首页 | 本学科首页   官方微博 | 高级检索  
文章检索
  按 检索   检索词:      
出版年份:   被引次数:   他引次数: 提示:输入*表示无穷大
  收费全文   251篇
  免费   35篇
  国内免费   1篇
  2023年   2篇
  2017年   1篇
  2015年   1篇
  2014年   1篇
  2013年   3篇
  2012年   9篇
  2011年   5篇
  2010年   14篇
  2009年   5篇
  2008年   11篇
  2007年   18篇
  2006年   16篇
  2005年   7篇
  2004年   12篇
  2003年   11篇
  2002年   11篇
  2001年   14篇
  2000年   8篇
  1999年   12篇
  1998年   3篇
  1997年   2篇
  1996年   6篇
  1995年   7篇
  1993年   4篇
  1992年   7篇
  1991年   6篇
  1990年   6篇
  1989年   7篇
  1988年   4篇
  1987年   3篇
  1986年   8篇
  1984年   5篇
  1983年   1篇
  1981年   1篇
  1980年   2篇
  1979年   5篇
  1978年   4篇
  1977年   8篇
  1976年   10篇
  1975年   1篇
  1974年   7篇
  1973年   3篇
  1972年   3篇
  1971年   1篇
  1970年   1篇
  1969年   3篇
  1966年   1篇
  1953年   1篇
  1891年   2篇
  1889年   1篇
排序方式: 共有287条查询结果,搜索用时 15 毫秒
1.
Photoreactions of bacteriorhodopsin at acid pH.   总被引:6,自引:3,他引:3       下载免费PDF全文
It has been known that bacteriorhodopsin, the retinal protein in purple membrane which functions as a light-driven proton pump, undergoes reversible spectroscopic changes at acid pH. The absorption spectra of various bacteriorhodopsin species were estimated from measured spectra of the mixtures that form at low pH, in the presence of sulfate and chloride. The dependency of these on pH and the concentration of Cl- fit a model in which progressive protonation of purple membrane produces "blue membrane", which will bind, with increasing affinity as the pH is lowered, chloride ions to produce "acid purple membrane." Transient spectroscopy with a multichannel analyzer identified the intermediates of the photocycles of these altered pigments, and described their kinetics. Blue membrane produced red-shifted KL-like and L-like products, but no other photointermediates, consistent with earlier suggestions. Unlike others, however, we found that acid purple membrane exhibited a very different photocycle: its first detected intermediate was not like KL in that it was much more red-shifted, and the only other intermediate detectable resembled the O species of the bacteriorhodopsin photocycle. An M-like intermediate, with a deprotonated Schiff base, was not found in either of these photocycles. There are remarkable similarities between the photoreactions of the acid forms of bacteriorhodopsin and the chloride transport system halorhodopsin, where the Schiff base deprotonation seems to be prevented by lack of suitable aspartate residues, rather than by low pH.  相似文献   
2.
The halorhodopsin chromoprotein, a retinal-protein complex with an apparent molecular mass of 20 kilo-daltons, exhibits all of the halide-dependent effects found for the chromophore of functional halorhodopsin in cell envelope vesicles. With increasing halide concentration (a) an alkali-dependent 580/410 nm chromophore equilibrium (attributed to reversible deprotonation of the retinal Schiff's base) is shifted toward the 580-nm chromophore and (b) the flash-induced photocycle proceeds increasingly via P520, rather than via P660. The halide-binding site(s) responsible for these effects must reside, therefore, in the chromoprotein. Chloride and bromide are about equivalent, but iodide is much less effective in these effects and in being transported. Several other anions, i.e. thiocyanate, nitrate, phosphate, and acetate, affect the absorption maximum of the chromophore but do not allow the production of P520 upon flash illumination and are not transported. However, these ions appear to compete with chloride in the flash experiments. These observations suggest that binding of anions to a relatively nonspecific site affects the protonation state of the Schiff's base in the chromophore. Either this site directly or a more specific site, connected to the first one by a sequential pathway, is involved with the photocycle intermediates and with chloride transport by halorhodopsin.  相似文献   
3.
The membrane-bound detergent-activated ATPase from Halobacterium saccharovorum was purified at a physiological salt concentration (4 M NaCl) in the presence of nonionic detergents. The preparation contains putative subunits of 110, 71, 31, 22, and 14 kDa. The enzyme activity required high salt concentration but was not dependent on any one specific monovalent cation or any anion. The hydrolysis of ATP was nonlinear with time; the data are consistent with a kinetic model where the enzyme is irreversibly converted from an initial into a final stable form during the first few minutes of the reaction. The model thus contains a rate constant (k) for the transition and hydrolytic rates, v1 and v2, for the two forms of the enzyme. We found that this hysteretic behavior was influenced differently by various conditions and inhibitors. The constant k was smaller with Mn2+ than with Mg2+ as the divalent cation, showed negative temperature dependence, and a distinct pH optimum between 7.5 and 8.5. Thiols decreased k, but nitrate, a specific inhibitor of archaebacterial ATPases, increased it. ADP showed competitive inhibition against both the initial and the final form of the enzyme. Nitrate reversibly inhibited only the latter and in a manner dependent on whether Mn2+ or Mg2+ was used. The kinetic data suggest that all agents tested, with effects on the hydrolytic activity, seem to act at or near the catalytic site of the enzyme.  相似文献   
4.
The bacteriorhodopsin photocycle contains more than five spectrally distinct intermediates, and the complexity of their interconversions has precluded a rigorous solution of the kinetics. A representation of the photocycle of mutated D96N bacteriorhodopsin near neutral pH was given earlier (Váró, G., and J. K. Lanyi. 1991. Biochemistry. 30:5008-5015) as BRhv-->K<==>L<==>M1-->M2--> BR. Here we have reduced a set of time-resolved difference spectra for this simpler system to three base spectra, each assumed to consist of an unknown mixture of the pure K, L, and M difference spectra represented by a 3 x 3 matrix of concentration values between 0 and 1. After generating all allowed sets of spectra for K, L, and M (i.e., M1 + M2) at a 1:50 resolution of the matrix elements, invalid solutions were eliminated progressively in a search based on what is expected, empirically and from the theory of polyene excited states, for rhodopsin spectra. Significantly, the average matrix values changed little after the first and simplest of the search criteria that disallowed negative absorptions and more than one maximum for the M intermediate. We conclude from the statistics that during the search the solutions strongly converged into a narrow region of the multidimensional space of the concentration matrix. The data at three temperatures between 5 and 25 degrees C yielded a single set of spectra for K, L, and M; their fits are consistent with the earlier derived photocycle model for the D96N protein.  相似文献   
5.
The pK(a) values of D85 in the wild-type and R82Q, as well as R82A recombinant bacteriorhodopsins, and the Schiff base in the D85N, D85T, and D85N/R82Q proteins, have been determined by spectroscopic titrations in the dark. They are used to estimate the coulombic interaction energies and the pK(a) values of the Schiff base, D85, and R82 during proton transfer from the Schiff base to D85, and the subsequent proton release to the bulk in the initial part of the photocycle. The pK(a) of the Schiff base before photoexcitation is calculated to be in effect only 5.3-5.7 pH units higher than that of D85; overcoming this to allow proton transfer to D85 requires about two thirds of the estimated excess free energy retained after absorption of a photon. The proton release on the extracellular surface is from an unidentified residue whose pK(a) is lowered to about 6 after deprotonation of the Schiff base (Zimanyi, L., G. Varo, M. Chang, B. Ni, R. Needleman, and J.K. Lanyi, 1992. Biochemistry. 31:8535-8543). We calculate that the pK(a) of the R82 is 13.8 before photoexcitation, and it is lowered after proton exchange between the Schiff base and D85 only by 1.5-2.3 pH units. Therefore, coulombic interactions alone do not appear to change the pK(a) of R82 as much and D85 only by 1.5-2.3 pH units. Therefore, coulombic interactions alone do not appear to change the pK(a) of R82 as much as required if it were the proton release group.  相似文献   
6.
The stages in the photocycle of bacteriorhodopsin (BR) involving the M and N intermediates are investigated using a double pulse excitation method. A first (cycling) pulse at 532 nm is followed, with an appropriate time delay, by a second pulse (337, 406, 446, or 470 nm) which induces the M-->BR back-photoreaction. After depletion by the second pulse a repopulation of M in the millisecond range is observed which is interpreted in terms of a thermal N-->M relaxation. It is thus concluded that a (thermal) M<-->N equilibrium accounts for the biphasic decay of M in the BR photocycle. Other models for this stage of the light-driven proton-pump are therefore unnecessary.  相似文献   
7.
8.
Following ovariectomy of five New Zealand white rabbits at day 25 of pregnancy, the intrauterine pressure (IUP) and uterine progesterone (P) and prostaglandin (PG) levels were measured sequentially at days 25, 26 and 27. At day 25, when the uterine P and PGE and PGF were high, massive intrauterine treatment with 500 μg PGF2α provoked only a sustained contracture on which only low level oscillation in IUP was superimposed. At day 26, when the P levels had decreased significantly (P<0.001) and the PG levels had not changed significantly, 50 μg PGF2α significantly increased cyclic IUP as compared with the day 25 value (P<0.001). At day 27, when the P levels decreased further, as little as 5 μg PGF2α provoked still higher cyclic IUP, in spite of a significant reduction in PG levels (P<0.05).Stretching the uterus of six post partum and six 26 days pregnant rabbits (after removing the uterine contents) significantly increased the uterine PGF levels (P<0.001). However, stretch increased only cyclic IUP of the post partum uterus and was without effect on the pregnant uterus, which still had high P levels. These results indicate that the myometrium activated by exogenous PG or stretch, regardless of whether the uterine PG levels increase, remain unchanged or even moderately decrease, provided that the uterine P levels are reduced to a critical value.  相似文献   
9.
Light-induced sodium extrusion from H halobium cell envelope vesicles proceeds largely through an uncoupler-sensitive pathway involving bacteriorhodopsin and a proton/sodium antiporter. Vesicles from bacteriorhodopsin-negative strains also extrude sodium ions during illumination, but this transport is not sensitive to uncouplers and has been proposed to involve a light-energized primary sodium pump. Proton uptake in such vesicles is passive, and under steady-state illumination the large electrical potential (negative inside) is just balanced by a pH difference (acid inside), so that the protonmotive force is near zero. Action spectra indicated that this effect of illumination is attributable to a pigment absorbing near 585 nm (of 568 for bacteriorhodopsin). Bleaching of the vesicles by prolonged illumination with hydroxylamine results in inactivation of the transport; retinal addition causes partial return of the activity. Retinal addition also causes the appearance of an absorption peak at 588 nm, while the absorption of free retinal decreases. The 588 nm pigment is present in very small quantities (0.13 nmole/mg protein), and behaves differently from bacteriorhodopsin in a number of respects. Vesicles can be prepared from bacteriorhodopsin-containing H halobium strains in which primary transport for both protons and sodium can be observed. Both pumps appear to cause the outward transport of the cations. The observations indicate the existence of a second retinal protein, in addition to bacteriorhodopsin, in H halobium, which is associated with primary sodium translocation. The initial proton uptake normally observed during illumination of whole H halobium cells may therefore be a passive flux in response to the primary sodium extrusion.  相似文献   
10.
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