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 328 毫秒
1.
We have developed a bioreactor vessel design which has the advantages of simplicity and ease of assembly and disassembly, and with the appropriately determined flow rate, even allows for a scaffold to be suspended freely regardless of its weight. This article reports our experimental and numerical investigations to evaluate the performance of a newly developed non-perfusion conical bioreactor by visualizing the flow through scaffolds with 45 degrees and 90 degrees fiber lay down patterns. The experiments were conducted at the Reynolds numbers (Re) 121, 170, and 218 based on the local velocity and width of scaffolds. The flow fields were captured using short-time exposures of 60 microm particles suspended in the bioreactor and illuminated using a thin laser sheet. The effects of scaffold fiber lay down pattern and Reynolds number were obtained and correspondingly compared to results obtained from a computational fluid dynamics (CFD) software package. The objectives of this article are twofold: to investigate the hypothesis that there may be an insufficient exchange of medium within the interior of the scaffold when using our non-perfusion bioreactor, and second, to compare the flows within and around scaffolds of 45 degrees and 90 degrees fiber lay down patterns. Scaffold porosity was also found to influence flow patterns. It was therefore shown that fluidic transport could be achieved within scaffolds with our bioreactor design, being a non-perfusion vessel. Fluid velocities were generally same of the same or one order lower in magnitude as compared to the inlet flow velocity. Additionally, the 90 degrees fiber lay down pattern scaffold was found to allow for slightly higher fluid velocities within, as compared to the 45 degrees fiber lay down pattern scaffold. This was due to the architecture and pore arrangement of the 90 degrees fiber lay down pattern scaffold, which allows for fluid to flow directly through (channel-like flow).  相似文献   

2.
Equatorial X-ray diffraction patterns from single skinned fibres from bony fish muscle (turbot) were obtained with the fibres at 6 degrees C bathed in relaxing solutions of 170 down to 26 mM ionic strength. Diffraction patterns from rigor fibres were also obtained as controls. Unlike fibres from rabbit muscle, which show very clear evidence of substantial crossbridge formation at low ionic strength in what is mechanically a rapid equilibrium ("weak-binding") state (Brenner et al., 1982), diffraction patterns from bony fish fibres showed only a small change in relative peak intensities at low ionic strength (26 mM) compared with normal (170 mM) ionic strength. However, there was a slight ordering of the filament lattice at low ionic strength. The specimen temperature used (about 6 degrees C) was not far from the normal physiological temperature of the fish. Likewise, only a small change was seen by Xu et al. (1987) in patterns from frog fibres at low ionic strength at 2 to 6 degrees C. (Rabbit fibres previously studied, where large changes were seen at temperatures of 5 to 20 degrees C, were about 17 to 32 degrees C below physiological.) The I11/I10 ratio for fish fibres at 26 mM ionic strength was actually lower than that for rabbit even at normal ionic strength. This may be associated with an intrinsic structural difference between these muscles or alternatively with the disordering of the crossbridge helix in rabbit muscle found at low temperature by Wray (1987), and could support the view that rabbit fibres at 5 degrees C and normal ionic strength may already have a significant population of weak-binding crossbridges.  相似文献   

3.
A cotton fiber was kept under slight tension and exposed locally to a stream of aqueous 1 N NaOH microdrops of 50 microm diameter. The resulting "macrodrop" of about 300 microm size was at the origin of the formation of Na-cellulose I domains extending about 550 microm from the center of the macrodrop along the fiber. The phase transformation zone between cellulose I and Na-cellulose I was mapped by scanning synchrotron radiation microdiffraction using a 300 nm x 300 nm beam. A stitching technique was used to limit radiation damage. Subsequent exposure of the NaOH containing macrodrop to a stream of H2O or HCl microdrops converted part of the Na-cellulose I back into cellulose I.  相似文献   

4.
Scanning synchrotron radiation microdiffraction with an approximately 1 x 1 microm(2) beam has been used as a novel method for characterizing the decay of several T'ang dynasty (618-907 AD) silk fabrics. The crystalline fraction could be visualized based on beta-sheet 210 reflection intensities, extracted by recursive peak fits from several thousand diffraction patterns recorded during mesh scans. The azimuthal width of the 210 reflection, which is related to the orientation distribution of the crystalline domains within nanofibrils and the macroscopic orientation of the fibers traversed by the beam, was found to be sensitive to the overall state of decay of the fabric. The fine structure of the histogram of azimuthal width was related to the fiber hierarchical microstructure and the fabric morphology. SAXS/WAXS analysis supports the assumption of an initial loss of the random chain network with decay. At a subsequent state of aging, decay proceeds into the nanofibrils and the silk fibers break up into even smaller fractions.  相似文献   

5.
Polarization-modulated second harmonic generation in collagen   总被引:6,自引:0,他引:6       下载免费PDF全文
Collagen possesses a strong second-order nonlinear susceptibility, a nonlinear optical property characterized by second harmonic generation in the presence of intense laser beams. We present a new technique involving polarization modulation of an ultra-short pulse laser beam that can simultaneously determine collagen fiber orientation and a parameter related to the second-order nonlinear susceptibility. We demonstrate the ability to discriminate among different patterns of fibrillar orientation, as exemplified by tendon, fascia, cornea, and successive lamellar rings in an intervertebral disc. Fiber orientation can be measured as a function of depth with an axial resolution of approximately 10 microm. The parameter related to the second-order nonlinear susceptibility is sensitive to fiber disorganization, oblique incidence of the beam on the sample, and birefringence of the tissue. This parameter represents an aggregate measure of tissue optical properties that could potentially be used for optical imaging in vivo.  相似文献   

6.
When light is diffracted by a single frog muscle fiber the intensities I kappa of the different orders kappa (kappa = 1,2,3) strongly depend on the angle between the axis of the incident beam and the fiber axis. Maximum intensity is not obtained with perpendicular incidence (omega = 0 degree) but at angles that can be calculated for each order number and sarcomere length using Bragg's formula. In analogy to techniques developed for x-ray structure analysis of mosaic crystals we have rotated the fiber around an axis perpendicular to the fiber axis and to the incident beam axis within an angular range delta omega = +/- 35 degrees and recorded the light intensities I kappa. Diffraction efficiencies defined as E kappa = integral of I kappa d omega were studied as a function of sarcomere length and during isometric contraction. The sarcomere length dependences of the efficiencies E kappa of the first three orders show characteristic trends. E1 increases with fiber stretch, E2 has a minimum at a sarcomere length near 2.8 micrometers, and E3 has a maximum near 2.5 micrometers. These trends as well as the observed efficiency ratios are in fairly good agreement with predictions by the intensity formula developed for x-ray structure analysis. During isometric contraction, the diffraction efficiencies of the fiber decrease, with the decreases becoming greater the higher the order number. These decreases might be caused by a longitudinal displacement of myofibrils of up to 0.4 micrometers. The efficiency of light diffraction strongly depends on the tonicity of the bathing fluid. Hypertonic (3/2 x normal) solution reduces E1 to less than half, hypotonic (2/3 x normal) solution increases E1 to almost twice the value obtained in normal Ringer's solution.  相似文献   

7.
The regulatory light chain (RLC) from chicken gizzard myosin was covalently modified on cysteine 108 with either the 5- or 6-isomer of iodoacetamidotetramethylrhodamine (IATR). Labeled RLCs were purified by fast protein liquid chromatography and characterized by reverse-phase high-performance liquid chromatography (HPLC), tryptic digestion, and electrospray mass spectrometry. Labeled RLCs were exchanged into the native myosin heads of single skinned fibers from rabbit psoas muscle, and the ATR dipole orientations were determined by fluorescence polarization. The 5- and 6-ATR dipoles had distinct orientations, and model orientational distributions suggest that they are more than 20 degrees apart in rigor. In the rigor-to-relaxed transition (sarcomere length 2.4 microm, 10 degrees C), the 5-ATR dipole became more perpendicular to the fiber axis, but the 6-ATR dipole became more parallel. This orientation change was absent at sarcomere length 4.0 microm, where overlap between myosin and actin filaments is abolished. When the temperature of relaxed fibers was raised to 30 degrees C, the 6-ATR dipoles became more parallel to the fiber axis and less ordered; when ionic strength was lowered from 160 mM to 20 mM (5 degrees C), the 6-ATR dipoles became more perpendicular to the fiber axis and more ordered. In active contraction (10 degrees C), the orientational distribution of the probe dipoles was similar but not identical to that in relaxation, and was not a linear combination of the orientational distributions in relaxation and rigor.  相似文献   

8.
BACKGROUND: Improving our knowledge of plant metal metabolism is facilitated by the use of analytical techniques to map the distribution of elements in tissues. One such technique is X-ray fluorescence (XRF), which has been used previously to map metal distribution in both two and three dimensions. One of the difficulties of mapping metal distribution in two dimensions is that it can be difficult to normalize for tissue thickness. When mapping metal distribution in three dimensions, the time required to collect the data can become a major constraint. In this article a compromise is suggested between two- and three-dimensional mapping using multi-angle XRF imaging. METHODS: A synchrotron-based XRF microprobe was used to map the distribution of K, Ca, Mn, Fe, Ni, Cu and Zn in whole Arabidopsis thaliana seeds. Relative concentrations of each element were determined by measuring fluorescence emitted from a 10 microm excitation beam at 13 keV. XRF spectra were collected from an array of points with 25 or 30 microm steps. Maps were recorded at 0 and 90 degrees , or at 0, 60 and 120 degrees for each seed. Using these data, circular or ellipsoidal cross-sections were modelled, and from these an apparent pathlength for the excitation beam was calculated to normalize the data. Elemental distribution was mapped in seeds from ecotype Columbia-4 plants, as well as the metal accumulation mutants manganese accumulator 1 (man1) and nicotianamine synthetase (nasx). CONCLUSIONS: Multi-angle XRF imaging will be useful for mapping elemental distribution in plant tissues. It offers a compromise between two- and three-dimensional XRF mapping, as far as collection times, image resolution and ease of visualization. It is also complementary to other metal-mapping techniques. Mn, Fe and Cu had tissue-specific accumulation patterns. Metal accumulation patterns were different between seeds of the Col-4, man1 and nasx genotypes.  相似文献   

9.
Accurate dose estimation under various inhalation conditions is important for assessing both the potential health effects of pollutant particles and the therapeutic efficacy of medicinal aerosols. We measured total deposition fraction (TDF) of monodisperse micrometer-sized particles [particle diameter (Dp) = 1, 3, and 5 microm in diameter] in healthy adults (8 men and 7 women) in a wide range of breathing patterns; tidal volumes (Vt) of 350-1500 ml and respiratory flow rates (Q) of 175-1,000 ml/s. The subject inhaled test aerosols for 10-20 breaths with each of the prescribed breathing patterns, and TDF was obtained by monitoring inhaled and exhaled aerosols breath by breath by a laser aerosol photometer. Results show that TDF varied from 0.12-0.25, 0.26-0.68, and 0.45-0.83 for Dp = 1, 3, and 5 microm, respectively, depending on the breathing pattern used. TDF was comparable between men and women for Dp = 1 microm but was greater in women than men for Dp = 3 and 5 microm for all breathing patterns used (P < 0.05). TDF increased with an increase in Vt regardless of Dp and Q used. At a fixed Vt TDF decreased with an increase in Q for Dp = 1 and 3 microm but did not show any significant changes for Dp = 5 microm. The varying TDF values, however, could be consolidated by a single composite parameter (omega) consisting of Dp, Vt, and Q. The results indicate that unifying empirical formulas provide a convenient means of assessing deposition dose of particles under varying inhalation conditions.  相似文献   

10.
Fan Y  Xu Z  Jiang W  Deng X  Wang K  Sun A 《Journal of biomechanics》2008,41(11):2498-2505
The development of distal end-to-side anastomotic intimal hyperplasia (IH) has been attributed to the flow disturbance and abnormal wall shear stress (WSS) distribution there. The geometry of the bypass has a strong influence on the flow pattern and WSS distribution. Using a canine model of end-to-side anastomosis, a 45 degrees S-type bypass was compared with 60 degrees , 45 degrees and 30 degrees conventional bypasses in the term of IH along the host artery floor. Numerical blood flow simulations were also carried out to characterize the flow patterns at the distal parts of the bypassed arteries for the 4 models. The results showed that the averaged intima thicknesses of the host artery floors for the 4 bypass models were 119.50+/-10.30 microm (60 degrees ), 65.56+/-6.53 microm (45 degrees ), 45.26+/-5.99 microm (30 degrees ) and 47.64+/-4.85 microm (S-type), respectively, vs. 9.81+/-1.88 microm in the control group (without bypass surgery). Compared with the control group, neointima thickness in all 4 bypass models was significantly increased, but the neointima thickness of the 45 degrees S-type bypass was apparently much better than its 45 degrees conventional counterpart, and was as good as the 30 degrees conventional bypass. The numerical simulation revealed an apparent swirling flow pattern in the S-type bypass, which was very different than the flow patterns in the 3 conventional bypass models. This swirling flow altered the overall flow pattern in the distal part of the bypassed artery and eliminated the low WSS zone along the host artery floor. The improvement in the term of IH for the S-type bypass is most likely due to the alteration of the overall flow pattern and WSS distribution by the geometrical configuration of the S-type bypass.  相似文献   

11.
Water microdrops of about 50 microm in diameter, generated by an ink-jet system, have been used to hydrate fragments of Pogonophora tubes. In situ X-ray microdiffraction with a beam size of 10 microm was used to follow the structural transformations that affected the crystalline beta-chitin part of the specimens. Starting from anhydrous chitin, the formation of a full beta-chitin dihydrate was observed within about 90 s. A disordered intermediate phase with variable d-spacing that could be due to a mixture of anhydrous and hydrated beta-chitin layers was also detected.  相似文献   

12.
A novel hydrogel was obtained by reticulation of chitosan with dextrin enzymatically linked to vinyl acrylate (dextrin-VA), without cross-linking agents. The hydrogel had a solid-like behaviour with G′ (storage modulus) >> G″ (loss modulus). Glucose diffusion coefficients of 3.9 × 10−6 ± 1.3 × 10−6 cm2/s and 2.9 × 10−6 ± 0.5 × 10−6 cm2/s were obtained for different substitution degrees of the dextrin-VA (20% and 70% respectively). SEM observation revealed a porous structure, with pores ranging from 50 μm to 150 μm.  相似文献   

13.
对采自青岛和芜湖两地的萼花臂尾轮虫在3种温度(20 ℃、25 ℃和30 ℃)和2种藻类食物浓度(1.0×106和5.0×106 cells·ml-1)下所产休眠卵的长径、短径和体积等形态特征进行了显微测量、计算和分析.结果表明,2种食物浓度下,培养温度以及培养温度和品系间的交互作用均对轮虫休眠卵的长径、短径和体积具有显著影响.当食物浓度分别为1.0×106和5.0×106 cells·ml-1时,轮虫在20 ℃下所产休眠卵的长径、短径和体积均最大;在25 ℃和30 ℃下所产休眠卵的短径和体积均最小.品系对轮虫休眠卵长径、短径和体积的影响也取决于食物浓度.当食物浓度为1.0×106 cells·ml-1时,芜湖品系轮虫的休眠卵长径、短径和体积(156.00 μm、99.95 μm和12 269.11 μm3)均显著大于青岛品系轮虫的休眠卵(145.13 μm、91.97 μm和10 498.19 μm3);而当食物浓度为5.0×106 cells·ml-1时,芜湖品系轮虫的休眠卵长径、短径和体积(155.68 μm、100.85 μm和12 348.59 μm3)均与青岛品系轮虫的休眠卵(156.63 μm、98.04 μm和12 054.20 μm3)之间无显著差异.两品系中,仅芜湖品系轮虫休眠卵的长径、短径和体积分别与温度呈曲线相关.同一温度下,两品系轮虫的休眠卵体积均随着食物浓度升高而增大;但30 ℃下芜湖品系轮虫所产休眠卵体积却随着食物浓度的升高而减小.  相似文献   

14.
In the present study, a shape-independent differential scanning calorimeter (DSC) technique was used to measure the dehydration response during freezing of sperm cells from diploid and tetraploid Pacific oysters, Crassostrea gigas. This represents the first application of the DSC technique to sperm cells from nonmammalian species. Volumetric shrinkage during freezing of oyster sperm cell suspensions was obtained at cooling rates of 5 and 20 degrees C/min in the presence of extracellular ice and 8% (v/v) concentration of dimethyl sulfoxide (DMSO), a commonly used cryoprotective agent (CPA). Using previously published data, sperm cells from diploid oysters were modeled as a two-compartment "ball-on-stick" model with a "ball" 1.66 microm in diameter and a "stick" 41 microm in length and 0.14 microm wide. Similarly, sperm cells of tetraploid oysters were modeled with a "ball" 2.14 microm in diameter and a "stick" 53 microm in length and 0.17 microm wide. Sperm cells of both ploidy levels were assumed to have an osmotically inactive cell volume, Vb, of 0.6 Vo, where Vo is the isotonic (or initial) cell volume. By fitting a model of water transport to the experimentally obtained volumetric shrinkage data, the best-fit membrane permeability parameters (Lpg and ELp) were determined. The combined-best-fit membrane permeability parameters at 5 and 20 degrees C/min for haploid sperm cells (or cells from diploid Pacific oysters) in the absence of CPAs were: Lpg = 0.30 x 10(-15) m(3)/Ns (0.0017 microm/min-atm) and ELp = 41.0 kJ/mole (9.8 kcal/mole). The corresponding parameters in the presence of 8% DMSO were: Lpg[cpa] = 0.27 x 10(-15) m(3)/Ns (0.0015 microm/min-atm) and ELp[cpa] = 38.0 kJ/mole (9.1 kcal/mole). Similarly, the combined-best-fit membrane permeability parameters at 5 and 20 degrees C/min for diploid sperm cells (or cells from tetraploid Pacific oysters) in the absence of CPAs were: Lpg = 0.34 x 10(-15) m(3)/Ns (0.0019 microm/min-atm) and ELp = 29.7 kJ/mole (7.1 kcal/mole). The corresponding parameters in the presence of 8% DMSO were: Lpg[cpa] = 0.34 x 10(-15) m(3)/Ns (0.0019 microm/min-atm) and ELp[cpa] = 37.6 kJ/mole (9.0 kcal/mole). The parameters obtained in this study suggest that optimal rates of cooling for Pacific oyster sperm cells range from 40 to 70 degrees C/min. These theoretical cooling rates are in close conformity with empirically determined optimal rates of cooling sperm cells from Pacific oysters, C. gigas.  相似文献   

15.
In the present study a well-established differential scanning calorimeter (DSC) technique is used to measure the water transport phenomena during freezing of stromal vascular fraction (SVF) and adipose tissue derived adult stem (ADAS) cells at different passages (Passages 0 and 2). Volumetric shrinkage during freezing of adipose derived cells was obtained at a cooling rate of 20 degrees C/min in the presence of extracellular ice and two different, commonly used, cryoprotective agents, CPAs (10% DMSO and 10% Glycerol). The adipose derived cells were modeled as spheres of 50 microm diameter with an osmotically inactive volume (Vb) of 0.6Vo, where Vo is the isotonic cell volume. By fitting a model of water transport to the experimentally obtained volumetric shrinkage data, the "best-fit" membrane permeability parameters (reference membrane permeability to water, Lpg or Lpg[cpa] and the activation energy, ELp or ELp[cpa]) were determined. The "best-fit" membrane permeability parameters for adipose derived cells in the absence and presence of CPAs ranged from: Lpg=23.1-111.5x10(-15) m3/Ns (0.135-0.652 microm/min-atm) and ELp=43.1-168.8 kJ/mol (9.7-40.4 kcal/mol). Numerical simulations of water transport were then performed under a variety of cooling rates (5-100 degrees C/min) using the experimentally determined membrane permeability parameters. And finally, the simulation results were analyzed to predict the optimal rates of freezing adipose derived cells in the presence and absence of CPAs.  相似文献   

16.
Two strains of Moorella thermoacetica, JW/B-2 and JW/DB-4, isolated as contaminants from autoclaved media for chemolithoautotrophic growth containing 0.1% (wt/vol) yeast extract, formed unusually heat-resistant spores. Spores of the two strains required heat activation at 100 degrees C of more than 2 min and up to 90 min for maximal percentage of germination. Kinetic analysis indicated the presence of two distinct subpopulations of heat-resistant spores. The decimal reduction time (D10-time=time of exposure to reduce viable spore counts by 90%) at 121 degrees C was determined for each strain using spores obtained under different conditions. For strains JW/DB-2 and JW/ DB-4, respectively, spores obtained at approximately 25 degrees C from cells grown chemolithoautotrophically had D10-times of 43 min and 23 min; spores obtained at 60 degrees C from cells grown chemoorganoheterotrophically had D10-times of 44 min and 38 min; spores obtained at 60 degrees C from cells grown chemolithoautotrophically had D10-times of 83 min and 111 min. The thickness of the cortex varied between 0.10 and 0.29 microm and the radius of the cytoplasm from 0.14 to 0.46 microm. These spores are amongst the most heat-resistant noted to date. Electron microscopy revealed structures within the exosporia of spores prior to full maturity that were assumed to be layers of the outer spore coat.  相似文献   

17.
The microvascularization of the brains of the hagfishes, Myxine glutinosa L. and Eptatretus stouti, were studied by scanning electron microscopy (SEM) of microvascular corrosion casts. Sections of these casts were used to determine the vascular territories of defined brain areas. Histological serial sections (10 microm) of the brains served for correlation of findings. Analysis of the microvascular casts of both species revealed that the blood supply to and from these brains arose ventrally and dorsally, respectively. Neither species possesses an arterial circle (Circulus Willisi) and both have similar microvascular patterns. The only difference between Myxine and Eptatretus was that the posterior cerebral artery in Myxine divides into mesencephalic and rhombencephalic branches, and in Eptatretus a third branch, termed telencephalic branch, arises from the posterior cerebral artery. 3D-morphometry revealed that luminal diameters of: 1) intracerebral arteries and arterioles range from 35.11 +/- 5.66 microm (mean +/- SEM) in the hypothalamus to 92.69 +/- 14.48 microm in the thalamus; 2) capillaries range from 17.8 +/- 0.44 microm in the olfactory bulb to 21.70 +/- 0.87 microm in the basal ganglia; and 3) intracerebral venules and veins range from 49.38 +/- 4.17 microm in the hypothalamus to 75.58 +/- 6.59 microm in the rhombencephalon. Interbranching distances of arteries and arterioles range from 179.19 +/- 11.32 microm in the optic tectum to 235.19 +/- 94.64 microm in the hypothalamus. Capillaries range from 91.07 +/- 6.22 microm in the hypothalamus to 116.15 +/- 9.45 microm in the thalamus, and venules and veins range from 137.30 +/- 18.11 microm in the hypothalamus to 189.83 +/- 17.47 microm in the optic tectum. Intervascular distances range from 70.58 +/- 3.58 microm in the olfactory bulb to 89.52 +/- 5.74 microm in the optic tectum. Branching angles of arteries and arterioles range from 38.39 +/- 10.9 degrees in the olfactory bulb to 100.73 +/- 9.4 degrees in the optic tectum, and the branching angles of capillaries range from 74.40 +/- 5.42 degrees in the optic tectum to 90.24 +/- 4.66 degrees in the olfactory bulb. Finally, the branching angles of the venules and veins range from 67.84 +/- 6.83 degrees in the tegmentum of the mesencephalon to 92.30 +/- 6.35 degrees in the optic tectum.  相似文献   

18.
Cholera toxin binds to its ganglioside GM1 receptor via its B-subunit, a pentameric assembly of identical subunits (Mr = 11,600). Diffraction quality crystals of cholera toxin B-subunit have been obtained at room temperature by vapor diffusion with polyethylene glycol in the presence of the nonionic detergent beta-octyl glucoside. The crystals have been characterized with x-radiation as monoclinic, space group P21, with unit cell dimensions a = 39.0 A, b = 94.3 A, c = 67.5 A, beta = 96.0 degrees. There are two molecules per unit cell, with one molecule (Mr = 58,000) in each asymmetric unit. Precession photographs (micron = 13 degrees) show that crystals diffract beyond 3.3-A resolution and are stable in the x-ray beam at room temperature for at least 40 h; thus, they can be used to collect three-dimensional crystallographic data.  相似文献   

19.
Ontogenetic changes in the organization and volume fraction of collagenous connective tissues were examined in the mantle of Sepioteuthis lessoniana, the oval squid. Outer tunic fiber angle (the angle of a tunic collagen fiber relative to the long axis of the squid) decreased from 33.5 degrees in newly hatched animals to 17.7 degrees in the largest animals studied. The arrangement of intramuscular collagen fiber systems 1 (IM-1) and 2 (IM-2) also changed significantly during ontogeny. Because of the oblique trajectory of the IM-1 collagen fibers, two fiber angles were needed to describe their organization: (1) IM-1(SAG), the angle of an IM-1 collagen fiber relative to the squid's long axis when viewed from a sagittal plane and (2) IM-1(TAN), the angle of an IM-1 collagen fiber relative to the squid's long axis when viewed from a plane tangential to the outer curvature of the mantle. The sagittal component (IM-1(SAG)) of the IM-1 collagen fiber angle was lowest in hatchling squid (32.7 degrees ) and increased exponentially during growth to 43 degrees in squid with a dorsal mantle length (DML) of 15 mm. In squid larger than 15 mm DML, IM-1(SAG) fiber angle did not change. The tangential component (IM-1(TAN)) of IM-1 collagen fiber angle was highest in hatchling squid (39 degrees ) and decreased to 32 degrees in the largest squid examined. IM-2 collagen fiber angle (the angle of an IM-2 collagen fiber relative to the outer surface of the mantle) was lowest in hatchling squid (34.6 degrees ) and increased exponentially to about 50 degrees in 15-mm DML animals. In squid larger than 15 mm DML, IM-2 fiber angle increased slightly with size. The volume fraction of collagen in IM-1 and IM-2 increased 68 and 36 times, respectively, during growth. The ontogenetic changes in the organization of collagen fibers in the outer tunic, IM-1, and IM-2 may lead to ontogenetic differences in the kinematics of mantle movement and in elastic energy storage during jet locomotion.  相似文献   

20.
The soluble proteins from bovine lens homogenate were separated on Sepharose CL-6B (2 X 200 cm) in 0.05 M tris-NaHSO3 pH 8.2 buffer containing 20 mM EDTA. Five sharp and defined fractions (HM alpha, alpha, beta H, beta L, gamma) were obtained. Each crystallin fraction was further purified by rechromatography on the same column. Each protein fraction was pure as judged by ultracentrifugation and SDS-gel electrophoresis. The molecular weights of the five fractions were 3.04 x 10(6), 5.83 x 10(5), 1.58 x 10(5) , 4.59 x 10(4), 2.14 x 10(4) as determined from sedimentation coefficient and intrinsic viscosity data by Scheraga-Mandelkern equation, which was in close agreement with that obtained by gel filtration. The polypeptide composition of crystallins as determined by SDS-gel electrophoresis revealed one band for high molecular weight alpha (HM alpha) and alpha, three for beta H, two for beta L and one for gamma. The gross CD patterns of crystallins were about the same in the peptide region (200 nm similar to or approximately 250 nm) with a minimum centered at about 217 nm, indicative of a beta-sheet structure in all crystallins. The [theta] values at 217 nm ranged from --1700 to --3700 degrees cm2 per decimole. The CD spectra of these crystallins in the aromatic region (250 nm similar to or approximately 300 nm) were different, reflecting the different contributions of aromatic amino acids to the tertiary structure of crystallins.  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