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 156 毫秒
1.
The first part of this review on entropic elastic processes in protein mechanisms (Urry, 1988) demonstrated with the polypentapeptide of elastin (Val1-Pro2-Gly3-Val4-Gly5)n that elastic structure develops as the result of an inverse temperature transition and that entropic elasticity is due to internal chain dynamics in a regular nonrandom structure. This demonstration is contrary to the pervasive perspective of entropic protein elasticity of the past three decades wherein a network of random chains has been considered the necessary structural consequence of the occurrence of dominantly entropic elastomeric force. That this is not the case provides a new opportunity for understanding the occurrence and role of entropic elastic processes in protein mechanisms. Entropic elastic processes are considered in two classes: passive and active. The development of elastomeric force on deformation is class I (passive) and the development of elastomeric force as the result of a chemical process shifting the temperature of a transition is class II (active). Examples of class I are elastin, the elastic filament of muscle, elastic force changes in enzyme catalysis resulting from binding processes and resulting in the straining of a scissile bond, and in the turning on and off of channels due to changes in transmembrane potential. Demonstration of the consequences of elastomeric force developing as the result of an inverse temperature transition are seen in elastin, where elastic recoil is lost on oxidation, i.e., on decreasing the hydrophobicity of the chain and shifting the temperature for the development of elastomeric force to temperatures greater than physiological. This is relevant in general to loss of elasticity on aging and more specifically to the development of pulmonary emphysema. Since random chain networks are not the products of inverse temperature transitions and the temperature at which an inverse temperature transition occurs depends on the hydrophobicity of the polypeptide chain, it now becomes possible to consider chemical processes for turning elastomeric force on and off by reversibly changing the hydrophobicity of the polypeptide chain. This is herein called mechanochemical coupling of the first kind; this is the chemical modulation of the temperature for the transition from a less-ordered less elastic state to a more-ordered more elastic state. In the usual considerations to date, development of elastomeric force is the result of a standard transition from a more-ordered less elastic state to a less-ordered more elastic state. When this is chemically modulated, it is herein called mechanochemical coupling of the second kind. For elastin and the polypentapeptide of elastin, since entropic elastomeric force results on formation of a regular nonrandom structure and thermal randomization of chains results in loss of elastic modulus to levels of limited use in protein mechanisms, consideration of regular spiral-like structures rather than ramdom chain networks or random coils are proposed for mechanochemical coupling of the second kind. Chemical processes to effect mechanochemical coupling in biological systems are most obviously phosphorylation-dephosphorylation and changes in calcium ion activity but also changes in pH. These issues are considered in the events attending parturition in muscle contraction and in cell motility.  相似文献   

2.
Numerous physical characterizations clearly demonstrate that the polypentapeptide of elastin (Val1-Pro2-Gly3-Val4-Gly5)n in water undergoes an inverse temperature transition. Increase in order occurs both intermolecularly and intramolecularly on raising the temperature from 20 to 40 degrees C. The physical characterizations used to demonstrate the inverse temperature transition include microscopy, light scattering, circular dichroism, the nuclear Overhauser effect, temperature dependence of composition, nuclear magnetic resonance (NMR) relaxation, dielectric relaxation, and temperature dependence of elastomer length. At fixed extension of the cross-linked polypentapeptide elastomer, the development of elastomeric force is seen to correlate with increase in intramolecular order, that is, with the inverse temperature transition. Reversible thermal denaturation of the ordered polypentapeptide is observed with composition and circular dichroism studies, and thermal denaturation of the crosslinked elastomer is also observed with loss of elastomeric force and elastic modulus. Thus, elastomeric force is lost when the polypeptide chains are randomized due to heating at high temperature. Clearly, elastomeric force is due to nonrandom polypeptide structure. In spite of this, elastomeric force is demonstrated to be dominantly entropic in origin. The source of the entropic elastomeric force is demonstrated to be the result of internal chain dynamics, and the mechanism is called the librational entropy mechanism of elasticity. There is significant application to the finding that elastomeric force develops due to an inverse temperature transition. By changing the hydrophobicity of the polypeptide, the temperature range for the inverse temperature transition can be changed in a predictable way, and the temperature range for the development of elastomeric force follows. Thus, elastomers have been prepared where the development of elastomeric force is shifted over a 40 degrees C temperature range from a midpoint temperature of 30 degrees C for the polypentapeptide to 10 degrees C by increasing hydrophobicity with addition of a single CH2 moiety per pentamer and to 50 degrees C by decreasing hydrophobicity.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

3.
Fibrous elastin of the biological elastic fiber is a cross-linked condensed state in which there is roughly one-half polypeptide and one-half water. The precursor protein tropoelastin, a chemical fragmentation product -elastin, and a sequential polypeptide (l·Val1-l·Pro2-Gly3-l·Val4-Gly5) n , which is a prominent primary structural feature of tropoelastin, are each soluble in all proportions in water at 20°C. On heating to physiological temperatures, each undergoes aggregation and forms a dense viscoelastic phase, which as the fiber itself, is about 60% water. This reversible heat-elicited condensed phase is called the coacervate. Circular dichroism studies show coacervation to be a process of increasing intramolecular order. Electron microscopy (light, scanning, and transmission) shows coacervation to be a process of increasing order intermolecularly. Thus a rise in temperature between 20 and 40°C results in an increase in order of the polypeptide. Coacervation is an inverse temperature transition, and the condensed state is anisotropic at the molecular level. Thermoelasticity studies in water on bovine ligamentum nuchae fibrous elastin and on -irradiation cross-linked polypentapeptide coacervates show increases in elastomeric force,f, over the same 20–40°C temperature range in which the inverse temperature transition gives rise to the coacervate, and the constancy off/T with temperature, once the transition is effectively completed, suggests a high-entropy component to the elastomeric force. Thus the data argue for an anisotropic-entropic elastomer.Detailed conformational studies on the polypentapeptide result in the development of a -spiral conformation in which there are regularly recurring -turns in loose helical array (a structure that forms on raising the temperature) and in which there are recurring dynamic suspended segments that are the focal point of large, low-energy oscillatory motions called librations. The structure gives rise to a librational entropy mechanism of elasticity wherein the amplitudes of the rocking motions become damped on stretching. This perspective is substantiated by dielectric relaxation studies on the coacervate state and by characterization of synthetic analogs of the polypentapeptide. Dielectric relaxation studies on a concentrated state of about 60% water show the development of a regular structure over the same temperature range as for the development of the coacervate state, and the development of the regular structure with increasing temperature is seen to parallel the development of elastomeric force with increasing temperature. Increasing elastomeric force coincides with increasing regularity of structure! Synthetic analogs of the polypentapeptide, designed to interfere with the librational processes of the suspended segment, impair elastic function, and an analog that makes the -turn more rigid results in increased elastic modulus. This development of a librational entropy mechanism for protein elasticity is a departure from the kinetic theory of rubber elasticity, the random network perspective that has dominated the traditional view of biological elasticity for the past several decades. The new perspective opens the way to insightful consideration of new elastomeric biomaterials with numerous biomedical applications.  相似文献   

4.
High molecular weight polytetrapeptide of elastin, (L.Val1-L.Pro2-Gly3-Gly4)n, was synthesized using activation of the (GGVP) permutation for polymerization. The temperature-dependence of aggregation was characterized as a function of concentration and the circular dichroism spectra were obtained in the 20 degrees to 70 degrees C temperature range. The latter showed an inverse temperature transition centered near 50 degrees C in which polypeptide order increased on raising the temperature. A concentration of 0.6 g of polytetrapeptide in 1 g of water was gamma irradiation cross-linked (20 Mrad) to form an elastomeric matrix. A study of the temperature-dependence of elastomeric force demonstrated a transition toward increased force on raising the temperature with a midpoint of the transition near 50 degrees C. Thus, there is a correlation between increase in intramolecular order and elastomeric force development. These results are compared to previous results on the polypentapeptide of elastin, (VPGVG)n and on an analog, (IPGVG)n, to demonstrate that the temperature of the transition is proportional to the hydrophobicity of the repeating unit. The point is noted that the elastomeric force development correlates better with intramolecular ordering than with intermolecular processes.  相似文献   

5.
Carbon-13 NMR longitudinal relaxation time and line-width studies are reported on the coacervate concentration (about 60% water by weight) of singly carbonyl carbon enriched polypentapeptides of elastin: specifically, (L-Val1-L-[1-13C]Pro2-Gly3-L-Val4-Gly5)n and (L-Val1-L-Pro2-Gly3-L-Val4-[1-13C]Gly5)n. On raising the temperature from 10 to 25 degrees C and from 40 to 70 degrees C, carbonyl mobility increases, but over the temperature interval from 25 to 40 degrees C, the mobility decreases. The results characterize an inverse temperature transition in the most fundamental sense of temperature being a measure of molecular motion. This transition in the state of the polypentapeptide indicates an increase in order of polypeptide on raising the temperature from 25 degrees C to physiological temperature. This fundamental NMR characterization corresponds with the results of numerous other physical methods, e.g., circular dichroism, dielectric relaxation, and electron microscopy, that correspondingly indicate an increase in order of the polypentapeptide both intramolecularly and intermolecularly for the same temperature increase from 25 to 40 degrees C. Significantly with respect to elastomeric function, thermoelasticity studies on gamma-irradiation cross-linked polypentapeptide coacervate show a dramatic increase in elastomeric force over the same interval that is here characterized by NMR as an inverse temperature transition. The temperature dependence of mobility above 40 degrees C indicates an activation energy of the order of 1.2 kcal/mol, which is the magnitude of barrier expected for elasticity.  相似文献   

6.
Elastomeric proteins are widespread in the animal kingdom, and their main function is to confer elasticity and resilience to organs and tissues. Besides common functional properties, elastomeric proteins share a common sequence design. They are usually constituted by repetitive sequences with a high content of glycine residues. From a conformational point of view, all the elastomeric proteins since now analyzed show a dynamic equilibria between folded (mainly beta-turns) and extended (polyproline II and beta-strands) conformations that could be at the origin of the high entropy of the relaxed state. As a matter of fact, elastin, lamprin, abductin, as well as the PEVK domain of titin share the same conformational ensemble, thus pointing to a common molecular mechanism as the origin of elasticity. CD spectroscopy represents the proper spectroscopic technique to be used overall because of its particular sensitivity to the presence of PPII structure. Its use in the molecular studies of elastin, abductin, and lamprin as well as the recently analyzed protein resilin will be presented.  相似文献   

7.
In wheat, the high-molecular weight (HMW) glutenin subunits are known to contribute to gluten viscoelasticity, and show some similarities to elastomeric animal proteins as elastin. When combining the sequence of a glutenin with that of elastin is a way to create new chimeric functional proteins, which could be expressed in plants. The sequence of a glutenin subunit was modified by the insertion of several hydrophobic and elastic motifs derived from elastin (elastin-like peptide, ELP) into the hydrophilic repetitive domain of the glutenin subunit to create a triblock protein, the objective being to improve the mechanical (elastomeric) properties of this wheat storage protein. In this study, we investigated an expression model system to analyze the expression and trafficking of the wild-type HMW glutenin subunit (GSW) and an HMW glutenin subunit mutated by the insertion of elastin motifs (GSM-ELP). For this purpose, a series of constructs was made to express wild-type subunits and subunits mutated by insertion of elastin motifs in fusion with green fluorescent protein (GFP) in tobacco BY-2 cells. Our results showed for the first time the expression of HMW glutenin fused with GFP in tobacco protoplasts. We also expressed and localized the chimeric protein composed of plant glutenin and animal elastin-like peptides (ELP) in BY-2 protoplasts, and demonstrated its presence in protein body-like structures in the endoplasmic reticulum. This work, therefore, provides a basis for heterologous production of the glutenin-ELP triblock protein to characterize its mechanical properties.  相似文献   

8.
Thermoelastic (force–temperature) measurements were carried out on elastin networks, swollen with a nonvolatile diluent, in elongation over the temperature range 8–35°C. The experiments were conducted at constant composition (thermodynamically closed system) rather than at swelling equilibrium (open system), thus avoiding the need for approximate corrections to account for the changes in swelling with temperature and with elongation. The results indicate that the elastic force is primarily of entropic origin and thus support the random-network model rather than the liquid-drop model for elastin in the rubberlike state. A significant energetic contribution to the force was observed, however, as is usually the case for a variety of elastomeric polymers.  相似文献   

9.
Elastin provides extensible tissues, including arteries and skin, with the propensity for elastic recoil, whereas amyloid fibrils are associated with tissue-degenerative diseases, such as Alzheimer's. Although both elastin-like and amyloid-like materials result from the self-organization of proteins into fibrils, the molecular basis of their differing physical properties is poorly understood. Using molecular simulations of monomeric and aggregated states, we demonstrate that elastin-like and amyloid-like peptides are separable on the basis of backbone hydration and peptide-peptide hydrogen bonding. The analysis of diverse sequences, including those of elastin, amyloids, spider silks, wheat gluten, and insect resilin, reveals a threshold in proline and glycine composition above which amyloid formation is impeded and elastomeric properties become apparent. The predictive capacity of this threshold is confirmed by the self-assembly of recombinant peptides into either amyloid or elastin-like fibrils. Our findings support a unified model of protein aggregation in which hydration and conformational disorder are fundamental requirements for elastomeric function.  相似文献   

10.
During the last half century, identification of an ideal (predominantly entropic) protein elastomer was generally thought to require that the ideal protein elastomer be a random chain network. Here, we report two new sets of data and review previous data. The first set of new data utilizes atomic force microscopy to report single-chain force-extension curves for (GVGVP)(251) and (GVGIP)(260), and provides evidence for single-chain ideal elasticity. The second class of new data provides a direct contrast between low-frequency sound absorption (0.1-10 kHz) exhibited by random-chain network elastomers and by elastin protein-based polymers. Earlier composition, dielectric relaxation (1-1000 MHz), thermoelasticity, molecular mechanics and dynamics calculations and thermodynamic and statistical mechanical analyses are presented, that combine with the new data to contrast with random-chain network rubbers and to detail the presence of regular non-random structural elements of the elastin-based systems that lose entropic elastomeric force upon thermal denaturation. The data and analyses affirm an earlier contrary argument that components of elastin, the elastic protein of the mammalian elastic fibre, and purified elastin fibre itself contain dynamic, non-random, regularly repeating structures that exhibit dominantly entropic elasticity by means of a damping of internal chain dynamics on extension.  相似文献   

11.
D W Urry  M M Long  R D Harris  K U Prasad 《Biopolymers》1986,25(10):1939-1953
The Ile1 analog of the polypentapeptide of elastin, (L · Ile1-L · Pro2-Gly3-L · Val4-Gly5)n, abbreviated as Ile1-PPP, was synthesized with n > 100 to determine the effect of the increased hydrophobicity of the pentamer resulting from Val1 replacement by Ile1 on the previously characterized inverse temperature transition of the polypentapeptide of elastin (PPP). Ile1-PPP, dissolved in water at 4°C, was found to aggregate, forming a viscoelastic coacervate on raising the temperature. The onset of aggregation was 8°C for Ile1-PPP, as compared to 24°C for PPP. Characterization by CD demonstrated an increase in intramolecular order on raising the temperature from 8°C to 25°C, and demonstrated similar conformations for PPP and Ile1-PPP before and after their respective transitions. The CD-characterized transition also occurred at a temperature some 15°C lower than that of PPP. By means of 20-Mrad γ-irradiation cross-linking of the Ile1-PPP coacervate, an elastomeric matrix was formed with an elastic modulus, similar to that of 20-Mrad cross-linked PPP. The temperature dependence of elastomeric force of cross-linked Ile1-PPP showed an abrupt increase from essentially zero at 8°C to three-quarters of full force at 10°C and essentially full force by 20–25°C. This development of elastomeric force for the more hydrophobic Ile1-PPP matrix, which parallels the increase in intramolecular order characterized by the CD studies, also occurs at a temperature some 15°C lower than that for the PPP matrix. Thus, in these elastomeric polypeptides, development of elastomeric force is coupled to an inverse temperature transition, the temperature of which depends inversely on the hydrophobicity of the constituent pentamer. It appears that a series of elastomeric polypeptide biomaterials are possible in which the temperature over which elastomeric force develops can be varied.  相似文献   

12.
The cyanogen bromide‐resistant proteins of lamprey cartilage are biochemically related to the mammalian elastic protein, elastin. This study investigates their mechanical properties and enquires whether, like elastin, long‐range elasticity arises in them from a combination of entropic and hydrophobic mechanisms. Branchial and pericardial proteins resembled elastin mechanically, with elastic moduli of 0.13–0.35 MPa, breaking strains of 50%, and low hysteresis. Annular and piston proteins had higher elastic moduli (0.27–0.75 MPa) and larger hysteresis. Exchanging solvent water for trifluoroethanol increased the elastic moduli, whereas increasing temperature lowered the elastic moduli. Raman microspectrometry showed small differences in side‐chain modes consistent with reported biochemical differences. Decomposition of the amide I band indicated that the secondary structures were like those of elastin, preponderantly unordered, which probably confer the conformational flexibility necessary for entropy elasticity. Piston and annular proteins showed the strongest interactions with water, suggesting, together with the mechanical testing data, a greater role of hydrophobic interactions in their mechanics. Two‐photon imaging of intrinsic fluorescence and dye injection experiments showed that annular and piston proteins formed closed‐cell honeycomb structures, whereas the branchial and pericardial proteins formed open‐cell structures, which may account for the differences in mechanical properties. © 2014 Wiley Periodicals, Inc. Biopolymers 103: 187–202, 2015.  相似文献   

13.
The structure and properties of gluten: an elastic protein from wheat grain   总被引:12,自引:0,他引:12  
The wheat gluten proteins correspond to the major storage proteins that are deposited in the starchy endosperm cells of the developing grain. These form a continuous proteinaceous matrix in the cells of the mature dry grain and are brought together to form a continuous viscoelastic network when flour is mixed with water to form dough. These viscoelastic properties underpin the utilization of wheat to give bread and other processed foods. One group of gluten proteins, the HMM subunits of glutenin, is particularly important in conferring high levels of elasticity (i.e. dough strength). These proteins are present in HMM polymers that are stabilized by disulphide bonds and are considered to form the 'elastic backbone' of gluten. However, the glutamine-rich repetitive sequences that comprise the central parts of the HMM subunits also form extensive arrays of interchain hydrogen bonds that may contribute to the elastic properties via a 'loop and train' mechanism. Genetic engineering can be used to manipulate the amount and composition of the HMM subunits, leading to either increased dough strength or to more drastic changes in gluten structure and properties.  相似文献   

14.
Temperature dependence of length of elastin and its polypentapeptide   总被引:1,自引:0,他引:1  
Comparison of the temperature dependence of elastomer length of the cross-linked protein, elastin, and of gamma-irradiation cross-linked poly(VPGVG), the polypentapeptide of elastin, with that of latex rubber demonstrate markedly dissimilar behaviors between a classical rubber and the protein and polypeptide elastomers. In the absence of a load latex rubber expands with increasing temperature as is known for classical rubbers comprised of a network of random chains whereas the protein and polypeptide elastomers markedly decrease in length. When under load with a constant applied force, as a classical rubber, latex linearly decreases length with increasing temperature whereas the decrease in length is very non-linear with temperature increase for the protein and polypeptide elastomers. The protein and polypeptide elastomers examined here do not exhibit the characteristic and fundamental temperature dependence of length considered typical of networks of random chains. Accordingly the more complex and even inverse behavior of elastin and the polypentapeptide of elastin in the absence of load require consideration of structural perspectives different from those of a random chain network with negligible interchain interactions.  相似文献   

15.
Z R Wasserman  F R Salemme 《Biopolymers》1990,29(12-13):1613-1631
A repetitive polypentapeptide organized as a connected chain of beta-bends is believed to be an important structural element of elastin, the major elastomer in biological systems. Molecular dynamics simulations were carried out on hydrated polymers of (Val-Pro-Gly-Val- Gly)18 at various extensions. Analysis of the fluctuations of backbone angles in relaxed elastin showed that particularly large-amplitude torsional motions occur in phi and psi angles of residues connecting sequentially adjacent hairpin bends. Many such motions reflect peptide plane librations that result from anticorrelated crankshaft rotations of psi i and phi i+1. These effects were much reduced in stretched polymer models. The conformational entropy of relaxed and stretched elastin models was estimated using a treatment due to Meirovitch, and gave a calculated decrease in entropy of about 1 cal/mol deg when the polymer was stretched to 1.75 times its original length. There are large changes in solvent-accessible surface area during the initial stages of elastin stretching. Collectively these results suggest that hydrophobic interactions make contributions to elastin entropy at low extensions, but that librational mechanisms make larger contributions to the elastic restoring force at longer extensions.  相似文献   

16.
17.
Fibrin fibers form the structural scaffold of blood clots. Thus, their mechanical properties are of central importance to understanding hemostasis and thrombotic disease. Recent studies have revealed that fibrin fibers are elastomeric despite their high degree of molecular ordering. These results have inspired a variety of molecular models for fibrin’s elasticity, ranging from reversible protein unfolding to rubber-like elasticity. An important property that has not been explored is the timescale of elastic recoil, a parameter that is critical for fibrin’s mechanical function and places a temporal constraint on molecular models of fiber elasticity. Using high-frame-rate imaging and atomic force microscopy-based nanomanipulation, we measured the recoil dynamics of individual fibrin fibers and found that the recoil was orders of magnitude faster than anticipated from models involving protein refolding. We also performed steered discrete molecular-dynamics simulations to investigate the molecular origins of the observed recoil. Our results point to the unstructured αC regions of the otherwise structured fibrin molecule as being responsible for the elastic recoil of the fibers.  相似文献   

18.
Fibrin fibers form the structural scaffold of blood clots. Thus, their mechanical properties are of central importance to understanding hemostasis and thrombotic disease. Recent studies have revealed that fibrin fibers are elastomeric despite their high degree of molecular ordering. These results have inspired a variety of molecular models for fibrin’s elasticity, ranging from reversible protein unfolding to rubber-like elasticity. An important property that has not been explored is the timescale of elastic recoil, a parameter that is critical for fibrin’s mechanical function and places a temporal constraint on molecular models of fiber elasticity. Using high-frame-rate imaging and atomic force microscopy-based nanomanipulation, we measured the recoil dynamics of individual fibrin fibers and found that the recoil was orders of magnitude faster than anticipated from models involving protein refolding. We also performed steered discrete molecular-dynamics simulations to investigate the molecular origins of the observed recoil. Our results point to the unstructured αC regions of the otherwise structured fibrin molecule as being responsible for the elastic recoil of the fibers.  相似文献   

19.
How nature tunes sequences of disordered protein to yield the desired coiling properties is not yet well understood. To shed light on the relationship between protein sequence and elasticity, we here investigate four different natural disordered proteins with elastomeric function, namely: FG repeats in the nucleoporins; resilin in the wing tendon of dragonfly; PPAK in the muscle protein titin; and spider silk. We obtain force-extension curves for these proteins from extensive explicit solvent molecular dynamics simulations, which we compare to purely entropic coiling by modeling the four proteins as entropic chains. Although proline and glycine content are in general indicators for the entropic elasticity as expected, divergence from simple additivity is observed. Namely, coiling propensities correlate with polyproline II content more strongly than with proline content, and given a preponderance of glycines for sufficient backbone flexibility, nonlocal interactions such as electrostatic forces can result in strongly enhanced coiling, which results for the case of resilin in a distinct hump in the force-extension curve. Our results, which are directly testable by force spectroscopy experiments, shed light on how evolution has designed unfolded elastomeric proteins for different functions.  相似文献   

20.
Elastin and silk spidroins are fibrous, structural proteins with elastomeric properties of extension and recoil. While elastin is highly extensible and has excellent recovery of elastic energy, silks are particularly strong and tough. This study describes the biophysical characterization of recombinant polypeptides designed by combining spider wrapping silk and elastin‐like sequences as a strategy to rationally increase the strength of elastin‐based materials while maintaining extensibility. We demonstrate a thermo‐responsive phase separation and spontaneous colloid‐like droplet formation from silk‐elastin block copolymers, and from a 34 residue disordered region of Argiope trifasciata wrapping silk alone, and measure a comprehensive suite of tensile mechanical properties from cross‐linked materials. Silk‐elastin materials exhibited significantly increased strength, toughness, and stiffness compared to an elastin‐only material, while retaining high failure strains and low energy loss upon recoil. These data demonstrate the mechanical tunability of protein polymer biomaterials through modular, chimeric recombination, and provide structural insights into mechanical design. © 2016 Wiley Periodicals, Inc. Biopolymers 105: 693–703, 2016.  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