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.
Cytosine-rich RNA copolymers facilitate the template-directed synthesis of complementary oligomers from mononucleotide 5′-phospho-2-methylimidazolides (Fig. 1). The efficiency of this reaction falls sharply as the ratio of cytosine to non-cytosine in the template is decreased. This is a severe limitation for self-replication because it means that any cytosine-rich polynucleotide that can serve as a good template will produce a cytosine-poor complementary strand that is unable to direct further rounds of synthesis. Studies with low-ratio random copolymer templates have shown that the efficiency can be increased by adjusting initial monomer concentrations and by providing additional activated monomers during later stages of the reaction. The oligomeric reaction products can be studied in detail using high performance liquid chromatography. It is possible to separate oligonucleotides on the basis of chain length and base composition. Thus a wealth of information is available to characterize the distribution of products over the course of the reaction and under a variety of reaction conditions.  相似文献   

2.
A K Mazur 《Biopolymers》1984,23(5):859-876
Several research groups have studied depolymerizing processes by enzyme cleavage using product distribution. Although only the shortest chain products can be measured experimentally, interpretation of much evidence requires an assumption of the nature of high polymer distribution. The present work analyzes the high polymer distribution of products on the basis of mathematical models suggested for the two most widespread α-amylase action-mechanisms. The most-probable distribution was found to be distorted by hydrolysis only for the shortest chain products. The size of the distorted region is directly dependent on the characteristics of a specific enzyme. We studied the distribution of products for substrates of varying composition to show that it approaches the most-probable distribution at depolymerization.  相似文献   

3.
酶反应速率方程的普适形式   总被引:1,自引:0,他引:1  
酶反应速率方程的普适形式是应用于相互关联的大规模代谢途径动力学建模的重要方法.把酶反应速率方程写成Michaelis-Menten-King-Altman方程形式可以使得动力学参数(或函数)容易与数据库中的实验数据相接轨,并可以处理任意数量的底物和产物,有利于大规模的计算.普适形式可以同时描述正、负反应方向,并能精确地用于准稳态条件.展示了在三类生物体系中广泛存在的酶反应机制中普适方程的严格推导过程,并讨论了普适方程的特点,针对不可逆反应酶反应产生的产物抑制效应可以自然消除,总结了在普适速率方程中体现调节剂的作用和协同作用.  相似文献   

4.
Hyaluronidase-catalyzed copolymerization was carried out with monomer combinations of 2-methyl (1a)/2-vinyl (1b), 2-methyl (1a)/2-ethyl (1c), 2-methyl (1a)/2-n-propyl (1d), and 2-vinyl (1b)/2-ethyl (1c) oxazoline derivatives of hyalobiuronate [GlcAbeta(1-->3)GlcN]. All copolymerization reactions proceeded successfully in a regio and stereoselective manner, giving rise to hyaluronan derivatives bearing different N-acyl groups at the C2 position of the glucosamine unit in the polymer chain. The composition of the N-acyl groups was controlled by varying the comonomer feed ratio. The copolymerization mechanism was also discussed.  相似文献   

5.
A new synthetic procedure for the monomer, 5(6)-vinylbenzimidazole, has been developed. Also, a new method for the formation of the benzimidazole moiety is reported. The monomer, 5(6)-vinylbenzimidazole, was subjected to a free radical solution polymerization and a solid state thermal polymerization. Poly[5(6)-vinyl-benzimidazole], prepared by the solution-free radical polymerization, was shown to be a linear addition polymer with pendant benzimidazole groups, whereas the thermally prepared polymer was shown to have a different structure. The solvolyses of negatively charged esters with varying aliphatic chain lengths catalyzed by poly[5(6)-vinylbenzimidazole] and compared to monomeric benzimidazole in 40% 1-propanol-water at 26°C revealed that the polymer was more efficient and that its activity was a function of the degree of ionization. Neutral-neutral and neutral-anionic benzimidazole interactions are suggested. The polymer conformation was found to have a significant effect on the solvolysis reactions.  相似文献   

6.
In this study, we synthesized a new cell immobilization support having poly(glycidyl methacrylate) as a graft polymer chain and used this support for cell cultivation. Base polymer particle was synthesized by suspension polymerization and epoxy polymer chain was extended from particle surface on graft polymerization. Produced polymer particles had broad particle size distribution ranging from 20 to 1000 μm and the degree of polymerization of grafted polymer chain was ranged from 500 to 1000. The effects of various factors, such as grafted polymer chain length and its surface density, composition of base polymer network and graft polymer chain, on the cell growth of murine fibroblast cell line (MS-5 cell) on polymer particle were studied. This polymer particle could cultivate not only fibroblast cell line but also epidermal cell line (HeLa cell), osteoblast cell line (MC3T3E1 cell), and chondrocyte cell line (ch-8 cell) on its surface. Growth rate is almost the same as that of cells using poly(styrene) tissue culture dish. To apply this cell cultivation system for examination of cell co-culture, HeLa cell immobilized on 100 μm of polymer particle was successfully co-cultured with MS-5 cell immobilized on 300 μm of polymer particle for four weeks.  相似文献   

7.
To gain better insight into mechanistic features of enzyme-catalyzed malolactonate polymerization, reactions with propyl malolactonate were analyzed while varying enzyme concentration, reaction media composition, and reaction temperature. Monomer conversion and product molecular weights were characterized by (1)H NMR and MALDI-TOF MS, respectively. A high extent of thermal polymerization of propyl malolactonate was observed, while the polymer chain length in all reactions was controlled by the elimination of alpha-hydrogen from propyl malolactonate with formation of a new initiator and the new chains. The most efficient enzymatic catalysis occurred in toluene (2.11 M monomer) at 60 degrees C. Candida rugosa lipase (10 wt %) accelerated polymerization 25-fold over the rate of thermal polymerization. The maximum poly(propyl malate) number-average molecular weight obtained was 5000 Da at 20 wt % enzyme with a polydispersity of 1.15. These values compare with 1800 Da and 1.5, respectively, in the absence of enzyme.  相似文献   

8.
Summary One of the most important sets of model prebiotic experiments consists of reactions that synthesize complementary oligonucleotides from preformed templates under nonenzymatic conditions. Most of these experiments are conducted at 4°C using 0.01–0.1 M concentrations of activated nucleotide monomer and template (monomer equivalent). In an attempt to extend the conditions under which this type of reaction can occur, we have concentrated the reactants by freezing at –18°C, which is close to the NaCl–H2O eutectic at –21°C.The results from this set of experiments suggest that successful syntheses can occur with poly(C) concentrations as low at 5×10–4 M and 2MeImpG concentrations at 10–3 M. It was also anticipated that this mechanism might allow the previously unsuccessful poly(A)-directed synthesis of oligo(U)s to occur. However, no template effect was seen with the poly(A) and ImpU system. The failure of these conditions to allow template-directed synthesis of oligo(U)s supports the previously proposed idea that pyrimidines may not have been part of the earliest genetic material.Because of the low concentrations of monomer and template that would be expected from prebiotic syntheses, this lower temperature could be considered a more plausible geologic setting for template-directed synthesis than the standard reaction conditions.  相似文献   

9.
We derive the analytical form of a rate-equilibrium free-energy relationship (with slope Phi) for a bounded, linear chain of coupled reactions having arbitrary connecting rate constants. The results confirm previous simulation studies showing that Phi-values reflect the position of the perturbed reaction within the chain, with reactions occurring earlier in the sequence producing higher Phi-values than those occurring later in the sequence. The derivation includes an expression for the transmission coefficients of the overall reaction based on the rate constants of an arbitrary, discrete, finite Markov chain. The results indicate that experimental Phi-values can be used to calculate the relative heights of the energy barriers between intermediate states of the chain but provide no information about the energies of the wells along the reaction path. Application of the equations to the case of diliganded acetylcholine receptor channel gating suggests that the transition-state ensemble for this reaction is nearly flat. Although this mechanism accounts for many of the basic features of diliganded and unliganded acetylcholine receptor channel gating, the experimental rate-equilibrium free-energy relationships appear to be more linear than those predicted by the theory.  相似文献   

10.
Polyelectrolyte solutions are often described by structural theories. These theories in some cases yield values for the counterion concentration at the charged monomer surface that exceed the saturation concentration. This means a change of the ion properties due to ion immobilization or ion condensation in close vicinity to the polymer chain. The extent of this counterion condensation (CIC) and the respective surface potential are calculated from the saturation concentrations of the electrolyte involved including the influence of ion hydration on the effective dielectric number. In this paper, we shall consider all these influences by a fundamental differential equation and a set of explicit formulae yielding quantitative expressions without linearization. All calculations are based on the abstraction of an idealized elementary cell.  相似文献   

11.
PH. Dumas  P. Sigwalt 《Chirality》1991,3(6):484-491
The polymerization of racemic methylthiirane in homogeneous phase, initiated by bis(isopropyl-S-cysteinato) cadmium is a living process. The resulting polymers are isotactic and optically active at partial conversion. The optical purity of the residual monomer may reach 27% at half conversion. The propagation occurs mainly on one valency of Cd, however oligomers grow slowly on the second valency. The stereoregularity of the polymer chain appears only when the length of the oligomer becomes high enough, making possible a bicoordination of the Cd counterions. The stereoregularity of the polymer is characterized by the molar fraction σ of isotactic diads which varies from 0.5 for atactic chains—formed at the beginning—to about one for isotactic segments formed for longer chains. The stereospecifictity also depends on temperature of propagation and on initiator concentration. The kinetics observed (zero order in monomer and one-half in Cd) are explained by monomer coordination before insertion and dimeric association of the thiolate end groups. The enantioasymmetric process observed results from an unbalance in the number of the two different types of active sites and possibly from a difference in their reactivities. Enantioasymmetry has been found to decrease significantly when the dielectric constant ε of the medium increases.  相似文献   

12.
J E Ferrell  K J Lee  W H Huestis 《Biochemistry》1985,24(12):2857-2864
The rate of phospholipid transfer from sonicated phospholipid vesicles to human erythrocytes has been studied as a function of membrane concentration and lipid acyl chain composition. Phospholipid transfer exhibits saturable first-order kinetics with respect to both cell and vesicle membrane concentrations. This kinetic behavior is consistent either with transfer during transient contact between cell and vesicle surfaces (but only if the fraction of the cell surface susceptible to such interaction is small) or with transfer of monomers through the aqueous phase. The acyl chain composition of the transferred phospholipid affects the transfer kinetics profoundly; for homologous saturated phosphatidylcholines, the rate of transfer decreases exponentially with increasing acyl chain length. This behavior is consistent with passage of phospholipid monomers through a polar phase, which might be the bulk aqueous phase( as in the monomer transfer model) or the hydrated head-group regions of a cell-vesicle complex (transient collision model). Collisional transfer also predicts that intercell transfer of phospholipids should be slow compared to cell-vesicle transfer, as surface charge and steric effects should prevent close apposition of donor and acceptor membranes. This is not found; dilauroylphosphatidylcholine transfers rapidly between red cells. Thus, the observed relationship between acyl chain length and intermembrane phospholipid transfer rates likely reflects the energetics of monomer transfer through the aqueous phase.  相似文献   

13.
The optical rotatory dispersion of copolymers of O-carbobenzoxy-L -tyrosine and benzyl L (or D )-glutamate as well as benzyl L -aspartate, dissolved in nonpolar solvent, has been studied. Moffitt's equation permits the determination of b0 coefficients whose variation, with varying composition in amino acid residues, suggests that the molecules of poly-O-carbobenzoxy-L -tyrosine have a helical structure similar to that of poly-(benzyl L -glutamate). Results obtained from infrared spectroscopy and x-ray diffraction show that the copolymers possess a helical conformation in the solid state, even when they are very rich in carbobenzoxy-L -tyrosine residues. The value of the b0, coefficient for poly-O-carbobenzoxy-L -tyrosine may be explained by a regular stacking of the chromophore groups around the helical backbone. The ordering of the molecules of this polymer in a purely helical structure seems favored by the insertion of a small number of foreign residues in the polypeptide chain.  相似文献   

14.
We have considered the partitioning of tubulin between monomer and polymer in the cell under conditions of dynamic instability. Dynamic instability adds to the on and off rate constant of steady-state dynamics’ new parameters: (1) the rate at which growing microtubules transit to a shrinking phase; and (2) the rate at which shrinking microtubules transit to the growing phase. Under these conditions the free-monomer concentration in the cell increases with total tubulin if the number of nucleating sites is fixed. If the number of nucleating sites increases at fixed total tubulin, subunits shift from the monomer to the polymer phase. These important properties deviate from the traditional equilibrium and steady-state theories and have important implications for the biosynthetic regulation of tubulin.  相似文献   

15.
The enzyme exonuclease I from Escherichia coli hydrolyzes successive nucleotides from the 3'-termini of single-stranded deoxyribonucleotide homopolymers. When the reaction is stopped after partial hydrolysis, only intact starting material and small oligomers can be isolated. The distribution of oligomeric products varies with the base composition of the polymer but the largest oligomer that can be isolated from the reaction of exonuclease I with homopolymers of deoxyadenylate, deoxythymidylate, or deoxycytidylate is a decamer. These results suggest a model in which exonuclease I possesses at least two nucleotide binding sites. When both sites are filled, with 11-mers and longer polymers, the enzyme does not dissociate from the polymer during hydrolysis. When, with smaller oligomers, only a single site is filled, the reaction partitions at each oligomer between hydrolysis and dissociation. The kinetics of the reactions of exonuclease I with purified polydeoxyriboadenylates of defined size distributions have been investigated. The maximum rates of hydrolysis are nearly independent of polymer size while the apparent Michaelis constants are inversely proportional to the polymer size. A simple steady state model yields a kinetic equation that is consistent with our results. Competition experiments indicate that the rate at which exonuclease I associates with the 3'-terminus of a polydeoxyribonucleotide is independent of the polymer's chain length.  相似文献   

16.
This work describes the generation of novel PHAs (named PHACOS) with a new monomer composition containing thioester groups in the side chain, which confers new properties and made them suitable for chemical modifications after their biosynthesis. We have analyzed the PHACOS production abilities of the wild-type strain Pseudomonas putida KT2442 vs. its derived strain P. putida KT42FadB, mutated in the fadB gene from the central metabolic β-oxidation pathway involved in the synthesis of medium-chain-length PHA (mcl-PHA). Different fermentation strategies based on one- or two-stage cultures have been tested resulting in PHACOS with different monomer composition. Using decanoic acid as inducer of the growth and polymer synthesis and 6-acetylthiohexanoic acid as PHA precursor in a two-stage strategy, the maximum yield was obtained by culturing the strain KT42FadB. Nuclear magnetic resonance and gas chromatography coupled to mass spectrometry showed that polymers obtained from the wild-type and KT42FadB strains, included 6-acetylthio-3-hydroxyhexanoic acid (OH-6ATH) and the shorter derivative 4-acetylthio-3-hydroxybutanoic acid (OH-4ATB) in their composition, although in different ratios. While the polymer obtained from KT42FadB strain contained mainly OH-6ATH monomer units, mcl-PHA produced by the wild-type strain contained OH-6ATH and OH-4ATB. Furthermore, polyesters showed differences in the OH-alkyl derivates moiety. The strain KT42FadB overproduced PHACOS when compared to the production rate of the control strain in one- and two-stage cultures. Thermal properties obtained by differential scanning calorimetry indicated that both polymers have different glass transition temperatures related to their composition.  相似文献   

17.
Walker SI  Grover MA  Hud NV 《PloS one》2012,7(4):e34166
Many models for the origin of life have focused on understanding how evolution can drive the refinement of a preexisting enzyme, such as the evolution of efficient replicase activity. Here we present a model for what was, arguably, an even earlier stage of chemical evolution, when polymer sequence diversity was generated and sustained before, and during, the onset of functional selection. The model includes regular environmental cycles (e.g. hydration-dehydration cycles) that drive polymers between times of replication and functional activity, which coincide with times of different monomer and polymer diffusivity. Template-directed replication of informational polymers, which takes place during the dehydration stage of each cycle, is considered to be sequence-independent. New sequences are generated by spontaneous polymer formation, and all sequences compete for a finite monomer resource that is recycled via reversible polymerization. Kinetic Monte Carlo simulations demonstrate that this proposed prebiotic scenario provides a robust mechanism for the exploration of sequence space. Introduction of a polymer sequence with monomer synthetase activity illustrates that functional sequences can become established in a preexisting pool of otherwise non-functional sequences. Functional selection does not dominate system dynamics and sequence diversity remains high, permitting the emergence and spread of more than one functional sequence. It is also observed that polymers spontaneously form clusters in simulations where polymers diffuse more slowly than monomers, a feature that is reminiscent of a previous proposal that the earliest stages of life could have been defined by the collective evolution of a system-wide cooperation of polymer aggregates. Overall, the results presented demonstrate the merits of considering plausible prebiotic polymer chemistries and environments that would have allowed for the rapid turnover of monomer resources and for regularly varying monomer/polymer diffusivities.  相似文献   

18.
A double nucleation mechanism for the polymerization of sickle hemoglobin is described. The mechanism accounts for all of the major kinetic observations: the appearance of a delay, the high concentration dependence of the delay time, and the stochastic behavior of slowly polymerizing samples in small volumes. The mechanism postulates that there are two pathways for polymer formation: polymerization is initiated by homogeneous nucleation in the solution phase, followed by nucleation of additional polymers on the surface of existing ones. This second pathway is called heterogeneous nucleation. Since the surface of polymers is continuously increasing with time, heterogeneous nucleation provides a mechanism for the extreme autocatalysis that is manifested as an apparent delay in the kinetic progress curves. In this mechanism, each spherulitic domain of polymers is considered to be initiated by a single homogeneous nucleation event. The mechanism explains the irreproducibility of the delay time for single domain formation as arising from stochastic fluctuations in the time at which the homogeneous nucleus for the first polymer is formed. Integration of the linearized rate equations that describe this model results in a simple kinetic form: A[cosh(Bt)-1] (Bishop & Ferrone, 1984). In the accompanying paper (Ferrone et al., 1985) it was shown that the initial 10 to 15% of progress curves, with delay times varying from a few milliseconds to over 10(5) seconds, is well fit by this equation. In this paper, we present an approximate statistical thermodynamic treatment of the equilibrium nucleation processes that shows how the nucleus sizes and nucleation equilibrium constants depend on monomer concentration. The equilibrium model results in expressions for B and B2A as a function of monomer concentration in terms of five adjustable parameters: the bimolecular addition rate of a monomer to the growing aggregate, the fraction of polymerized monomers that serve as heterogeneous nucleation sites, the free energy of intermolecular bonding within the polymer, and two parameters that describe the free energy change as a function of size for the bonding of the heterogeneous nucleus to a polymer surface. This model provides an excellent fit to the data for B and B2A as a function of concentration using physically reasonable parameters. The model also correctly predicts the time regime in which stochastic behavior is observed for polymerization in small volumes.  相似文献   

19.
G F Joyce 《Origins of life》1984,14(1-4):613-620
Cytosine-rich RNA copolymers facilitate the template-directed synthesis of complementary oligomers from mononucleotide 5'-phospho-2-methylimidazolides (Fig. 1). The efficiency of this reaction falls sharply as the ratio of cytosine to non-cytosine in the template is decreased. This is a severe limitation for self-replication because it means that any cytosine-rich polynucleotide that can serve as a good template will produce a cytosine-poor complementary strand that is unable to direct further rounds of synthesis. Studies with low-ratio random copolymer templates have shown that the efficiency can be increased by adjusting initial monomer concentrations and by providing additional activated monomers during later stages of the reaction. The oligomeric reaction products can be studied in detail using high performance liquid chromatography. It is possible to separate oligonucleotides on the basis of chain length and base composition. Thus a wealth of information is available to characterize the distribution of products over the course of the reaction and under a variety of reaction conditions.  相似文献   

20.
Effects of proteolytic modifications of the DNase-I-binding loop (residues 39-51) in subdomain 2 of actin on F-actin dynamics were investigated by measuring the rates of the polymer subunit exchange with the monomer pool at steady state and of ATP hydrolysis associated with it, and by determination of relative rate constants for monomer addition to and dissociation from the polymer ends. Cleavage of actin between Gly-42 and Val-43 by protease ECP32 resulted in enhancement of the turnover rate of polymer subunits by an order of magnitude or more, in contrast to less than a threefold increase produced by subtilisin cleavage between Met-47 and Gly-48. Probing the structure of the modified actins by limited digestion with trypsin revealed a correlation between the increased F-actin dynamics and a change in the conformation of subdomain 2, indicating a more open state of the filament subunits relative to intact F-actin. The cleavage with trypsin and steady-state ATPase were cooperatively inhibited by phalloidin, with half-maximal effects at phalloidin to actin molar ratio of 1:8 and full inhibition at a 1:1 ratio. The results support F-actin models in which only the N-terminal segment of loop 39-51 is involved in monomer-monomer contacts, and suggest a possibility of regulation of actin dynamics in the cell through allosteric effects on this segment of the actin polypeptide chain.  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