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.
Abstract. Irradiation of closed circular phage Λ DNA in vivo at 365 nm results in the induction of single-strand breaks and alkali-labile lesions at rates of 1.1 × 10-14, and 0.2 × 10-14/dalton/J/m2, respectively. The sum of the induction rates is similar to the rate of induction of single-strand breaks plus alkali-labile lesions (1 × 10-14/dalton/J/m2) observed in the E. coli genome. Postirradiation incubation of wild-type cells in buffer results in rapid repair of the breaks (up to 80% repaired in 10 min). No repair was observed in a DNA polymerase I-deficient mutant of E. coli.  相似文献   

2.
Abstract —Irradiation at 365 nm results in the induction of approximately 2–4 times 10-6 and 1-2times 10-6 single-strand breaks (alkali-labile bonds) per 108 daltons per J m-2 in extracted phage T4 DNA and in Escherichia coli bacterial DNA, respectively. The rate of break induction in DNA of intact phage is approximately one-fourth that for extracted phage DNA. 2-aminoethylisothiouronium bromide-HBr protects against break induction in both phage systems. No breaks are induced in the DNA of bacteria irradiated under anaerobic conditions over the dose range tested. Possible induction mechanisms are suggested. Consideration is given to the relative importance of pyrimidine dimers and single-strand breaks in the bactericidal action of 365 nm radiation.  相似文献   

3.
Abstract— The synchronously developing aggregates of the cellular slime mold, Dictyostelium discoideum NC-4, were disaggregated into individual cells and irradiated with 254 nm UV light at preaggregation (0h), late interphase (6h), late aggregation (12 h), and preculmination (18 h). When assayed for replica-tive ability (colony formation), the developing cells at 0, 6, 12, and 18h showed the same sensitivity as vegetative cells; the 10% survival dose (D10) was 160 J/m2. The spores were more sensitive, with D10 of 70 J/m2. Excision repair of the nuclear DNA of the developing cells was studied by alkaline sucrose gradients. UV-induced single-strand breakage and rejoining of the DNA occurred to the same final extent in the cells from the 0, 6, 12 and 18 h stages of development, but a longer time was required for the completion of rejoining at the later stages (for example, at 54 J/m2, 6.6 h for preculmination cells, 3.3 h for preaggregation cells). When the cells irradiated at various stages were required to redevelop, as measured by the relative numbers of spores produced, their sensitivity for completing this development increased the later the stage from which they were taken. The D10s for spore production were 200, 130, 100 and 70 J/m2 for cells at the 0, 6, 12 and 18 h stages, respectively. The fractional viability among the spores that appeared after this treatment was the same independent of the stage at which the cells were irradiated; the D10 for this viability was 160 J/m2, the same as if the cells had been plated immediately with no intervening developmental sequence. We conclude that DNA excision repair as related to replicative ability is retained at all stages of development; however, development seems independent of replicative ability and depends upon DNA and/or non-DNA damage in a more complex way.  相似文献   

4.
THE PHOTODYNAMIC EFFECT OF HEMATOPORPHYRIN ON DNA   总被引:1,自引:0,他引:1  
Abstract— Breakage of DNA in vitro and inside E. colt cells has been determined after exposure to monochromatic 365 nm light in the presence of 10 µM hematoporphyrin. When measured by alkaline sucrose sedimentation, the yields of breaks were 1.4 × 10-12 per dalton and per J/m2 for Col El-DNA in vitro and 5.9 × 10-3 per dalton and per J/m2 for superinfecting phage Λ DNA inside E. coli cells made permeable by toluene. No breaks were found by neutral sucrose sedimentation, demonstrating that the lesions represent alkali-labile bonds. The majority of the alkali-labile bonds were induced by singlet oxygen, as evidenced by the several-fold higher yield obtained in D2O-containing buffer.  相似文献   

5.
Abstract DNA isolated from Escherichia coli and substituted with various amounts of iododeoxyuridine (12-95%) was irradiated at either 265. 300 or 313 nm and the frequency of chain breakage measured by sedimentation in either neutral or alkaline sucrose gradients. The wavelength dependence of the photochemical cross section for chain breaks paralleled that for iodine loss, and the average frequency of chain breakage per halogen loss was 0.50 ± 0.07. Approximately 80-90% of the breaks observed in alkali are alkali-labile bonds and are not observed under neutral conditions. The presence of ethanol during irradiation reduced the frequency of chain breakage by more than an order of magnitude. These results are interpreted in terms of a model in which photo-induced deiodination leads to the formation of a uracilyl radical which then abstracts a hydrogen atom from either a nearby sugar moiety or from another hydrogen donor such as ethanol. The resulting modified sugar can then rearrange to form either a clean chain break or an alkali-labile bond.  相似文献   

6.
Radical cations and dications of two carotenoids astaxanthin and canthaxanthin were prepared by oxidation with FeCl3 in fluorinated alcohols at room temperature. Absorption and electroabsorption (Stark effect) spectra were recorded for astaxanthin cations in mixed frozen matrices at temperatures about 160 K. The D0→D2 transition in cation radical is at 835 nm. The electroabsorption spectrum for the D0→D2 transition exhibits a negative change of molecular polarizability, Δα=−1.2·10−38 C·m2/V (−105 A3), which seems to originate from the change in bond order alternation in the ground state rather than from the electric field-induced interaction of D1 and D2 excited states. Absorption spectrum of astaxanthin dication is located at 715–717 nm, between those of D0→D2 in cation radical and S0→S2 in neutral carotenoid. Its shape reflects a short vibronic progression and strong inhomogeneous broadening. The polarizability change on electronic excitation, Δα=2.89·10−38 C·m2/V (260 A3), is five times smaller than in neutral astaxanthin. This value reflects the larger energetic distance from the lowest excited state to the higher excited states than in the neutral molecule.  相似文献   

7.
Abstract— Irradiation at 440 + 360 nm and a fluence rate of 3.8 kJm-2 min-1, of both complexes previously formed between proflavine and either øX circular single-stranded (ss) DNA or øX supercoiled duplex (RFI)DNA, induces single-strand scissions in the two DNAs under consideration. Linear øXSS DNA molecules are detected by sedimentation through alkaline sucrose gradients. After treatment of the øXRFI DNA, however, the degree of degradation is the same whether it is measured under neutral or alkaline conditions, indicating that alkaline-labile bonds are not created; moreover, double-strand breaks can only be detected after accumulation of single-strand breaks. In addition to the amount of proflavine bound to the DNA and the duration of irradiation, the following factors are shown to influence the nicking activity of the treatment: (1) the DNA structure (the øXRFI DNA is much more sensitive than the øXss DNA); (2) the ionic strength of the medium during irradiation (a high value of 0.5 leads to a markedly increased efficiency); (3) the addition of cysteamine (this latter compound decreases the reaction rate) and (4) the irradiation wavelength (after irradiation at 440 nm alone, the reaction occurs at a reduced rate and is sensitive to NaN3). The kinetics of the nicking reaction does not follow a single-hit curve showing that at least one primary lesion occurs prior to strand breakage. On the other hand, strand scission cannot be detected after irradiation of the proflavine-DNA complexes at the low fluence rate causing a decrease in the infectivity of both øXSS and øXRFI DNAs. Similarly. the sedimentation pattern of the DNA extracted from treated øx174 phages 99.9% inactivated, is identical to that of the control ss DNA, although more drastic treatments are susceptible to induce single strand breaks inside the phage head. Finally, the unknown lesion (s) that is biologically important does not prevent the treated DNAs from penetrating into the hostcells.  相似文献   

8.
Two complexes [Ln(e,a-cis-1,4-chdc)(e,a-cis-1,4-Hchdc)(phen)(H2O)]2?10H2O (Ln = Eu, 1; Tb, 2, 1,4-H2chdc = 1,4-cyclohexanedicarboxylic acid; phen = 1,10-phenanthroline) have been synthesized and structurally characterized by single-crystal X-ray diffraction. Both complexes are doubly e,a-cis-1,4-chdc-bridged dimers. The e,a-cis-1,4-Hchdc, phen, and water molecules bond to Ln3+, forming nine-coordinate complexes. 3-D supramolecular frameworks are constructed by hydrogen bonds and π–π stacking interactions. Luminescence spectra exhibit the 5D07F J (J = 0–4) and 5D47F J (J = 6–3) transitions of Eu3+ for 1 and Tb3+ ion for 2, respectively.  相似文献   

9.
The interaction between the buffer 2-amino-2-(hydroxymethyl)propane-1,3-diol (Tris) and the Eu(III) ion has been studied by luminescence spectroscopy in D2O. Emission and excitation spectra (5D07F0 transition) indicate an interaction with both [TrisH]+ and neutral Tris species. The former is weak and probably of the outer-sphere type. The latter is of inner-sphere type and corresponds to the formation of the [Eu(Tris)]3+ species (estimated logK1 = 2.3 ± 0.3). Buffer Tris is also demonstrated to prevent the formation of an Eu-hydroxo species in the pD range of 7–8. Potentiometric measurements in H2O allowed a more precise calculation of the stability constant: logK1 = 2.44 ± 0.07. Comparison with the data for aliphatic amines and other metal ions lead to the conclusion that the Eu/Tris interaction is mainly achieved through the amino group. 1H-NMR spectra in presence of Tb(III) ions confirmed both this assumption and the presence of a weak interaction with TrisH+. Quantitative determinations of association constants between lanthanide ions and macromolecules of biological interest performed in presence of Tris should, therefore, be corrected for the Eu/Tris interaction.  相似文献   

10.
Cis- and trans-buten-2-yl free radicals are shown to react with butene-2 cis (Bc) and D2S in the following metathetical steps: giving rise to butenes-1 (B1). The initial formations of butene-1,3 d1 and total butene-1 in D2S? Bc mixtures have been studied in the initial pressure range 20–200 torr for B c, 0–41 torr for D2S and at 717–817 K. The main initiation and termination steps are shown to be: Assuming a rapid equilibrium between cis- and trans-C4H7?, ki ? 1015.5 exp(?85.5/RT) s?1 (RT in kcal mol?1) and kt ? 1013 mol?1 cm3 s?1, we get: k2c = 1.1 k2t = 1012.0 exp[(?9.25 ± 2)/RT] mol?1 cm3 s?1 and k2 + 1.5 k2′ ? 1012.1 exp[(?15.2 ± 2)/RT] mol?1 cm3 s?1. Isotopic effects relating to processes (2c′) + (2t′) and to (i′) have been evaluated.  相似文献   

11.
The γ‐irradiation induced stepwise genesis of dimer, trimer, tetramer, and higher agglomerates in 1% ovalbumin aqueous solution was quantitatively followed by a SDS‐PAGE analytical method. The molecular mass distributions obtained in the dose interval form 0 to 2 kGy were used to calculate the gel dose by using Good's theory of cascade processes. The difference between the calculated value (3.5 kGy) and the measured value (8.5 kGy) is attributed to the competition between the agglomeration and the interradical conversion processes, which prevails in the higher dose region. It was estimated from the gel dose that 1 · 1020 of intermolecular bonds per dm3 were necessary to form a gel from 1% ovalbumin solution. A dynamic viscosity of the solution was also measured simultaneously. A simple semiempirical equation was developed containing only one parameter—gel dose, Dg—and it fits the viscosity–dose data fairly well. © 2000 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 38: 1729–1733, 2000  相似文献   

12.
DFT (B3LYP, M06‐2X) and MP2 methods are applied to the design of a wide series of the potentially 10‐C‐5 neutral compounds based on 6‐azabicyclotetradecanes: XC1(YCH2CH2CH2)3N 1 – 3 , XC1(YC6H4CH2)3N 4 – 6 , XC1[Y(tBuC6H3)CH2]3N 7 – 9 and carbatranophanes 10 – 25 (X=Me, F, Cl; Y=O, NH, CH2, SiH2; Z=O, CH2, (CH2)2, (CH2)3). Carbatranophanes 10 – 25 are characterized by a sterical compression of their axial 3c–4e XC1←N fragment with respect to that in the parent molecules 4 – 6 . A magnitude of the revealed effect depends on a valence surrounding of the central carbon atom C1, the size and the nature of the side chains (Z) that link the “π‐electron cap” with a tetradecane backbone. This circumstance allowed us to obtain 10‐C‐5 structures with the configuration of the bonds around the C1 atom, which corresponds to practically an ideal trigonal bipyramid. In these compounds, the values of the covalence ratio χ of approximately 0.6 for the coordination C1←N contacts with a covalent contribution (atoms in molecules (AIM) and natural bond orbital (NBO)) are record in magnitude. These values lie close to a low limit of the interval of the χSi←D change (0.6–0.9) being characteristic of the dative and ionic‐covalent (by nature) Si←D bond (D=N, O) in the known 10‐Si‐5 silicon compounds.  相似文献   

13.
Six α, β, β-trifluorostyrenes with the following substituents, viz., p-MeO, p-Me, m-Me, p-Cl, m-Cl, and m-CF3, were synthesized by the reaction of the corresponding Grignard reagents with tetrafluoroethylene in tetrahydrofuran. Similarly, α-and β-trifluoroethenylnaphthalenes were prepared. The substituent electronic effects on the 19F-NMR parameters were investigated for the trifluorostyrenes (I). Linear correlations between the Hammett σ constants and the following 19F-NMR parameters were established, namely, chemical shifts δ. (F1) and δ (F2), coupling constants J12, differences of chemical shifts Δδ3-1 (δ (F3)—δ(f1) or Δδ3-2. The results are consistent with previous expectations based on the simple concept of “distorted π-electron clouds”. Facts are presented which indicate that the Δδ3-1 (or Δδ3-2) values may serve as empirical measures of the degree of polarization of the π bonds of these fluoroolefins.  相似文献   

14.
Twelve 2,3′-bisindolylmethanes with various substituents were investigated using electrospray ionization quadrupole time-of-flight tandem mass spectrometry in positive ion mode. A retro-[3+2] reaction was observed in the collision-induced dissociation spectra of protonated 2,3′-bisindolylmethanes for the first time. The mechanism of retro-[3+2] reaction was concerted or stepwise. For the concerted pathway, carbon–carbon bonds of a protonated compound simultaneously cracked and the m/z 208 ion ([C15H10D2N]+) was observed with hydrogen–deuterium exchange labeling. The stepwise pathway goes through 1,3-hydrogen migration twice and the m/z 208 ion ([C15H10D2N]+) and m/z 207 ion ([C15H11DN]+) were detected with deuterium labeling. In the deuterium-labeled tandem mass spectrum for one compound, only the peak at m/z 208 was present at high abundance, suggesting that the concerted pathway is more likely. In addition, the substituents have no obvious trends on the ratios of the product intensity to the base intensity, further supporting the concerted pathway.  相似文献   

15.
We developed an experimental method for the determination of the tracer diffusivity Dtr in ultrathin polymer films, and the changes in the segmental mobility of tracer molecules while they diffuse through matrices of different thickness and get adsorbed onto a target substrate. Dtr starts decreasing already at 120–150 nm and drops to 1% of its bulk value in films as thin as 7.5 nm. We discuss the results highlighting a strong decoupling between the reduction in mass transport at the nanoscale and the increase in the glass transition temperature determined via capacitive dilatometry together with a breakdown of the Stokes–Einstein relation between orientational and translational degrees of freedom.

  相似文献   


16.
Ab initio configuration interaction wavefunctions and energies are reported for 16 doublet states of the anion radical of ethyl bacteriopheophorbide a (Et-BPheo a), and are employed in an analysis of the electronic absorption spectrum. The lowest excited doublet state D1 is predicted to lie 8601 cm-1 above the ground state D0; the D1← D0 transition is nearly forbidden, with a computed oscillator strength f= 0.002. The visible absorption spectrum is shown to consist of transitions to three 2(π, π*) states, D2, D3, and D4. The D4← D0 transition (y-polarized, f= 0.91) appears to account for observed intense absorption at 15 800 cm-1. The Soret band of Et-BPheo a is shown to consist of transitions to several 2(π,π*) states, D7-D15. Transitions of particularly high intensity include D7← D0 (y-polarized, f= 0.72), D10← D0 (y-polarized, f= 1.1), D12← D0 (xy-polarized, f= 0.86) and D15← D0 (y-polarized, f= 0.83). Spin density data and plots are used to describe and compare the general features of the unpaired spin distributions in D0 and D1, which are in reasonable agreement with other reported calculated values and available experimental data for D0.  相似文献   

17.
A dinuclear copper(II) complex, [CuII2(L)2] is afforded by the reaction of CuCl2 · 2H2O with a triazenido ligand, 1-[(2-carboxymethyl) benzene]-3-[2-carboxybenzene] triazene (H2L). Structural investigation shows that the copper-copper distance [2.3985(7) Å] is significantly shorter than the sum of the van der Waals radii of Cu (1.40 Å), suggesting that there are metal-metal bonds in [CuII2(L)2]. In solid, there is a strong antiferromagnetic interaction between copper(II) ions (J = –135.6 cm–1). In homogeneous environment, [CuII2(L)2] shows electrocatalytic activities for hydrogen generation both from acetic acid with a turnover frequency (TOF) of 32 mol of hydrogen per mole of catalyst per hour [mol(H2) · mol–1(catalyst) · h–1] at an overpotential (OP) of 941.6 mV, and neutral buffer with a TOF of 512 mol(H2) · mol–1(catalyst) · h–1 at an OP of 836.7 mV.  相似文献   

18.
Low-angle electron diffraction (LAED) was used to study the microstructure of crazes produced at different temperatures T and strain rates in thin films of monodisperse polystyrene (PS). At a slow strain rate of 4.1 × 10?6 s?1 both the fibril diameter D and the fibril spacing D0 of crazes in 1800k molecular weight PS remained constant with temperature up to T ≈ 70°C and then sharply increased as T approaches Tg. At a higher strain rate of ~ 10?2 s?1, both D and D0 increase only slightly with T. The values of D and D0 over a range of temperature are in very good agreement with those values obtained in bulk samples using small-angle x-ray scattering. The crazing stress was measured as a function of temperature in the thin films of the 1800k molecular weight PS strained at the same slow strain rate used for the LAED measurements. These measurements were analyzed using a simple model of craze growth to reveal the temperature and strain rate dependence of the craze surface energy Γ. At room temperature Γ ≈ 0.076 J/m2 (versus Γ ≈ 0.087 J/m2 predicted) and was observed to remain constant up to T ≈ 70°C and then decrease by approximately a factor of two at T = 90°C. This decrease in Γ is believed to result from chain disentanglement to form fibril surfaces at sufficiently high temperatures and occurs in the same temperature range in which the craze fibril extension ratio λ was observed to increase.  相似文献   

19.
Using density functional calculations, we demonstrate that the planarity of the nonclassical planar tetracoordinate carbon (ptC) arrangement can be utilized to construct new families of flat, tubular, and cage molecules which are geometrically akin to graphenes, carbon nanotubes, and fullerenes but have fundamentally different chemical bonds. These molecules are assembled with a single type of hexagonal blocks called starbenzene (D6h C6Be6H6) through hydrogen‐bridge bonds that have an average bonding energy of 25.4–33.1 kcal mol?1. Starbenzene is an aromatic molecule with six π electrons, but its carbon atoms prefer ptC arrangements rather than the planar trigonal sp2 arrangements like those in benzene. Various stability assessments indicate their excellent stabilities for experimental realization. For example, one starbenzene unit in an infinite two‐dimensional molecular sheet lies on average 154.1 kcal mol?1 below three isolated linear C2Be2H2 (global minimum) monomers. This value is close to the energy lowering of 157.4 kcal mol?1 of benzene relative to three acetylene molecules. The ptC bonding in starbenzene can be extended to give new series of starlike monocyclic aromatic molecules (D4h C4Be4H42?, D5h C5Be5H5?, D6h C6Be6H6, D7h C7Be7H7+, D8h C8Be8H82?, and D9h C9Be9H9?), known as starenes. The starene isomers with classical trigonal carbon sp2 bonding are all less stable than the corresponding starlike starenes. Similarly, lithiated C5Be5H5 can be assembled into a C60‐like molecule. The chemical bonding involved in the title molecules includes aromaticity, ptC arrangements, hydrogen‐bridge bonds, ionic bonds, and covalent bonds, which, along with their unique geometric features, may result in new applications.  相似文献   

20.
The instantaneous elastic moduli for a nylon-6 monofilament were derived on strain recoveries right after creep, stress relaxation, and rapid elongation,E c ,E s andE e , respectively. It was found that during strain recoveryE s (>E e ) andE e increase monotonically with increasing load,m 1, on the sample. The extrapolated value of Es atm 1=0 g is almost equal to Young's modulus, 4.06 GPa. The value ofE c also increased with increasingm 1, and atm 1=600 g (1.93 t cm–2) reached about 14 GPa. The endothermic heat change right after creep, stress relaxation or rapid elongation,Q, was negligibly small. For comparison,E s ,E c andQ were also investigated for silicone rubber. It was found thatE s (53.8 M Pa at the draw ratioD=1.2) decreased abruptly atD=1.3. In the range ofD=1.4–1.9,E s was only 22.6 MPa. In the case of stress relaxation,Q increased with increasingD from 4 J mol–1 (atD=1.2) to 56 J mol–1 (atD=1.9). FurthermoreE c (5.58 MPa atm 1=133.8 g (429.4 kg cm–2)) increased gradually with increasing m1 and attained 16.6 MPa atm 1=548.4 g (1.76 t cm–2). In the case of creep,Q was in the range of 0–11.5 J mol–1 and larger when larger loads,m 2 were removed during the later stages of creep.Dedicated to Professor Bernhard Wunderlich on the occasion of his 65th birthdayThe author wishes to thank Mr. Keizi Igarashi and Mr. Tetsuya Yasui for helping in the experiments.  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