) and Cys(209) thiols of soluble TF. Moreover, PAO (phenylarsonous acid), which cross-links two cysteine thiols that are in close proximity, and MMTS (methyl methanethiolsulfonate), at concentrations where it oxidizes closely spaced cysteine residues to a cystine residue, were efficient activators of cellular TF. These findings further support a role for Cys(186) and Cys(209) in TF activation.  相似文献   

11.
Six rhodopsin mutants containing disulfide cross-links between different cytoplasmic regions were prepared: disulfide bond 1, between Cys65 (interhelical loop I-II) and Cys316 (end of helix VII); disulfide bond 2, between Cys246 (end of helix VI) and Cys312 (end of helix VII); disulfide bond 3, between Cys139 (end of helix III) and Cys248 (end of helix VI); disulfide bond 4, between Cys139 (end of helix III) and Cys250 (end of helix VI); disulfide bond 5, between Cys135 (end of helix III) and Cys250 (end of helix VI); and disulfide bond 6, between Cys245 (end of helix VI) and Cys338 (C-terminus). The effects of local restrictions caused by the cross-links on transducin (G(T)) activation and phosphorylation by rhodopsin kinase (RK) following illumination were studied. Disulfide bond 1 showed little effect on either G(T) activation or phosphorylation by RK, suggesting that the relative motion between interhelical loop I-II and helix VII is not crucial for recognition by G(T) or by RK. In contrast, disulfide bonds 2-5 abolished both G(T) activation and phosphorylation by RK. Disulfide bond 6 resulted in enhanced G(T) activation but abolished phosphorylation by RK, suggesting the structure recognized by G(T) was stabilized in this mutant by cross-linking of the C-terminus to the cytoplasmic end of helix VI. Thus, the consequences of the disulfide cross-links depended on the location of the restriction. In particular, relative motions of helix VI, with respect to both helices III and VII upon light activation, are required for recognition of rhodopsin by both G(T) and RK. Further, the conformational changes in the cytoplasmic face that are necessary for protein-protein interactions need not be cooperative, and may be segmental.  相似文献   

12.
13.
The effect of decreased protein flexibility on the stability and calcium binding properties of calbindin D9k has been addressed in studies of a disulfide bridged calbindin D9k mutant, denoted (L39C + P43M + I73C), with substitutions Leu 39-->Cys, Ile 73-->Cys, and Pro 43-->Met. Backbone 1H NMR assignments show that the disulfide bond, which forms spontaneously under air oxidation, is well accommodated. The disulfide is inserted on the opposite end of the protein molecule with respect to the calcium sites, to avoid direct interference with these sites, as confirmed by 113Cd NMR. The effect of the disulfide bond on calcium binding was assessed by titrations in the presence of a chromophoric chelator. A small but significant effect on the cooperativity was found, as well as a very modest reduction in calcium affinity. The disulfide bond increases Tm, the transition midpoint of thermal denaturation, of calcium free calbindin D9k from 85 to 95 degrees C and Cm, the urea concentration of half denaturation, from 5.3 to 8.0 M. Calbindins with one covalent bond linking the two EF-hand subdomains are equally stable regardless if the covalent link is the 43-44 peptide bond or the disulfide bond. Kinetic remixing experiments show that separated CNBr fragments of (L39C + P43M + I73C), each comprising one EF-hand, form disulfide linked homodimers. Each homodimer binds two calcium ions with positive co-operativity, and an average affinity of 10(6) M-1. Disulfide linkage dramatically increases the stability of each homodimer. For the homodimer of the C-terminal fragment Tm increases from 59 +/- 2 without covalent linkage to 91 +/- 2 degrees C with disulfide, and Cm from approximately 1.5 to 7.5 M. The overall topology of this homodimer is derived from 1H NMR assignments and a few key NOEs.  相似文献   

14.
Sacred lotus (Nelumbo nucifera) regulates temperature in its floral chamber to 32°C to 35°C across ambient temperatures of 8°C to 40°C with heating achieved through high alternative pathway fluxes. In most alternative oxidase (AOX) isoforms, two cysteine residues, Cys1 and Cys2, are highly conserved and play a role in posttranslational regulation of AOX. Further control occurs via interaction of reduced Cys1 with α-keto acids, such as pyruvate. Here, we report on the in vitro regulation of AOX isolated from thermogenic receptacle tissues of sacred lotus. AOX protein was mostly present in the reduced form, and only a small fraction could be oxidized with diamide. Cyanide-resistant respiration in isolated mitochondria was stimulated 4-fold by succinate but not pyruvate or glyoxylate. Insensitivity of the alternative pathway of respiration to pyruvate and the inability of AOX protein to be oxidized by diamide suggested that AOX in these tissues may lack Cys1. Subsequently, we isolated two novel cDNAs for AOX from thermogenic tissues of sacred lotus, designated as NnAOX1a and NnAOX1b. Deduced amino acid sequences of both confirmed that Cys1 had been replaced by serine; however, Cys2 was present. This contrasts with AOXs from thermogenic Aroids, which contain both Cys1 and Cys2. An additional cysteine was present at position 193 in NnAOX1b. The significance of the sequence data for regulation of the AOX protein in thermogenic sacred lotus is discussed and compared with AOXs from other thermogenic and nonthermogenic species.

Thermogenesis in Sacred Lotus

Sacred lotus (Nelumbo nucifera) is a thermogenic plant that regulates the temperature of its floral chamber between 32°C and 35°C for up to 4 d (Seymour and Schultze-Motel, 1996). Heating of plant tissues has been described as an adaptation to attract insect pollinators either by volatilization of scent compounds (Meeuse, 1975) or by providing a heat reward (Seymour et al., 1983), protect floral parts from low temperatures (Knutson, 1974), or provide the optimum temperature for floral development (Ervik and Barfod, 1999; Seymour et al., 2009). In sacred lotus, heat is produced by high rates of alternative pathway respiration (Watling et al., 2006; Grant et al., 2008); however, the mechanisms of heat regulation, which likely occur at a cellular level, remain unclear.

Alternative Oxidase

Alternative pathway respiration is catalyzed by the alternative oxidase protein (AOX), which acts as a terminal oxidase in the electron transport chain but, unlike the energy conserving cytochrome pathway (COX), complexes III and IV are bypassed and energy is released as heat. Traditionally, AOX activity was measured using oxygen consumption of tissue, cells, or isolated mitochondria in the presence or absence of AOX and COX inhibitors. However, this method does not accurately measure activity in vivo but does indicate the capacity of the alternative pathway (Ribas-Carbo et al., 1995; Day et al., 1996). The only method to date to accurately determine AOX activity, that is, flux of electrons through the AOX pathway in vivo, is to use oxygen isotope discrimination techniques (for review, see Robinson et al., 1995). Determining AOX activity in vivo is important because heat production in plants could be due to activity of either the AOX and/or plant uncoupling proteins. Using oxygen fractionation techniques, we have shown that flux through the AOX pathway is responsible for heating in sacred lotus (Watling et al., 2006; Grant et al., 2008). Furthermore, we were unable to detect any uncoupling protein in these tissues (Grant et al., 2008). AOX protein content within the sacred lotus receptacle increases markedly prior to thermogenesis, but it remains constant during heating (Grant et al., 2008), suggesting that regulation of heating occurs through posttranslational modification of the protein.

Posttranslational Regulation of AOX Protein

The plant AOX is a cyanide-insensitive dimeric protein located in the inner mitochondrial membrane (Day and Wiskich, 1995). The dimer subunits (monomers) can be linked via a noncovalent association (reduced protein) or covalently through the formation of a disulfide bridge (oxidized protein; Umbach and Siedow, 1993). The reduced protein when run on SDS-PAGE has a molecular mass of approximately 30 to 35 kD and the oxidized protein 60 to 71 kD; this holds true for AOX from a number of species, including soybean (Glycine max) roots and cotyledons (Umbach and Siedow, 1993), tobacco (Nicotiana tabacum) leaf (Day and Wiskich, 1995), and the thermogenic spadix of Arum maculatum (Hoefnagel and Wiskich, 1998).Regulation of AOX has been well studied in nonthermogenic plant species, and two mechanisms have been identified. Most AOX isoforms have two highly conserved Cys residues, Cys1 and Cys2 (defined in Berthold et al., 2000 and Holtzapffel et al., 2003), located near the N-terminal hydrophilic domain of the protein. In these isoforms, Cys1 can either be reduced on both subunits of the AOX dimer, or the Cys1 sulfhydryl groups can be oxidized to form a disulfide bridge (Rhoads et al., 1998). Reduction/oxidation modulation of AOX in vitro can be achieved using the sulfhydryl reductant dithiothreitol (DTT) to reduce the protein or diamide to oxidize the Cys residues. The reduced dimer can be further activated via the interaction of Cys1 with α-keto acids, principally pyruvate (Rhoads et al., 1998; see McDonald [2008] for a model of posttranslational regulation of AOX). In addition, Cys2 may also be involved in regulating AOX activity through interaction with the α-keto acid glyoxylate (which can also stimulate activity at Cys1; Umbach et al., 2002).Recently, however, AOX proteins with different regulatory properties have been reported. Naturally occurring AOX proteins without the two regulatory Cys residues have been identified and, along with site-directed mutagenesis studies, used to further elucidate the specific roles of Cys1 and Cys2. The LeAOX1b isoform from tomato (Lycopersicon esculentum), which has a Ser residue at the position of Cys1 and thus does not form disulfide linked dimers, is also activated by succinate rather than pyruvate when expressed in Saccharomyces cerevisiae (Holtzapffel et al., 2003). In Arabidopsis (Arabidopsis thaliana), uncharged or hydrophobic amino acid substitutions of either Cys result in an inactive enzyme, while positively charged substitutions produce an enzyme with higher than wild type basal activity but that is insensitive to pyruvate or succinate (Umbach et al., 2002). Single substitutions at Cys1 or Cys2 have revealed that glyoxylate can activate AOX via both Cys residues, but only one is needed for glyoxylate stimulation (Umbach et al., 2002, 2006). Double substitution mutants were not stimulated by either pyruvate or glyoxylate (Umbach et al., 2006).Previously, we determined that thermogenesis via the AOX pathway in the sacred lotus receptacle is precisely regulated through changes in AOX flux rather than changes to protein content (Grant et al., 2008). In this study, we investigated the nature of this regulation in mitochondria isolated from heating receptacles. Our aim was to elucidate the reduction/oxidation behavior of the AOX protein and the mechanisms of activation of cyanide-resistant respiration in sacred lotus receptacles to provide insights into the mechanism(s) of heat regulation in this species. We further investigated AOX regulation by determining the amino acid sequence of two novel AOX genes isolated from thermogenic receptacle tissue of sacred lotus.  相似文献   

15.
The alpha-ketoglutate (alpha-KG)-dependent dioxygenases are a large class of mononuclear non-heme iron enzymes that require Fe(II), alpha-KG and dioxygen for catalysis, with the alpha-KG cosubstrate supplying the two additional electrons required for dioxygen activation. A sub-class of these enzymes exists in which the alpha-keto acid is covalently attached to the substrate, including (4-hydroxy)mandelate synthase (HmaS) and (4-hydroxyphenyl)pyruvate dioxygenase (HPPD) which utilize the same substrate but exhibit two different general reactivities (H-atom abstraction and electrophilic attack). Previous kinetic studies of Streptomyces avermitilis HPPD have shown that the substrate analog phenylpyruvate (PPA), which only differs from the normal substrate (4-hydroxyphenyl)pyruvate (HPP) by the absence of a para-hydroxyl group on the aromatic ring, does not induce a reaction with dioxygen. While an Fe(IV)O intermediate is proposed to be the reactive species in converting substrate to product, the key step utilizing O(2) to generate this species is the decarboxylation of the alpha-keto acid. It has been generally proposed that the two requirements for decarboxylation are bidentate coordination of the alpha-keto acid to Fe(II) and the presence of a 5C Fe(II) site for the O(2) reaction. Circular dichroism and magnetic circular dichroism studies have been performed and indicate that both enzyme complexes with PPA are similar with bidentate alpha-KG coordination and a 5C Fe(II) site. However, kinetic studies indicate that while HmaS reacts with PPA in a coupled reaction similar to the reaction with HPP, HPPD reacts with PPA in an uncoupled reaction at an approximately 10(5)-fold decreased rate compared to the reaction with HPP. A key difference is spectroscopically observed in the n-->pi( *) transition of the HPPD/Fe(II)/PPA complex which, based upon correlation to density functional theory calculations, is suggested to result from H-bonding between a nearby residue and the carboxylate group of the alpha-keto acid. Such an interaction would disfavor the decarboxylation reaction by stabilizing electron density on the carboxylate group such that the oxidative cleavage to yield CO(2) is disfavored.  相似文献   

16.
Sites for Cys substitutions to form a disulfide bond were chosen in subtilisin E from Bacillus subtilis, a cysteine-free bacterial serine protease, based on the structure of aqualysin I of Thermus aquaticus YT-1 (a thermophilic subtilisin-type protease containing two disulfide bonds). Cys residues were introduced at positions 61 (wild-type, Gly) and 98 (Ser) in subtilisin E by site-directed mutagenesis. The Cys-61/Cys-98 mutant subtilisin appeared to form a disulfide bond spontaneously in the expression system used and showed a catalytic efficiency equivalent to that of the wild-type enzyme for hydrolysis of a synthetic peptide substrate. The thermodynamic characteristics of these enzymes were examined in terms of enzyme autolysis (t1/2) and thermal stability (Tm). The half-life of the Cys-61/Cys-98 mutant was found to be 2-3 times longer than that of the wild-type enzyme. Similar results were obtained by differential scanning calorimetry. The disulfide mutant showed a Tm of 63.0 degrees C, which was 4.5 degrees C higher than that observed for the wild-type enzyme. Under reducing conditions, however, the characteristics of the mutant enzyme were found to revert to those of the wild-type enzyme. These results strongly suggest that the introduction of a disulfide bond by site-directed mutagenesis enhanced the thermostability of subtilisin E without changing the catalytic efficiency of the enzyme.  相似文献   

17.
Chicken cystatin (cC) mutant I66Q is located in the hydrophobic core of the protein and increases the propensity for amyloid formation. Here, we demonstrate that under physiological conditions, the replacement of Ile with the Gln in the I66Q mutant increases the susceptibility for the disulfide bond Cys71–Cys81 to be reduced when compared to the wild type (WT) cC. Molecular dynamics (MD) simulations under conditions favoring cC amyloid fibril formation are in agreement with the experimental results. MD simulations were also performed to investigate the impact of disrupting the Cys71–Cys81 disulfide bond on the conformational stability of cC at the atomic level, and highlighted major disruption to the cC appendant structure. Domain swapping and extensive unfolding has been proposed as one of the possible mechanisms initiating amyloid fibril formation by cystatin. Our in silico studies suggest that disulfide bond formation between residues Cys95 and Cys115 is necessary to maintain conformational stability of the I66Q mutant following breakage of the Cys71–Cys81 disulfide bridge. Subsequent breakage of disulfide bond Cys95–Cys115 resulted in large structural destabilization of the I66Q mutant, which increased the α–β interface distance and expanded the hydrophobic core. These experimental and computational studies provide molecular-level insight into the relationship between disulfide bond formation and progressive unfolding of amyloidogenic cC mutant I66Q.

An animated Interactive 3D Complement (I3DC) is available in Proteopedia at http://proteopedia.org/w/Journal:JBSD:23  相似文献   

18.
An insertion of residues in the third extracellular loop and a disulfide bond linking this loop to the N-terminal domain were identified in a structural model of a G-protein coupled receptor specific to angiotensin II (AT1 receptor), built in homology to the seven-transmembrane-helix bundle of rhodopsin. Both the insertion and the disulfide bond were located close to an extracellular locus, flanked by the second extracellular loop (EC-2), the third extracellular loop (EC-3) and the N-terminal domain of the receptor; they contained residues identified by mutagenesis studies to bind the angiotensin II N-terminal segment (residues D1 and R2). It was postulated that the insertion and the disulfide bond, also found in other receptors such as those for bradykinin, endothelin, purine and other ligands, might play a role in regulating the function of the AT1 receptor. This possibility was investigated by assaying AT1 forms devoid of the insertion and with mutations to Ser on both positions of Cys residues forming the disulfide bond. Binding and activation experiments showed that abolition of this bond led to constitutive activation, decay of agonist binding and receptor activation levels. Furthermore, the receptors thus mutated were translocated to cytosolic environments including those in the nucleus. The receptor form with full deletion of the EC-3 loop residue insertion, displayed a wild type receptor behavior.  相似文献   

19.
We prepared two dissected fragments of hen lysozyme and examined whether or not these two fragments associated to form a native-like structure. One (Fragment I) is the peptide fragment Asn59-homoserine-105 containing Cys64-Cys80 and Cys76-Cys94. The other (Fragment II) is the peptide fragment Lys1-homoserine-58 connected by two disulfide bridges, Cys6-Cys127 and Cys30-Cys115, to the peptide fragment Asn106-Leu129. It was found that the Fragment I immobilized in the cuvette formed an equimolar complex with Fragment II (K(d) = 3.3x10(-4) M at pH 8 and 25 degrees C) by means of surface plasmon resonance. Moreover, from analyses by circular dichroism spectroscopy and ion-exchange chromatography of the mixture of Fragments I and II at pH 8 under non-reducing conditions, it was suggested that these fragments associated to give the native-like structure. However, the mutant Fragment I in which Cys64-Cys80 and Cys76-Cys94 are lacking owing to the mutation of Cys to Ala, or the mutant fragment in which Trp62 is mutated to Gly, did not form the native-like species with Fragment II, because the mutant Fragment I derived from mutant lysozymes had no local conformation due to mutations. Considering our previous results where the preferential oxidation of two inside disulfide bonds, Cys64-Cys80 and Cys76-Cys94, occurred in the refolding of the fully reduced Fragment I, we suggest that the peptide region corresponding to Fragment I is an initiation site for hen lysozyme folding.  相似文献   

20.
Feng YH  Saad Y  Karnik SS 《FEBS letters》2000,484(2):133-138
Dithiothreitol (DTT) treatment of angiotensin II (Ang II) type 2 (AT(2)) receptor potentiates ligand binding, but the underlying mechanism is not known. Two disulfide bonds proposed in the extracellular domain were examined in this report. Based on the analysis of ligand affinity of cysteine (Cys, C) to alanine (Ala, A) substitution mutants, we provide evidence that Cys(35)-Cys(290) and Cys(117)-Cys(195) disulfide bonds are formed in the wild-type AT(2) receptor. Disruption of the highly conserved Cys(117)-Cys(195) disulfide bond linking the second and third extracellular segments leads to inactivation of the receptor. The Cys(35)-Cys(290) bond is highly sensitive to DTT. Its breakage results in an increased binding affinity for both Ang II and the AT(2) receptor-specific antagonist PD123319. Surprisingly, in the single Cys mutants, C35A and C290A, a labile population of receptors is produced which can be re-folded to high-affinity state by DTT treatment. These results suggest that the free -SH group of Cys(35) or Cys(290) competes with the disulfide bond formation between Cys(117) and Cys(195). This Cys-disulfide bond exchange results in production of the inactive population of the mutant receptors through formation of a non-native disulfide bond.  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