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.
The N-terminal amino acid sequence of human C \(\bar l\) s b chain has been extended to 52 residues. The histidine residue involved in the charge-relay system is located at position 38, whereas the ‘histidine-loop’ disulphide bridge is missing. So far, human complement subcomponents C \(\bar l\) r and C \(\bar l\) s are the only known mammalian serine proteinases lacking this disulphide bridge.  相似文献   

2.
The behaviour of similar coupled non-linear oscillators of the type \(\dot x\) =f(x, y, µ \(\dot y\) =g(x, y, µ is to be investigated. The oscillators are assumed to be coupled by diffusion gradients. If some conditions on the magnitude of the diffusion coefficients are satisfied, it is proved that: 1) if the oscillators have the same period (identical value of the parameter μ) and different phases before coupling, after coupling they tend to synchronize the phases; 2) if the periods of the oscillators are not too different (in terms of the values of the parameter μ) before coupling, after coupling they tend to oscillate with the same period. It is suggested the possible role of diffusion as a synchronizing mechanism in some biological phenomena.  相似文献   

3.
Kinetic models of the F0F1-ATPase able to transport H+ or/and Na+ ions are proposed. It is assumed that (i) H+ and Na+ compete for the same binding sites, (ii) ion translocation through F0 is coupled to the rate-limiting step of the F1-catalyzed reaction. The main characteristics of the dependences of ATP synthesis and hydrolysis rates on Δφ, ΔpH, and ΔpNa are predicted for various versions of the coupling model. The mechanism of the switchover from \(\Delta \bar \mu _{H^ + } \) -dependent synthesis to the \(\Delta \bar \mu _{Na^ + } \) -dependent one is demonstrated. It is shown that even with a drastic drop in \(\Delta \bar \mu _{H^ + } \) , ATP hydrolysis by the proton mode of catalysis can be effectively inhibited by Δφ and ΔpNa. The results obtained strongly support the possibility that the same F0F1-ATPase in bacterial cells can utilize both \(\Delta \bar \mu _{H^ + } \) and \(\Delta \bar \mu _{Na^ + } \) for ATP synthesis underin vivo conditions.  相似文献   

4.
Metabolic, body temperature, and cardiorespiratory responses of 16 healthy middle-aged (40–57 years) men, 9 nonsmokers and 7 smokers, were obtained during tests of maximal aerobic power at ambient environmental temperatures of 25 ± 0.5 and 35 ± 0.5°C and 20% relative humidity under four conditions: (a) filtered air, FA; (b) 50 ppm carbon monoxide in filtered air, CO; (c) 0.27 ppm peroxyacetylnitrate in filtered air, PAN; and (d) a combination of all three mixtures, PANCO. There was no significant change in maximum aerobic power \(\left( {\dot VO2max} \right)\) related to the presence of air pollutants, although total working time was lowered in the 25°C environment while breathing CO. Older nonsmokers did have a decrement in \(\left( {\dot VO2max} \right)\) while breathing 50 ppm CO, while older smokers failed to show any change. This difference was related to the initial COHb levels of the smokers, who, when breathing this level of ambient CO, had only a 14% increase in COHb over their initial levels in contrast to the 200% increase in the nonsmokers. Smoking habits were the most influential factor affecting the cardiorespiratory responses of these older men to maximal exercise. Regardless of ambient conditions, smokers had a significantly lower (27%) aerobic power than nonsmokers, were breathing closer to their maximal breathing capacities throughout the walk, and had a higher respiratory exchange ratio. While the \(\left( {\dot VO2max} \right)\) of nonsmokers was only 6% less than that of younger nonsmoking males ( \(\bar x\) age = 25 years) working under similar conditions, the aerobic power of the older smokers was 26% lower than that of young smokers ( \(\bar x\) age = 24 years).  相似文献   

5.
There have been two contrasting doctrines concerning learning, more generally about acquisition of knowledge: empiricism and rationalism. The theory of learning in such a field as artificial intelligence seems to fall within the empiricist framework. On the hand, N. Chomsky and his followers have discussed, during the last decade, concerning learning, especially about language learning, from the rationalist point of view (Chomsky, 1965). The main feature in the rationalist approach toward a theory of learning lies in the speculation that in order to acquire knowledge it is indispensable for a learner to be endowed with “innate ideas”, and that “experience” in the external world are merely subsidiary types of information for the learner. If this is acceptable, we can inquire: Under what kind of innate ideas can the learner understand the structure of the external world? In our previous paper (Uesaka, Aizawa, Ebara, and Ozeki, 1973), we formalized this by introducing the mathematical notion of “learnability”, and gave a partial answer to the above inquiry. In this formalization we assumed that the set F of objects to be learned consists of mappings of N to itself, where N is the set of positive integers. Then, constructing a topological space (F, \(\mathcal{O}\) ) by an appropriate family \(\mathcal{O}\) of open sets, we observed that the notion of learnability can be well described in terms of topological properties of the learning space (F, \(\mathcal{O}\) ). Many problems must be solved, however, before we raise the theory to a complete model of the rationalist theory of learning. The topological study of the space (F, \(\mathcal{O}\) ) is, we believe, the first step toward this approach. In this context, we discuss the topological aspects of this space. Now we define \(\mathcal{O}\) as follows: By N 2 we mean the direct product of two N's. Let s be a subset of N 2. If, for any (x, y), (x′, y′) in s, x=x′ implies y=y′, then we say that s is single-valued. Let fF, If, for any (x, y) in s, y=f(x), then f is said to be on s, denoted as \(f\underline \supseteq s\) . Let \(\pi \left( s \right) = \left\{ {g;g \in F,g\underline \supseteq s} \right\}\) . A single-valued finite subset of N 2 is called datum. Let D denote the family of all data. Let \(\mathcal{O}* = \left\{ \phi \right\} \cup \left\{ {\pi \left( d \right);d \in D} \right\}\) , and \(\mathcal{O}\) denote the family of all subsets of F, each of which is written as \(\mathop \cup \limits_\alpha W_{\alpha }\) , where W α is in \(\mathcal{O}*\) . Then, it is easily seen that \(\mathcal{O}\) satisfies the axiom of the open system of a topological space. It is shown that the learning space (F, \(\mathcal{O}\) ) has the following properties:
  1. It satisfies the first and the second countability axioms.
  2. It is separable and is totally disconnected.
  3. It is a Hausdorff space and, further, is regular and normal.
  4. It is neither compact nor locally compact.
  5. It is metrizable, or more precisely there exists a complete but not totally bounded metric space which is homeomorphic to learning space.
  6. Any of its subspace can be embedded into its special subspace.
  相似文献   

6.
We postulate that the biomass distribution function for an ecological population may be derived from the condition that the biomas diversity functional is maximal subject to an energetic constraint on the total biomass. This leads to a biomass distribution of the form \(p(m) = \bar m^{ - 1} \exp ( - m/\bar m)\) , where \(\bar m\) is the mean biomass per individual. The same condition yields a unique value for the biomass diversity functional. These predictions are tested against fishery data and found to be in good agreement. It is argued that the existence of a unique value for biomass diversity may provide a preliminary theoretical foundation for the observed upper limit to species diversity.  相似文献   

7.
8.
Effects of a peptide hormone—human recombinant erythropoietin (EPO)—on transmembrane potential (TMP) and a number of active mitochondria in rat thymocytes were studied in vitro using the fluorescent cationic probe 4-(p-dimethylaminostyryl)-1-methylpyridinium (DSM). It was established that EPO changes electric potentials on the surface of cellular membranes of rat thymocytes. Changes in fluorescent signals of DSM were found to depend on the EPO concentration and physiological status of thymus cells. EPO concentrations sufficient to increase average fluorescence intensity of DSM ( $\tilde F$ ) in thymus cell mitochondria were established in vitro. These effects were stimulated by increasing the average number of energized mitochondria (Ñ m ) able to accumulate the fluorescent cationic probe. These changes can also be due to elevation of proton potential on the mitochondrial membrane and/or of electric potential of the plasma membrane. Experimental values of positive correlation coefficients between mean values of $\tilde F$ and Ñ m in experimental (EPO) thymus cell samples differed from those in control. In the presence of EPO, $\tilde F$ increased nonlinearly with Ñ m due to different responsiveness of thymocytes to EPO and change in polarization of the outer mitochondrial membrane in some EPO-stimulated cells. In the majority of cases, the peak in the distribution histogram of $\tilde F$ was shifted towards augmentation of the DSM signal. The EPO response of cells isolated from different thymuses depended on the initial level of the average fluorescent signal of DSM in mitochondria, which testifies to differences in physiological statuses of animal thymuses and experimental animals at large.  相似文献   

9.
10.
Over the years numerous models of \(SIS\) (susceptible \(\rightarrow \) infected \(\rightarrow \) susceptible) disease dynamics unfolding on networks have been proposed. Here, we discuss the links between many of these models and how they can be viewed as more general motif-based models. We illustrate how the different models can be derived from one another and, where this is not possible, discuss extensions to established models that enables this derivation. We also derive a general result for the exact differential equations for the expected number of an arbitrary motif directly from the Kolmogorov/master equations and conclude with a comparison of the performance of the different closed systems of equations on networks of varying structure.  相似文献   

11.
Augmentation of the mechanical properties of connective tissue using ultraviolet (UV) radiation—by targeting collagen cross-linking in the tissue at predetermined UV exposure time \((t)\) and wavelength \((\lambda )\) —has been proposed as a therapeutic method for supporting the treatment for structural-related injuries and pathologies. However, the effects of \(\lambda \) and \(t\) on the tissue elasticity, namely elastic modulus \((E)\) and modulus of resilience \((u_\mathrm{Y})\) , are not entirely clear. We present a thermomechanical framework to reconcile the \(t\) - and \(\lambda \) -related effects on \(E\) and \(u_\mathrm{Y}\) . The framework addresses (1) an energy transfer model to describe the dependence of the absorbed UV photon energy, \(\xi \) , per unit mass of the tissue on \(t\) and \(\lambda \) , (2) an intervening thermodynamic shear-related parameter, \(G\) , to quantify the extent of UV-induced cross-linking in the tissue, (3) a threshold model for the \(G\) versus \(\xi \) relationship, characterized by   \(t_\mathrm{C}\) —the critical \(t\) underpinning the association of \(\xi \) with \(G\) —and (4) the role of \(G\) in the tissue elasticity. We hypothesized that \(G\) regulates \(E\) (UV-stiffening hypothesis) and \(u_\mathrm{Y}\) (UV-resilience hypothesis). The framework was evaluated with the support from data derived from tensile testing on isolated ligament fascicles, treated with two levels of \(\lambda \) (365 and 254 nm) and three levels of \(t\) (15, 30 and 60 min). Predictions from the energy transfer model corroborated the findings from a two-factor analysis of variance of the effects of \(t\) and \(\lambda \) treatments. Student’s t test revealed positive change in \(E\) and \(u_\mathrm{Y}\) with increases in \(G\) —the findings lend support to the hypotheses, implicating the implicit dependence of UV-induced cross-links on \(t\) and \(\lambda \) for directing tissue stiffness and resilience. From a practical perspective, the study is a step in the direction to establish a UV irradiation treatment protocol for effective control of exogenous cross-linking in connective tissues.  相似文献   

12.
The parameters of the reciprocal function (“Reziprokfunktion”) for the growth of herrings(Clupea harengus) are calculated from older and more recent measurements. The logarithmic expression of the proposed reciprocal function is as follows: \(\log y_x = \log y_{max} - \frac{1}{{\chi + \xi }}\log N\) . Values less than 1 are found for the additive age (ξ). In further calculations 0.4 is used as the estimated mean value. Measurements made before the second world war yield ca. 30 cm for the maximum value (Lmax). After this period the maximum values increase to ca. 34 cm. The Scandinavian and Atlantic herrings differ from North Sea herrings by higher maximum values. The values for the constant of velocity (log N) may be different for identical ξ and Ymax values. The velocity constant determines the position of the inflection point of the growth curve. The dimension, which is only dependent on the maximum value, is at the inflection point: \(\frac{{Y_{max} }}{{7,389}}\) . From the results ofSchumacher (1967) on the growth of 3 herring populations from the North Sea it was calculated that the values for the constant of velocity rise from northern to southern areas. A low value for the constant of velocity marks an early inflection point and a high velocity of growth before this point and vice versa. The growth of the 3 populations tends to almost the same maximum value; consequently, a high velocity before the inflection point is compensated by a lower velocity after this point and vice versa. The maximum velocity of linear growth at the point of inflection is given by the expression \(\frac{{Y_{max} }}{{4,25 \cdot \log N}}\) . This expression may possibly be a useful device for quantitative comparisons of growth processes.  相似文献   

13.
14.

Background

The basic RNA secondary structure prediction problem or single sequence folding problem (SSF) was solved 35 years ago by a now well-known \(O(n^3)\)-time dynamic programming method. Recently three methodologies—Valiant, Four-Russians, and Sparsification—have been applied to speedup RNA secondary structure prediction. The sparsification method exploits two properties of the input: the number of subsequence Z with the endpoints belonging to the optimal folding set and the maximum number base-pairs L. These sparsity properties satisfy \(0 \le L \le n / 2\) and \(n \le Z \le n^2 / 2\), and the method reduces the algorithmic running time to O(LZ). While the Four-Russians method utilizes tabling partial results.

Results

In this paper, we explore three different algorithmic speedups. We first expand the reformulate the single sequence folding Four-Russians \(\Theta \left(\frac{n^3}{\log ^2 n}\right)\)-time algorithm, to utilize an on-demand lookup table. Second, we create a framework that combines the fastest Sparsification and new fastest on-demand Four-Russians methods. This combined method has worst-case running time of \(O(\tilde{L}\tilde{Z})\), where \(\frac{{L}}{\log n} \le \tilde{L}\le min\left({L},\frac{n}{\log n}\right)\) and \(\frac{{Z}}{\log n}\le \tilde{Z} \le min\left({Z},\frac{n^2}{\log n}\right)\). Third we update the Four-Russians formulation to achieve an on-demand \(O( n^2/ \log ^2n )\)-time parallel algorithm. This then leads to an asymptotic speedup of \(O(\tilde{L}\tilde{Z_j})\) where \(\frac{{Z_j}}{\log n}\le \tilde{Z_j} \le min\left({Z_j},\frac{n}{\log n}\right)\) and \(Z_j\) the number of subsequence with the endpoint j belonging to the optimal folding set.

Conclusions

The on-demand formulation not only removes all extraneous computation and allows us to incorporate more realistic scoring schemes, but leads us to take advantage of the sparsity properties. Through asymptotic analysis and empirical testing on the base-pair maximization variant and a more biologically informative scoring scheme, we show that this Sparse Four-Russians framework is able to achieve a speedup on every problem instance, that is asymptotically never worse, and empirically better than achieved by the minimum of the two methods alone.
  相似文献   

15.
16.
Mark A. Chappell 《Oecologia》1983,56(1):126-131
Temperature regulation and oxygen consumption were examined in two species of grasshoppers: Melanoplus sanguinipes from cold alpine tundra at elevation 3,800 m, and Trimerotropis pallidipennis from hot desert habitats at elevation 250 m. Both species utilized behavioral thermoregulation to keep body temperature (T b ) more constant than environmental temperatures (T e ) during the day. The difference in average T b in the two species was much less than the difference in T e 's. Microclimate measurements indicate that temperature regulation is not difficult for M. sanguinipes, but T. pallidipennis must restrict activity for much of the day to avoid heat stress and can easily overheat if it moves into sunlit areas. Oxygen consumption ( \(\dot V{\text{O}}_{\text{2}} \) ) at average T b and total daily energy expenditures are higher in M. sanguinipes than in T. pallidipennis, as is the Q10 for \(\dot V{\text{O}}_{\text{2}} \) . These differences may be related to different strategies for energy utilization and predator avoidance.  相似文献   

17.
The number ( \(\bar X\) =2.4) ofEucelatoria sp. maggots that completed development in 4th- or 5th-instar larvae of the tobacco budworm (TBW),Heliothis virescens (F.), was significantly greater (P<0.05) than the number ( \(\bar X\) =1.2) that completed development in 3rd-instar larvae. Maggot development time decreased with increasing number of maggots per host larva. It also decreased with advancing larval instars. The range was 6.9±1.1 days in early 3rd-instar TBW larvae and 5.0±0.8 days in early 5th-instar TBW larvae. Unparasitized 3rd- or 4th-instar TBW larvae consumed significantly more food than did similar aged larvae parasitized byEucelatoria sp., but larvae parasitized during the early 5th-instar consumed more food than did similar aged unparasitized larvae. Consumption by 4th- or 5th-instar larvae increased significantly as maggot densities increased from 1 to 3 per host larva, but decreased at a density of 4 or more maggots per host larva. Although body weight gain and consumption were both significantly reduced 48 and 120 h after parasitization of late 3rd-instar larvae (6 days old), the approximate digestibility (AD) value was significantly greater for parasitized than for unparasitized larvae. Unparasitized larvae were more efficient in converting digested food to body substance (ECD) than parasitized larvae, but the efficiency in conversion of ingested food to body substance (ECI) was similar for both parasitized and unparasitized larvae.  相似文献   

18.
All of the cells of the upper (adaxial) epidermis of the leaves ofOxalis carnosa are transformed into large bladders, while in the lower epidermis the bladder cells are interrupted by “normal” cells with stomata. The epidermal bladders contain a high concentration of free oxalic acid (pH approx. 1). Water-relations parameters of these epidermal bladder cells have been determined using the pressure probe. Original cell turgor (P0) of the closely packed bladders of theupper epidermis was P0=0.7 to 2.9 bar ( \(\overline {P_0 } = 1.7 \pm 0.5 bar\) ; mean±SD;N=25 cells) and lower than that in the club-shaped bladders of thelower epidermis (P0=1.3 to 3.7 bar; \(\overline {P_0 } = 2.5 \pm 0.7 bar\) ;N=25 cells). Large differences in the elastic modulus (ε) and the hydraulic conductivity (Lp) of the two different types of cells were observed. For the lower epidermal bladders, ε=18 to 166 bar and was similar to that of other higher plant cells. Also, for these cells it was found that ε was increasing with both, cell turgor and cell volume. By contrast, ε of the cells of the upper epidermis was by one order of magnitude smaller (ε=1.9 to 17.0 bar) and no dependence of ε on cell volume could be detected. The Lp values of the cell membranes were also different (lower epidermis: \(\overline {Lp} = (2.3 \pm 1.6) \cdot 10^{ - 5} cm s^{ - 1} bar^{ - 1}\) ; upper epidermis: \(\overline {Lp} = (3.8 \pm 2.4) \cdot 10^{ - 6} cm s^{ - 1} bar^{ - 1}\) ). These differences seem to be too large to be caused by errors in determining the exchange area for water (A) between cells and adjacent tissue. The half-times of water exchange between bladders and leaf (T1/2) were, on average, somewhat longer for the upper than for the lower epidermis (lower epidermis: T1/2=7 to 38 s; upper epidermis: T1/2=22 to 213 s), but the differences in the T1/2 values were not as distinct as for ε and Lp. This is because of the compensatory effects of ε, Lp and the different ratios of volume to exchange area. Since the bladders make up about 75% of the entire volume of the leaf, it is assumed that the rate of response of the leaf to changes in the water potential should be similar to that of the bladder cells. The results are discussed in terms of a possible function of the bladders in the leaf.  相似文献   

19.
In this paper, a mathematical model is derived to describe the transmission and spread of vector-borne diseases over a patchy environment. The model incorporates into the classic Ross–MacDonald model two factors: disease latencies in both hosts and vectors, and dispersal of hosts between patches. The basic reproduction number \(\mathcal{R }_0\) is identified by the theory of the next generation operator for structured disease models. The dynamics of the model is investigated in terms of \(\mathcal{R }_0\) . It is shown that the disease free equilibrium is asymptotically stable if \(\mathcal{R }_0<1\) , and it is unstable if \(\mathcal{R }_0>1\) ; in the latter case, the disease is endemic in the sense that the variables for the infected compartments are uniformly persistent. For the case of two patches, more explicit formulas for \(\mathcal{R }_0\) are derived by which, impacts of the dispersal rates on disease dynamics are also explored. Some numerical computations for \(\mathcal{R }_0\) in terms of dispersal rates are performed which show visually that the impacts could be very complicated: in certain range of the parameters, \(\mathcal{R }_0\) is increasing with respect to a dispersal rate while in some other range, it can be decreasing with respect to the same dispersal rate. The results can be useful to health organizations at various levels for setting guidelines or making policies for travels, as far as malaria epidemics is concerned.  相似文献   

20.
Science increasingly involves complex modeling. Here we describe a model for cell electroporation in which membrane properties are dynamically modified by poration. Spatial scales range from cell membrane thickness (5 nm) to a typical mammalian cell radius (10  \(\upmu\) m), and can be used with idealized and experimental pulse waveforms. The model consists of traditional passive components and additional active components representing nonequilibrium processes. Model responses include measurable quantities: transmembrane voltage, membrane electrical conductance, and solute transport rates and amounts for the representative “long” and “short” pulses. The long pulse—1.5 kV/cm, 100  \(\upmu\) s—evolves two pore subpopulations with a valley at \({\sim}\) 5 nm, which separates the subpopulations that have peaks at \({\sim}\) 1.5 and \({\sim}\) 12 nm radius. Such pulses are widely used in biological research, biotechnology, and medicine, including cancer therapy by drug delivery and nonthermal physical tumor ablation by causing necrosis. The short pulse—40 kV/cm, 10 ns—creates 80-fold more pores, all small ( \(<\) 3 nm; \(\sim\) 1 nm peak). These nanosecond pulses ablate tumors by apoptosis. We demonstrate the model’s responses by illustrative electrical and poration behavior, and transport of calcein and propidium. We then identify extensions for expanding modeling capability. Structure-function results from MD can allow extrapolations that bring response specificity to cell membranes based on their lipid composition. After a pulse, changes in pore energy landscape can be included over seconds to minutes, by mechanisms such as cell swelling and pulse-induced chemical reactions that slowly alter pore behavior.  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