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 rhizomorph branching habit in soil, competitive saprophytic ability and virulence were determined for 15 species of Armillaria from Europe, North America and Australia and New Zealand. In soil, rhizomorphs of northern hemisphere species branched either monopodially or dichotomously, whereas all five species from Australia and New Zealand branched dichotomously. The dry weight of rhizomorphs produced in soil by isolates of a species and by species was very variable. Species with monopodially branched rhizomorphs had significantly higher saprophytic colonization scores than dichotomously branched species and scores were significantly higher in Garry oak than Douglas‐fir segments and in fresh than autoclaved segments. The damage to Douglas‐fir seedlings caused by isolates of most dichotomously branched species was significantly greater than that caused by monopodially branched species. Species producing dichotomously branched rhizomorphs were more aggressive than monopodially branched species, killing 80% (vs. 17%) of seedlings that died during the first year of the 2‐year experiment.  相似文献   

2.
Weights of rhizomorphs of undetermined Armillaria sp. obtained using ring-trench and core soil sampling methods around aspen stumps in stands at different intervals after harvest were correlated. The core method was used to sample a second series of stands twice (at an interval of two growing seasons) at radial distances of 0.5, 1.0, and 1.5 meters from stump centers. Stands were uncut (at both samplings) or 3, 6, 9, 12, or 15 years after harvest (at the time of the initial sampling). Rhizomorphs were obtained from less than half the cores from the uncut stand, and the quantities (weights and lengths) were relatively small from both samplings. Mean numbers of soil cores yielding rhizomorphs and the mean quantities obtained generally increased as a function of the interval after harvest (and from the first to the second sampling), except for the oldest stand.  相似文献   

3.
Information about the entry of Armillaria into first-rotation pine and spruce stands was obtained by searching for infected stumps, rhizomorph systems or trees that had been killed. In pines Armillaria foci were very rare. In pure Norway spruce Armillaria lutea and A. mellea were detected in stumps but rhizomorphs did not extend into the soil; in Norway spruce mixed with oak, by contrast, A. lutea sometimes produced extensive rhizomorph systems. In Sitka spruce small groups of trees had been killed by A. ostoyae. All foci investigated in conifers contained different genotypes of Armillaria and probably originated from spore infection of stumps created by thinning. Some implications of these findings are discussed.  相似文献   

4.
In managed spruce forests, Armillaria cepistipes and A. ostoyae are efficient stump colonizers and may compete for these resources when they co‐occur at the same site. The aim of this experiment was to quantify the mutual competitive ability of the two Armillaria species in producing rhizomorphs and in colonizing Norway spruce (Picea abies) stumps. Five isolates of A. cepistipes and two isolates of A. ostoyae were simultaneously inoculated pair‐wise into pots containing a 4‐year‐old spruce seedling. For comparison, each isolate was also inoculated alone. One year after inoculation, stumps were created by cutting down the seedlings. Six months after creation of the stumps, rhizomorph production and stump colonization were assessed. Armillaria spp. were identified from 347 rhizomorphs and 48 colonized stumps. Armillaria cepistipes dominated both as rhizomorphs in the soil and on the stumps. Nevertheless, A. ostoyae was relatively more frequent on the stumps than in the soil and A. cepistipes was relatively more frequent in the soil than on the stumps. In both species, the ability to colonize the stumps in simultaneous inoculations was significantly reduced compared with single inoculations. In respect to rhizomorph production, simultaneous co‐inoculations had a slightly stimulatory effect on A. cepistipes and no significant effect on A. ostoyae. Our study suggests a rather neutralistic co‐existence of A. cepistipes and A. ostoyae as rhizomorphs in the soil. Concerning the ability to colonize stumps, the two species experience a mutual negative effect from the interaction, probably because of interspecific competition.  相似文献   

5.
The density of Armillaria gallica epiphytic rhizomorphs on tree collar was mapped in a young stand of pedunculate oaks (Quercus robur). A quick method to estimate the rhizomorph density was developed, investigating 19 trees, and was used in the subsequent study. The method consists of measuring the frequency of rhizomorphs present on a grid applied to an exposed small area of the tree collar. The pattern in tree collar rhizomorph density was closely related to the pattern of stump colonization by Armillaria and stumps were therefore important as inoculum sources. Rhizomorph density on tree collar showed a strong aggregated spatial pattern which could not be explained by variability in the soil characteristics, in particular waterlogging, or by a spatial pattern in tree vigour. Moreover, rhizomorph density on tree collar did not depend on their dominance status.  相似文献   

6.
Although several Armillaria species have been reported in Turkey, there is little information about their ecology in Turkish forests. In this study, we investigated five forest stands, approximately 5–74 ha in size, in Kastamonu province in the Black Sea Region of Turkey for the presence of Armillaria species in stumps and logs. The stands were mixed Abies nordmanniana ssp. bornmülleriana and Pinus sylvestris forests managed using a selective cuttings system; the proportion of fir in the total number of stems and stumps ranged from 36 to 98%. Based on sequence analysis of the internal transcribed spacer and intergenic spacer regions of the rDNA, all rhizomorphs sampled from the stumps and logs were of Armillaria ostoyae. The size of the genets was estimated with random amplified microsatellites analysis of the isolates and ranged from single stumps to approximately 450 m2. One to seven genets were found in each stand. These results indicate that the genets had arisen from spores and vegetative spread was limited on most sites.  相似文献   

7.
Armillaria spp. are some of the most important forest pathogens in mixed hardwood forests of southern New England, yet their role as prominent disturbance agents is still not fully appreciated. We investigated the distribution of Armillaria species across eight separate stands of northern hardwood and mixed oak forests in western Massachusetts. We were specifically interested in the Armillaria species distribution from live, symptomatic hosts and not in determining overall incidence in the forest. From 32 plots (16 within each forest type), 320 isolates were collected. Armillaria was routinely encountered causing disease of live trees. In total, 89% (286/320) of all isolations came from live hosts exhibiting symptoms of root and butt rot. Overall, A. gallica was the dominant species in each forest type, making up 88/160 (55%) isolates from northern hardwood and 153/160 (96%) of all isolations from mixed oak stands. However, northern hardwood forests showed much greater species diversity, as A. calvescens, A. gemina, A. ostoyae, and A. sinapina were all found. At one site, a northern hardwood forest surrounding a high elevation spruce-fir forest, A. ostoyae was the most abundant species encountered. All five Armillaria species were found causing disease of live hosts, including A. gemina, a species considered by some as weakly virulent. Armillaria gallica was found on 22/23 tree species’ sampled, and was found most often causing butt rot.  相似文献   

8.
Wightman  Kevyn Elizabeth  Shear  Ted  Goldfarb  Barry  Haggar  Jeremy 《New Forests》2001,22(1-2):75-96
Seedlings of three economically important and ecologicallydifferent native hardwoods, Cordia alliodora (Boraginaceae),Hyeronima alchorneoides (Euphorbiaceae), and Calophyllumbrasiliense (Clusiaceae), were grown in Rootrainers® (abook-type container), paper pots, and plastic bags filled witheither soil, soil with fertilizer, or compost substrates. Aftertransplanting in the field, treatments with and withoutfertilizer and herbicide were applied to all nursery stock types.In the nursery, species responded primarily to substrate type.Cordia grew better in bags of soil with NPK fertilizer andcompost than in unamended soil, probably responding to highernitrogen availability. Despite large treatment differences atplanting, there were no significant differences in plant sizeafter one year in the field between book containers and bags. Theexception were stump plants that were shorter and had highermortality. Hyeronima grew better in compost than in soil with orwithout fertilizer, probably responding to higher phosphorusavailability and lower bulk density of the compost. Plantsproduced in compost were also bigger after one year's fieldgrowth. Plants produced with soil or in paper pots had highermortality. Calophyllum grew less in compost compared to soil andgrew better when micronutrients were added to the compost andsoil. In the field, seedling produced in soil or withmicronutrients had higher survival or growth, respectively. Ingeneral, species grew better with herbicide and fertilizerapplication after transplanting. However, there were nointeractions with nursery treatments. Responses to fieldtreatments were independent and thus additive to the nurserytreatments. Differences in species response can be related tobiomass allocation patterns and ecology of the species.  相似文献   

9.
The basidiomycetes Armillaria cepistipes and Armillaria ostoyae frequently occur in the same forest stand. In this study, we determined the virulence of 20 isolates of A. cepistipes and 16 isolates of A. ostoyae on four different provenances of 2‐year‐old Norway spruce (Picea abies). Within 30 months after inoculation, 1.1 and 19.1% of the seedlings inoculated with A. cepistipes and A. ostoyae, respectively, had died or were dying. The incidence of dead and dying seedlings varied between 3 and 49% among the A. ostoyae isolates. The virulence of an isolate was positively correlated to its ability to produce rhizomorphs. One Norway spruce provenance showed significantly lower susceptibility to A. ostoyae than the other three. Rhizomorphs of both Armillaria species were attached to the root surface. The attached rhizomorphs of A. ostoyae, however, were associated with significantly more lesions. The virulence of the isolates was not correlated with their wood‐degrading capability for either of the Armillaria species.  相似文献   

10.
Isozyme patterns were examined to characterize biological species ofArmillaria in Japan. Of 25 different enzymes tested, glutamate dehydrogenase (GDH), lactate dehydrogenase (LDH), and aspartate aminotransferase (AAT) showed the most stable and reliable activities. These isozyme patterns showed little variability between haploid and diploid hyphae even under different cultural conditions among isolates tested. From the results, six Japanese biological species were clearly characterized through combination of GDH and LDH isozymes. There were no differences on the whole in isozyme patterns among European biological species ofArmillaria.  相似文献   

11.
Armillaria species from Japan were characterized using polymerase chain reaction-restriction fragment length polymorphism (PCR-RFLP) of the intergenic spacer region-1 (IGS-1) of ribosomal DNA (rDNA). Eleven different digestion patterns by restriction endonuclease Alu I were found among 70 isolates of seven Armillaria species in Japan. Isolates within Armillaria nabsnona, A. ostoyae, A. cepistipes, and Japanese biological species E showed the same Alu I digestion patterns. Five Alu I patterns were detected for A. gallica, three patterns for A. mellea, and two patterns for A. tabescens. Seven Armillaria species in Japan were clearly distinguished by using the profiles obtained when PCR products were digested with Alu I, Msp I, and Hae III restriction enzymes. There was considerable variability of Alu I restriction sites within the IGS-1 between the isolates of five Armillaria species, A. gallica, A. nabsnona, A. cepistipes, A. mellea, and A. tabescens, in Japan and those of their European and North American counterparts.  相似文献   

12.
The distribution of Armillaria species was investigated in Serbian forest ecosystems, in relation to the main host species attacked, forest‐types, geography and altitude. In total, 388 isolates were identified from 36 host species in 47 sites. Armillaria gallica was the most commonly observed species with the widest distribution and with an altitudinal range of 70–1450 m, it was the dominating Armillaria species in lowland alluvial forests and in Quercus and Fagus forests at higher elevations. Armillaria mellea occurred in Quercus spp. – dominated forests in the north and central regions at 70–1050 m. Sixty‐eight per cent of the A. mellea isolates were collected from living hosts, most commonly in declining conifer plantations. Armillaria ostoyae was distributed in the cooler coniferous forest types and plantations in the Dinaric Alps in the south of Serbia, at 850–1820 m. Armillaria cepistipes was found in the eastern and southern hilly and mountainous regions of the country, at 600–1900 m. Most isolates were obtained from conifers and rhizomorphs in the soil around decaying stumps. Armillaria tabescens was found only on dead oak material in the northern and eastern regions of the country at altitudes lower than 600 m.  相似文献   

13.
Abstract

Stem bases of 210 Fraxinus excelsior trees of three different health categories were sampled by the means of an increment borer in declining ash stands in northern Lithuania. From this number, 15 sound-looking, 132 declining and 63 dead trees from three discrete plots yielded 352 isolates, representing 75 fungal species. In addition, mycelial fans and rhizomorphs typical of Armillaria spp. from 205 and 20 trees, respectively, were sampled and subjected to fungal isolations. Species richness was similar in trees from each health category, but community structures differed, indicating that species composition of wood-inhabiting fungi in stems changes along with the changes in tree health condition. Armillaria cepistipes was the most common species (86 isolates from 210 wood samples, or 41.0%), isolated more frequently and consistently than any other potential tree pathogen. It also showed abundant occurrence on a majority of trees in the form of mycelial fans and rhizomorphs, from which 64 and 14 isolates of the fungus were obtained, respectively. The population structure of A. cepistipes revealed the presence of 53–93 genets per hectare, some of which extended up to 30–55?m.  相似文献   

14.
Turnover of nitrogen-rich root nodules follows the pruning of legume trees, forming a potentially important yet little studied way of N release to the soil. The effects of soil moisture, herbivory by soil mesofauna and microbial decomposition on the disappearance rate of woody legume nodules was studied in two tree/grass forage production associations. Litter bags containing nodules of Erythrina variegata L. (Papilionoideae: Phaseoleae) were incubated for four weeks in grass-covered alleys between Gliricidia sepium (Jacq.) Walp. (Papilionoideae: Robinieae) hedgerows, established on a deep alluvial Oxisol under a humid tropical climate and on a shallow Vertisol under a subhumid tropical climate in Guadeloupe, French Antilles. Soil moisture was regulated by irrigating or covering small plots from natural rainfall. Fine nylon mesh bags were used to study the rate of microbial decomposition, and open-ended perforated cylinders were used to estimate nodule herbivory. The chemical traits, especially the lignin: nitrogen ratio, of E. variegata and G. sepium nodules were similar (lignin: N 1.70 and 1.55, respectively), and suggest that the results are probably also applicable to the G. sepium nodules in the associations. Both soil moisture and decomposing agent (microbes or mesofauna) had a significant effect on the nodule disappearance rate, but soil type did not have any apparent effect. The nodule half-life varied from three to seven days under different treatments. The N release rate from the nodules was high, with N half-life varying from three to five days. Herbivory accounted for ca. 10% of total mass and N loss from nodules during the four-week field incubation period, but its importance increased towards the end of the incubation, especially in Vertisol, after the most easily decomposable part of the nodules had decayed. After pruning, the nodule N is released to soil more rapidly than from mulch. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

15.
[目的]分析凉水国家级自然保护区不同林型天然红松混交林林隙大小、凋落物放置位置和采样时间对土壤微生物碳(SMBC)的影响,揭示影响本地区SMBC变化的因素,为天然红松混交林生态系统碳循环的研究提供基础数据。[方法]在天然红松混交林3种林型的大、中、小林隙内不同位置的土壤表层放置装有红松、椴树、枫桦枯叶的分解袋,并以各自的郁闭林分为对照,在2012年植物生长季的6—9月,每月采集枯叶分解袋下0 10 cm土层土样,采用氯仿熏蒸-K2SO4浸提法测定SMBC。[结果]在椴树红松混交林(TP)内,林隙大小对SMBC的影响依次为小林隙大林隙中林隙;在云冷杉红松混交林(PAP)内,依次为中林隙大林隙小林隙;在枫桦红松混交林(BP)内,依次为大林隙中林隙小林隙。3种林型下,采样时间(月份)对SMBC均有显著的影响(P0.05);林隙大小对其影响均不显著(P0.05);枯叶分解袋放置位置对大、中、小林隙内SMBC的影响均不显著(P0.05)。[结论]不同林型下林隙大小对SMBC的影响排列顺序不同;枯叶分解袋放置位置对天然红松混交林3种林型大、中和小林隙内SMBC的影响均未达到显著水平。  相似文献   

16.
The effect of allicin (a stabilized garlic extract product) at five different concentrations (0, 20, 30, 50 and 100 mg/l) was studied in vitro on the growth rate of 100 isolates of Armillaria gallica and A. mellea. Isolates were obtained from 41 host genera growing in gardens located in 39 counties in the United Kingdom. Agar plugs of the actively growing Armillaria isolates were added to the centre of malt agar plates infused with allicin, and radial mycelium growth was measured on days 7, 14 and 21. The total number of rhizomorphs and length of rhizomorphs were also measured. Relative growth rates were calculated as the growth rate relative to the controls (0 mg/ l). The relative growth of each isolate at each allicin concentration was used to estimate EC50 values for A. mellea and A. gallica populations as well as individual isolates. EC50 values for both Armillaria spp. increased over time. The mean EC50 values for A. mellea of 16.0, 26.4 and 102 mg/l (days 7, 14 and 21, respectively) were higher than those for A. gallica (8.8, 7.9 and 11.0 mg/l) and probably relate to the more aggressive nature of A. mellea. Isolates with higher EC50 values were also more likely to produce more rhizomorphs. At allicin concentrations of 20 and 30 mg/l, the production of rhizomorphs and the growth rates of A. mellea isolates were stimulated, when compared to the control treatments. From this study's findings, it appears that the field use potential of allicin is limited, due to better inhibition of the less virulent A. gallica, than the more aggressive A. mellea.  相似文献   

17.
Symptoms and signs associated with root rot caused by Heterobasidion annosum or Armillaria ostoyae in mountain pines (Pinus mugo ssp. uncinata) were investigated in the Swiss Alps. A sample of dying or recently dead mountain pine trees (≥12 cm d.b.h.) and saplings (<1.3 m height) was assessed for root pathogen infection by taking root samples followed by isolations in the laboratory. From a subsample, an additional core was taken from the butt of each tree and evaluated in the same fashion. A total of 157 dying or recently dead mountain pine trees and 184 saplings with roots infected by either of the two pathogens or which lacked infection were analyzed using logistic regression models. The main objectives were to determine the most prominent symptoms induced by the fungi (resinosis), signs of the fungi (mycelia, fruiting bodies and rhizomorphs), and tree characteristics (d.b.h./height and evidence of wounds) that would allow an easy and reliable determination of H. annosum and/or A. ostoyae infection of mountain pines in the field. Heterobasidion annosum caused both root and butt rot on mountain pine, whereas A. ostoyae was mostly restricted to the root systems of the trees sampled. The most discriminating sign for the presence of A. ostoyae infection was the presence of characteristic mycelial fans, and for H. annosum root rot the presence of H. annosum mycelia (sheets of paper‐thin mycelium and mycelial pustules). In addition, resinosis was a powerful predictor for A. ostoyae in trees. Symptoms and signs indicating A. ostoyae or H. annosum infections were more reliable for saplings than for mature trees. Armillaria rhizomorphs were not useful in detecting A. ostoyae infection and, if present, were often formed by saprophytic Armillaria species. Heterobasidion annosum fruiting bodies were rarely observed and poorly reflected the widespread occurrence of this pathogen in the mountain pine forests.  相似文献   

18.
Penetration of root bark tissues of Picea sitchensis by Armillaria ostoyae, Armillaria mellea and Heterobasidion annosum was examined in the absence of wounds, in superficial wounds (rhytidome tissues removed to expose the secondary phloem) and in wounds to the depth of the vascular cambium (deep wounding). Both species of Armillaria penetrated bark without prior wounding, but neither species formed rhizomorphs in this treatment. Armillaria ostoyae penetrated to 39 cell layers in depth by 48 days after inoculation of unwounded bark, whereas A. mellea penetrated 25 cell layers in the same time. Armillaria mellea penetrated superficial wounds significantly more rapidly than did A. ostoyae. Both species produced rhizomorphs within wounded host tissues. Inoculation of deep wounds with Armillaria resulted in a greater depth of bark necrosis with A. mellea than with A. ostoyae. In the absence of wounding, H. annosum failed to penetrate root bark tissues, but in both superficial and deep wounds hyphae penetrated beyond the ligno–suberized boundary zone (LSZ) by 12 days after inoculation. Where no inoculations were made, superficial or deep wounding led within 25 days to the restoration of a structurally continuous LSZ, and by day 48 the wound periderm (WP) was fully differentiated. In inoculated wounds, however, formation of the LSZ and WP was delayed or inhibited in most trees, particularly following inoculation with A. ostoyae or A. mellea. Suberization in the LSZ and WP remained diffuse and discontinuous 48 days after inoculation. Moreover, the presence of WP did not prevent further penetration of the tissues by the pathogens. Variations between trees in the depth of pathogen penetration were noted, possibly indicating differing susceptibilities of individual host genotypes. The possible host factors involved in resistance to penetration of root bark tissues by Armillaria and Heterobasidion are discussed.  相似文献   

19.
Contour hedgerows of multipurpose tree species in the sloping tea lands of Sri Lanka are expected to reduce soil erosion and also add significant amounts of plant nutrients to the soil via periodic prunings. The objective of this experiment was to characterize the biomass decomposition pattern and quantify the amount of nutrients added through prunings of six tree species (Calliandra calothyrsus, Senna spectabilis, Euphatorium innulifolium, Flemingia congesta, Gliricidia sepium and Tithonia diversifolia) currently being used in hedgerows associated with tea. Withered leaf and stem prunings (50 g) were enclosed in 2-mm litter bags, placed at 5-cm depth and retrieved after one, three, six, nine and 12 weeks. Loss of initial dry weight, N, P and K was measured. Single exponential decay function adequately described both dry weight and nutrient loss. Tree species differed significantly in their rate of breakdown with decomposition constants (k) varying from 0.0299 to 0.2006 week−1 for leaves and from 0.0225 to 0.0633 week−1 for stems. Gliricidia showed the highest k for leaves with the rest in the following descending order: Senna > TithoniaEuphatorium > Calliandra > Flemingia. A similar pattern was observed for loss of all nutrients with Calliandra and Flemingia always having lower k values than the rest. Although N immobilization was not observed, immobilization of P and K was observed during the first week of incubation in some species, particularly in stem prunings. Annual biomass of prunings differed significantly between tree species in the following descending order: Calliandra > Senna > Flemingia > Tithonia > Gliricidia > Euphatorium. Calliandra added the greatest amount of nutrients annually to the soil with Euphatorium adding the least. Calliandra prunings provided the annual total K requirement and 49% of the N requirement of mature tea. However, none of the species provided more than 5% of the P requirement. It is concluded that among the tree species tested, Calliandra and Flemingia are the most suitable for contour hedgerows in tea plantations of this agroclimatic region because of their higher soil nutrient enrichment capacity and slower decomposition rates which would minimize leaching losses. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

20.
Occurrence of species and clones of Armillaria in spruce stands, mixed stands and hardwood stands in close neighbourhood . From Armillarid rhizomorphs (collected around trees) and from spruce butt rots, isolates of the diploids were made. In pairings between the diploid isolates and haploid testers from the five (European) biological species (BULLER phenomenon) the mating reactions often were not clear enough to identify the diploids. So carpophores were raised from the isolates and single spore cultures were obtained. In pairings with the haploid testers Armillaria borealis, A. bulbosa and A. bulbosa were identified. Usually more than one Armillaria species and from each species more than one clone occurred in each stand.  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