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.
Lipoprotein(a), atherosclerosis and thrombosis   总被引:1,自引:0,他引:1  
Lipoprotein(a) denotes cholesterol-rich particles similar to low density lipoproteins but characterized by an extra large hydrophilic glycoprotein, Apo(a), added to low density lipoproteins. Apolipoprotein(a) is bound to ApoB-100 by a disulfide bridge. Eleven different Apo(a) isoforms of varying sizes coded for by alleles at the Apo(a) gene locus on chromosome 6 have been identified, ranging in Mr between roughly 400-800 kDa. The level of lipoprotein(a) is inversely correlated with isoform size. A strong independent association between high lipoprotein(a) levels and atherosclerotic disorders is documented. Lipoprotein(a) is selectively retained in the intima and engulfed by macrophages in unmodified form. Human Apo(a) is very similar to plasminogen, which suggests that lipoprotein(a) represents a link between atherosclerosis and thrombosis.  相似文献   

2.
We attempted to reconcile three microbial maintenance models (Herbert, Pirt, and Compromise) through a theoretical reassessment. We provided a rigorous proof that the true growth yield coefficient (Y(G) ) is the ratio of the specific maintenance rate (a in Herbert) to the maintenance coefficient (m in Pirt). Other findings from this study include: (1) the Compromise model is identical to the Herbert for computing microbial growth and substrate consumption, but it expresses the dependence of maintenance on both microbial biomass and substrate; (2) the maximum specific growth rate in the Herbert (μ(max,H) ) is higher than those in the other two models (μ(max,P) and μ(max,C) ), and the difference is the physiological maintenance factor (m(q) =?a); and (3) the overall maintenance coefficient (m(T) ) is more sensitive to m(q) than to the specific growth rate (μ(G) ) and Y(G) . Our critical reassessment of microbial maintenance provides a new approach for quantifying some important components in soil microbial ecology models.  相似文献   

3.
A series of heterobinuclear cofacial porphyrin-corrole dyads containing a Co(IV) corrole linked by one of four different spacers in a face-to-face arrangement with an Fe(III) or Mn(III) porphyrin have been examined as catalysts for the electroreduction of O(2) to H(2)O and/or H(2)O(2) when adsorbed on the surface of a graphite electrode in air-saturated aqueous solutions containing 1M HClO(4). The examined compounds are represented as (PCY)M(III)ClCo(IV)Cl where P is a porphyrin dianion, C is a corrole trianion and Y is a biphenylene (B), 9,9-dimethylxanthene (X), dibenzofuran (O) or anthracene (A) spacer. The catalytic behavior of the seven investigated dyads in the two heterobimetallic (PCY)MClCoCl series of catalysts is compared on one hand to what was previously reported for related dyads with a single Co(III) corrole macrocycle linked to a free-base porphyrin with the same set of linking bridges, (PCY)H(2)Co, and on the other hand to dicobalt porphyrin-corrole dyads of the form (PCY)Co(2) which were shown to efficiently electrocatalyze the four electron reduction of O(2) at a graphite electrode in acid media. Comparisons between the four series of porphyrin-corrole dyads, (PCY)Co(2), (PCY)H(2)Co, (PCY)FeClCoCl and (PCY)MnClCoCl, show that in all cases the biscobalt dyads catalyze O(2) electroreduction at potentials more positive by an average 110mV as compared to the related series of compounds containing a Co(III) or Co(IV) corrole macrocycle linked to a free-base metalloporphyrin or a metalloporphyrin with an Fe(III) or Mn(III) central metal ion. The data indicates that the E(1/2) values where electrocatalysis is initiated is related to the initial site of electron transfer, which is the Co(III)/Co(II) porphyrin reduction process in the case of (PCY)Co(2) and the Co(IV)/Co(III) corrole reduction in the case of (PCY)MnClCoCl, (PCY)FeClCoCl and (PCY)H(2)Co. The overall data also suggests that the catalytically active form of the biscobalt dyad in (PCY)Co(2) contains a Co(II) porphyrin and a Co(IV) corrole.  相似文献   

4.
Glucose oxidase (GO) and copper amine oxidase (CAO) catalyze the reduction of molecular oxygen to hydrogen peroxide. If a closed-shell cofactor (like FADH(2) in GO and topaquinone (TPQ) in CAO) is electron donor in dioxygen reduction, the formation of a closed-shell species (H(2)O(2)) is a spin forbidden process. Both in GO and CAO, formation of a superoxide ion that leads to the creation of a radical pair is experimentally suggested to be the rate-limiting step in the dioxygen reduction process. The present density functional theory (DFT) studies suggest that in GO, the creation of the radical pair induces a spin transition by spin orbit coupling (SOC) in O(2)(-)(rad), whereas in CAO, it is induced by exchange interaction with the paramagnetic metal ion (Cu(II)). In the rate-limiting step, this spin-transition is suggested to transform the O(2)(-)(rad)-FADH(2)(+)(rad) radical pair in GO and the Cu(II)-TPQ (triplet) species in CAO, from a triplet (T) to a singlet (S) state. For CAO, a mechanism for the O[bond]O cleavage step in the biogenesis of TPQ is also suggested.  相似文献   

5.
6.
Multidrug resistance protein 1 (MRP1) is capable of actively transporting a wide range of conjugated and unconjugated organic anions. The protein can also transport additional conjugated and unconjugated compounds in a GSH- or S-methyl GSH-stimulated manner. How MRP1 binds and transports such structurally diverse substrates is not known. We have used [(3)H]leukotriene C(4) (LTC(4)), a high affinity glutathione-conjugated physiological substrate, to photolabel intact MRP1, as well as fragments of the protein expressed in insect cells. These studies revealed that: (i) LTC(4) labels sites in the NH(2)- and COOH-proximal halves of MRP1, (ii) labeling of the NH(2)-half of MRP1 is localized to a region encompassing membrane-spanning domain (MSD) 2 and nucleotide binding domain (NBD) 1, (iii) labeling of this region is dependent on the presence of all or part of the cytoplasmic loop (CL3) linking MSD1 and MSD2, but not on the presence of MSD1, (iv) labeling of the NH(2)-proximal site is preferentially inhibited by S-methyl GSH, (v) labeling of the COOH-proximal half of the protein occurs in a region encompassing transmembrane helices 14-17 and appears not to require NBD2 or the cytoplasmic COOH-terminal region of the protein, (vi) labeling of intact MRP1 by LTC(4) is strongly attenuated in the presence of ATP and vanadate, and this decrease in labeling is attributable to a marked reduction in LTC(4) binding to the NH(2)-proximal site, and (vii) the attenuation of LTC(4) binding to the NH(2)-proximal site is a consequence of ATP hydrolysis and trapping of Vi-ADP exclusively at NBD2. These data suggest that MRP1-mediated transport involves a conformational change, driven by ATP hydrolysis at NBD2, that alters the affinity with which LTC(4) binds to one of two sites composed, at least in part, of elements in the NH(2)-proximal half of the protein.  相似文献   

7.
Flavin-binding Kelch repeat F-box (FKF1) protein plays important roles in the photoregulation of flowering in Arabidopsis. FKF1 has a light, oxygen, and voltage (LOV) sensing domain binding a flavin mononucleotide (FMN) as a chromophore noncovalently. Photoreaction of the FKF1-LOV polypeptide was studied by low-temperature absorption spectroscopy. Upon blue light irradiation, a ground state, D(450), is converted to S(390) known as a cysteinyl-flavin adduct intermediate in the photoreaction of phototropin. Below 150 K, bleaching of D(450) was much reduced and a new photoproduct, Z(370), appeared as well as S(390) formation. The calculated absorption spectrum for Z(370) is very similar to those of flavoproteins in an anion radical state. On the basis of the results that S(390) formation proceeds to Z(370) formation and that Z(370) formed at low temperatures reverts to D(450) upon temperature increase, Z(370) is concluded to be not an intermediate from D(450) to S(390). Z(370) is suggested to be formed from the biradical triplet-excited state after relaxing to the ground state with the FMN anion radical trapped at the low temperature, in which the SH of the cysteine is in the wrong position that is able to produce a radical pair but unable to form the cysteinyl-flavin adduct. The counter SH in the cationic radical state may revert to the ground state by extracting an electron from the unidentified amino acid residue. Interestingly, S(390) that has been thought to be irreversible to D(450) was revealed to revert to D(450) very slowly with a half-life time of 62.5 h in solution at 298 K. The photoreaction mechanism is discussed in reference to the calculated activation energy of the reaction processes.  相似文献   

8.
Salts inhibit the activity of sweet almond beta-glucosidase. For cations (Cl(-) salts) the effectiveness follows the series: Cu(+2), Fe(+2)>Zn(+2)>Li(+)>Ca(+2)>Mg(+2)>Cs(+)>NH(4)(+)>Rb(+)>K(+)>Na(+) and for anions (Na(+) salts) the series is: I(-)>ClO(4)(-)>(-)SCN>Br(-) approximately NO(3)(-)>Cl(-) approximately (-)OAc>F(-) approximately SO(4)(-2). The activity of the enzyme, like that of most glycohydrolases, depends on a deprotonated carboxylate (nucleophile) and a protonated carboxylic acid for optimal activity. The resulting pH-profile of k(cat)/K(m) for the beta-glucosidase-catalyzed hydrolysis of p-nitrophenyl glucoside is characterized by a width at half height that is strongly sensitive to the nature and concentration of the salt. Most of the inhibition is due to a shift in the enzymic pK(a)s and not to an effect on the pH-independent second-order rate constant, (k(cat)/K(m))(lim). For example, as the NaCl concentration is increased from 0.01 M to 1.0 M the apparent pK(a1)increases (from 3.7 to 4.9) and the apparent pK(a2)decreases (from 7.2 to 5.9). With p-nitrophenyl glucoside, the value of the pH-independent (k(cat)/K(m))(lim) (=9 x 10(4) M(-1) s(-1)) is reduced by less than 4% as the NaCl concentration is increased. There is a similar shift in the pK(a)s when the LiCl concentration is increased to 1.0 M. The results of these salt-induced pK(a) shifts rule out a significant contribution of reverse protonation to the catalytic efficiency of the enzyme. At low salt concentration, the fraction of the catalytically active monoprotonated enzyme in the reverse protonated form (i.e., proton on the group with a pK(a) of 3.7 and dissociated from the group with a pK(a) of 7.2) is very small ( approximately 0.03%). At higher salt concentrations, where the two pK(a)s become closer, the fraction of the monoprotonated enzyme in the reverse protonated form increases over 300-fold. However, there is no increase in the intrinsic reactivity, (k(cat)/K(m))(lim), of the monoprotonated species. For other enzymes which may show such salt-induced pK(a) shifts, this provides a convenient test for the role of reverse protonation.  相似文献   

9.
Meyer T  Kieseritzky G  Knapp EW 《Proteins》2011,79(12):3320-3332
The solvent accessible surface area (SASA) algorithm is conventionally used to characterize protein surfaces in electrostatic energy computations of proteins. Unfortunately, it often fails to find narrow cavities inside a protein. As a consequence pK(a) computations based on this algorithm perform badly. In this study a new cavity-algorithm is introduced, which solves this problem and provides improved pK(a) values. The procedure is applied to 20 pK(a) values of titratable groups introduced as point mutations in SNase variants, where crystal structures are available. The computations of these pK(a)s are particular challenging, since they are placed in a rather hydrophobic environment. For nine mutants, where the titratable residue is in contact with a large cavity, the RMSD(pKa) between computed and measured pK(a) values is 2.04, which is a considerable improvement as compared to the original results obtained with Karlsberg(+) (http://agknapp.chemie.fu-berlin.de/karlsberg/) that yielded an RMSD(pKa) of 8.8. However, for 11 titratable residues the agreement with experiments remains poor (RMSD(pKa) = 6.01). Considering 15 pK(a)s of SNase, which are in a more conventional less hydrophobic protein environment, the RMSD(pKa) is 2.1 using the SASA-algorithm and 1.7 using the new cavity-algorithm. The agreement is reasonable but less good than what one would expect from the general performance of Karlsberg(+) indicating that SNase belongs to the more difficult proteins with respect to pK(a) computations. We discuss the possible reasons for the remaining discrepancies between computed and measured pK(a)s.  相似文献   

10.
Apoptosis is characterized by the proteolytic cleavage of hundreds of proteins. One of them, the type 1 inositol-1,4,5-trisphosphate receptor (IP(3) R-1), a multimeric receptor located on the endoplasmic reticulum (ER) membrane that is critical to calcium homeostasis, was reported to be cleaved during staurosporine (STS) induced-apoptosis in Jurkat cells. Because the reported cleavage site separates the IP(3) binding site from the channel moiety, its cleavage would shut down a critical signaling pathway that is common to several cellular processes. Here we show that IP(3) R-1 is not cleaved in 293 cells treated with STS, TNFα, Trail, or ultra-violet (UV) irradiation. Further, it is not cleaved in Hela or Jurkat cells induced to undergo apoptosis with Trail, TNFα, or UV. In accordance with previous reports, we demonstrate that it is cleaved in a Jurkat cell line treated with STS. However its cleavage occurs only after poly(ADP-ribose) polymerase (PARP), which cleavage is a hallmark of apoptosis, and p23, a poor caspase-7 substrate, are completely cleaved, suggesting that IP(3) R-1 is a relatively late substrate of caspases. Nevertheless, the receptor is fully accessible to proteolysis in cellulo by ectopically overexpressed caspase-7 or by the tobacco etch virus (TEV) protease. Finally, using recombinant caspase-3 and microsomal fractions enriched in IP(3) R-1, we show that the receptor is a poor caspase-3 substrate. Consequently, we conclude that IP(3) R-1 is not a key death substrate.  相似文献   

11.
Histidinol phosphate aminotransferase (HPAT) is a pyridoxal 5'-phosphate (PLP)-dependent aminotransferase classified into Subgroup I aminotransferase, in which aspartate aminotransferase (AspAT) is the prototype. In order to expand our knowledge on the reaction mechanism of Subgroup I aminotransferases, HPAT is an enzyme suitable for detailed mechanistic studies because of having low sequence identity with AspAT and a unique substrate recognition mode. Here we investigated the spectroscopic properties of HPAT and the effect of the C4-C4' strain of the PLP-Lys(214) Schiff base on regulating the Schiff base pK(a) in HPAT. Similar to AspAT, the PLP-form HPAT showed pH-dependent absorption spectral change with maxima at 340 nm at high pH and 420 nm at low pH, having a low pK(a) of 6.6. The pK(a) value of the methylamine-reconstituted K214A mutant enzyme was increased from 6.6 to 10.6. Mutation of Asn(157) to Ala increased the pK(a) to 9.2. Replacement of Arg(335) by Leu increased the pK(a) to 8.6. On the other hand, the pK(a) value of the N157A/R335L double mutant enzyme was 10.6. These data indicate that the strain of the Schiff base is the principal factor to decrease the pK(a) in HPAT and is crucial for the subsequent increase in the Schiff base pK(a) during catalysis, although the electrostatic effect of the arginine residue that binds the negatively charged group of the substrate is larger in HPAT than that in AspAT. Our findings also support the idea that the strain mechanism is common to Subgroup I aminotransferases.  相似文献   

12.
The purpose of this paper is to investigate the theoretical basis for the pressure-distension behavior of the urinary bladder. A finite strain theory is developed for hollow spherical structures and it is shown that the Treloar model is a good prototype only for rubber balloons. The pressure-extension ratio relationship is inverted to lead a general form of strain energy function, and fitted by an empirical relation involving one exponential. The following form of strain energy function is derived: W(lambda, lambda, lambda -2) = C1 (P(1), a) + P(1)C2 (a, lambda)ea(lambda -1). Where C1(P(1), a) is a constant (N m-2), P(1) is the initial pressure, a is the rate of pressure increase and C2 (a, lambda) a third degree polynomial relation. P(1) and a are experimentally determined through volumetric pressure-distension data. It is verified that this type of energy function is also valid for uniaxial loading experiments by testing strips coming from the same bladder for which P(1) and a were computed. There is a good agreement between the experimental points and the theoretical stress-strain relation. Finally, the strain energy function is plotted as a function of the first strain invariant and appears to be of an exponential nature.  相似文献   

13.
Membrane-bound heme-copper oxidases catalyze the reduction of O(2) to water. Part of the free energy associated with this process is used to pump protons across the membrane. The O(2) reduction reaction results in formation of high-pK(a) protonatable groups at the catalytic site. The free energy associated with protonation of these groups is used for proton pumping. One of these protonatable groups is OH(-), coordinated to the heme and Cu(B) at the catalytic site. Here we present results from EPR experiments on the Rhodobacter sphaeroides cytochrome c oxidase, which show that at high pH (9) approximately 50% of oxidized heme a(3) is hydroxide-ligated, while at low pH (6.5), no hydroxide is bound to heme a(3). The kinetics of hydroxide binding to heme a(3) were investigated after dissociation of CO from heme a(3) in the enzyme in which the heme a(3)-Cu(B) center was reduced while the remaining redox sites were oxidized. The dissociation of CO results in a decrease of the midpoint potential of heme a(3), which results in electron transfer (tau approximately equal 3 micros) from heme a(3) to heme a in approximately 100% of the enzyme population. At pH >7.5, the electron transfer is followed by proton release from a H(2)O molecule to the bulk solution (tau approximately equal 2 ms at pH 9). This reaction is also associated with absorbance changes of heme a(3), which on the basis of the results from the EPR experiments are attributed to formation of hydroxide-ligated heme a(3). The OH(-) bound to heme a(3) under equilibrium conditions at high pH is also formed transiently after O(2) reduction at low pH. It is proposed that the free energy associated with electron transfer to the binuclear center and protonation of this OH(-) upon reduction of the recently oxidized enzyme provides the driving force for the pumping of one proton.  相似文献   

14.
15.
Rhesus monkey apolipoprotein(a). Sequence, evolution, and sites of synthesis   总被引:11,自引:0,他引:11  
Human lipoprotein(a) is a low density lipoprotein-like lipoprotein whose concentration in plasma is correlated with atherosclerosis. The characteristic protein component of lipoprotein(a) is apolipoprotein(a) (apo(a)) which is disulfide-linked to apolipoprotein B-100. Sequencing of rhesus monkey apo(a) cDNA suggests that this protein, like human apo(a), is highly similar to plasminogen. Sequence data suggests that a plasminogen-like protease activity and kringle 1-, 2-, 3-, and 5-like domains are unnecessary for apo(a) function, but a highly repeated kringle four-like domain is important. Liver is the major site of apo(a) RNA synthesis; reduced amounts of message were also found in testes and brain. Co-expression with apoB-100 and plasminogen in rhesus tissues is not mandatory.  相似文献   

16.
Arginine 54 in subunit I of cytochrome c oxidase from Paracoccus denitrificans interacts with the formyl group of heme a. Mutation of this arginine to methionine (R54M) dramatically changes the spectral properties of heme a and lowers its midpoint redox potential [Kannt et al. (1999) J. Biol. Chem. 274, 37974-37981; Lee et al. (2000) Biochemistry 39, 2989-2996; Riistama et al. (2000) Biochim. Biophys. Acta 1456, 1-4]. During anaerobic reduction of the mutant enzyme, a small fraction of heme a is reduced first along with heme a(3), while most of heme a is reduced later. This suggests that electron transfer is impaired thermodynamically due to the low redox potential of heme a but that it still takes place from Cu(A) via heme a to the binuclear site as in wild-type enzyme, with no detectable bypass from Cu(A) directly to the binuclear site. Consistent with this, the proton translocation efficiency is unaffected at 1 H(+)/e(-) in the mutant enzyme, although turnover is strongly inhibited. Time-resolved electrometry shows that when the fully reduced enzyme reacts with O(2), the fast phase of membrane potential generation during the P(R )()--> F transition is unaffected by the mutation, whereas the slow phase (F --> O transition) is strongly decelerated. In the 3e(-)-reduced mutant enzyme heme a remains oxidized due to its lowered midpoint potential, whereas Cu(A) and the binuclear site are reduced. In this case the reaction with O(2) proceeds via the P(M) state because transfer of the electron from Cu(A) to the binuclear site is delayed. The single phase of membrane potential generation in the 3e(-)-reduced mutant enzyme, which thus corresponds to the P(M)--> F transition, is decelerated, but its amplitude is comparable to that of the P(R)--> F transition. From this we conclude that the completely (4e(-)) reduced enzyme is fully capable of proton translocation.  相似文献   

17.
Eukaryotic cytochrome c oxidase (CcO) and homologous prokaryotic forms of Rhodobacter and Paraccocus differ in the EPR spectrum of heme a. It was noted that a histidine ligand of heme a (H102) is hydrogen bonded to serine in Rhodobacter (S44) and Paraccocus CcOs, in contrast to glycine in the bovine enzyme. Mutation of S44 to glycine shifts the heme a EPR signal from g(z) = 2.82 to 2.86, closer to bovine heme a at 3.03, without modifying other properties. Mutation to aspartate, however, results in an oppositely shifted and split heme a EPR signal of g(z) = 2.72/2.78, accompanied by lower activity and drastically inhibited intrinsic electron transfer from CuA to heme a. This intrinsic rate is biphasic; the proportion that is slow is pH dependent, as is the relative intensity of the two EPR signal components. At pH 8, the heme a EPR signal at 2.72 is most intense, and the electron transfer rate (CuA to heme a) is 10-130 s(-1), compared to wild-type at 90,000 s(-1). At pH 5.5, the signal at 2.78 is intensified, and a biphasic rate is observed, 50% fast (approximately wild type) and 50% slow (90 s(-1)). The data support the prediction that the hydrogen-bonding partner of the histidine ligand of heme a is one determinant of the EPR spectral difference between bovine and bacterial CcO. We further demonstrate that the heme a redox potential can be dramatically altered by a nearby carboxyl, whose protonation leads to a proton-coupled electron transfer process.  相似文献   

18.
The last biosynthetic step for 2-methylthio-N(6)-isopentenyl-adenosine (ms(2)i(6)A), present at position 37 in some tRNAs, consists of the methylthiolation of the isopentenyl-adenosine (i(6)A) precursor. In this work we have reconstituted in vitro the conversion of i(6)A to ms(2)i(6)A within a tRNA substrate using the iron-sulfur MiaB protein, S-adenosylmethionine (AdoMet), and a reducing agent. We show that a synthetic i(6)A-containing RNA corresponding to the anticodon stem loop of tRNA(Phe) is also a substrate. This study demonstrates that MiaB protein is a bifunctional system, involved in both thiolation and methylation of i(6)A. In this process, one molecule of AdoMet is converted to 5'-deoxyadenosine, probably through reductive cleavage and intermediate formation ofa5'-deoxyadenosyl radical as observed in other "Radical-AdoMet" enzymes, and a second molecule of AdoMet is used as a methyl donor as shown by labeling experiments. The origin of the sulfur atom is discussed.  相似文献   

19.
Around 1970, the author proposed a general theoretical approach to multiple decision problems (MDPs) of which multiple comparison problems (MCPs) are special cases. Suppose that a sample space Χ is given together with a set of probability distributions P = {P(θ), θ ∈ Ω} defined over Χ. Let a finite partition of the parameter space Ω = cupω(a), a ∈ A be given. Based on the observation X ∈ X, an MDP is to decide, which ω(a) the true parameter θ belongs to. An MD confidence procedure is a mapping ψ from X to the class of subsets of A, such that the probability that cupω(a), ω(a) C ψ(X) includes the true parameter θ is not smaller than 1-α(θ) . Here, 1-α(θ) is called the level of the confidence procedures and may vary depending on θ∈ω(a) . The MP confidence procedures are derived from the following proposition. When the ω(a) 's are mutually disjoint, there is a one-to-one correspondence between an MD confidence procedure ψ and a collection of (non-randomized) tests for the hypotheses H(a)?: θ∈ω(a) with level α(a) by rejecting the hypothesis H(a) if ω(a) ? ψ(X). In this paper we discuss in detail the problems of determining the signs or the orderings of normal means. The resulting confidence procedures from the LR tests are seen to be too complicated and difficult to understand. We therefore propose simplified, less powerful methods. We define an overlapping partition of Ω into simple sets, such that the original ω(a) 's can be expressed as an intersection of such simple sets. For each such set we define rejection regions corresponding to the levels α, α/2,...,α/k. Then we obtain the acceptance regions for H(a) :?θ∈ω(a) given as the intersection of all acceptance regions for the simple sets containing ω(a) at the level α/k, if there are k such simple sets. This method can be extended to obtain sequential confidence procedures.  相似文献   

20.
For a long time, NMR chemical shifts have been used to identify protein secondary structures. Currently, this is accomplished through comparing the observed (1)H(alpha), (13)C(alpha), (13)C(beta), or (13)C' chemical shifts with the random coil values. Here, we present a new protocol, which is based on the joint probability of each of the three secondary structural types (beta-strand, alpha-helix, and random coil) derived from chemical-shift data, to identify the secondary structure. In combination with empirical smooth filters/functions, this protocol shows significant improvements in the accuracy and the confidence of identification. Updated chemical-shift statistics are reported, on the basis of which the reliability of using chemical shift to identify protein secondary structure is evaluated for each nucleus. The reliability varies greatly among the 20 amino acids, but, on average, is in the order of: (13)C(alpha)>(13)C'>(1)H(alpha)>(13)C(beta)>(15)N>(1)H(N) to distinguish an alpha-helix from a random coil; and (1)H(alpha)>(13)C(beta) >(1)H(N) approximately (13)C(alpha) approximately (13)C' approximately (15)N for a beta-strand from a random coil. Amide (15)N and (1)H(N) chemical shifts, which are generally excluded from the application, in fact, were found to be helpful in distinguishing a beta-strand from a random coil. In addition, the chemical-shift statistical data are compared with those reported previously, and the results are discussed. A JAVA User Interface program has been developed to make the entire procedure fully automated and is available via http://ccsr3150-p3.stanford.edu.  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