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.
The quality and quantity of light reflected from Nicotinia tabacum L. internodes was monitored as the angle of incidence was varied. Reflectance of incident light, which was either normal or longitudinal to the internode axis, was investigated. Increasing the angle of incidence caused a greater proportion of incident light to be reflected. Light striking N. tabacum internodes was always modified prior to reflection, and smaller incident angles produced greater modification of the reflected light quality. At larger angles, interactions with the internode were reduced. As a parallel investigation, the extension growth rate of light-grown Sinapis alba L. seedlings was monitored continuously using sensitive linear displacement transducers. When the angle of illumination by fibre-optic probes, presenting far-red light to the growing internode, was changed from 0° to 45°, an increase was observed in stem extension rate. There is a possibility that this rate increase was the result of a fall in the red:far-red ratio (R:FR) present inside the plant internode as the angle of incidence was altered. However, it is more likely that it was due to the larger surface area of stem being illuminated. The consequences of these observations are discussed in relation to the potential influence of such modified reflections on canopy light environments and resultant shade responses which may occur when light of known R:FR impinges on plant internodes at angles other than 0°. The possibility is discussed that plants may perceive the quality of reflected radiation from neighbouring plants to be substantially different dependent upon the angle at which it is reflected.  相似文献   

2.
Materials for nanoporous coatings that exploit optimized chemistries and self‐assembly processes offer capabilities to reach ≈98% transmission efficiency and negligible scattering losses over the broad wavelength range of the solar spectrum from 350 nm to 1.5 µm, on both flat and curved glass substrates. These nanomaterial anti‐reflection coatings also offer wide acceptance angles, up to ±40°, for both s‐ and p‐polarization states of incident light. Carefully controlled bilayer films have allowed for the fabrication of dual‐sided, gradient index profiles on plano‐convex lens elements. In concentration photovoltaics platforms, the resultant enhancements in the photovoltaics efficiencies are ≈8%, as defined by experimental measurements on systems that use microscale triple‐junction solar cells. These materials and their applications in technologies that require control over interface reflections have the potential for broad utility in imaging systems, photolithography, light‐emitting diodes, and display technologies.  相似文献   

3.
Mechanical properties of the adventitia are largely determined by the organization of collagen fibers. Measurements on the waviness and orientation of collagen, particularly at the zero-stress state, are necessary to relate the structural organization of collagen to the mechanical response of the adventitia. Using the fluorescence collagen marker CNA38-OG488 and confocal laser scanning microscopy, we imaged collagen fibers in the adventitia of rabbit common carotid arteries ex vivo. The arteries were cut open along their longitudinal axes to get the zero-stress state. We used semi-manual and automatic techniques to measure parameters related to the waviness and orientation of fibers. Our results showed that the straightness parameter (defined as the ratio between the distances of endpoints of a fiber to its length) was distributed with a beta distribution (mean value 0.72, variance 0.028) and did not depend on the mean angle orientation of fibers. Local angular density distributions revealed four axially symmetric families of fibers with mean directions of 0°, 90°, 43° and ?43°, with respect to the axial direction of the artery, and corresponding circular standard deviations of 40°, 47°, 37° and 37°. The distribution of local orientations was shifted to the circumferential direction when measured in arteries at the zero-load state (intact), as compared to arteries at the zero-stress state (cut-open). Information on collagen fiber waviness and orientation, such as obtained in this study, could be used to develop structural models of the adventitia, providing better means for analyzing and understanding the mechanical properties of vascular wall.  相似文献   

4.
Stages of Metaseiulus occidentalis sensitive to photoperiod induction of diapause were determined by transferring various stadia into diapause-inducing conditions, and rearing them until adult females could be scored for reproductive condition. When eggs were transferred to 10 hr light at 19°C from 24 hr light at 25°C and the mites reared to adults, 92 per cent entered diapause. When larvae and all subsequent stages were kept under the inductive conditions, 62 per cent of adult females diapaused. Mites transferred as protonymphs into inductive conditions yielded only 10 per cent in diapause, and mites transferred as deutonymphs or newly emerged females did not enter diapause.However, adult females reared from eggs at 19°C under 12 hr light (which is near the critical photophase of 11·2 hr at 19°C) showed an unexpected sensitivity to photoperiod. Some newly emerged females oviposited upon transfer to an 8 hr photophase at 19°C. Some then stopped ovipositing and apparently entered diapause; these females resumed ovipositing after intervals ranging from 34 to 100 days. This was termed ‘switching’ into diapause. Some females reared under a 16 hr photophase at 19°C ‘switched’ also upon transfer as adults to shorter photophases—either 8 or 12 hr at 19°C. Thus, ‘switching’ may be due to transfer to shorter photophases. Promptness of mating vs delayed mating allowed ‘switching’ to be more easily detected.  相似文献   

5.
The feeding biomechanics of premetamorphic, metamorphic, and postmetamorphic southern flounder, Paralichthys lethostigma, were investigated to better understand the origin and design of adult pleuronectiform feeding mechanisms. Larval P. lethostigma were sampled from culture tanks every day from first feeding through metamorphosis. Fish were then fixed, cleared, and double stained for cartilage and bone. Postmetamorphic juvenile and adult fish were obtained from aquaculture facilities, fixed, and the muscles and bones of the head dissected. All fish were digitally photographed from both sides of the head. Measurements from digital images included head depth, head length, and quadratal angle (a measure of articular‐quadrate position). Measurements were also made of closing in‐lever, opening in‐lever, and out‐lever moment arm lengths for the determination of lower jaw opening and closing mechanical advantage. In premetamorphic larvae, quadratal angle increased from 40° to 80°, opening lever ratio increased from 0.10 to 0.37, and closing lever ratio increased from 0.06 to 0.40. From these measurements and observations of cleared and double‐stained specimens, it was determined that lower jaw depression and elevation changed from a hyoid‐based to an opercular‐based mechanism prior to the onset of metamorphosis. With migration of the right eye to the left side of the head, quadratal angle remained relatively unchanged at 72° to 84°, opening lever ratio decreased from a high of 0.32 to a low of 0.14, and closing lever ratio decreased to as low as 0.17. Postmetamorphic fish exhibited little change with a quadratal angle of 83° to 84°, an opening lever ratio of 0.19, and a closing lever ratio of 0.17 to 0.19. Paired measurements made on the left (ocular) and right (blind) sides of the head indicated that quadratal angle was asymmetrical during metamorphosis (P = 0.003, α = 0.017). Mechanical advantage for lower jaw elevation was also bilaterally asymmetrical following metamorphosis (P = 0.002, α = 0.013). Because mechanical advantage for lower jaw depression was not directionally asymmetrical in metamorphic or postmetamorphic P. lethostigma, functional asymmetry (lateral jaw flexion) is not predicted for jaw opening. These results suggest differences in the design and function of feeding mechanisms for premetamorphic, metamorphic, and postmetamorphic P. lethostigma. J. Morphol., 2008. © 2008 Wiley‐Liss, Inc.  相似文献   

6.
The compound eye of the housefly Musca domestica L. contains two different types of receptors. The visual acuity of the eye is determined by the divergence angle Δ? between the optical axes of neighbouring ommatidia. Δ? and its dependence on the mean pattern brightness is determined by an evaluation of the optomotor responses elicited from various test patterns. Based on the assumption that the visual fields of both types of receptors approximate the shape of a spatial Gaussian distribution they can be characterized by their half-width, designated as the acceptance angle ΔQ. The contrast transfer from the optical environment onto the receptor cells is limited by ΔQ. It is shown experimentally that ΔQ depends on the mean environmental brightness. The characteristic values Δ? and ΔQ constitute the limiting factors for the light flux received by the receptors. The light flux Φ exciting the receptor cells is proportional to (ΔQ·Δ?)2. If the product ΔQ·Δ? is kept constant, there exists a certain ratio \(\frac{{\Delta _\rho }}{{\Delta _\varphi }}\) that leads to an optimal combination of both, contrast transfer and resolution. The ratio \(\frac{{\Delta _\rho }}{{\Delta _\varphi }}\) is experimentally determined and compared with the optimal condition. The torque exerted by fixed flying Muscae has been used as a measure of the reaction strength of the optomotor response elicited by the rotation of cylindrical patterns consisting of periodic distributions of surface brightness. The responses were investigated under different spatial wavelengths, contrasts, contrast frequencies and mean pattern brightness. Detailed results are:
  1. The visual acuity (optical resolution power) of the compound eye of Musca is determined by the divergence angle Δ ? between the optical axes of those adjacent ommatidia which are not positioned in the same horizontally oriented row but — closer together — in neighboured rows.
  2. Δ? and consequently also the visual acuity do not depend on the mean environmental brightness.
  3. The acceptance angle ΔQ changes with the mean brightness of the environment. According to experimental conditions only the minimal acceptance angle Δ min can be experimentally determined. Δ min decreases with increasing mean pattern brightness from 3.6°–4.1° to 1.7°.
  4. The decrease of ΔQ min with increasing mean pattern brightness is not caused by a change of the acceptance angles of single receptors. The present tentative explanation is that the centrally located receptors No. 7 and 8 are participating in the uptake of relevant visual information at a critical brightness level.
  5. Near the optomotor threshold the large acceptance angle ΔQ min=3.6° at very dim light would thus be associated with the receptors No. 1 to 6, whereas the smaller acceptance angle ΔQ min=1.7° with the receptors No. 7 and 8.
  6. Due to a sample spacing of Δ?=2°, the acceptance angles of neighbouring receptors No. 1 to 6 show a considerable overlap.
  7. Based on anatomical data, the difference in absolute light sensitivity for both receptor systems is calculated. It is predicted that the absorption rate of light quanta in the less sensitive system of the receptors No. 7 and 8 should be reduced by a factor of 24–48 compared to the more sensitive system of the receptors No. 1 to 6. This factor nicely meets the experimentally determined brightness thresholds of both receptor systems.
  8. The optimal condition \(\frac{{\Delta _\rho }}{{\Delta _\varphi }}\) is nearly fulfilled by the receptor system No. 7 and 8 of Musca. The experimentally determined ratio amounts to \(\frac{{\Delta _\rho }}{{\Delta _\varphi }}\) =0.83. For the receptor system No. 1 to 6 one finds \(\frac{{\Delta _\rho }}{{\Delta _\varphi }}\) =1.86; in that system the transfer of spatial wavelengths is mainly limited by the reduced contrast transfer which drops to low values before the optical resolution limit is reached.
  9. Based on the hypothesis that movement perception of the fly Musca is due to a correlation of sensory data one would expect an optomotor peak reaction at a spatial wavelength of λ max=8° and a decrease of the optomotor response towards longer spatial wavelengths. The experimental data are in conflict with these predictions. The present notion is that the absence of the expected reaction decrease is not likely to be caused by a saturation effect in the reaction but rather is explainable in terms of a receptor system consisting of larger numbers of receptor types No. 1 to 6 whose excitations being summed before a correlation evaluation takes place.
  相似文献   

7.
The visual properties of the stemmata of fourth and fifth instar sawfly larvae have been investigated by means of intracellular electrophysiological recordings, interference-, light-, scanning, and transmission electron microscopy, and ray tracing techniques. Each larva possesses one pair of lateral ocelli consisting of biconvex lens, a clear-zone of corneagenous cells devoid of pigment, and a retina made up of groups of 8 retinula cells with a central and fused rhabdom. Migration of pigment during the night is predominantly radial, i.e. directed towards the periphery of the retinula cell. Vertical pigment movement occurs after exposure to bright sunlight for several hours. A circadian rhythm controls the sensitivity of single units, which at night show an increased sensitivity by 2 to 3 log units. The difference between the mean values of acceptance angles for light-adapted units (11·46 ± 5·1° S.D.) and dark-adapted ones (13·83 ± 6·8°) is not statistically significant. The wide range of acceptance angles, with 4·5° being the narrowest and 31° the widest, is explained by the optics of the single lens: there is a small region of highest resolution where light is well focused to a spot, but towards the outer edge of the cup-shaped eye the focusing becomes less accurate. Sawfly larvae turn their heads towards an approaching object if it subtends at least 4° of arc. This limit does not change over a range of 3 log units. A polarization sensitivity of up to 10 : 1 was determined electrophysiologically. Electron microscope studies of the rhabdom show a system of highly oriented microvilli which is thought to be responsible for the polarization sensitivity. Two different waveforms, both occurring after resting potentials of 40 to 70 mV, were found in electrophysiological recordings: (a) hyperpolarizations and (b) depolarizations. Throughout the study only depolarizing units were taken into consideration. These showed characteristics of both the compound eye and the ocellus.  相似文献   

8.
Low-magnification photographs of glabrous upper surfaces using white light show that no matter whether the visual appearance is shiny or matte, the outer cuticles are specularly reflective. In all cases, the individual epidermal cells are seen as bright spots of reflected light, due to their convex outer surfaces. This shows that at least some of the apparently diffuse reflection from leaves is specularly reflected from the outer surface. Quantitative data were obtained for both upper and lower surfaces of a variety of leaves, using light of 632.8 nm incident at 60° from the normal. Light departing the leaf was detected in two directions: normal to the leaf (“diffuse”), and at 60° (“specular”), and compared with the reflection from a standard white block. By this criterion, some leaves were quite shiny, confirming the visual impression, but others were not. In only two cases did the “diffuse” reflection exceed the “specular” (as would be expected from a true diffuse reflector); these were leaves with thick coatings of hairs. Less than 10% of the incident light is reflected by the cuticle of a glabrous leaf.  相似文献   

9.
Low-angle light scattering results reported previously demonstrated that measurements on high molecular weight native DNA must be made at angles below 30° in order to obtain correct molecular weights. Earlier light-scattering data obtained on denaturated DNA at angles above 30° showed no change in molecular weight upon denaturation, even though other techniques clearly showed that strand separation occurred. This paper reports low-angle measurements on solutions of calf thymus and T7 DNA denatured under acidic conditions. The results demonstrate that a halving of molecular weight consistent with strand separation is detected by light scattering only when low-angle data are used to obtain correct molecular weights for native material. As expected from theoretical considerations, the scattering from denatured DNA is a linear function of sin2(θ/2), where θ is the angle of observation. This result shows that anticipated experimental artifacts have no significant effect on the low-angle measurements and demonstrates that the curvature in the scattering envelope observed for native DNA below 30° is an inherent property of the native molecule.  相似文献   

10.
The orientation of the 568 nm transition dipole moment of the retinal chromophore of bacteriorhodopsin has been determined in purple membranes from Halobacterium halobium and in reconstituted vesicles. The angle between the 568 nm transition dipole moment and the normal to the plane of the membrane was measured in two different ways.In the first method the angle was obtained from transient dichroism measurements on bacteriorhodopsin incorporated into large phosphatidylcholine vesicles. Following flash excitation with linearly polarized light, the anisotropy of the 568 nm ground-state depletion signal first decays but then reaches a time-independent value. This result, obtained above the lipid phase transition, is interpreted as arising from rotational motion of bacteriorhodopsin which is confined to an axis normal to the plane of the membrane. It is shown that the relative amplitude of the time-independent component depends on the orientation of the 568 nm transition dipole moment. From the data an angle of 78 ° ± 3 ° is determined.In the second method the linear dichroism was measured as a function of the angle of tilt between the oriented purple membranes and the direction of the light beam. The results were corrected for the angular distribution of the membranes within the oriented samples, which was determined from the mosaic spread of the first-order lamellar neutron diffraction peak. In substantial agreement with the results of the transient dichroism method, linear dichroism measurements on oriented samples lead to an angle of 71 ° ± 4 °.No significant wavelength dependence of the dichroic ratio across the 568 nm band was observed, implying that the exciton splitting in this band must be substantially smaller than the recently suggested value of 20 nm (Ebrey et al., 1977).The orientation of the 568 nm transition dipole moment, which coincides with the direction of the all-trans polyene chain of retinal, is not only of interest in connection with models for the proton pump, but can also be used to calculate the inter-chromophore distances in the purple membrane.  相似文献   

11.
The refractive indices of the bilayer-electrolyte system allow the membrane to operate as a light-guide. This system is then able to monitor, optically, the flow of ions across the bilayer. The light is coupled into and decoupled from a spherically bulged bilayer by means of optical, single mode fibers. The light wave travels along the curved bilayer for several millimeters. This light transmission depends critically on the angle of incidence between the fiber axis and the tangent to the film. Three transmission peaks were observed when the angle of incidence was varied between 0° and 90°. The transmitted light intensity can be modulated by the application of an electric potential upon the bilayer. The center peak, with maximum light transmission, appears at an angle of incidence which is defined by the launching geometry. A quadratic field dependence (independent of the polarity) is observed, which originates from changes in the shape of the torus transition region. The transmission of the satellite peaks, which appear just before and after the central peak, can also be modulated by an external potential. This modulation signal reflects a linear dependence on the polarity of the external voltage. The phase of the modulation signal changes its sign at each satellite peak. It is shown that this modulation signal originates from the bimolecular area of the lipid film. We present evidence that this transmission modulation occurs as a result of ion transport through the lipid film. This provides the basis for the use of wave-guide spectroscopy to investigate membrane ionic fluxes.  相似文献   

12.
SUMMARY 1. Based on the findings that some dragonflies prefer either ‘dark’ or ‘bright’ water (as perceived by the human eye viewing downwards perpendicularly to the water surface), while others choose both types of water bodies in which to lay their eggs, the question arises: How can dragonflies distinguish a bright from a dark pond from far away, before they get sufficiently close to see it is bright or dark? 2. Our hypothesis is that certain dragonfly species may select their preferred breeding sites from a distance on the basis of the polarisation of reflected light. Is it that waters viewed from a distance can be classified on the basis of the polarisation of reflected light? 3. Therefore we measured, at an angle of view of 20° from the horizontal, the reflection‐polarisation characteristics of several ponds differing in brightness and in their dragonfly fauna. 4. We show that from a distance, at which the angle of view is 20° from the horizontal, dark water bodies cannot be distinguished from bright ones on the basis of the intensity or the angle of polarisation of reflected light. At a similar angle of view, however, dark waters reflect light with a significantly higher degree of linear polarisation than bright waters in any range of the spectrum and in any direction of view with respect to the sun. 5. Thus, the degree of polarisation of reflected light may be a visual cue for the polarisation‐sensitive dragonflies to distinguish dark and bright water bodies from far away. Future experimental studies should prove if dragonflies do indeed use this cue for habitat selection.  相似文献   

13.
In order to model laryngeal aerodynamics from a quasi-steady point of view[1], both the dynamic distri-bution of intraglottal air pressures that act upon the vocal folds and the tissue properties of the vocal folds are required[2]. Concerning the first po…  相似文献   

14.
Vocal fold geometry plays an important role in human phonation. The intraglottal quasi-steady pressure and velocity distributions depend upon the shape, size, and diameter of the glottis. This study reports the effects of the variation of glottal shapes on intraglottal pressures and velocities using a Plexiglas model with a glottis having nine symmetric glottal angles (uniform, as well as convergent and divergent 5°, 10°, 20° and 40°), while the minimal glottal diameter was held constant at 0.06 cm. The empirical data were supported by penalty finite element computational results. The results suggest that larger convergent glottal angles correspond to increased pressures and decreased velocities in the glottis upstream of the minimum glottal location, with a reversal of this pattern at the minimal glottal diameter location. The pressure dip near the glottal entrance for divergent glottal angles was greatest for the 10° divergence angle condition, and was sequentially less for 5°, 20°, and 40°. Flow resistance was greater for a convergent angle than a divergent angle of the same value, and least for the 10° divergent condition. Pressure recovery in the glottis suggested that the optimal glottal diffuser angle was near 10°. Results suggest that the glottal geometry has a critical relationship with phonation (especially for vocal efficiency), and therefore important significance to understanding artistic voice and clinical voice management.  相似文献   

15.
Seedlings of Norway spruce (Picea abies L.) have been found to synthesize anthocyanins in the root tips as well as in the hypocotyls upon irradiation with white light when kept at 4°C for 6–8 days. In addition, it has also been found that the elongation and the geotropic curvature of spruce roots are dependent on the light conditions. The course of the geotropic curvature in spruce roots containing anthocyanins has been followed during a period of 5 h, in which the seedlings were geotropically stimulated continuously in the horizontal position. When the stimulation was performed in white light and in darkness at 21°C, significantly larger curvatures were observed in the roots pretreated at 4°C in darkness than in the roots containing anthocyanins. The specific curvature (curvature in degrees per mm elongation), however, was approximately the same in both types of roots stimulated in white light. This was due to a retarded elongation of the roots pretreated with light at 4°C and containing anthocyanins. A smaller difference in elongation rate between roots with and without anthocyanins was observed in the dark than in the light, but even in the dark the anthocyanin-containing roots grew more slowly than roots without anthocyanins. In order to find out if it is the anthocyanin content or the illumination which affects the elongation and geotropic curvature in the roots, a series of similar experiments was performed using cress seedlings grown at 4°C in light or darkness. Roots of cress seedlings cultivated under conditions which would induce anthocyanin formation in spruce roots exhibited the highest geotropic responses both in light and darkness as compared to cress seedlings grown at 4°C in darkness. As in the case of spruce roots an increase in elongation was observed in cress roots illuminated during the geotropic stimulation. These similarities in the behaviour made it relevant to compare the development of the geotropic curvature in cress and spruce roots.  相似文献   

16.
The title compound, Co(NH3)5PO4, prepared by a modified literature procedure, was used to study the inhibition of Na+-K+ ATPase and to serve as a structural model for ML4(nucleotide) complexes. The structure was determined by single crystal X-ray diffraction techniques. The crystals are monoclinic, space group P21/n, a = 8.638(3), b = 14.517(2), c = 9.145(2) Å, and β = 112.71(2)°. The structure, solved by the heavy atom method to an R value of 3.3% for 1924 reflections, consists of a slightly distorted octahedron with the cobalt bound to the five amines and a monodentate phosphate. Solution structural data is taken from 31P NMR measurements. From comparison with other metal phosphate complexes it is concluded that multiple monodentate coordination of a di- or triphosphate closely resembles the coordination of a monophosphate This is based on the similarity of the MO bond angle which is 129.6° in the present example.  相似文献   

17.
The structure of the complex [Pt(trans-1,2-di- aminocyclohexane) (acetate)2]·H2O has been determined by X-ray diffraction. This racemic compound is orthorhombic, space group Aba2, a = 20.813(9), b = 7.926(5), c = 17.296(8) Å, Z = 8. The structure was refined on 1214 nonzero Cu Kα reflections to R = 0.028. The square planar environment of Pt includes the amino groups of the diamine in cis positions and oxygens from two monodentate acetates. The PtN and PtO distances average 2.00(3) and 2.02(3) Å, respectively. The bite of the diamine ligand imposes a NPtN angle of 85(1)°, whereas the small OPtO angle of 85(1)° probably results from packing effects. The average plane through the puckered cyclohexyl ring makes an angle of 19° with the PtN2O2 plane. The molecules are stacked by pairs along the b axis. The two molecules of each pair are 180° apart about the stacking axis, and form altogether four NH···O hydrogen bonds.  相似文献   

18.
Leaf phyllotaxis: Does it really affect light capture?   总被引:1,自引:0,他引:1  
The intriguing mathematical properties of leaf phyllotaxis still attract scientific attention after centuries of research. Phyllotaxis, and in particular the divergence angle between successive leaves, have been frequently interpreted in terms of maximization of light capture, although certain model simulations of light capture by vertical shoots revealed minor effects of phyllotaxis in comparison with the effect of other morphological features of the plant. However, these simulations assumed a number of simplifications, did not take into account diffuse light, and were not based on real plants with their natural range of morphological variation. This study was aimed at filling these gaps by examining the influence on light harvesting of shoot architecture and divergence angle in four species with spiral phyllotaxis (Quercus ilex, Arbutus unedo, Heteromeles arbutifolia and Daphne gnidium) with a realistic 3-D model (Y-plant). A wide range of divergence angles (from 100° to 154°) was observed within each species, with 144° being the most frequent one. These different divergence angles rendered very different vertical projections of the shoot due to contrasting patterns of leaf overlap as seen from above, but they rendered indistinguishable light interception efficiencies (Ea). Setting the leaves with an opposite-decussate phyllotaxis led, however, to a 40–50% decrease of Ea. The interplay of internode length, leaf size and shape, and leaf elevation angle led to significant species differences in Ea. Thus, only particular phyllotaxis (e.g., decussate) might be functionally inefficient under certain combinations of the various morphological variables that influence light capture of a shoot. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

19.
A finite element model was used to investigate the counter-intuitive experimental observation that some regions of the aponeuroses of a loaded and contracting muscle may shorten rather than undergo an expected lengthening. The model confirms the experimental findings and suggests that pennation angle plays a significant role in determining whether regions of the aponeuroses stretch or shorten. A smaller pennation angles (25°) was accompanied by aponeurosis lengthening whereas a larger pennation angle (47°) was accompanied by mixed strain effects depending upon location along the length of the aponeurosis. This can be explained by the Poisson effect during muscle contraction and a Mohr’s circle analogy. Constant volume constraint requires that fiber cross sectional dimensions increase when a fiber shortens. The opposing influences of these two strains upon the aponeurosis combine in proportion to the pennation angle. Lower pennation angles emphasize the influence of fiber shortening upon the aponeurosis and thus favor aponeurosis compression, whereas higher pennation angles increase the influence of cross sectional changes and therefore favor aponeurosis stretch. The distance separating the aponeuroses was also found to depend upon pennation angle during simulated contractions. Smaller pennation angles favored increased aponeurosis separation larger pennation angles favored decreased separation. These findings caution that measures of the mechanical properties of aponeuroses in intact muscle may be affected by contributions from adjacent muscle fibers and that the influence of muscle fibers on aponeurosis strain will depend upon the fiber pennation angle.  相似文献   

20.
The crystal and molecular structure of nitrosyltris-(trimethylphosphine)nickel(O) hexafluorophosphate, {Ni(NO)(PMe3)3}PF6, has been determined from three dimensional single crystal X-ray analysis. The compound crystallizes in the orthorhombic space group Pnma with Z = 4 and a unit cell of dimensions: a = 16.253(3), b = 10.536(1) and c = 12.228(2) Å. The structure was solved by conventional heavy atom techniques and refined by least-squares methods to R1 = 0.036 and R2 = 0.048 respectively for 1085. independent reflections. The coordination geometry around the nickel is a slightly distorted tetrahedron with an average PNiP angle of 105.63° and PNiN angle 113.03°. The nickel nitrosyl group is slightly bent with an NiNO angle of 175.4(5)°. The bending occurs in the ClPlNiNO plane toward Pl. The structure is compared with other tetrahedral {MNO}10 phosphine complexes and the MNO bonding is discussed.  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