首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Comparative energetics of chicken latissimus dorsi muscles, tonic anterior (ALD) and phasic posterior (PLD), were investigated by measuring initial heat production. Heat components were analyzed in terms of the equation: E = A + W + αF(L) + f(P, t) As the muscles were stretched by increments, heat produced in isometric twitches and tetani decreased in a linear fashion. Two processes are involved: one tension independent, the activation heat, or A; and the other tension dependent, Wi + αF(L) + f(P, t). In twitches, A, per unit tension, is equivalent in the PLD and ALD. Tension-dependent heat, per unit tension, is greater in the PLD due to Wi; but tension-time-related heat, f(P, t), per unit tension, is similar in both muscles. In tetanic contractions, differences in A and f(P, t), per unit tension, are attributed to the greater Vmax in the PLD. The differences in the energetics of isometric contractions in the PLD and ALD, therefore, can be explained by inherent differences in tension development, compliance, and myosin and reticular ATPase activities. Data from isotonic twitches were quantified by means of the equivalent tension technique. Both muscles exhibited an extra heat associated with shortening, αF(L). In the PLD, the ratio αF/Pot is greater; it is load independent and ½ the value of a/Po in both muscles. Enthalpy efficiency, We + Wi/E, is comparable in both muscles. A Fenn effect is observed only when isotonic energy liberation is compared to a decreasing isometric energy expenditure base line.  相似文献   

2.
A method is developed for analyzing in a unified manner both uniaxial and uniform biaxial strain data obtained from nearly isotropic tissues. The formulation is a direct application of nonlinear elasticity theory pertaining to large deformations. The general relation between Eulerian stress (σ) and extension ratio (λ) in soft isotropic elastic bodies undergoing uniform deformation takes the simple form: σ = ((λ3 - 1)/λ) f(λ), where f(λ) must be determined for each material. The extension ratio may be either greater than 1.0 (uniaxial elongation), or lie between zero and 1.0 (uniform biaxial extension). Simple analytical functions for f(λ) are most readily found for each tissue by plotting all data as (λ3 - 1)/λσ vs. λ. Of those tissues investigated in this way (dog pericardium and pleura, and cat mesentery and dura), all but pleura could be adequately described by a parabola: 1/f(λ) = 1/k{[(λM - λ)(λ - λm)]/[λM - λm}. In these instances, three material constants per tissue (K, λM, λm) served to predict approximately the stresses attained during both small and large deformations, in strips and sheets alike. It was further found that the uniaxial strain asymptote (λM) was linearly related to the biaxial strain asymptote (ΛM), thus effectively reducing the number of constants by one.  相似文献   

3.
We took a discovery approach to explore the actions of cAMP and two of its analogs, one a cAMP mimic ((Sp)-adenosine cyclic 3′:5′-monophosphorothioate ((Sp)-cAMPS)) and the other a diastereoisomeric antagonist ((Rp)-cAMPS), on a model system of the type Iα cyclic AMP-dependent protein kinase holoenzyme, RIα(91–244)·C-subunit, by using fluorescence spectroscopy and amide H/2H exchange mass spectrometry. Specifically, for the fluorescence experiments, fluorescein maleimide was conjugated to three cysteine single residue substitution mutants, R92C, T104C, and R239C, of RIα(91–244), and the effects of cAMP, (Sp)-cAMPS, and (Rp)-cAMPS on the kinetics of R-C binding and the time-resolved anisotropy of the reporter group at each conjugation site were measured. For the amide exchange experiments, ESI-TOF mass spectrometry with pepsin proteolytic fragmentation was used to assess the effects of (Rp)-cAMPS on amide exchange of the RIα(91–244)·C-subunit complex. We found that cAMP and its mimic perturbed at least parts of the C-subunit interaction Sites 2 and 3 but probably not Site 1 via reduced interactions of the linker region and αC of RIα(91–244). Surprisingly, (Rp)-cAMPS not only increased the affinity of RIα(91–244) toward the C-subunit by 5-fold but also produced long range effects that propagated through both the C- and R-subunits to produce limited unfolding and/or enhanced conformational flexibility. This combination of effects is consistent with (Rp)-cAMPS acting by enhancing the internal entropy of the R·C complex. Finally, the (Rp)-cAMPS-induced increase in affinity of RIα(91–244) toward the C-subunit indicates that (Rp)-cAMPS is better described as an inverse agonist because it decreases the fractional dissociation of the cyclic AMP-dependent protein kinase holoenzyme and in turn its basal activity.Cyclic AMP-dependent protein kinase (PKA)1 plays a crucial role in a plethora of cellular functions. All isoforms of PKA are composed of two catalytic (C) subunits and homodimeric regulatory (R) subunits (13). As the name implies, cAMP is a major PKA regulator (4). Much progress has been made in the last decade in delineating the molecular basis of action of cAMP. An important tactic in this endeavor has been through the comparison of the effects of cAMP with those of two phosphorothioate cAMP analogs: (Sp)-cAMPS (a cAMP mimic) and (Rp)-cAMPS (an antagonist and a diastereoisomer of (Sp)-cAMPS). Although the importance of geometry of the sulfur substitution is critical in determining the pharmacological properties of the two phosphorothioate cAMP analogs, the molecular basis for this behavior is not fully understood. To date, these comparisons have only been made using either wild-type or truncated mutants of the type Iα regulatory subunit (RIα) that are free in solution, not complexed to the C-subunit. X-ray spectroscopic examination of ligand-bound RIα(92–379) complexes reveals few differences between ligand-bound complexes, but the (Rp)-cAMPS complex is structurally “looser” with higher thermal factors than complexes formed with either cAMP or (Sp)-cAMPS (5). This is consistent with the observation that both cAMP and (Sp)-cAMPS, but not (Rp)-cAMPS, raise the urea concentration required for wild-type RIα unfolding (6). Further insight into the structural basis of cAMP action stems from NMR spectroscopic comparison of the effects of (Rp)-cAMPS, cAMP, and (Sp)-cAMPS on chemical shifts and 15N relaxation of the RIα(119–244) mutant (7). In addition to producing fewer significant chemical shift changes than either cAMP or (Sp)-cAMPS, (Rp)-cAMPS binding is associated with enhanced millisecond to microsecond time scale backbone motions of a β-turn (β2,3 loop) and around the phosphate-binding cassette (PBC) (7).Further insight into the molecular basis of actions of cAMP and its analogs should come from the analysis of ligand-bound R·C complexes. Unfortunately, the large size of even the heterodimeric R·C complex (∼95 kDa) and the difficulty of preparing (Rp)-cAMPS·R·C-subunit crystals currently preclude the use of both NMR spectroscopy and x-ray crystallography. Consequently, we took two alternative lower resolution approaches to this issue. One approach involves the use of site-directed labeling combined with fluorescence spectroscopy to examine both the effects of cAMP and its analogs on R-C subunit binding kinetics and on the conformational dynamics of RIα(91–244). RIα(91–244) includes the “A” cyclic nucleotide binding (CNB) domain, the pseudosubstrate, and linker domains and represents the minimal segments necessary for high affinity C-subunit binding (Fig. 1) (8). The other approach involves an examination of the effects of cAMP and its analogs on solvent exposure/conformational flexibility of RIα(91–244)·C-subunit complex using H/2H amide exchange measured with a combination of mass spectrometry (ESI-Q-TOF) and proteolytic fragmentation. In the first approach, fluorescein maleimide (FM) was conjugated to three cysteine substitution mutants with the substitution sites located near or within the pseudosubstrate sequence, the linker domain, or αC (R92C, T104C, and R239C, respectively) of RIα(91–244) (Fig. 1). The time-resolved fluorescence anisotropy results suggest that cAMP and (Sp)-cAMPS reduce the interaction of the RIα linker domain and αC with the two peripheral R-C interaction sites on the C-subunit (so-called Sites 2 and 3) without affecting the interaction of the pseudosubstrate sequence with the active site cleft (so-called Site 1). Because of limitations of the amide H/2H exchange experiments, only the effects of (Rp)-cAMPS on H/2H amide exchange in RIα(91–244)·C-subunit complex could be investigated. The results showed that (Rp)-cAMPS induces a relatively widespread increase in amide exchange, indicating limited unfolding and/or enhanced conformational flexibility that is propagated almost globally through the C-subunit and, at least, part of RIα. These conformational changes were accompanied by a 5-fold increase in the affinity of RIα(91–244) toward C-subunit, suggesting that, at least, some of the (Rp)-cAMPS effects are mediated by an increase in internal entropy. Finally, the (Rp)-cAMPS-induced increase in R-C affinity indicates that (Rp)-cAMPS is better described as an inverse agonist because the basal activity of the PKA holoenzyme should be decreased by (Rp)-cAMPS.Open in a separate windowFig. 1.Overview of PKA structure and cAMP analogs. A, domain organization of RIα showing the domain boundaries of RIα(91–244) where the pseudosubstrate in green is connected to CNB-A domain in blue by a linker segment. B, structure of RIα(91–244) in the C-subunit-bound conformation (Protein Data Bank code 1U7E (23)) showing the pseudosubstrate in green, linker in yellow, and helical subdomain comprising helices αN, αA, αB, and αC in blue and β-subdomain in tan. The PBC is in red. C, structure of the C·RIα(91–244) holoenzyme showing the C-subunit in tan and RIα(91–244) in blue. Sites for introduction of cysteines by site-directed mutagenesis are represented by red circles. The cAMP binding site (PBC) is in red. D, structure of cAMP showing the 2′-OH group and 3′–5′ phosphodiester bond. The exocyclic oxygens upon replacement with sulfur atoms to generate the (Sp)-cAMPS and (Rp)-cAMPS diastereomers are highlighted.  相似文献   

4.
5.
After a short time interval of length δt during microbial growth, an individual cell can be found to be divided with probability Pd(tt, dead with probability Pm(tt, or alive but undivided with the probability 1 − [Pd(t) + Pm(t)]δt, where t is time, Pd(t) expresses the probability of division for an individual cell per unit of time, and Pm(t) expresses the probability of mortality per unit of time. These probabilities may change with the state of the population and the habitat''s properties and are therefore functions of time. This scenario translates into a model that is presented in stochastic and deterministic versions. The first, a stochastic process model, monitors the fates of individual cells and determines cell numbers. It is particularly suitable for small populations such as those that may exist in the case of casual contamination of a food by a pathogen. The second, which can be regarded as a large-population limit of the stochastic model, is a continuous mathematical expression that describes the population''s size as a function of time. It is suitable for large microbial populations such as those present in unprocessed foods. Exponential or logistic growth with or without lag, inactivation with or without a “shoulder,” and transitions between growth and inactivation are all manifestations of the underlying probability structure of the model. With temperature-dependent parameters, the model can be used to simulate nonisothermal growth and inactivation patterns. The same concept applies to other factors that promote or inhibit microorganisms, such as pH and the presence of antimicrobials, etc. With Pd(t) and Pm(t) in the form of logistic functions, the model can simulate all commonly observed growth/mortality patterns. Estimates of the changing probability parameters can be obtained with both the stochastic and deterministic versions of the model, as demonstrated with simulated data.  相似文献   

6.
An analytical solution is obtained for the three-dimensional spatial distribution of potential inside a flat cell, such as the layer of horizontal cells, as a function of its geometry and resistivity characteristics. It was found that, within a very large range of parameter values, the potential is given by [Formula: see text] where r = ρ/ρ0, = z0, ρ = (Ri/Rm)·ρ0, δ = h0; K is a constant; J is the assumed synaptic current; ρ, z are cylindrical coordinates; ρ0 is the radius of the synaptic area of excitation; h is the cell thickness; and Ri, Rm are the intracellular and membrane resistivities, respectively. Formula A closely fits data for the spatial decay of potential which were obtained from the catfish internal and external horizontal cells. It predicts a decay which is exponential down to about 40% of the maximum potential but is much slower than exponential below that level, a characteristic also exhibited by the data. Such a feature in the decay mode allows signal integration over the large retinal areas which have been observed experimentally both at the horizontal and ganglion cell stages. The behavior of the potential distribution as a function of the flat cell parameters is investigated, and it is found that for the range of the horizontal cell thicknesses (10-50 μ) the decay rate depends solely on the ratio Rm/Ri. Data obtained from both types of horizontal cells by varying the diameter of the stimulating spot and for three widely different intensity levels were closely fitted by equation A. In the case of the external horizontal cell, the fit for different intensities was obtained by varying the ratio Rm/Ri; in the case of the internal horizontal cell it was found necessary, in order to fit the data for different intensities, to vary the assumed synaptic current J.  相似文献   

7.
The extra heat liberation accompanying muscular shortening, the force-determined shortening heat, is defined as the difference between the heat produced when shortening occurs and that produced in an isometric contraction developing the same amount of force and performing the same amount of internal work. Based on this definition, the initial energy production in twitches and tetanic contractions (E) is given by E = A + f (P, t) + αFx + W, where A is the activation heat, f(P, t), the tension-related heat (a heat production associated with the development and maintenance of tension), αFx, the force-determined shortening heat, and W, the external work. It is demonstrated that this equation accurately accounts for the time-course of heat evolution and the total initial energy production in both twitches and tetani at 0°C. The force-determined shortening heat is liberated, during shortening, in direct proportion to (a) the distance shortened, and (b) the force against which shortening occurs. The normalized value of the force-determined shortening heat coefficient, αF/Po, is the same in both the twitch and the tetanus. Finally, this formulation of the muscle's energy production also accounts for the total energy production in afterload isotonic twitches at 20°C, where a Fenn effect is not demonstrable.  相似文献   

8.
The ribosome is an ancient macromolecular machine responsible for the synthesis of all proteins in all living organisms. Here we demonstrate that the ribosomal peptidyl transferase center (PTC) is supported by a framework of magnesium microclusters (Mg2+-μc's). Common features of Mg2+-μc's include two paired Mg2+ ions that are chelated by a common bridging phosphate group in the form Mg(a)2+–(O1P-P-O2P)–Mg(b)2+. This bridging phosphate is part of a 10-membered chelation ring in the form Mg(a)2+–(OP-P-O5′-C5′-C4′-C3′-O3′-P-OP)–Mg(a)2+. The two phosphate groups of this 10-membered ring are contributed by adjacent residues along the RNA backbone. Both Mg2+ ions are octahedrally coordinated, but are substantially dehydrated by interactions with additional RNA phosphate groups. The Mg2+-μc's in the LSU (large subunit) appear to be highly conserved over evolution, since they are unchanged in bacteria (Thermus thermophilus, PDB entry 2J01) and archaea (Haloarcula marismortui, PDB entry 1JJ2). The 2D elements of the 23S rRNA that are linked by Mg2+-μc's are conserved between the rRNAs of bacteria, archaea and eukarya and in mitochondrial rRNA, and in a proposed minimal 23S-rRNA. We observe Mg2+-μc's in other rRNAs including the bacterial 16S rRNA, and the P4–P6 domain of the tetrahymena Group I intron ribozyme. It appears that Mg2+-μc's are a primeval motif, with pivotal roles in RNA folding, function and evolution.  相似文献   

9.
10.
11.
We previously found that mice fed lutein accumulated its oxidative metabolites (3′-hydroxy-ε,ε-caroten-3-one and ε,ε-carotene-3,3′-dione) as major carotenoids, suggesting that mammals can convert xanthophylls to keto-carotenoids by the oxidation of hydroxyl groups. Here we elucidated the metabolic activities of mouse liver for several xanthophylls. When lutein was incubated with liver postmitochondrial fraction in the presence of NAD+, (3′R,6′R)-3′-hydroxy-β,ε-caroten-3-one and (6RS,3′R,6′R)-3′-hydroxy-ε,ε-caroten-3-one were produced as major oxidation products. The former accumulated only at the early stage and was assumed to be an intermediate, followed by isomerization to the latter. The configuration at the C3′ and C6′ of the ε-end group in lutein was retained in the two oxidation products. These results indicate that the 3-hydroxy β-end group in lutein was preferentially oxidized to a 3-oxo ε-end group via a 3-oxo β-end group. Other xanthophylls such as β-cryptoxanthin and zeaxanthin, which have a 3-hydroxy β-end group, were also oxidized in the same manner as lutein. These keto-carotenoids, derived from dietary xanthophylls, were confirmed to be present in plasma of normal human subjects, and β,ε-caroten-3′-one was significantly increased by the ingestion of β-cryptoxanthin. Thus, humans as well as mice have oxidative activity to convert the 3-hydroxy β-end group of xanthophylls to a 3-oxo ε-end group.  相似文献   

12.
We report herein an investigation into dynamic magnetic clustering that occurs during immunoassays as biofunctionalized magnetic nanoparticles (BMNs) become associated with biotargets. We measure the dynamic effective relaxation time τ eff(t) and use scanning electron microscopy (SEM) and transmission electron microscopy (TEM) to investigate the C-reactive protein (CRP) as it associates with the BMN Fe3O4-antiCRP to form the magnetic cluster Fe3O4-antiCRP-CRP. The results indicate that τ eff(t) increases with increasing association time. In addition, the ration Δτ eff/τ 0 as a function of CRP concentration follows a characteristic logistic function, which provides a basis for estimating the quantity of biomolecules with a detection sensitivity close to 0.1 ppm. After the association, SEM and TEM images show that CRP and Fe3O4-antiCRP conjugate to form Fe3O4-antiCRP-CRP clusters hundreds of nanometers in size. The SEM and TEM images provide direct evidence of the formation of magnetic clustering.  相似文献   

13.
C. Chevalet  M. Gillois    R. F. Nassar 《Genetics》1977,86(3):697-713
Properties of identity relation between genes are discussed, and a derivation of recurrent equations of identity coefficients in a random mating, diploid dioecious population is presented. Computations are run by repeated matrix multiplication. Results show that for effective population size (Ne) larger than 16 and no mutation, a given identity coefficient at any time t can be expressed approximately as a function of (1—f), (1—f)3 and (1— f)6, where f is the mean inbreeding coefficient at time t. Tables are presented, for small Ne values and extreme sex ratios, showing the pattern of change in the identity coefficients over time. The pattern of evolution of identity coefficients is also presented and discussed with respect to N eu, where u is the mutation rate. Applications of these results to the evolution of genetic variability within and between inbred lines are discussed.  相似文献   

14.
The genomes of positive-strand RNA viruses undergo conformational shifts that complicate efforts to equate structures with function. We have initiated a detailed analysis of secondary and tertiary elements within the 3′ end of Turnip crinkle virus (TCV) that are required for viral accumulation in vivo. MPGAfold, a massively parallel genetic algorithm, suggested the presence of five hairpins (H4a, H4b, and previously identified hairpins H4, H5, and Pr) and one H-type pseudoknot (Ψ3) within the 3′-terminal 194 nucleotides (nt). In vivo compensatory mutagenesis analyses confirmed the existence of H4a, H4b, Ψ3 and a second pseudoknot (Ψ2) previously identified in a TCV satellite RNA. In-line structure probing of the 194-nt fragment supported the coexistence of H4, H4a, H4b, Ψ3 and a pseudoknot that connects H5 and the 3′ end (Ψ1). Stepwise replacements of TCV elements with the comparable elements from Cardamine chlorotic fleck virus indicated that the complete 142-nt 3′ end, and subsets containing Ψ3, H4a, and H4b or Ψ3, H4a, H4b, H5, and Ψ2, form functional domains for virus accumulation in vivo. A new 3-D molecular modeling protocol (RNA2D3D) predicted that H4a, H4b, H5, Ψ3, and Ψ2 are capable of simultaneous existence and bears some resemblance to a tRNA. The related Japanese iris necrotic ring virus does not have comparable domains. These results provide a framework for determining how interconnected elements participate in processes that require 3′ untranslated region sequences such as translation and replication.  相似文献   

15.
Ni BR  Bradford KJ 《Plant physiology》1992,98(3):1057-1068
Mathematical models were developed to characterize the physiological bases of the responses of tomato (Lycopersicon esculentum Mill. cv T5) seed germination to water potential (ψ) and abscisic acid (ABA). Using probit analysis, three parameters were derived that can describe the germination time courses of a seed population at different ψ or ABA levels. For the response of seed germination to reduced ψ, these parameters are the mean base water potential (¯ψb, MPa), the standard deviation of the base water potential among seeds in the population (σψb, MPa), and the “hydrotime constant” (θH, MPa·h). For the response to ABA, they are the log of the mean base ABA concentration ([unk]ABAb, m), the standard deviation of the base ABA concentration among seeds in the population (σABAb, log[m]), and the “ABA-time constant” (θABA, log[m]·h). The values of ¯ψb and [unk]ABAb provide quantitative estimates of the mean sensitivity of germination rate to ψ or ABA, whereas σψb and σABAb account for the variation in sensitivity among seeds in the population. The time constants, θH and θABA, indicate the extent to which germination rate will be affected by a given change in ψ or ABA. Using only these parameters, germination time courses can be predicted with reasonable accuracy at any medium ψ according to the equation probit(g) = [ψ - (θH/tg) - ¯ψb]/σψb, or at any ABA concentration according to the equation probit(g) = [log[ABA] - (θABA/tg) - log[[unk]ABAb]]/σABAb, where tg is the time to radicle emergence of percentage g, and ABA is the ABA concentration (m) in the incubation solution. In the presence of both ABA and reduced ψ, the same parameters can be used to predict seed germination time courses based upon strictly additive effects of ψ and ABA in delaying the time of radicle emergence. Further analysis indicates that ABA and ψ can act both independently and interactively to influence physiological processes preparatory for radicle growth, such as the accumulation of osmotic solutes in the embryo. The models provide quantitative values for the sensitivity of germination to ABA or ψ, allow evaluation of independent and interactive effects of the two factors, and have implications for understanding how ABA and ψ may regulate growth and development.  相似文献   

16.
《Mathematical biosciences》1987,83(2):199-225
Unidirectional extraction of a substrate S in the capillaries following the arterial injection of a bolus containing S and a reference tracer R is assumed to follow first-order kinetics. If CR and CS denote normalized venous effluent concentrations of R and S, respectively, let L(t)=ln[CR(t)⧸CS(t)]. We derive a formula which expresses the experimental L(t) data in terms of the mean μ(t) and variance of the transit times of those capillaries which are contributing indicators at each sample time t. We examine the information thus contained in the L data about capillary and noncapillary transit times under several kinematic assumptions. We show that if the capillary and noncapillary transit times are stochastically independent with frequency functions hc(t) and hav(t), respectively, then the shapes of the graphs of L(t) and μ(t) depend on the variances and skewnesses of hc(t) and hav(t). Specifically, let r2 be the ratio of the variance of hc(t) to the variance of hav(t), and let r3 be the ratio of skewnesses in the same order. Then the graph of μ(t) is concave downward if r2r3 > 1, concave upward if r2r3< 1, and linear if r2r3 = 1. If the fraction of S extracted is not too large, L(t) has nearly the same shape as μ(t), and therefore, L(t) contains information about hc(t) and hav(t).  相似文献   

17.
The kinetic parameters Km, Vmax, Tt (turnover time), and v (natural velocity) were determined for H2 and acetate conversion to methane by Wintergreen Lake sediment, using short-term (a few hours) methods and incubation temperatures of 10 to 14°C. Estimates of the Michaelis-Menten constant, Km, for both the consumption of hydrogen and the conversion of hydrogen to methane by sediment microflora averaged about 0.024 μmol g−1 of dry sediment. The maximal velocity, Vmax, averaged 4.8 μmol of H2 g−1 h−1 for hydrogen consumption and 0.64 μmol of CH4 g−1 h−1 for the conversion of hydrogen to methane during the winter. Estimated natural rates of hydrogen consumption and hydrogen conversion to methane could be calculated from the Michaelis-Menten equation and estimates of Km, Vmax, and the in situ dissolved-hydrogen concentration. These results indicate that methane may not be the only fate of hydrogen in the sediment. Among several potential hydrogen donors tested, only formate stimulated the rate of sediment methanogenesis. Formate conversion to methane was so rapid that an accurate estimate of kinetic parameters was not possible. Kinetic experiments using [2-14C]acetate and sediments collected in the summer indicated that acetate was being converted to methane at or near the maximal rate. A minimum natural rate of acetate conversion to methane was estimated to be about 110 nmol of CH4 g−1 h−1, which was 66% of the Vmax (163 nmol of CH4 g−1 h−1). A 15-min preincubation of sediment with 5.0 × 10−3 atm of hydrogen had a pronounced effect on the kinetic parameters for the conversion of acetate to methane. The acetate pool size, expressed as the term Km + Sn (Sn is in situ substrate concentration), decreased by 37% and Tt decreased by 43%. The Vmax remained relatively constant. A preincubation with hydrogen also caused a 37% decrease in the amount of labeled carbon dioxide produced from the metabolism of [U-14C]valine by sediment heterotrophs.  相似文献   

18.
Two dimensional (2D) NMR and molecular dynamics simulations have been used to determine the three dimensional (3D) structure of a hairpin DNA, d-CTA-GAGGATCC-TUTT-GGATCCT (22mer; abbreviated as U2-hairpin), which has uracil at the second position from the 5′ end of the tetraloop. The 1H resonances of this hairpin have been assigned almost completely. NMR restrained molecular dynamics and energy minimization procedures have been used to describe the 3D structure of U2-hairpin. This study establishes that the stem of the hairpin adopts a right-handed B-DNA conformation, while the T12 and T15 nucleotides stack upon 3′ and 5′ ends of the stem, respectively. Further, T14 stacks upon both T12 and T15. Though U13 partially stacks upon T14, no stacking interaction is observed between U13 and T12. All the individual nucleotide bases belonging to the stem and T12 and T15 of the loop adopt ‘anti’ conformation with respect to their sugar moiety, while the U13 and T14 of the loop are in ‘syn’ conformation. The turning phosphate in the loop is located between T13 and T14. This study and a concurrent NMR structural study on yet another hairpin DNA d-CTAGAGGAATAA-TTTU-GGATCCT (22mer; abbreviated as U4-hairpin), with uracil at the fourth position from the 5′ end of the tetraloop throw light upon various interactions which have been reported between Escherichia coli uracil DNA glycosylase (UDG) and uracil containing DNA. The of T12 and α, β, γ, and ζ of U13 and γ of T14, which partially influence the local conformation of U13 in U2-hairpin are all locked in ‘trans’ conformation. Such stretched out backbone conformation in the vicinity of U13 could be the reason as to why the U2-hairpin is found to be the poor substrate for its interaction with UDG compared to the other substrates in which the uracil is at first, third and fourth positions of the tetraloop from its 5′ end, as reported earlier by Vinay and Varshney. This study shows that UDG actively promotes the flipping of uracil from a stacked conformation and rules out the possibility of UDG recognizing the flipped out uracil bases.  相似文献   

19.
Sphingomonas herbicidovorans MH was able to completely degrade both enantiomers of the chiral herbicide dichlorprop [(RS)-2-(2,4-dichlorophenoxy)propanoic acid], with preferential degradation of the (S) enantiomer over the (R) enantiomer. These results are in agreement with the recently reported enantioselective degradation of mecoprop [(RS)-2-(4-chloro-2-methylphenoxy)propanoic acid] by this bacterium (C. Zipper, K. Nickel, W. Angst, and H.-P. E. Kohler, Appl. Environ. Microbiol. 62:4318–4322, 1996). Uptake of (R)-dichlorprop, (S)-dichlorprop, and 2,4-D (2,4-dichlorophenoxyacetic acid) was inducible. Initial uptake rates of cells grown on the respective substrate showed substrate saturation kinetics with apparent affinity constants (Kt) of 108, 93, and 117 μM and maximal velocities (Vmax) of 19, 10, and 21 nmol min−1 mg of protein−1 for (R)-dichlorprop, (S)-dichlorprop, and 2,4-D, respectively. Transport of (R)-dichlorprop, (S)-dichlorprop, and 2,4-D was completely inhibited by various uncouplers and by nigericin but was only marginally inhibited by valinomycin and by the ATPase inhibitor N,N′-dicyclohexylcarbodiimine. Experiments on the substrate specificity of the putative transport systems revealed that (R)-dichlorprop uptake was inhibited by (R)-mecoprop but not by (S)-mecoprop, (S)-dichlorprop, or 2,4-D. On the other hand, the (S)-dichlorprop transport was inhibited by (S)-mecoprop but not by (R)-mecoprop, (R)-dichlorprop, or 2,4-D. These results provide evidence that the first step in the degradation of dichlorprop, mecoprop, and 2,4-D by S. herbicidovorans is active transport and that three inducible, proton gradient-driven uptake systems exist: one for (R)-dichlorprop and (R)-mecoprop, another for (S)-dichlorprop and (S)-mecoprop, and a third for 2,4-D.  相似文献   

20.
We demonstrate a new, efficient and easy-to-use method for enzymatic synthesis of (stereo-)specific and segmental 13C/15N/2H isotope-labeled single-stranded DNA in amounts sufficient for NMR, based on the highly efficient self-primed PCR. To achieve this, new approaches are introduced and combined. (i) Asymmetric endonuclease double digestion of tandem-repeated PCR product. (ii) T4 DNA ligase mediated ligation of two ssDNA segments. (iii) In vitro dNTP synthesis, consisting of in vitro rNTP synthesis followed by enzymatic stereo-selective reduction of the C2′ of the rNTP, and a one-pot add-up synthesis of dTTP from dUTP. The method is demonstrated on two ssDNAs: (i) a 36-nt three-way junction, selectively 13C9/15N3/2H(1′,2″,3′,4′,5′,5″)-dC labeled and (ii) a 39-nt triple-repeat three-way junction, selectively 13C9/15N3/2H(1′,2″,3′,4′,5′,5″)-dC and 13C9/15N2/2H(1′,2″,3′,4′,5′,5″)-dT labeled in segment C20-C39. Their NMR spectra show the spectral simplification, while the stereo-selective 2H-labeling in the deoxyribose of the dC-residues, straightforwardly provided assignment of their C1′–H2′ and C2′–H2′ resonances. The labeling protocols can be extended to larger ssDNA molecules and to more than two segments.  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