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.
In a metabolic study of [1-(14)C]geranylgeranial involving rat thymocytes, the radioactivity was mainly incorporated into two metabolites, Z1 and Z2, the latter moving slower than the former on a silica-gel thin-layer plate. The time course of Z1 and Z2 formation superficially suggested a precursor-product relationship between Z1 and Z2. The two metabolites were chemically converted to their methyl esters on treatment with trimethylsilyl diazomethane. Z1 was cochromatographed with E,E,E-geranylgeranoic acid (GGA). Z2 was prepared in a large quantity from geranylgeranial using thymocytes, and purified by TLC followed by ESI (negative ion mode) or EI mass-spectrometry. The observation of a negative ion at m/z 305 on ESI and a molecular ion at m/z 306 (C(20)H(34)O(2)) with fragments similar to GGA on EI implied that Z2 was dihydroGGA, which has been detected in the urine and serum of patients with Refsum disease. The EI mass spectrum of (R)-2,3-dihydroGGA was identical to that of Z2. The diastereomeric amide synthesized from metabolite Z2 with (R)-1-(1-naphtyl)ethylamine was cochromatographed with (R acid, R) amide, not with (S acid, R) amide, which were similarly synthesized from (R)- and (S)-2,3-dihydroGGAs, respectively. In another metabolic study on [1-(14)C]geranylgeraniol (GGOH), the radioactivity was similarly incorporated into a metabolite corresponding to (R)-2,3-dihydroGGA. (R)-2,3-DihydroGGA induced DNA ladder formation with a maximum at 15 mciroM in thymocytes. However, 2,3-dihydrofarnesoic acid did not induce it at all.  相似文献   

2.
A high affinity monoclonal antibody, 8G2 B9, was used to assess the expression of a 72,000 m.w. glycoprotein ( GP72 ) in isoenzyme-typed T. cruzi strains ( zymodemes ). Western blotting analysis of T. cruzi clones showed that 8G2 B9 bound strongly to GP72 and also suggested that this antigen was absent or weakly detectable in T. cruzi zymodeme 1 (Z1) strains. Purified 8G2 B9 was radiolabeled with 125I and used in an inhibition radioimmune binding assay to compare the quantities of GP72 in different zymodemes . Ninety-six T. cruzi strains were assayed, of which 36 were Z1, 36 were Z2, five were Z3 , and 19 were Z2 (heterozygous). Most (64%) Z1 strains lacked detectable GP72 , whereas this antigen was always detected Z2 and Z2 (heterozygous) strains. There was an 18-fold difference between geometric mean values for the quantities of GP72 (expressed as nanograms per milligram total cell protein) in Z1 and Z2 strains (Z1, 36 ng/mg; Z2, 639 ng/mg; p less than 0.001). There were also significant differences between the geometric mean values for Z2 and Z2 (heterozygous) strains, i.e., 639 ng/mg and 1648 ng/mg, respectively (p less than 0.001). GP72 was detected in four of the Z3 strains in quantities ranging from 740 to 3640 ng/mg. The absolute amounts of antigen in GP72 -positive strains were low, comprising less than 1% of the total cell protein. The specificities of two other anti- GP72 monoclonal antibodies, 7C6 D7 and WIC 29.26, were compared with 8G2 B9. Both antibodies completely inhibited the binding of 8G2 B9 to GP72 in solid phase immunoassays, suggesting that they reacted with the same antigenic determinants. The results show that monoclonal antibody-based assessments of the expression of GP72 correlate with zymodeme classification, and they also suggest that the monoclonal antibodies recognize major antigenic determinants on GP72 . It should be possible to use 8G2 B9 as an immunologic marker to additionally investigate the clinical significance of T. cruzi zymodemes and the biologic significance of GP72 .  相似文献   

3.
Water oxidation at photosystem II Mn-cluster is mediated by the redox-active tyrosine Y(Z). We calculated the redox potential (E(m)) of Y(Z) and its symmetrical counterpart Y(D), by solving the linearized Poisson-Boltzmann equation. The calculated E(m)(Y( )/Y(-)) were +926 mV/+694 mV for Y(Z)/Y(D) with the Mn-cluster in S2 state. Together with the asymmetric position of the Mn-cluster relative to Y(Z/D), differences in H-bond network between Y(Z) (Y(Z)/D1-His(190)/D1-Asn(298)) and Y(D) (Y(D)/D2-His(189)/D2-Arg(294)/CP47-Glu(364)) are crucial for E(m)(Y(Z/D)). When D1-His(190) is protonated, corresponding to a thermally activated state, the calculated E(m)(Y(Z)) was +1216 mV, which is as high as the E(m) for P(D1/D2). We observed deprotonation at CP43-Arg(357) upon S-state transition, which may suggest its involvement in the proton exit pathway. E(m)(Y(D)) was affected by formation of P(D2)(+) (but not P(D1)(+)) and sensitive to the protonation state of D2-Arg(180). This points to an electrostatic link between Y(D) and P(D2).  相似文献   

4.
Affibody molecules constitute a class of engineered binding proteins based on the 58-residue three-helix bundle Z domain derived from staphylococcal protein A (SPA). Affibody proteins are selected as binders to target proteins by phage display of combinatorial libraries in which typically 13 side-chains on the surface of helices 1 and 2 in the Z domain have been randomized. The Z(Taq):anti-Z(Taq) affibody-affibody complex, consisting of Z(Taq), originally selected as a binder to Taq DNA polymerase, and anti-Z(Taq), selected as binder to Z(Taq), is formed with a dissociation constant K(d) approximately 100 nM. We have determined high-precision solution structures of free Z(Taq) and anti-Z(Taq), and the Z(Taq):anti-Z(Taq) complex under identical experimental conditions (25 degrees C in 50 mM NaCl with 20 mM potassium phosphate buffer at pH 6.4). The complex is formed with helices 1 and 2 of anti-Z(Taq) in perpendicular contact with helices 1 and 2 of Z(Taq). The interaction surface is large ( approximately 1670 A(2)) and unusually non-polar (70 %) compared to other protein-protein complexes. It involves all varied residues on anti-Z(Taq), most corresponding (Taq DNA polymerase binding) side-chains on Z(Taq), and several additional side-chain and backbone contacts. Other notable features include a substantial rearrangement (induced fit) of aromatic side-chains in Z(Taq) upon binding, a close contact between glycine residues in the two subunits that might involve aliphatic glycine Halpha to backbone carbonyl hydrogen bonds, and four hydrogen bonds made by the two guanidinium N(eta)H(2) groups of an arginine side-chain. Comparisons of the present structure with other data for affibody proteins and the Z domain suggest that intrinsic binding properties of the originating SPA surface might be inherited by the affibody binders. A thermodynamic characterization of Z(Taq) and anti-Z(Taq) is presented in an accompanying paper.  相似文献   

5.
Geijer P  Morvaridi F  Styring S 《Biochemistry》2001,40(36):10881-10891
Here we report an EPR signal that is induced by a pH jump to alkaline pH in the S(3) state of the oxygen-evolving complex in photosystem II. The S(3) state is first formed with two flashes at pH 6. Thereafter, the pH is changed in the dark prior to freezing of the sample. The EPR signal is 90-100 G wide and centered around g = 2. The signal is reversibly induced with a pK = 8.5 +/- 0.3 and is very stable with a decay half-time of 5-6 min. If the pH is changed in the dark from pH 8.6 to 6.0, the signal disappears although the S(3) state remains. We propose that the signal arises from the interaction between the Mn cluster and Y(Z), resulting in the spin-coupled S(2)Y(Z)(*) signal. Our data suggest that the potential of the Y(Z)(*)/Y(Z) redox couple is sensitive to the ambient pH in the S(3) state. The alkaline pH decreases the potential of the Y(Z)(*)/Y(Z) couple so that Y(Z) can give back an electron to the S(3) state, thereby obtaining the S(2)Y(Z)(*) EPR signal. The tyrosine oxidation also involves proton release from Y(Z), and the results support a mechanism where this proton is released to the bulk medium presumably via a close-lying base. Thus, the equilibrium is changed from S(3)Y(Z) to S(2)Y(Z)(*) by the alkaline pH. At normal pH (pH 5.5-7), this equilibrium is set strongly to the S(3)Y(Z) state. The results are discussed in relation to the present models of water oxidation. Consequences for the relative redox potentials of Y(Z)(*)/Y(Z) and S(3)/S(2) at different pH values are discussed. We also compare the pH-induced S(2)Y(Z)(*) signal with the S(2)Y(Z)(*) signal from Ca(2+)-depleted photosystem II.  相似文献   

6.
A pathogen-inducible oxygenase in tobacco leaves and a homologous enzyme from Arabidopsis were recently characterized (Sanz, A., Moreno, J. I., and Castresana, C. (1998) Plant Cell 10, 1523-1537). Linolenic acid incubated at 23 degrees C with preparations containing the recombinant enzymes underwent alpha-oxidation with the formation of a chain-shortened aldehyde, i.e., 8(Z),11(Z), 14(Z)-heptadecatrienal (83%), an alpha-hydroxy acid, 2(R)-hydroxy-9(Z),12(Z),15(Z)-octadecatrienoic acid (15%), and a chain-shortened fatty acid, 8(Z),11(Z),14(Z)-heptadecatrienoic acid (2%). When incubations were performed at 0 degrees C, 2(R)-hydroperoxy-9(Z),12(Z),15(Z)-octadecatrienoic acid was obtained as the main product. An intermediary role of 2(R)-hydroperoxy-9(Z), 12(Z),15(Z)-octadecatrienoic acid in alpha-oxidation was demonstrated by re-incubation experiments, in which the hydroperoxide was converted into the same alpha-oxidation products as those formed from linolenic acid. 2(R)-Hydroperoxy-9(Z),12(Z), 15(Z)-octadecatrienoic acid was chemically unstable and had a half-life time in buffer of about 30 min at 23 degrees C. Extracts of cells expressing the recombinant oxygenases accelerated breakdown of the hydroperoxide (half-life time, about 3 min at 23 degrees C), however, this was not attributable to the recombinant enzymes since the same rate of hydroperoxide degradation was observed in the presence of control cells not expressing the enzymes. No significant discrimination between enantiomers was observed in the degradation of 2(R,S)-hydroperoxy-9(Z)-octadecenoic acid in the presence of recombinant oxygenases. A previously studied system for alpha-oxidation in cucumber was re-examined using the newly developed techniques and was found to catalyze the same conversions as those observed with the recombinant enzymes, i.e. enzymatic alpha-dioxygenation of fatty acids into 2(R)-hydroperoxides and a first order, non-stereoselective degradation of hydroperoxides into alpha-oxidation products. It was concluded that the recombinant enzymes from tobacco and Arabidopsis were both alpha-dioxygenases, and that members of this new class of enzymes catalyze the first step of alpha-oxidation in plant tissue.  相似文献   

7.
Loricariid catfishes show a predominance of homomorphism in sex chromosomes, but cases of simple and multiple systems were also found. Here we describe two cases of multiple sex chromosome systems in loricariids from Brazilian Amazonia. Males of Ancistrus sp.1 "Balbina" have a modal number of 2n = 39 chromosomes, fundamental number (FN) of 78, and karyotypic formula of 27 m + 10 sm + 2 st; females have 2n = 38 chromosomes, FN = 76, and 26 m + 10 sm + 2 st. Ancistrus sp.2 "Barcelos" has 2n = 52 chromosomes for both sexes, FN = 80 for males and FN = 79 for females. Karyotypic formula is 12 m + 12 sm + 4 st + 24a for males and 11 m + 12 sm + 4st + 25a for females. The two species show different arrangements of constitutive heterochromatin blocks, which are coincident with NORs and absent in sex chromosomes. We suggest a XX/XY(1)Y(2) mechanism for Ancistrus sp.1 "Balbina", and a Z(1)Z(1)Z(2)Z(2)/Z(1)Z(2)W(1)W(2) mechanism for Ancistrus sp.2 "Barcelos". The XX/XY(1)Y(2) mechanism here reported is the second known occurrence of this type of multiple sex chromosomes for Loricariidae and the third for Neotropical fishes; the mechanism Z(1)Z(1)Z(2)Z(2)/Z(1)Z(2)W(1)W(2) represents the first record among fishes. The presence of different sex chromosome systems in Ancistrus indicates a probable independent origin and suggests that the differentiation of sex chromosomes is evolutionarily recent among species in this genus.  相似文献   

8.
Analysis of the grass Zingeria trichopoda (2n = 8, 2C = 5.3 pg) revealed a dynamic evolution with the following characteristics. (i) Genomic in situ hybridization (GISH) demonstrates that Z. trichopoda evolved from an interspecific hybrid involving a species like contemporary Zingeria biebersteiniana (2n = 4) and a second species with a similar low number of chromosomes. The nucleus of Z. trichopoda is spatially organized at the genome level and the two parental genomes occupy distinct and separate domains of lateral arrangements. (ii) The copy number of the Z. biebersteiniana specific pericentromeric tandem repeat family Zbcen1 is drastically reduced in Z. trichopoda. (iii) GISH in combination with labeled rDNA sequences simultaneously discriminated the two parental genomes and the corresponding 5S and 45S rDNA sites. Hence, following allopolyploidization of Z. trichopoda the Z. biebersteiniana like parental chromosomes probably underwent drastic loss of 45S rDNA. This could have arisen either through the loss of Z. biebersteiniana derived 45S rDNA or through Z. trichopoda genome-wide homogenization of Z. biebersteiniana type 45S rDNA and subsequent elimination of 45S rDNA loci from Z. biebersteiniana derived chromosomes. Finally, 5S rDNA loci are present in both subgenomes of Z. trichopoda and the chromosomal position of these loci is similar for both Z. biebersteiniana and the Z. biebersteiniana like parental genome of Z. trichopoda.  相似文献   

9.
Wang F  Li F  Ganguly M  Marky LA  Gold B  Egli M  Stone MP 《Biochemistry》2008,47(27):7147-7157
Site-specific insertion of 5-(3-aminopropyl)-2'-deoxyuridine (Z3dU) and 7-deaza-dG into the Dickerson-Drew dodecamers 5'-d(C (1)G (2)C (3)G (4)A (5)A (6)T (7)T (8)C (9) Z (10)C (11)G (12))-3'.5'-d(C (13)G (14)C (15)G (16)A (17)A (18)T (19)T (20)C (21) Z (22)C (23)G (24))-3' (named DDD (Z10)) and 5'-d(C (1)G (2)C (3)G (4)A (5)A (6)T (7) X (8)C (9) Z (10)C (11)G (12))-3'.5'-d(C (13)G (14)C (15)G (16)A (17)A (18)T (19) X (20)C (21) Z (22)C (23)G (24))-3' (named DDD (2+Z10)) (X = Z3dU; Z = 7-deaza-dG) suggests a mechanism underlying the formation of interstrand N+2 DNA cross-links by nitrogen mustards, e.g., melphalan and mechlorethamine. Analysis of the DDD (2+Z10) duplex reveals that the tethered cations at base pairs A (5).X (20) and X (8).A (17) extend within the major groove in the 3'-direction, toward conserved Mg (2+) binding sites located adjacent to N+2 base pairs C (3).Z (22) and Z (10).C (15). Bridging waters located between the tethered amines and either Z (10) or Z (22) O (6) stabilize the tethered cations and allow interactions with the N + 2 base pairs without DNA bending. Incorporation of 7-deaza-dG into the DDD (2+Z10) duplex weakens but does not eliminate electrostatic interactions between tethered amines and Z (10) O (6) and Z (22) O (6). The results suggest a mechanism by which tethered N7-dG aziridinium ions, the active species involved in formation of interstrand 5'-GNC-3' cross-links by nitrogen mustards, modify the electrostatics of the major groove and position the aziridinium ions proximate to the major groove edge of the N+2 C.G base pair, facilitating interstrand cross-linking.  相似文献   

10.

Background  

APOBEC3 (A3) proteins deaminate DNA cytosines and block the replication of retroviruses and retrotransposons. Each A3 gene encodes a protein with one or two conserved zinc-coordinating motifs (Z1, Z2 or Z3). The presence of one A3 gene in mice (Z2–Z3) and seven in humans, A3A-H (Z1a, Z2a-Z1b, Z2b, Z2c-Z2d, Z2e-Z2f, Z2g-Z1c, Z3), suggests extraordinary evolutionary flexibility. To gain insights into the mechanism and timing of A3 gene expansion and into the functional modularity of these genes, we analyzed the genomic sequences, expressed cDNAs and activities of the full A3 repertoire of three artiodactyl lineages: sheep, cattle and pigs.  相似文献   

11.
Telethonin interacts specifically with the two Z-disk IG-like domains (Z1Z2) at the N terminus of titin, the largest presently known protein. Analytical ultracentrifugation and synchrotron radiation x-ray scattering were employed to study the solution structures of Z1Z2 and its complexes with telethonin, and low resolution models were constructed ab initio from the scattering data. A seven residues-long polyhistidine tag was localized at the tip of the Z1 domain by comparison of independent models of native and His-tagged versions of Z1Z2. The stoichiometry and shape of the complex between the telethonin construct lacking the C terminus and Z1Z2 indicate antiparallel association of two Z1Z2 molecules with telethonin acting as a central linker. The complex of full-length telethonin with Z1Z2 appears to also have a 1:2 stoichiometry at concentrations below 1 mg/ml but dimerizes at higher concentrations. These results suggest a possible role of telethonin in linking titin filaments at the Z-disk periphery.  相似文献   

12.
The Y(Z)-tyrosine radical was trapped by freezing immediately after illumination in Ca(2+)-depleted Photosystem II (PS II) membranes and the pH-dependent characteristics of the radical were investigated using CW-EPR and pulsed ENDOR. The spectrum of the Y*(Z) radical trapped in the Y*(Z)S(1) state at pH 5.5 was cation-like as reported in Mn-depleted PS II (H. Mino et al., Spectrochim. Acta A 53 (1997) 1465-1483). By illuminating the PS II-retaining S(2) state, the Y*(Z) radical and a broad doublet signal formed in the g approximately 2 region were trapped concomitantly. The spectrum of the trapped Y*(Z) radical in the Y*(Z)S(2) state was cation-like at pH 5.5 but the pulsed ENDOR measurements reveals the involvement of the neutral Y*(Z) radical in the doublet signal. At pH 7.0, the resulting Y*(Z) signal was the mixture of the cation-like and neutral radical spectra, and considerably different from the neutral radical found in Mn-depleted PS II. pH-Dependent changes in the properties of the Y*(Z) radical are discussed in relation to the redox events occurring in Ca(2+)-depleted PS II.  相似文献   

13.
Electron paramagnetic resonance (EPR) and electron nuclear double resonance (ENDOR) were performed to investigate the difference in microenvironments and functions between tyrosine Z (Y(Z)) and tyrosine D (Y(D)). Mn-depletion or Ca(2+)-depletion causes extension of the lifetime of tyrosine radical Y(Z)(*), which can be trapped by rapid freezing after illumination at about 250 K. Above pH 6.5, Y(Z)(*) radical in Mn-depleted PS II shows similar EPR and ENDOR spectra similar to that of Y(D)(*) radical, which are ascribed to a typical neutral tyrosine radical. Below pH 6.5, Y(Z)(*) radical shows quite different EPR and ENDOR spectra. ENDOR spectra show the spin density distribution of the low-pH form of Y(Z)(*) that has been quite different from the high-pH form of Y(Z)(*). The spin density distribution of the low-pH Y(Z)(*) can be explained by a cation radical or the neutral radical induced by strong electrostatic interaction. The pH dependence of the activation energy of the recombination rate between Y(Z)(*) and Q(A)(-) shows a gap of 4.4 kJ/mol at pH 6.0-6.5. In the Ca(2+)-depleted PS II, Y(Z)(*) signal was the mixture of the cation-like and normal neutral radicals, and the pH dependence of Y(Z)(*) spectrum in Ca(2+)-depleted PS II is considerably different from the neutral radical found in Mn-depleted PS II. Based on the recent structure data of cyanobacterial PS II, the pH dependence of Y(Z)(*) could be ascribed to the modification of the local structure and hydrogen-bonding network induced by the dissociation of ASP170 near Y(Z).  相似文献   

14.
The delta 5,9 fatty acids (5Z,9Z)-5,9-hexadecadienoic acid, (5Z,9Z)-5,9-nonadecadienoic acid, and (5Z,9Z)-5,9-eicosadienoic acid were synthesized for the first time in four steps (9-12% overall yield) starting from commercially available 2-(2-bromoethyl)-1,3-dioxolane. The synthetic approach provided enough material to corroborate the structure and stereochemistry of (5Z,9Z)-5,9-nonadecadienoic acid which was recently identified in the flowers of Malvaviscus arboreus (Malvaceae). The novel phospholipids 1-hexadecanoyl-2-[(5Z,9Z)-5,9-eicosadienoyl]-sn-glycer o-3-phosphocholine and 1-octadecanoyl-2-[(5Z,9Z)-5,9-eicosadienoyl]-sn- glycero-3-phosphocholine were also synthesized from commercially available L-alpha-phosphatidylcholine (egg yolk) and characterized by positive ion electrospray mass spectrometry. These are the first examples of unsymmetrical phospholipids with saturated fatty acids at the sn-1 position and delta 5,9 fatty acids at the sn-2 position.  相似文献   

15.
Zbed3, a BED finger domain-containing protein was found to promote cancer proliferation by regulating β-catenin expression through interacting with Axin. But whether and how BED finger domain function in regulating cancer proliferation is unknown. We constructed five mutants of Zbed3, which lacks the Axin-Zbed3 binding site, and the 43 to 52, 69 to 77, 87 to 92, and 97 to 104 sequences in BED finger domain, respectively and named them as Z-A, Z1, Z2, Z3, and Z4. Transfection of both wild-type of Zbed3 and the mutants Z1, Z3, and Z4 (P < 0.05), but not Z2 (P > 0.05) significantly upregulated β-catenin expression in NCI-H1299 cells. Overexpression of both wild-type of Zbed3 and the mutants Z1, Z3, and Z4 (P < 0.05) but not Z2 (P > 0.05) significantly promoted cancer cell proliferation and invasion. The ability of proliferation (P < 0.05) but not invasion (P < 0.05) of cancer cells transfected with Z1 and Z4 was significantly lower than that with wild-type Zbed3 and Z3. Overexpression of wild-type Zbed3 (P < 0.05) but not the mutant Z-A, which lacks the binding site with Axin and Z2 (P > 0.05) significantly upregulated the interaction of Axin and Zbed3, β-catenin expression and the activity of Wnt signaling. Both overexpression of wild-type Zbed3 and the mutant Z1 and Z4 significantly upregulated the activity of Wnt signaling and promoted cancer cell proliferation (P < 0.05) but only overexpression of wild-type Zbed3 (P < 0.05), but not the mutant Z1, and Z4 (P > 0.05), significantly upregulated the expression of proliferating cell nuclear antigen (PCNA) in NCI-H1299 cells. These results indicate that Zbed3 may promote lung cancer cell proliferation through regulating PCNA expression besides regulating β-catenin expression and BED finger domain can impact on this function.  相似文献   

16.
Human seminal fluid was recently found to contain 18,19-dehydroprostaglandins E1 and E2 (E. H. Oliw, H. Sprecher, and M. Hamberg, (1986) J. Biol. Chem. 261, 2675-2683). In the present study, the cis and trans isomers of 18,19-dehydroprostaglandins E1 and E2 were prepared by incubation of microsomes of ram vesicular glands and glutathione with the precursor fatty acids, 8(Z),11(Z),14(Z),18(E/Z)-eicosatetraenoic acids, and 5(Z),8(Z),11(Z),14(Z),18(E/Z)-eicosapentaenoic acids, and used as references to characterize the 18,19-dehydroprostaglandins of human seminal fluid. Based on separation by reversed-phase high-performance liquid chromatography, capillary gas chromatography-mass spectrometry, and ozonolysis of the (-)-menthoxycarbonyl derivatives and on comparison with the authentic compounds, human seminal fluid was found to contain both the cis and trans isomers of 18,19-dehydroprostaglandins E1 and E2. Furthermore, human seminal fluid contained two related compounds, viz. 19,20-dehydroprostaglandins E1 and E2. The structures of these compounds were established by conversion into the corresponding prostaglandin B compounds, by mass spectrometric analysis and by chemical degradation by oxidative ozonolysis, which afforded, inter alia, 2(S)-hydroxy-adipic acid.  相似文献   

17.
The essential histone variant H2A.Z localises to both active and silent chromatin sites. In embryonic stem cells (ESCs), H2A.Z is also reported to co-localise with polycomb repressive complex 2 (PRC2) at developmentally silenced genes. The mechanism of H2A.Z targeting is not clear, but a role for the PRC2 component Suz12 has been suggested. Given this association, we wished to determine if polycomb functionally directs H2A.Z incorporation in ESCs. We demonstrate that the PRC1 component Ring1B interacts with multiple complexes in ESCs. Moreover, we show that although the genomic distribution of H2A.Z co-localises with PRC2, Ring1B and with the presence of CpG islands, H2A.Z still blankets polycomb target loci in the absence of Suz12, Eed (PRC2) or Ring1B (PRC1). Therefore we conclude that H2A.Z accumulates at developmentally silenced genes in ESCs in a polycomb independent manner.  相似文献   

18.
小偃麦附加系Z1和Z2中外源染色体2Ai-2的结构组成@张增燕$中国农业科学院作物育种栽培研究所!北京100081@辛志勇$中国农业科学院作物育种栽培研究所!北京100081@陈孝$中国农业科学院作物育种栽培研究所!北京100081小偃麦;;附加系;;染色体  相似文献   

19.
He XH  Zhang YM 《PloS one》2011,6(9):e24575
Epistasis plays an important role in genetics, evolution and crop breeding. To detect the epistasis, triple test cross (TTC) design had been developed several decades ago. Classical procedures for the TTC design use only linear transformations Z(1), Z(2) and Z(3), calculated from the TTC family means of quantitative trait, to infer the nature of the collective additive, dominance and epistatic effects of all the genes. Although several quantitative trait loci (QTL) mapping approaches in the TTC design have been developed, these approaches do not provide a complete solution for dissecting pure main and epistatic effects. In this study, therefore, we developed a two-step approach to estimate all pure main and epistatic effects in the F(2)-based TTC design under the F(2) and F(∞) metric models. In the first step, with Z(1) and Z(2) the augmented main and epistatic effects in the full genetic model that simultaneously considered all putative QTL on the whole genome were estimated using empirical Bayes approach, and with Z(3) three pure epistatic effects were obtained using two-dimensional genome scans. In the second step, the three pure epistatic effects obtained in the first step were integrated with the augmented epistatic and main effects for the further estimation of all other pure effects. A series of Monte Carlo simulation experiments has been carried out to confirm the proposed method. The results from simulation experiments show that: 1) the newly defined genetic parameters could be rightly identified with satisfactory statistical power and precision; 2) the F(2)-based TTC design was superior to the F(2) and F(2:3) designs; 3) with Z(1) and Z(2) the statistical powers for the detection of augmented epistatic effects were substantively affected by the signs of pure epistatic effects; and 4) with Z(3) the estimation of pure epistatic effects required large sample size and family replication number. The extension of the proposed method in this study to other base populations was further discussed.  相似文献   

20.
Polycomb repressive complex 2 (PRC2) is an essential chromatin-modifying enzyme that implements gene silencing. PRC2 methylates histone H3 on lysine-27 and is conserved from plants to flies to humans. In Drosophila melanogaster, PRC2 contains four core subunits: E(Z), SU(Z)12, ESC, and NURF55. E(Z) bears a SET domain that houses the enzyme active site. However, PRC2 activity depends upon critical inputs from SU(Z)12 and ESC. The stimulatory mechanisms are not understood. We present here functional dissection of the SU(Z)12 subunit. SU(Z)12 contains two highly conserved domains: an ∼140-amino-acid VEFS domain and a Cys2-His2 zinc finger (ZnF). Analysis of recombinant PRC2 bearing VEFS domain alterations, including some modeled after leukemia mutations, identifies distinct elements needed for SU(Z)12 assembly with E(Z) and stimulation of histone methyltransferase. The results define an extensive VEFS subdomain that organizes the SU(Z)12-E(Z) interface. Although the SU(Z)12 ZnF is not needed for methyltransferase in vitro, genetic rescue assays show that the ZnF is required in vivo. Chromatin immunoprecipitations reveal that this ZnF facilitates PRC2 binding to a genomic target. This study defines functionally critical SU(Z)12 elements, including key determinants of SU(Z)12-E(Z) communication. Together with recent findings, this illuminates PRC2 modulation by conserved inputs from its noncatalytic subunits.  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