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 46 毫秒
1.
Abstact  A reaction of high-temperature solid-phase catalytic isotope exchange (HSCIE) was studied for the preparation of tritium- and deuterium-labeled ligands of glutamate and dopamine receptors. Tritium-labeled (5S,10R)-(+)-5-methyl-10,11-dihydro-5H-dibenzo[a,d]cyclopenten-5,1-imine ([G-3H]MK-801) and R(+)-7-hydroxy-N,N-di-n-propyl-2-aminotetraline ([G-3H]-7-OH-DPAT) were obtained with a specific activity of 210 and 120 Ci/mol, respectively. The isotopomeric distribution of deuterium-labeled ligands was studied using time-of-flight mass-spectrometer MX 5310 (ESI-o-TOF) with electrospray and orthogonal ion injection. Mean deuterium incorporation per ligand molecule was 11.09 and 3.21 atoms for [G-3H]MK-801 and [G-3H]-7-OH-DPAT, respectively. The isotope label was shown to be distributed all over the ligand molecule. The radioreceptor binding of tritium-labeled ligands [G-3H]MK-801 and [G-3H]-7-OH-DPAT was analyzed using the brain structure of Vistar rats. It was demonstrated that [G-3H]MK-801 specifically binds to hippocampus membranes with K d 8.3 ± 1.4 nM, B max being 3345 ± 300 fmol/mg protein. The [G-3H]-7-OH-DPAT ligand specifically binds to rat striatum membranes with K d 10.01 ± 0.91 nM and B max 125 ± 4.5 fmol/mg protein. It was concluded that the HSCIE reaction can be used for the preparation of highly tritium-labeled (+)-MK-801 and 7-OH-DPAT with retention of their physiological activities. Original Russian Text ? Yu.A. Zolotarev, Yu.Yu. Firsova, A. Abaimov, A.K. Dadayan, V.S. Kosik, A. V. Novikov, N.V. Krasnov, B. V. Vaskovskii, I.V. Nazimov, G.I. Kovalev, N.F. Myasoedov, 2009, published in Bioorganicheskaya Khimiya, 2009, Vol. 35, No. 3, pp. 323–333.  相似文献   

2.
Melting parameters of 2U:1A complexes formed by polyuridylic acid [poly(U)] and three adenine dinucleotides, diribonucleoside monophosphonate ApA and diastereoisomers of dideoxyribonucleoside methyl phosphonate [(dApA)1 and (dApA)2], in 1M NaCl and at a number of dinucleotide concentrations were obtained from differential scanning microcalorimetric data and interpreted in terms of the theory of helix–coil equilibrium in oligonucleotide–polynucleotide systems. The apparent binding constant, 1/cm, at 39°C and melting temperatures, Tm, at 1 × 10?3 M dinucleotide concentration indicate the following order of thermodynamic stability of the complexes: 2 poly(U) · (dApA)2 (2.27 × 103M?1, 44.2°C) > 2 poly(U) · (dApA)1 (9.9 × 102M1, 39.2°C) > 2 poly(U) · (ApA) (5.9 × 102M?1, 35.8°C). Corresponding calorimetric enthalpies of melting, ΔHm: 13.5, 12.7, and 12.8 kcal/mol (UUA base triplets) were found to be considerably lower than the van't Hoff enthalpies, ΔHapp: 29.4, 16.2, and 16.2 kcal/mol, respectively, evaluated from the dependence of the melting temperatures on dinucleotide concentration. Self-association of dinucleotides and their simultaneous binding as monomers, dimers, and higher-order associated species is suggested as the most probable cause of the differences between ΔHm and ΔHapp values. The differences in thermodynamic properties of the complexes formed by (dApA)1 and (dApA)2 diastereoisomers are discussed in connection with their known conformational properties. The higher and essentially enthalpic stability of the 2 poly(U) · (dApA)2 complex correlates with a lower degree of intramolecular stacking of the (dApA)2 isomer. The hydrophobically enhanced strong self-association of the latter greatly influences the thermodynamics of its complex formation with poly(U) and results in ΔHappHm = 2.3.  相似文献   

3.
The conformational properties of 5-fluorouracil derivatives are compared to uracil derivatives. FUrd, 5′-FUMP, and poly(FU) are studied as a function of pH and temperature by 19F- and 1H-nmr spectroscopy, and the corresponding uracil derivatives by 1H-nmr spectroscopy. FUrd exhibits no significant conformational changes with solution pH (5–10). In contrast, at low pH (6–7) 5′-FUMP and 5′-UMP show similar conformational features, while at high pH (9) 5′-FUMP shows significant conformational alterations. Also, poly(U) and poly(FU) are conformationally similar at low pH, but increasing pH induces changes in poly(FU). These changes are observed in the backbone [γ(C4′-C5′)], furanose, and furanose-base conformations. The apparent pKa of N3-H ionization of the FUra base is determined by 1H- and 19F-nmr to range from 7.5 to 8.2 [FUrd < 5′-FUMP < 5′-FUDP < poly(FU)]. These observations are interpreted as a result of electrostatic interactions generated between the ionized phosphate group and the negatively charged base moiety as the pH is raised. The interaction properties of poly(FU) with ApA are studied by 1H- and 19F-nmr spectroscopy, and these properties compared to those published for poly(U). Poly(FU) forms a complex with ApA inducing upfield 1H-shifts in both components, and downfield 19F- shifts in poly(FU). The base stoichiometry of the complex for poly(U)·ApA is 2U:1A at various U/A ratios. In contrast, the base stoichiometry of the poly(FU)·ApA complex appears to be dependent on the FU/A ratio. At high FU/A ratio, the complex is 2FU:1A, and as the FU/A ratio approaches unity the complex becomes 1FU:1A.  相似文献   

4.
The interaction between poly (I) and poly (C) in acid medium has been studied by potentiometric titration, mixing curves and thermal denaturation. Phase diagramms as a function of ionic strength, pH, and temperature have been established. From these data it is shown that the acid titration of the complex poly (I) · poly (C) passes through a triple-stranded intermediate poly (I) · poly (C) · poly (C+) to yield finally the protonated double-helical complex poly (I) · poly (C+). The mixing curves indicate the sole presence of the three-stranded complex in the intermediate zone. On the basis of the pK's the coexistence between the three-stranded complex with the neighboring double-stranded structure is demonstrated in a narrow rang of pH and ionic strength. The geometry of the base arrangements, their conformation and the sense of the strands are discussed in the light of the data presented. A Hoogsteen-type pairing between the bases for poly (I) · poly (C+) is favored, although the reverse Hoogsteen pair cannot be excluded.  相似文献   

5.
We examined three sonicated, specific-seqiemce polydeoxynucleotides in solution over a wide range of concentrations of several salts by 13P-nmr spectroscopy, and we found that the alternating copolymer poly(dAdT)·poly(dAdT) exhibits a dinucleotide repeat unit in all five salts and at all concentrations studied, as indicated by the presence of a doubled in its 31P-nmr spectra. The two components of the doublet show selective shift effects. The upfield component is assigned to dApdT in the gauche?-gauche? conformation and shifts upfield in all four monovalent salts used, relative to a single-stranded oligonucleotide control. The downfield component is assigned to dTpdA in the trans-gauche? conformation and shifts downfield with increasing CsF concentration but remains essentially constant in LiCl, NaCl, and CsCl. These changes indicate a fast noncooperative transition for poly(dAdT)·poly-(dAdT) from a presumed right-handed dinucleotide-repeat B-form to another conformation with a dinucleotide-repeat structure, via a continuum of structures that may differ in the extent of the winding of the double helix. Ethanol causes the upfield component to collapse into the other component, indicating conversion to a structure with a mononucleotide repeat unit and a trans-gauche? conformation. Up to 1M Mg2+ appears to have no significant effect on the phosphodiester conformations of poly(dAdT)·poly(dAdT). By contrast, poly-(dGdC)·poly(dGdC) gives a slow cooperative transition from what is considered to be a right-handed regular B-form to a left-handed Z-form on increasing MgCl2 and NaCl concentrations, although we observed no changes in chemical shifts below the transition points. The homopolymer poly(dA)·poly(dT) exhibits no unusual shift effects or transitions upon the addition of salts when compared to the oligonucleotide control and is considered to be a regular B-form with a gauche?-gauche? phosphodiester backbone conformation. These differences emphasize the distinct secondary structures of DNAs of different sequences and their selective responses to changes in solution conditions.  相似文献   

6.
 As an extension of our earlier discoveries that ZnII-cyclen complex (1) (cyclen=1,4,7,10-tetraazacyclododecane) and ZnII-acridine-pendant cyclen complex ZnII-N-(9-acridin)ylmethyl-cyclen (3) are the first compounds to selectively recognize thymidine and uridine nucleosides in aqueous solution at physiological pH, the interaction of these and a relevant complex, bis(ZnII-cyclen) (7), has been investigated with a series of polynucleotides, single-stranded poly(U) and poly(G), and double-stranded poly(A)·poly(U), poly(dA)·poly(dT) and poly(dG)·poly(dC). These ZnII-cyclen complexes interact with the imide-containing nucleobases in the single-stranded poly(U), unperturbed by the presence of the anionic phosphodiester backbone. The affinity constant of 1 for each N(3)-deprotonated uracil base in poly(U) is determined to be log K= 5.1 by a kinetic measurement, which is almost the same as log K=5.2 for the interaction of 1 with uridine. Thus, they disrupt the A-U (or A-T) hydrogen bonds to unzip the duplex of poly(A)·poly(U) or poly(dA)·poly(dT), as demonstrated by lowering of the melting temperatures (T m) of poly(A)·poly(U) and poly(dA)·poly(dT) in 5 mM Tris-HCl buffer (pH 7.6, 10 mM NaCl) with increase in their concentrations. The order of the denaturing efficiency is well correlated with that of the 1 : 1 affinity constants for each complex with uracil or thymine;7>3>1. The comparison of circular dichroism (CD) spectra for poly(A)·poly(U), poly(A), and poly(U) in the presence of 3 has revealed a structural change from poly(A)·poly(U) to two single strands, poly(A) and poly(U), caused by 3 binding exclusively to uracils in poly(U). On the other hand, the acridine-pendant cyclen complex 3, which earlier was found to associate with guanine by the ZnII coordinating with guanine N(7), in addition to the π-π stacking, interacts with guanine in the double helix of poly(dG)·poly(dC) from outside and stabilized the double-stranded structure, as indicated by higher T m. Received: 31 December 1997 / Accepted: 23 February 1998  相似文献   

7.
The solution behavior of the single-stranded CCAAT-containing octamer 1 , d(AGCCAATA), that comprises part of the nuclear factor I (NF-I) recognition site at the origin of replication of human adenovirus has been studied by nmr spectroscopy at 500 and 600 MHz. Proton resonance assignments for 1 were aided by selective 13C enrichment at C1′ of A1 or A5. High-resolution 13C-1H heteronuclear multiple-bond coherence spectra of the 13C-labeled oligomers permitted the selective detection of furanosyl ring protons within each labeled residue due to short- and long-range 13C-1H couplings to the enriched C1′. The resulting assignments provided firm starting points in the interpretation of double quantum filtered correlated spectra, yielding information supplemented by total correlated spectroscopy (TOCSY) and rotating frame nuclear Overhauser effect spectroscopic data to completely assign the 1H-nmr spectrum of 1 and extract 3JHH values for furanose con-formational analysis. Several 13C-1H spin-coupling constants within the 13C-enriched A1 or A5 residues were measured from cross-peak shifts in TOCSY spectra, and their signs determined by inspection of the relative orientations of these shifts. 1H-2-H and 13C-1H spin-couplings both indicate a preference (> 75%) for south (C2′-endo) conformations by the furanosyl rings of 1 . © 1994 John Wiley & Sons, Inc.  相似文献   

8.
31P- and 1H-nmr and laser Raman spectra have been obtained for poly[d(G-T)]·[d(C-A)] and poly[d(A-T)] as a function of both temperature and salt. The 31P spectrum of poly[d(G-T)]·[d(C-A)] appears as a quadruplet whose resonances undergo separation upon addition of CsCl to 5.5M. 1H-nmr measurements are assigned and reported as a function of temperature and CsCl concentration. One dimensional nuclear Overhauser effect (NOE) difference spectra are also reported for poly[d(G-T)]·[d(C-A)] at low salt. NOE enhancements between the H8 protons of the purines and the C5 protons of the pyrimidines, (H and CH3) and between the base and H-2′,2″ protons indicate a right-handed B-DNA conformation for this polymer. The NOE patterns for the TH3 and GH1 protons in H2O indicate a Watson–Crick hydrogen-bonding scheme. At high CsCl concentrations there are upfield shifts for selected sugar protons and the AH2 proton. In addition, laser Raman spectra for poly[d(A-T)] and poly[d(G-T)]·[d(C-A)] indicate B-type conformations in low and high CsCl, with predominantly C2′-endo sugar conformations for both polymers. Also, changes in base-ring vibrations indicate that Cs+ binds to O2 of thymine and possibly N3 of adenine in poly[d(G-T)]·[d(C-A)] but not in poly[d(A-T)]. Further, 1H measurements are reported for poly[d(A-T)] as a function of temperature in high CsCl concentrations. On going to high CsCl there are selective upfield shifts, with the most dramatic being observed for TH1′. At high temperature some of the protons undergo severe changes in linewidths. Those protons that undergo the largest upfield shifts also undergo the most dramatic changes in linewidths. In particular TH1′, TCH3, AH1′, AH2, and TH6 all undergo large changes in linewidths, whereas AH8 and all the H-2′,2″ protons remain essentially constant. The maximum linewidth occurs at the same temperature for all protons (65°C). This transition does not occur for d(G-T)·d(C-A) at 65°C or at any other temperature studied. These changes are cooperative in nature and can be rationalized as a temperature-induced equilibrium between bound and unbound Cs+, with duplex and single-stranded DNA. NOE measurements for poly[d(A-T)] indicate that at high Cs+ the polymer is in a right-handed B-conformation. Assignments and NOE effects for the low-salt 1H spectra of poly[d(A-T)] agree with those of Assa-Munt and Kearns [(1984) Biochemistry 23 , 791–796] and provide a basis for analysis of the high Cs+ spectra. These results indicate that both polymers adopt a B-type conformation in both low and high salt. However, a significant variation is the ability of the phosphate backbone to adopt a repeat dependent upon the base sequence. This feature is common to poly[d(G-T)]·[d(C-A)], poly[d(A-T)], and some other pyr–pur polymers [J. S. Cohen, J. B. Wouten & C. L Chatterjee (1981) Biochemistry 20 , 3049–3055] but not poly[d(G-C)].  相似文献   

9.
Interaction between polylysine and DNA's of varied G + C contents was studied using thermal denaturation and circular dichroism (CD). For each complex there is one melting band at a lower temperature tm, corresponding to the helix–coil transition of free base pairs, and another band at a higher temperature tm, corresponding to the transition of polylysine-bound base pairs. For free base pairs, with natural DNA's and poly(dA-dT) a linear relation is observed between the tm and the G + C content of the particular DNA used. This is not true with poly(dG)·poly(dC), which has a tm about 20°C lower than the extrapolated value for DNA of 100% G + C. For polylysine-bound base pairs, a linear relation is also observed between the tm and the G + C content of natural DNA's but neither poly(dA-dT) nor poly(dG)·poly(dC) complexes follow this relationship. The dependence of melting temperature on composition, expressed as dtm/dXG·C, where XG·C is the fraction of G·C pairs, is 60°C for free base pairs and only 21°C for polylysine-bound base pairs. This reduction in compositional dependence of Tm is similar to that observed for pure DNA in high ionic strength. Although the tm of polylysine-poly(dA-dT) is 9°C lower than the extrapolated value for 0% G + C in EDTA buffer, it is independent of ionic strength in the medium and is equal to the tm0 extrapolated from the linear plot of tm against log Na+. There is also a noticeable similarity in the CD spectra of polylysine· and polyarginine·DNA complexes, except for complexes with poly(dA-dT). The calculated CD spectrum of polylysine-bound poly(dA-dT) is substantially different from that of polyarginine-bound poly(dA-dT).  相似文献   

10.
The effects of adding D2SO4, and p-toluenesulfonic acid-d to D-cellobiose dissolved in D2O were investigated at 23 °C by plotting 13C NMR chemical shift changes (Δδ) against the acid to D-cellobiose molar ratio. 13C Chemical shifts of all 18 carbon signals from α and β anomers of D-cellobiose showed gradual decreases due to increasing acidity in aqueous D2SO4 medium. The C-1 of the α anomer showed a slightly higher response to increasing D+ concentration in the surrounding. In the aqueous p-toluenesulfonic acid-d medium, C-6′ and C-4′ carbons of both α, and β anomeric forms of D-cellobiose are significantly affected by increasing the sulfonic acid concentrations, and this may be due to a 1:1 interaction of p-toluenesulfonic acid-d with the C-6′, C-4′ region of the cellobiose molecule.  相似文献   

11.
G · C complex formation was studied by infrared spectroscopy for a system that has been shown by Inoue & Orgel (1982) to give efficient, template-dependent synthesis of oligo(G). Guanosine-5′-phosphor-2-methylimiazolide (2-MeImpG) exhibits rapid formation with poly(C) of a G · C double helix at pD ~ 8 and of a C · G · CH+ triple helix at pD ~ 6.5 in the presence of Na+. Significant oligo(G) synthesis does not occur under these conditions. In the presence of synthetically effective concentrations of Mg2+ G · C complex formation is much slower but eventually goes to completion. The rate of complex formation parallels that of chemical synthesis. Infrared spectra and melting curves confirm that oligo(G) of high molecular weight is formed in high yield. The bulk of the G · C complex at any given time during the reaction is composed of G residues that have already been polymerized and not of the monomer 2-MeImpG. Evidence indicates that synthesis proceeds primarily at growing points at the ends of the G · C helical regions and not randomly on a fully occupied template.  相似文献   

12.
To identify and assess the specificity of the 1,25-dihydroxyvitamin D3 chick intestinal cytoplasmic receptor's nucleotide binding site, a competitive DNA-cellulose binding assay was utilized. Unlike other steroid hormone receptors, the 1,25-dihydroxyvitamin D3-receptor complex binds homologous DNA at 4 °C and does not appear to undergo thermal- or salt-induced activation. Data are presented which suggest that receptor binding discriminates between double-stranded DNA and RNA but is not specific with respect to DNA base sequences. However, DNA base sequence selectivity by 1,25-dihydroxyvitamin D3-receptor complexes is observed using synthetic polydeoxyribonucleotides, particularly, poly(dA-dT) · poly(dA-dT) and poly(dA) · poly(dT). Preference for double-stranded over single-stranded DNA was also observed. Consistent with this finding, both actinomycin D and ethidium bromide caused a dose-dependent inhibition of receptor binding to DNA-cellulose. It is concluded that the 1,25-dihydroxyvitamin D3-receptor complex has specificity for AT-rich segments of double-stranded DNA and that this interaction is not merely electrostatic, but also involves hydrophobic interaction with the major and/or minor grooves of the DNA helix.  相似文献   

13.
The use of desorption chemical ionization (D/CI) mass spectrometry for the structural characterization of the abscisic acid (ABA)-conjugated metabolites, abscisic acid-γ-d-glucosyl ester (ABAGE), 4′-O-dihydrophaseic acid-γ-d-glucoside (DPAGS) and γ-hydroxy-γ-methylglutaryl-hydroxyabscisic acid (HMG-HOABA), was studied. The effects of the D/CI source temperature on the spectral features are described. Enhanced quasi-molecular ion intensities (e.g., [M·NH4]+) were observed in the D/CI(NH3) spectrum at a low source temperature (170°C). When a higher source temperature (250°C) was used, more extensive fragmentation occurred, and structurally diagnostic fragment ions appeared in the spectra. The spectra obtained at two degrees of source temperature were complementary to each other for the structural characterization.  相似文献   

14.
The H-2L d allele has been identified by others as the sole Ir gene in the H-2 d haplotype for the cytotoxic T lymphocyte (CTL) response to mouse lymphocytic choriomeningitis virus (LCMV). The BALB/c-H-2 dm2 (C-H-2 dm2 ) mutant lacks H-2L d , and thus should be ideal for assessing the contribution of virus-immune CTL to LCM immunopathology. Comparison of the C-H-2 dm2 mice with congenic BALB/c mice revealed that there is a delay of about 24 h in the onset of severe inflammatory process and symptoms in the mutant strain, but the absence of H-2L d did not prevent the later development of fatal disease in mice injected intracerebrally (i.e.) with neurotropic LCMV. This could indicate that virus-immune CTL are not the major mediators of clinical LCM. Spleen cells from LCMV-primed BALB/c mice did not show CTL activity for LCMV-infected C3H.OH, C-H-2 dm2 , or (CBA × C-H-2 dm2 )F1 target cells. However, immune lymphocytes from both the mutant and the F1 strains lyse virus-infected BALB/c cells. Furthermore, BtO.HTG and, in some experiments, B10.A(5R) mice generated CTL lytic for LCMV-infected BALB/c, C-H-2 dm2 , and (CBA × CH-2 dm2 )F1 macrophages. Apparently H-2L d is immunodominant in the H-2d restricted response to LCMV. However, in the absence of H-2L d , it seems that H-2K d and, to a lesser extent, H-2D d also serve as Ir genes for the CTL response in this infection. Even so, the absence of the H-2Ld-restricting element results in a disease process which is either delayed in onset or less severe.  相似文献   

15.
Isotherms of the EtBr adsorption on native and denatured poly(dA)poly(dT) in the temperature interval 20–70°C were obtained. The EtBr binding constants and the number of binding sites were determined. The thermodynamic parameters of the EtBr intercalation complex upon changes of solution temperature 20–48°C were calculated: 1.0·106 M−1K≤1.4·106 M−1, free energy ΔG o=−8.7±0.3 kcal/mol, enthalpy ΔH o≅0, and entropy ΔS o=28±0.5 cal/(mol deg). UV melting has shown that the melting temperature (T m) of EtBr-poly(dA)poly(dT) complexes (μ=0.022,4.16·10−5 M EtBr) increased by 17°C as compared with the ΔT m of free homopolymer, whereas the half-width of the transition (T m) is not changed. It was shown for the first time that EtBr forms complexes of two types on single-stranded regions of poly(dA)poly(dT) denatured at 70°C: strong (K 1=1.7·105 M−1; ΔG o=−8.10±0.03 kcal/mol) and weak (K 2=2.9·103 M−1; ΔG o=−6.0±0.3 kcal/mol).The ΔG o of the strong and weak complexes was independent of the solution ionic strength, 0.0022≤μ≤0.022. A model of EtBr binding with single-stranded regions of poly(dA)poly(dT) is discussed.  相似文献   

16.
The binding of [G-3H]nitrobenzylthioinosine to intact Chinese hamster ovary cells has been studied kinetically and thermodynamically. The association of nitrobenzylthioinosine with cells is a second-order process which proceeds at 24°C with a rate constant of 2·107 M?1·s?1. Dissociation of the complex was characterized as a simple first-order process with rate constant on the order of 7·10?3 s?1. The quotient of these is comparable to the dissociation constant as measured in equilibrium binding studies, 2.2·10?10 M. The temperature dependence of the rate of association indicated an Arrhenius activation energy of 8.4 kcal·mol?1, while that of the equilibrium constant for dissociation indicated a standard enthalpy change of 8.8 kcal·mol?1. The large increase in affinity of nitrobenzylthioinosine as compared to natural nucleosides is attributable to an entropy-driven interaction with the binding site. Thymidine, dipyridamole and papaverine each decrease the apparent dissociation constant for the nitrobenzylthioinosine-cell complex; the latter, inhibitors of nucleoside transport, decrease the rate of dissociation of the complex.  相似文献   

17.
The interaction between poly (G) and poly (C) was investigated in neutral and acid medium by optical methods. Three main points arise from this investigation. (1) The formation of poly (G)·poly (C) was complete only above an ionic strength of about 0.6M [Na+]. Lowering the ionic strength increased the amounts of free poly (G) and free poly (C) that could be detected. (2) When titrating towards acid pH values a transition took place which was characterized by potentiometry, mixing curves, and circular dichroism: a three-stranded poly (G)·poly (C)·poly (C+) complex was formed analogous to the transition observed for the acid titration of poly (I)·poly (C). (3) Even when the poly (G)·poly (C) complex was incompletely formed (at low ionic strength) in neutral medium all poly (C) entered the triple-stranded complex.  相似文献   

18.
Abstract

This experiment investigates the effect of increasing the dietary content of bacterial protein meal (BPM) on the protein and energy metabolism, and carcass chemical composition of growing chickens. Seventy-two Ross male chickens were allocated to four diets, each in three replicates with 0% (D0), 2% (D2), 4% (D4), and 6% BPM (D6), BPM providing up to 20% of total dietary N. Five balance experiments were conducted when the chickens were 3 – 7, 10 – 14, 17 – 21, 23 – 27, and 30 – 34 days old. During the same periods, 22-h respiration experiments (indirect calorimetry) were performed with groups of 6 chickens (period 1), 5 chickens (period 2), and one chicken (periods 3 – 5). After each balance period, one chicken in each cage was killed and the carcass weight was recorded. Chemical analyses were performed on the carcasses from periods 1, 3, and 5. Weight gain, feed intake, and feed conversion rate were found to be similar for all diets. Chickens on D0 retained 1.59 g N · kg?0.75 · d?1, significantly more than chickens on D2, D4, and D6, which retained 1.44 g, 1.52 g, and 1.50 g N · kg?0.75 · d?1, respectively. This was probably caused by the higher nitrogen content of D0. Neither the HE (p = 0.92) nor the retention of energy (p = 0.88) were affected by diet. Carcass composition was similar between diets, in line with the values for protein and energy retention found in the balance and respiration experiments. It was concluded that the overall protein and energy metabolism as well as carcass composition were not influenced by a dietary content of up to 6% BPM corresponding to 20% of dietary N.  相似文献   

19.
Summary The cytochemical properties of a guanine-specific synthetic fluorescent analogue of actinomycin D, 7-amino-actinomycin D, have been studied in fixed and living preparations of L cells and polytene chromosomes of salivary glands ofChironomus thummi thummi andDrosophila lummei (Hackman).7-Amino-actinomycin D has been shown to bind to DNA-containing structures, thereby inducing in them a bright red fluorescence. No specific fluorescence has been found in RNA-containing structures treated with this fluorescent probe.The fluorescence pattern of some regions of polytene chromosomes with a known nucleotide composition was analysed. It has been established that 7-amino-actinomycin D induces a very weak fluorescence in GC-poor chromosome regions of theDrosophila lummei toromere structure. Data indicating a nonlinear dependence between the fluorescence intensity of a stained chromosome region and the GC content in its DNA have been obtained. The influence of DNA nucleotide composition in a chromosome region on the fluorescence of 7-amino-actinomycin D is discussed. In combination with quinacrine staining and the Feulgen fluorescence reaction, treatment with 7-amino-actinomycin D provides useful information about the distribution of GC base pairs in the chromosome region under study.  相似文献   

20.
S P Fodor  P A Starr  T G Spiro 《Biopolymers》1985,24(8):1493-1500
Raman spectra have been recorded for poly(dG-dT) · poly(dA-dC) and poly(dA-dT) · poly(dA-dT) in low salt and at high concentrations of CsF. Poly(dG-dT) · poly(dA-dC) shows no change in the 682-cm?1 guanine mode, demonstrating the absence of the Z-structure at high salt. The 790-cm?1 phosphodiester symmetric stretch, however, shifts up 5 cm?1 in 4.3M CsF, suggesting a slight conformational change, associated with ion binding or hydration changes. Poly(dA-dT) · poly(dA-dT) shows an additional broad band at 816 cm?1, attributed to the phosphodiester modes associated with the C3′-endo deoxyribose units in the alternating B-structure. In this case, both the 841- and the 816-cm?1 asymmetric phosphodiester stretches, associated with the C2′- and C3′-endo units, shift down on addition of CsF in a sequential manner. Correlation of this sequence with that previously observed for the two 31P-nmr resonances, establishes that the phosphodiester stretching frequencies depend on the conformation of the 5′-sugar, and not on the 3′-sugar.  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