首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The ability of stem bark to resist bending forces was examined by testing in bending segments of Acer saccharum, Fraxinus americana, and Quercus robur branches with and without their bark. For each species, the bark contributed significantly to the ability of stem segments differing in age to resist bending forces, but its contribution was age-dependent and differed among the three species. The importance of the mechanical role of the bark decreased basipetally with increasing age of F. americana and Q. robur stem segments and was superceded by that of the wood for segments ≥ 6 yr old. A. saccharum bark was as mechanically important as the wood for stem segments 7 yr old but was not a significant stiffening agent for younger or older portions of stems. On average, the stiffness of the bark from all three species was 50% that of the wood. However, the geometric contribution to the flexural rigidity of stems made by the bark (i.e., the bark's second moment of area) was sufficiently large to offset its lower stiffness (Young's modulus) relative to that of the wood. A simple model is presented that shows that the bark must be as mechanically important as the wood when its radial thickness equals 32% that of the wood and its stiffness is 50% that of the wood. Based on this model, which is shown to comply with the data from three species purported to have stiff woods, it is evident that the role of the bark cannot be neglected when considering the mechanical behavior of juvenile woody stems subjected to externally applied bending forces.  相似文献   

2.
The ratio of erythro and threo forms of beta-O-4 structures in tension wood lignin was investigated by ozonation analysis of wood meal taken from various positions in the stem of yellow poplar (Liriodendron tulipifera). The proportion of the erythro form was higher in tension wood than in opposite wood, and the methoxyl group content showed a similar trend. The proportion of the erythro form and the methoxyl group content in the 7 positions in the stem lignin was correlated (correlation coefficient R=0.98), suggesting that the type of aromatic ring, syringyl or guaiacyl, is one of the factors which stereochemically controls the ratio of erythro and threo forms of beta-O-4 structures during lignin formation.  相似文献   

3.
Righting of two tilted white pine (Pinus strobus L.) stem leaders by compression wood formation was followed for 16 weeks. The natural curves and three deflection curves under added end loads were determined from weekly field photographs. Data for self-loading and cross sectional diameters were interpolated from original estimated and final measurements. A mechanical-mathematical model was developed to predict curves under zero gravity for each stem each week. The model estimated stiffness of the leaders independently for each week, and the stiffnesses were consistent throughout the experiment. A second model was developed to simulate the deflection curves assumed when the zero gravity curves were subjected to different end loads. These predicted curves were nearly identical to the observed curves from the photographs, thus verifying the assumptions in the first model. Data from this study will be used to investigate the mechanical aspects of compression wood induction and action as the stem is bent upward toward the vertical.  相似文献   

4.
During tree growth, hardwoods can initiate the formation of tension wood, which is a strongly stressed wood on the upper side of the stem and branches. In Eucalyptus globulus, tension wood presents wider and thicker cell walls with low lignin, similar glucan and high xylan content, as compared to opposite wood. In this work, tension and opposite wood of E. globulus trees were separated and evaluated for the production of bioethanol using ethanol/water delignification as pretreatment followed by simultaneous saccharification and fermentation (SSF). Low residual lignin and high glucan retention was obtained in organosolv pulps of tension wood as compared to pulps from opposite wood at the same H-factor of reaction. The faster delignification was associated with the low lignin content in tension wood, which was 15% lower than in opposite wood. Organosolv pulps obtained at low and high H-factor (3,900 and 12,500, respectively) were saccharified by cellulases resulting in glucan-to-glucose yields up to 69 and 77%, respectively. SSF of the pulps resulted in bioethanol yields up to 35 g/l that corresponded to 85–95% of the maximum theoretical yield on wood basis, considering 51% the yield of glucose to ethanol conversion in fermentation, which could be considered a very satisfactory result compared to previous studies on the conversion of organosolv pulps from hardwoods to bioethanol. Both tension and opposite wood of E. globulus were suitable raw materials for organosolv pretreatment and bioethanol production with high conversion yields.  相似文献   

5.
The forage brassicas are a useful model system for the study of wood formation because the thickened cell walls of their vascular tissue can vary widely in lignin content. Solid-state 13C NMR spectroscopy was used to quantify lignin, and determine features of its structure, in the vascular cell walls of forage rape (Brassica napus L.), and Thousandhead and marrowstem cultivars of kale (Brassica oleracea L. var. acephala). During the first season of vegetative growth, lignin levels in these cell walls remained low in the upper part of the stems despite the physical resemblance of this tissue to wood. The extended flowering stems produced in the following year were thinner and their vascular tissue contained much more strongly lignified cell walls. The structure of the lignin was typical of angiosperm wood. It showed only small variations in syringyl/guaiacyl ratio, but this ratio increased with lignin content and thus with the proportion of the lignin that was associated with secondary cell-wall layers.  相似文献   

6.
? Reduced lignin content in perennial crops has been sought as a means to improve biomass processability for paper and biofuels production, but it is unclear how this could affect wood properties and tree form. ? Here, we studied a nontransgenic control and 14 transgenic events containing an antisense 4-coumarate:coenzyme A ligase (4CL) to discern the consequences of lignin reduction in poplar (Populus sp.). During the second year of growth, trees were grown either free-standing in a field trial or affixed to stakes in a glasshouse. ? Reductions in lignin of up to 40% gave comparable losses in wood strength and stiffness. This occurred despite the fact that low-lignin trees had a similar wood density and up to three-fold more tension wood. In free-standing and staked trees, the control line had twice the height for a given diameter as did low-lignin trees. Staked trees had twice the height for a given diameter as free-standing trees in the field, but did not differ in wood stiffness. ? Variation in tree morphogenesis appears to be governed by lignin x environment interactions mediated by stresses exerted on developing cells. Therefore our results underline the importance of field studies for assessing the performance of transgenic trees with modified wood properties.  相似文献   

7.
To identify the chromosomal regions affecting wood quality traits, we conducted a genome-wide quantitative trait locus (QTL) analysis of wood quality traits in Eucalyptus nitens. This information is important to exploit the full potential of the impending Eucalyptus genome sequence. A three generational mapping population consisting of 296 progeny trees was used to identify QTL associated with several wood quality traits in E. nitens. Thirty-six QTL positions for cellulose content, pulp yield, lignin content, density, and microfibril angle (MFA) were identified across different linkage groups. On linkage groups (LG)2 and 8, cellulose QTL cluster with pulp yield and extractives QTL while on LG4 and 10 cellulose and pulp yield QTLs cluster together. Similarly, on LG4, 5, and 6 QTL for lignin traits were clustered together. At two positions, QTL for MFA, a physical trait related to wood stiffness, were clustered with QTL for lignin traits. Several cell wall candidate genes were co-located to QTL positions affecting different traits. Comparative QTL analysis with Eucalyptus globulus revealed two common QTL regions for cellulose and pulp yield. The QTL positions identified in this study provide a resource for identifying wood quality genes using the impending Eucalyptus genome sequence. Candidate genes identified in this study through co-location to QTL regions may be useful in association studies.  相似文献   

8.
A. J. Peel 《Planta》1983,158(6):512-516
Uniformly labelled [14C]glucose was introduced into the xylem of segments of willow stem. Forty-eight hours later sieve-tube sucrose was collected via servered aphid stylets, and the distribution of radioactivity in the hexose moieties of this was compared with the distribution in those of sucrose extracted from the segment. Very little correlation was found between the two sets of values, indicating possible inversion during loading. This lack of correlation could not be attributed to contributions to the sieve-tube sucrose from pools of labelled hexoses in the segment. Further experiments, however, showed quite high degrees of correlation between sieve-tube sucrose and sucrose extracted from the wood, indicating that the latter tissue was a major source of sieve-tube sucrose. This conclusion was substantiated in experiments in which sieve-tube exudate, obtained from stem segments, was compared with exudate obtained from the isolated bark of the segment. In other experiments, stylets were established on stem segments, then on isolated pieces of bark obtained from these segments. Sucrose and potassium exudation rates fell by as much as 50% on removing the bark from the segment. It was not possible to formulate a precise figure for the contribution of the wood to stylet exudation owing to injury effects and the complexity of the experimental system. No firm evidence could be found in support of the view that sucrose is inverted during loading of sieve elements from the storage cells of the stem.  相似文献   

9.
Sap flow from excised maple stems collected over the winter (1986/87) was correlated with stem water content. Stem water content was high in the fall (>0.80) and decreased rapidly during 2 weeks of continuous freezing temperatures in late winter (<0.60). Exudation of sap from stem segments subjected to freeze/thaw cycles was small (<10 mL/kg) in the fall, but substantial exudation (45-50 mL/kg) occurred following the decline in water content. These observations are consistent with Milburn's and O'Malley's models (J.A. Milburn, P.E.R. O'Malley [1984] Can J Bot 62: 2101-2106; P.E.R. O'Malley, J.A. Milburn [1983] Can J Bot 61:3100-3106) of sap absorption into gas-filled fibers during freezing. Exudation volume was increased 200 to 300% in maple stems originally at high water content (>0.80) after perfusion with sucrose and dehydration at −12°C. Sap flow was also induced in butternut stem segments after the same treatment. Thus, sap flow may not be unique to maples. Sap flow could not be increased in stem segments dehydrated at 4°C. Migration of water molecules from small ice crystals in fibers to larger crystals in vessels while stems were frozen may account for increase exudation after dehydration at −12°C. This would result in preferential dehydration of fibers and a distribution of gas and sap favorable for stem-based sap flow.  相似文献   

10.
Prior explicit allometric models are extended to predict the scaling relationship between the ability of plants with leafless stems to harvest sunlight H and total standing plant biomass M(T) (which equals the sum of standing stem and root biomass, M(S) and M(R)). Provided that H scales in a directly proportional manner (isometrically) with respect to either stem surface area (i.e.H proportional, variant SA(S) ) or total stem biomass (i.e. H proportional, variant M(S)), the allometric model presented here predicts that SA(S) proportional, variant M(T)(3/4) or M(S) proportional, variant M(T)(3/4), respectively. These alternative predictions are tested empirically using data for standing stem and root biomass gathered for the large columnar cactus species Pachycereus pringlei. Statistical comparisons between observed and predicted scaling relationships indicate that SA(S) proportional, variant M(T)(3/4), whereas M(S) proportional, variant M(T)(3/4) is mathematically inconsistent with the observation that stem biomass scales nearly isometrically with respect to root biomass. The contention that the H of leafless stems scales isometrically with respect to stem surface area is thus reasonable both theoretically and empirically.  相似文献   

11.
We report on the root system of the large columnar cactus species Pachycereus pringlei to explore the hypothesis that increasing plant size decreases the ability to resist wind-throw but increases the capacity to absorb and store nutrients in roots (i.e., plant size limits the performance of these functions and may shift the performance of one function in favor of another as size increases). Based on 18 plants differing in size, the root system is characterized by a broad and deep bayonet-like root central to a shallow and extensive lateral system of root elements bearing sinker roots near the stem base. All root types have a living secondary cortex and contain wood with a large volume fraction of ray tissues that increases toward the stem base. Wood stiffness and tensile strength are correlated negatively with the ray tissue volume fraction and thus decrease toward the stem base in lateral and bayonet roots. Calculations show that the ability of the bayonet and proximal lateral root elements to resist wind-throw decreases with increasing plant size, whereas the nutrient absorption/storage capacity of the total root system increases with plant size (i.e., a size-dependent shift between these two root functions occurs).  相似文献   

12.
Stem segments of the maize (Zea mavs L.) hybrids LGH, Eta Ipho (EI) and a brown midrib mutant. INRA 260 bm3 (bm3) were freeze-dried, ground and analysed for cell wall content, hemicellulose, cellulose, lignin and in vitro cell wall degradability with rumen fluid. Stem cross-sections, stained with acid phloroglucinol (AP) and chlorine sulphite (CS) showed an increased intensity in staining during maturation, but no considerable difference in staining intensity was observed between genotypes. The lignin content increased during maturation with evidently less lignin in bm3 than in EI and LGH. However, cell wall degradability did not differ between the older stem segments of EI and bm3, although the amount of lignin in LGH was twice that of bm3.

It can be concluded that an increase in lignin content occurs simultancously with a decrease in cell wall degradability within a genotype. However, between different genotypes the lignin content is not an indicator of degree of cell wall degradability.  相似文献   


13.
Creosote bush clones in the Mojave Desert develop by irregular radial growth, stem segmentation and the production of new stems at the outer edge of stem segments. The resulting circular clone encloses a central bare area as the central dead wood rots away. Old clones become elliptical and may exceed 20 m in length. Modern growth rates estimated from annual increments in stem wood of seedlings (0.73 mm/yr) and young clones (0.82 mm/yr) approximate those estimated for radiocarbon-dated wood samples (0.66 mm/yr). Assuming comparable growth rates through time, the extrapolated age of the largest known clone (average radius = 7.8 m) may approach 11,700 years. If growth rates have changed, that clone's age may be somewhat less.  相似文献   

14.
Chronic hypoxic pulmonary hypertension (HPH) is associated with large pulmonary artery (PA) stiffening, which is correlated with collagen accumulation. However, the mechanisms by which collagen contributes to PA stiffening remain largely unexplored. Moreover, HPH may alter mechanical properties other than stiffness, such as pulse damping capacity, which also affects ventricular workload but is rarely quantified. We hypothesized that collagen content and cross-linking differentially regulate the stiffness and damping capacity of large PAs during HPH progression. The hypothesis was tested with transgenic mice that synthesize collagen type I resistant to collagenase degradation (Col1a1(R/R)). These mice and littermate controls (Col1a1(+/+)) were exposed to hypoxia for 10 days; some were treated with β-aminopropionitrile (BAPN), which prevents new cross-link formation. Isolated PA dynamic mechanical tests were performed, and collagen content and cross-linking were measured. In Col1a1(+/+) mice, HPH increased both collagen content and cross-linking, and BAPN treatment prevented these increases. Similar trends were observed in Col1a1(R/R) mice except that collagen content further increased with BAPN treatment. Mechanical tests showed that in Col1a1(+/+) mice, HPH increased PA stiffness and damping capacity, and these increases were impeded by BAPN treatment. In Col1a1(R/R) mice, HPH led to a smaller but significant increase in PA stiffness and a decrease in damping capacity. These mechanical changes were not affected by BAPN treatment. Vessel-specific correlations for each strain showed that the stiffness and damping capacity were correlated with the total content rather than cross-linking of collagen. Our results suggest that collagen total content is critical to extralobar PA stiffening during HPH.  相似文献   

15.
16.
In Cactaceae, wood anatomy is related to stem morphology in terms of the conferred support. In species of cacti with dimorphic wood, a unique process occurs in which the cambium stops producing wide-band tracheids (WBTs) and produces fibers; this is associated with the aging of individuals and increases in size. Stem support and lignification have only been studied in fibrous tree-like species, and studies in species with WBTs or dimorphic wood are lacking. In this study, we approach this process with a chemical focus, emphasizing the role of wood lignification. We hypothesized that the degree of wood lignification in Cactaceae increases with height of the species and that its chemical composition varies with wood anatomy. To test this, we studied the chemical composition (cellulose, hemicellulose, and lignin content) in 13 species (2 WBTs wood, 3 dimorphic, and 8 fibrous) with contrasting growth forms. We also analyzed lignification in dimorphic and fibrous species to determine the chemical features of WBTs and fibers and their relationship with stem support. The lignin contents were characterized by Fourier transform infrared spectroscopy and high performance liquid chromatography. We found that 11 species have a higher percentage (>35%) of lignin in their wood than other angiosperms or gymnosperms. The lignin chemical composition in fibrous species is similar to that of other dicots, but it is markedly heterogeneous in non-fibrous species where WBTs are abundant. The lignification in WBTs is associated with the resistance to high water pressure within cells rather than the contribution to mechanical support. Dimorphic wood species are usually richer in syringyl lignin, and tree-like species with lignified rays have more guaiacyl lignin. The results suggest that wood anatomy and lignin distribution play an important role in the chemical composition of wood, and further research is needed at the cellular level.  相似文献   

17.
Compression wood conifer tracheids show different swelling and stiffness properties than those of usual normal wood, which has a practical function in the living plant: when a conifer shoot is moved from its vertical position, compression wood is formed in the under part of the shoot. The growth rate of the compression wood is faster than in the upper part resulting in a renewed horizontal growth. The actuating and load-carrying function of the compression wood is addressed, on the basis of its special ultrastructure and shape of the tracheids. As a first step, a quantitative model is developed to predict the difference of moisture-induced expansion and axial stiffness between normal wood and compression wood. The model is based on a state space approach using concentric cylinders with anisotropic helical structure for each cell-wall layer, whose hygroelastic properties are in turn determined by a self-consistent concentric cylinder assemblage of the constituent wood polymers. The predicted properties compare well with experimental results found in the literature. Significant differences in both stiffness and hygroexpansion are found for normal and compression wood, primarily due to the large difference in microfibril angle and lignin content. On the basis of these numerical results, some functional arguments for the reason of high microfibril angle, high lignin content and cylindrical structure of compression wood tracheids are supported.  相似文献   

18.
When humans hopin place or run forward, leg stiffness is increased to offsetreductions in surface stiffness, allowing the global kinematics andmechanics to remain the same on all surfaces. The purpose of thepresent study was to determine the mechanism for adjusting legstiffness. Seven subjects hopped in place on surfaces of differentstiffnesses (23-35,000 kN/m) while force platform, kinematic, andelectromyographic data were collected. Leg stiffness approximatelydoubled between the most stiff surface and the least stiff surface.Over the same range of surfaces, ankle torsional stiffness increased1.75-fold, and the knee became more extended at the time of touchdown(2.81 vs. 2.65 rad). We used a computer simulation to examine thesensitivity of leg stiffness to the observed changes in ankle stiffnessand touchdown knee angle. Our model consisted of four segments (foot,shank, thigh, head-arms-trunk) interconnected by three torsionalsprings (ankle, knee, hip). In the model, an increase in anklestiffness 1.75-fold caused leg stiffness to increase 1.7-fold. A changein touchdown knee angle as observed in the subjects caused legstiffness to increase 1.3-fold. Thus both joint stiffness and limbgeometry adjustments are important in adjusting leg stiffness to allow similar hopping on different surfaces.

  相似文献   

19.
ATP-binding cassette (ABC) proteins contain two nucleotide-binding domains (NBDs) and two transmembrane (TM) domains (TMDs). Interdomain interactions and packing of the TM segments are critical for function, and disruption by genetic mutations contributes to disease. P-glycoprotein (P-gp) is a useful model to identify mechanisms that repair processing defects because numerous arginine suppressor mutations have been identified in the TM segments. Here, we tested the prediction that a mechanism of arginine rescue was to promote intradomain interactions between TM segments and restore interdomain assembly. We found that suppressor W232R(TM4/TMD1) rescued mutants with processing mutations in any domain and restored defective NBD1-NBD2, NBD1-TMD2, and TMD1-TMD2 interactions. W232R also promoted packing of the TM segments because it rescued a truncation mutant lacking both NBDs. The mechanism of W232R rescue likely involved intradomain hydrogen bond interactions with Asn296(TM5) since only N296A abolished rescue by W232R and rescue was only observed when Trp232 was replaced with hydrogen-bonding residues. In TMD2, suppressor T945R(TM11) also promoted packing of the TM segments because it rescued the truncation mutant lacking the NBDs and suppressed formation of alternative topologies. We propose that T945R rescue was mediated by interactions with Glu875(TM10) since T945E/E875R promoted maturation while T945R/E875A did not.  相似文献   

20.
Radiation-induced wheat-rye chromosome translocation lines resistant to Hessian fly, Mayetiola destructor (say), were analyzed by in situ hybridization using total genomic and highly repetitive rye DNA probes pSc119 and pSc74. In situ hybridization analysis revealed the exact locations of the translocation breakpoints and allowed the estimation of the sizes of the transferred rye segments. T6BS·6BL-6RL and T4BS· 4BL-6RL are terminal translocations with either most of the complete long arm of rye chromosome 6R or only the distal 57% of the 6RL arm attached to the long arms of wheat chromosomes 6B and 4B, respectively. The breakpoint in T6BS·6BL-6RL is located at a fraction length (FL) of 0.11 in the long arm of T6BS 6BL-6RL and at FL 0.46 in the long arm of T4BS·4BL-6RL. Ti4AS·4AL-6RL-4AL is an intercalary translocation with the breakpoint located at FL 0.06 in the long arm of wheat chromosome 4A. The inserted 6RL segment, with the Hessian fly resistance gene, has a size of 0.7 m, and is the smallest and, so far, the first radiation-induced intercalary translocation identified in wheat.by R. Apples  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