Journal articles on the topic 'Armillaria'

To see the other types of publications on this topic, follow the link: Armillaria.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Armillaria.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Kile, G. A., and R. Watling. "Armillaria bulbosa." Transactions of the British Mycological Society 84, no. 1 (January 1985): 173. http://dx.doi.org/10.1016/s0007-1536(85)80237-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Frontz, T. M., D. D. Davis, B. A. Bunyard, and D. J. Royse. "Identification of Armillaria species isolated from bigtooth aspen based on rDNA RFLP analysis." Canadian Journal of Forest Research 28, no. 1 (January 1, 1998): 141–49. http://dx.doi.org/10.1139/x97-197.

Full text
Abstract:
Restriction fragment length polymorphism analysis (RFLP) of the intergenic region (IGR-1) between the 3 ' end of the 26S ribosomal RNA gene and the 5 ' end of the 5S rRNA gene was used to identify 39 isolates of Armillaria species collected from live or recently dead bigtooth aspen (Populus grandidentata Michx.) trees and sucker sprouts in the Tioga State Forest, Pennsylvania. The unknown isolates were identified by comparing their restriction fragment patterns with 18 isolates of known Armillaria species common to the northeastern United States. Twenty of the unknown isolates (50%) were identified as either Armillaria gallica or Armillaria calvescens. Eighteen (46%) of the isolates were identified as Armillaria ostoyae. One isolate of Armillaria sinapina was obtained from a recently dead aspen tree. One isolate of Armillaria mellea, considered to be the most divergent of the Armillaria species, was obtained from basidiomes fruiting on a recently dead aspen tree near Berwick, Pennsylvania. In some instances, amplification of DNA was possible by adding mycelial scrapes directly to the polymerase chain reaction (PCR) mix, thus precluding the need for DNA extraction. Advancements in RFLP analysis may offer a method able to provide rapid and precise identification of most North American and European Armillaria isolates.
APA, Harvard, Vancouver, ISO, and other styles
3

McLaughlin, J. A. "Distribution, hosts, and site relationships of Armillaria spp. in central and southern Ontario." Canadian Journal of Forest Research 31, no. 9 (September 1, 2001): 1481–90. http://dx.doi.org/10.1139/x01-084.

Full text
Abstract:
This study investigated the species, geographic distribution, host range, site relationships, and impacts of Armillaria in central and southern Ontario. Rhizomorphs and infected wood samples were collected at 110 of 111 sites. Six species were identified by polymerase chain reaction or diploid–haploid pairings. Armillaria gallica Marxmuller & Romagn. was most commonly isolated and had the broadest host range. It was seldom isolated from conifers but often from oaks. It was the species most often found on moist sites and showed strong preference for calcareous soils. Armillaria calvescens Bérubé & Dessureault was rarely isolated from conifers but often from maples, where it commonly caused butt rot. It was found most often on coarse loamy or fine, well-drained, fresh sites. Armillaria ostoyae (Romagn.) Herink. had the second broadest host range. It was seldom found on sugar maple (Acer saccharum Marsh.) but dominated on conifers, especially on dry–fresh, rapidly drained sandy to coarse loamy sites. It was not found on sites with finer soils. Armillaria sinapina Bérubé & Dessureault and Armillaria gemina Bérubé & Dessureault were found in more northerly parts of the study area on noncalcareous sites. Armillaria sinapina often caused butt rot and was often found on poorly drained sites. Armillaria gemina was found only on hardwoods. Armillaria mellea (Vahl:Fr.) Kummer s.st. was found on dead hardwoods at four locations.
APA, Harvard, Vancouver, ISO, and other styles
4

Chen, Liqiong, Bettina Bóka, Orsolya Kedves, Viktor Dávid Nagy, Attila Szűcs, Simang Champramary, Róbert Roszik, et al. "Towards the Biological Control of Devastating Forest Pathogens from the Genus Armillaria." Forests 10, no. 11 (November 13, 2019): 1013. http://dx.doi.org/10.3390/f10111013.

Full text
Abstract:
Research Highlights: A large scale effort to screen, characterize, and select Trichoderma strains with the potential to antagonize Armillaria species revealed promising candidates for field applications. Background and Objectives: Armillaria species are among the economically most relevant soilborne tree pathogens causing devastating root diseases worldwide. Biocontrol agents are environment-friendly alternatives to chemicals in restraining the spread of Armillaria in forest soils. Trichoderma species may efficiently employ diverse antagonistic mechanisms against fungal plant pathogens. The aim of this paper is to isolate indigenous Trichoderma strains from healthy and Armillaria-damaged forests, characterize them, screen their biocontrol properties, and test selected strains under field conditions. Materials and Methods: Armillaria and Trichoderma isolates were collected from soil samples of a damaged Hungarian oak and healthy Austrian spruce forests and identified to the species level. In vitro antagonism experiments were performed to determine the potential of the Trichoderma isolates to control Armillaria species. Selected biocontrol candidates were screened for extracellular enzyme production and plant growth-promoting traits. A field experiment was carried out by applying two selected Trichoderma strains on two-year-old European Turkey oak seedlings planted in a forest area heavily overtaken by the rhizomorphs of numerous Armillaria colonies. Results: Although A. cepistipes and A. ostoyae were found in the Austrian spruce forests, A. mellea and A. gallica clones dominated the Hungarian oak stand. A total of 64 Trichoderma isolates belonging to 14 species were recovered. Several Trichoderma strains exhibited in vitro antagonistic abilities towards Armillaria species and produced siderophores and indole-3-acetic acid. Oak seedlings treated with T. virens and T. atrobrunneum displayed better survival under harsh soil conditions than the untreated controls. Conclusions: Selected native Trichoderma strains, associated with Armillaria rhizomorphs, which may also have plant growth promoting properties, are potential antagonists of Armillaria spp., and such abilities can be exploited in the biological control of Armillaria root rot.
APA, Harvard, Vancouver, ISO, and other styles
5

Rizzo, D. M., E. C. Whiting, and R. B. Elkins. "Spatial Distribution of Armillaria mellea in Pear Orchards." Plant Disease 82, no. 11 (November 1998): 1226–31. http://dx.doi.org/10.1094/pdis.1998.82.11.1226.

Full text
Abstract:
Pears have traditionally been considered to be highly resistant to Armillaria root disease (causal agent: Armillaria mellea). In recent years, however, the incidence of Armillaria root disease in pears has increased in California. To determine the spatial distribution of Armillaria root disease in the field, a total of 156 isolates of Armillaria were collected from dead and dying pear trees located within two orchards in Lake County. All isolates from these two orchards, as well as from an additional 10 pear orchards, were identified as Armillaria mellea sensu stricto. Based on pairings among 102 Armillaria isolates, four somatic incompatibility groups (SIGs) were identified at orchard 1. Three of the four SIGs at this site were over 100 m in length; the largest SIG was at least 200 m in length. Pairings among 54 isolates identified five SIGs at orchard 2. The SIGs at orchard 2 were generally smaller than those detected at orchard 1 and ranged from 20 to 60 m in length. The size of the SIGs points toward long-term establishment of the fungus on the two sites, most likely predating the establishment of the pear orchards. Extensive root excavations of 19 trees indicated that the primary means of secondary spread of Armillaria was via rhizomorphs, as opposed to root-to-root contact.
APA, Harvard, Vancouver, ISO, and other styles
6

Kromroy, K. W., R. A. Blanchette, and D. F. Grigal. "Armillaria species on small woody plants, small woody debris, and root fragments in red pine stands." Canadian Journal of Forest Research 35, no. 6 (June 1, 2005): 1487–95. http://dx.doi.org/10.1139/x05-067.

Full text
Abstract:
The incidence of Armillaria on small woody plants, small woody debris, and root fragments was estimated in red pine (Pinus resinosa Ait.) stands in northeastern Minnesota. Soil core samples 10 cm in diameter, and extending to a depth of either 16 or 25 cm, were collected from 13 stands belonging to three age-classes. Half of the youngest stands had been treated using herbicide. Mycelial fans or rhizomorphs of Armillaria were observed on 13% of the small woody plants and isolated from 8% of them. Including small woody debris and root fragments, 38% of 0–16 cm deep samples had Armillaria. Armillaria was observed on 3% and isolated from 1% of individual substrate units from 0 to 25 cm deep samples. Within a single stand, 0%–67% of the samples and 0%–9% of the individual units had evidence of Armillaria. All but one isolate were Armillaria ostoyae (Romagn.) Herink. Herbicide-treated and untreated red pine stands had similar Armillaria incidence, and there was a trend of incidence inversely related to stand age-class. Large numbers of small woody plants, woody debris, and root fragments were found in red pine stands; varying percentages of these substrates were contributing to the survival of Armillaria and could also be serving as sources of root disease inoculum.
APA, Harvard, Vancouver, ISO, and other styles
7

Warwell, Marcus, Geral McDonald, John Hanna, Mee-Sook Kim, Bradley Lalande, Jane Stewart, Andrew Hudak, and Ned Klopfenstein. "Armillaria altimontana Is Associated with Healthy Western White Pine (Pinus monticola): Potential in Situ Biological Control of the Armillaria Root Disease Pathogen, A. solidipes." Forests 10, no. 4 (March 28, 2019): 294. http://dx.doi.org/10.3390/f10040294.

Full text
Abstract:
Research Highlights: Two genets of Armillaria altimontana Brazee, B. Ortiz, Banik, and D.L. Lindner and five genets of Armillaria solidipes Peck (as A. ostoyae [Romagnesi] Herink) were identified and spatially mapped within a 16-year-old western white pine (Pinus monticola Doug.) plantation, which demonstrated distinct spatial distribution and interspecific associations. Background and Objectives: A. solidipes and A. altimontana frequently co-occur within inland western regions of the contiguous USA. While A. solidipes is well-known as a virulent primary pathogen that causes root disease on diverse conifers, little has been documented on the impact of A. altimontana or its interaction with A. solidipes on growth, survival, and the Armillaria root disease of conifers. Materials and Methods: In 1971, a provenance planting of P. monticola spanning 0.8 ha was established at the Priest River Experimental Forest in northern Idaho, USA. In 1987, 2076 living or recently dead trees were measured and surveyed for Armillaria spp. to describe the demography and to assess the potential influences of Armillaria spp. on growth, survival, and the Armillaria root disease among the study trees. Results: Among the study trees, 54.9% were associated with Armillaria spp. The genets of A. altimontana and A. solidipes comprised 82.7% and 17.3% of the sampled isolates (n = 1221) from the study plot, respectively. The mapped distributions showed a wide, often noncontiguous, spatial span of individual Armillaria genets. Furthermore, A. solidipes was found to be uncommon in areas dominated by A. altimontana. The trees colonized by A. solidipes were associated with a lower tree growth/survival and a substantially higher incidence of root disease than trees colonized only by A. altimontana or trees with no colonization by Armillaria spp. Conclusions: The results demonstrate that A. altimontana was not harmful to P. monticola within the northern Idaho planting. In addition, the on-site, species-distribution patterns suggest that A. altimontana acts as a long-term, in situ biological control of A. solidipes. The interactions between these two Armillaria species appear critical to understanding the Armillaria root disease in this region.
APA, Harvard, Vancouver, ISO, and other styles
8

Bérubé, J. A., and M. Dessureault. "Morphological characterization of Armillaria ostoyae and Armillaria sinapina sp.nov." Canadian Journal of Botany 66, no. 10 (October 1, 1988): 2027–34. http://dx.doi.org/10.1139/b88-277.

Full text
Abstract:
In Quebec, the root rot fungus Armillaria mellea (Vahl: Fr.) Kummer in the broad sense was found to be composed of three intersterile groups or biological species by using mating tests with standard voucher strains. Monosporous cultures of our specimens were compatible with strains of groups I, V, and VI. Groups I and V corresponding to A. ostoyae (Romagn.) Herink and A. sinapina sp.nov., respectively, are described and their occurrence and ecology documented. Morphological characteristics of fruiting bodies and of vegetative isolates can be used to differentiate A. ostoyae, A. sinapina, and A. mellea s.str. Armillaria ostoyae and A. sinapina are mild pathogens or saprotrophs on declining trees or stumps, whereas A. mellea s.str. appears to be an aggressive pathogen.
APA, Harvard, Vancouver, ISO, and other styles
9

Devkota, Pratima, and Raymond Hammerschmidt. "The infection process of Armillaria mellea and Armillaria solidipes." Physiological and Molecular Plant Pathology 112 (December 2020): 101543. http://dx.doi.org/10.1016/j.pmpp.2020.101543.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Burrill, E. A., J. J. Worrall, P. M. Wargo, and S. V. Stehman. "Effects of defoliation and cutting in eastern oak forests on Armillaria spp. and a competitor, Megacollybia platyphylla." Canadian Journal of Forest Research 29, no. 3 (March 1, 1999): 347–55. http://dx.doi.org/10.1139/x99-001.

Full text
Abstract:
Gypsy moth (Lymantria dispar L.) and Armillaria root rot interact to cause extensive mortality in eastern oak forests. Defoliation by gypsy moth weakens trees and increases their susceptibility to Armillaria root rot. Partial cutting prior to defoliation has been proposed as a management technique because it may increase tree vigor and the ability to withstand defoliation stress. However, cutting could also increase inoculum potential of Armillaria by providing a resource, the residual stumps. Megacollybia platyphylla (Pers.:Fr.) Kotl. & Pouz. is a native, cord-forming, saprobic fungus that may compete with Armillaria for resources such as stumps, snags and debris. A factorial treatment design with three levels of cutting and three levels of defoliation was used to examine the effects of cutting and defoliation on the two fungi. Among uncut stands, defoliated stands had significantly greater colonization of resource units by Armillaria than nondefoliated stands. However, stands that were cut prior to defoliation had significantly less Armillaria colonization and significantly more M. platyphylla colonization than those that were not cut. Armillaria colonized snags better than stumps and colonized least well in debris, where M. platyphylla showed its best colonizing performance. The data suggest that cutting mitigates the effects of defoliation on colonization by Armillaria and are consistent with the hypothesis that M. platyphylla plays a role in such mitigation.
APA, Harvard, Vancouver, ISO, and other styles
11

Wahlström, Kjell, Jan-Olof Karlsson, Ottmar Holdenrieder, and Jan Stenlid. "Pectinolytic activity and isozymes in European Armillaria species." Canadian Journal of Botany 69, no. 12 (December 1, 1991): 2732–39. http://dx.doi.org/10.1139/b91-343.

Full text
Abstract:
Pectinolytic activities in Armillaria ostoyae, Armillaria mellea, Armillaria gallica, Armillaria borealis, and Armillaria cepistipes were assessed by measuring the ability of fungal strains to reduce the viscosity of their pectic growth media. Isozyme patterns of pectin esterases and polygalacturonases were determined directly from the culture filtrate. A total of 94 strains, representing isolations from various parts of Europe, were analyzed for their isozyme patterns. Armillaria mellea and A. borealis caused a 50% reduction in viscosity within 7 and 9 days, respectively. Growth medium from the other species were slower to reach the 50% level, i.e., means were 13 days for A. ostoyae and 17 days for A. cepistipes and A. gallica. All species produced more isozymes on spruce wood than on citrus pectin medium, and pectic isozyme patterns differed between media. The pectic isozyme pattern for A. mellea differed distinctly from those of the other four species by having two bands of polygalacturonase not found in the others. The pectic isozyme patterns of the other four species were separated using multivariate analysis. The value of such analyses for use in distinguishing between European Armillaria species is discussed, as is the relation between enzyme activity and fungal pathogenicity. Key words: root rot, diagnostic tests, Agaricales, polygalacturonase, pectin esterase.
APA, Harvard, Vancouver, ISO, and other styles
12

McLaughlin, J. A., and T. Hsiang. "Identification protocol for sixArmillariaspecies from northeastern North America." Canadian Journal of Forest Research 40, no. 3 (March 2010): 536–48. http://dx.doi.org/10.1139/x10-015.

Full text
Abstract:
DNA sequences (~3 kb long) extending from the intergenic spacer 1 (IGS1) region to the 18S gene were obtained for isolates of Armillaria ostoyae , Armillaria calvescens , Armillaria gallica , and Armillaria sinapina . Additional investigation of 16 A. ostoyae, 11 Armillaria gemina , 21 A. calvescens, 18 A. gallica, and 15 A. sinapina isolates produced 117 sequences spanning the 3′ end of the IGS1 through the 5S gene and into the 5′ end of the IGS2 region. Additional sequences spanning the 3′ IGS2 to 5′ 18S gene region were obtained for two A. ostoyae, three A. gemina, two A. calvescens, two A. gallica, and three A. sinapina isolates. This is the first report of complete IGS2 sequences from Armillaria spp. A species identification protocol involving species-specific primers and restriction fragment length polymorphism analysis was devised based on species-specific polymorphisms. The protocol successfully identified all 16 A. ostoyae, 11 A. gemina, three of three Armillaria mellea , 18 A. gallica, 14 of 15 A. sinapina (11/12 diploid and 3/3 haploid), and 14 of 21 A. calvescens (13/15 diploid and 1/6 haploid) isolates included in this study. To the best of our knowledge, this success rate has not been matched by other methods.
APA, Harvard, Vancouver, ISO, and other styles
13

Yang, Yan Hui, Guo Qiang Zheng, Juan Tang, Yue Meng Wang, Chuan Wang Zhu, Hai Yu Ji, Xiao Ming Xu, and An Jun Liu. "Effect of Armillaria mellea on Blood Lipid Levels and Antioxidant Enzymes Activity in High Fat Mice." Advanced Materials Research 884-885 (January 2014): 423–28. http://dx.doi.org/10.4028/www.scientific.net/amr.884-885.423.

Full text
Abstract:
The effect of Armillaria mellea on blood lipid levels and oxidative stress in mice fed on high-fat diet was investigated. Animals were allocated to the Armillaria mellea polysaccharides-treatment groups (I, II) and Armillaria mellea oligosaccharides-treatment groups (I, II). All mice were fed with high-fat diet for 40 days but control mice with basic diet. TC, TG, HDL-c, LDL-c were measured by enzymatic and colorimetric methods. The same, MDA,SOD, GSH-PX were measured. Results showed that administration of Armillaria mellea polysaccharides and oligosaccharides significantly increased antioxidant enzymes GSH-Px activities and decreased TC, TG, LDL-c, MDA level in mice (P < 0.05) compared with model group. In conclusion Armillaria mellea polysaccharides and oligosaccharides were able to protect mices antioxidative and improve abnormal blood lipid levels.
APA, Harvard, Vancouver, ISO, and other styles
14

Keca, Nenad. "The test of eight tree species resistance to the attack of Armillaria mellea and A. ostoyae by artificial infection." Bulletin of the Faculty of Forestry, no. 102 (2010): 41–56. http://dx.doi.org/10.2298/gsf1002041k.

Full text
Abstract:
In the forest ecosystems in Serbia five Armillaria species are present. Understanding differences in the pathogenicity of Armillaria species to the tree species is of a great importance for the foresters. The aim of study was to test susceptibility of eight forest tree species to attack of Armillaria mellea and A. ostoyae. The sticks of Hazel previously infected with mycelium of two Armillaria were placed next to the root collar of two years old seediling. In the period of eighteen months health status of tested seedlings was observed. Differences in susceptibility among tested tree species was observed, while there was no difference in the pathogenicity between Armillaria mellea and A. ostoyae. The most susceptible species were Serbian Spruce, Common Fir, Scots and Austrian Pine, following by Spruce and Douglas Fir, while more resistant were Pedunculate and Sessile oak.
APA, Harvard, Vancouver, ISO, and other styles
15

Prodorutti, D., T. Vanblaere, D. Gobbin, A. Pellegrini, C. Gessler, and I. Pertot. "Genetic Diversity of Armillaria spp. Infecting Highbush Blueberry in Northern Italy (Trentino Region)." Phytopathology® 99, no. 6 (June 2009): 651–58. http://dx.doi.org/10.1094/phyto-99-6-0651.

Full text
Abstract:
Armillaria spp. are the causal agents of root rots of several woody plants, including highbush blueberry. Since 2003, highbush blueberry plants infected by Armillaria spp. have been found in Valsugana Valley, Trentino region, northern Italy. Our aim was to identify the Armillaria spp. involved in these infections, as well as possible sources of inoculum in blueberry fields. Samples of Armillaria spp. were collected from diseased blueberry plants in 13 infected blueberry fields, from bark spread along the blueberry rows, from infected trees in the vicinity of the fields, and from four forest locations. The identification of Armillaria spp. was accomplished using a species-specific multiplex polymerase chain reaction method and by sequencing the rDNA at a specific locus. The differentiation between genotypes was performed by using simple-sequence repeat analysis. Armillaria mellea and A. gallica were the most frequently observed species infecting blueberry in the Valsugana Valley. Three to eight Armillaria genotypes were identified in each blueberry field. No individual genotypes were found in more than one blueberry field. Two-thirds of the genotypes found colonizing trees in the immediate vicinity of infected fields and two-thirds of the genotypes found colonizing the bark spread in blueberry rows were also isolated from blueberry plants in the field, indicating that bark used as mulch and infected trees surrounding the fields may be important sources of inoculum.
APA, Harvard, Vancouver, ISO, and other styles
16

L Smith-White, J., and B. A Summerell. "Armillaria root rot." Microbiology Australia 24, no. 3 (2003): 31. http://dx.doi.org/10.1071/ma03331.

Full text
Abstract:
Armillaria luteobubalina is a fungal phytopathogen endemic to Australia. First described by Podger et al, this species affects a wide range of plants in horticultural and native environments of temperate regions within Australia, colonising root and trunk tissue. This colonisation causes tissue necrosis and ultimately death of the host, giving it the disease name of Armillaria root rot. This disease has brought about considerable economic loss to horticultural, forestry and amenity plantings. To date, control options are limited, with removal of the infected material as the only proven successful management procedure.
APA, Harvard, Vancouver, ISO, and other styles
17

Loreto, Francesco, Harold H. Burdsall, and Alfio Tirro'. "Armillaria Infection and Water Stress Influence Gas-exchange Properties of Mediterranean Trees." HortScience 28, no. 3 (March 1993): 222–24. http://dx.doi.org/10.21273/hortsci.28.3.222.

Full text
Abstract:
The effect of inoculating seedlings of Mediterranean cultivated trees grown under greenhouse conditions with North American isolates of Armillaria mellea (Vahl: Fr) Kumm. and A. ostoyae (Romagn.) Herink on net photosynthesis (A), stomatal conductance (gs), and water potential was examined. The effect of water stress was determined also on the same plant species independently and in combination with Armillaria infection. Red oak (Quercus rubra L.) was used as a control to indicate Armillaria virulence on North American trees. Carob (Ceratonia siliqua L.) was resistant to infection. Infection was successful in sour orange (Citrus aurantium L.), but A, gs, and water potential were unchanged over the 60-day experiment. In olive (Olea europea L.) and oak, A and gs were reduced following inoculation with A. mellea. A and gs of all species but carob were reduced under water stress. Olive and oak responses to water stress and Armillaria infection were quantitatively similar; however, the two stresses combined did not reduce A and gs further. Red oak was strongly susceptible to A. ostoyae infection, but Mediterranean trees were not infected by the same Armillaria isolate. Our results show that Armillaria infection may reduce A and gs in susceptible species.
APA, Harvard, Vancouver, ISO, and other styles
18

Szewczyk, Wojciech, and Małgorzata Mańka. "From the investigations on Armillaria root rot occurrence in young Scots pine stands in Zielonka Forest District." Acta Agrobotanica 55, no. 1 (2013): 319–24. http://dx.doi.org/10.5586/aa.2002.030.

Full text
Abstract:
Armillaria root rot, one of the most dangerous diseases in our forests, is caused in Poland mainly by <i>Armillaria ostoyae</i>, especially severe in young Scots pine stands, established after broadleaved stands or with participation of broadleaved species. In Forest District Zielonka young stands are severly affected by Armillaria root rot. Only one species, A.ostoyae, was found in the young (8-14 yrs) Scots pine stands, despite the presence of other <i>Armillaria</i> species in the district. The pathogen's frequent occurrence may be due, <i>inter alia</i>, to favouring environmental factors.
APA, Harvard, Vancouver, ISO, and other styles
19

Chapman, Bill, and Bruce Schellenberg. "An exploration of ringbarking to reduce the severity of Armillaria root disease in logged areas in British Columbia." Canadian Journal of Forest Research 45, no. 12 (December 2015): 1803–5. http://dx.doi.org/10.1139/cjfr-2015-0143.

Full text
Abstract:
Ringbarking is a girdling technique that is used prior to timber harvesting to reduce losses to Armillaria root disease in some parts of the world. The technique had not previously been evaluated in British Columbia, Canada. Small plots of primarily Douglas-fir (Pseudotsuga menziesii (Mirb.) Franco) trees that were ringbarked prior to timber harvesting had approximately 50% lower levels of Armillaria root disease caused tree mortality in young trees after 15 years than plots of trees that were not ringbarked. Ringbarking did not reduce Armillaria root disease in this trial as much as has been reported in other research. This could be attributable to the centre of the small plots being within 5–10 m of live and dead Armillaria-infected trees in the surrounding forest. The treatment did reduce the severity of the disease by both statistically and biologically significant amounts and, therefore, warrants further investigation as a possible treatment where timber harvesting is conducted in Armillaria root disease affected stands.
APA, Harvard, Vancouver, ISO, and other styles
20

Keca, Nenad, and Dragan Karadzic. "Role of Armillaria species on tree dying in Turkey oak and Hungarian oak forest in Lipovica." Bulletin of the Faculty of Forestry, no. 94 (2006): 151–58. http://dx.doi.org/10.2298/gsf0694151k.

Full text
Abstract:
The species and population structure of Armillaria species were studied in Turkey oak and Hungarian oak forest. Two species were observed, Armillaria gallica and A. mellea. Armillaria mellea was found on only one tree, and A. gallica was found on seven trees. Four gewets of A. gallica were observed of which two were represented only by one isolate each, while two covered the area of 5 and 9 areas respectively.
APA, Harvard, Vancouver, ISO, and other styles
21

McLaughlin, J. A. "Impact of Armillaria root disease on succession in red pine plantations in southern Ontario." Forestry Chronicle 77, no. 3 (June 1, 2001): 519–24. http://dx.doi.org/10.5558/tfc77519-3.

Full text
Abstract:
Armillaria root disease created openings in southern Ontario red pine plantations that are gradually succeeding to hardwood-dominated mixedwoods through natural regeneration. A study of 13 root disease centres found several tree and shrub species colonizing the openings. Black cherry was the most important hardwood and white pine the most important conifer colonizer. Mortality of black cherry and white pine regeneration was greater than for other species. Long-term survival of conifers in the centres is doubtful, and high mortality of black cherry is expected. Other hardwood species may fare better but with growth and yield losses due to Armillaria infection. Key words: Armillaria root disease, red pine plantations, succession, disturbance ecology, Armillaria ostoyae
APA, Harvard, Vancouver, ISO, and other styles
22

Bérubé, Jean A. "Armillaria species in Newfoundland." Canadian Journal of Forest Research 30, no. 3 (March 1, 2000): 507–12. http://dx.doi.org/10.1139/x99-211.

Full text
Abstract:
Crosses with standard testers were used to identify 96 haploid Armillaria isolates from 34 collections made in Newfoundland. Isoenzyme patterns were also used to identify 36 diploid and 3 haploid Armillaria isolates from 39 other collections. Diagnostic electromorphs at Rf 0.70 for esterases, Rf 0.32 for succinate dehydrogenase, and Rf 0.31 for 6-phosphogluconate dehydrogenase permitted positive identification of Armillaria ostoyae (Romagn.) Herink for all isolates tested by electrophoresis, and the method worked with haploid and diploid isolates. Crosses of the 96 haploid isolates collected on hardwoods and conifers were all positive for A. ostoyae, except for eight isolates. These eight isolates from three collections made on or near an American mountain-ash (Sorbus americana Marsh.) in a St. John's city park, where numerous exotic tree species are present, were found to be Armillaria sinapina Bérubé & Dessureault. These isolated collections in an area of limited natural hardwood presence may be the result of an introduction from the mainland; thus, A. ostoyae appears to be the only Armillaria species found in natural habitats in Newfoundland.
APA, Harvard, Vancouver, ISO, and other styles
23

Camprubi, Amelia, Jimena Solari, Paolo Bonini, Francesc Garcia-Figueres, Fabrizio Colosimo, Veronica Cirino, Luigi Lucini, and Cinta Calvet. "Plant Performance and Metabolomic Profile of Loquat in Response to Mycorrhizal Inoculation, Armillaria mellea and Their Interaction." Agronomy 10, no. 6 (June 24, 2020): 899. http://dx.doi.org/10.3390/agronomy10060899.

Full text
Abstract:
A greenhouse experiment was established with loquat plants to investigate the role of arbuscular mycorrhizal fungi (AMF) in the control of the white root rot fungus Armillaria mellea and to determine the changes produced in the plant metabolome. Plants inoculated with two AMF, Rhizoglomus irregulare and a native AMF isolate from loquat soils, were infected with Armillaria. Although mycorrhization failed to control the Armillaria root infection, the increased growth of infected plants following inoculation with the native mycorrhizal isolate suggests an initial tolerance towards Armillaria. Overall, metabolomics allowed highlighting the molecular basis of the improved plant growth in the presence of Armillaria following AMF colonization. In this regard, a wide and diverse metabolic response was involved in the initial tolerance to the pathogen. The AMF-mediated elicitation altered the hormone balance and modulated the production of reactive oxygen species (mainly via the reduction of chlorophyll intermediates), possibly interfering with the reactive oxygen species (ROS) signaling cascade. A complex modulation of fucose, ADP-glucose and UDP-glucose, as well as the down-accumulation of lipids and fatty acids, were observed in Armillaria-infected plants following AMF colonization. Nonetheless, secondary metabolites directly involved in plant defense, such as DIMBOA and conjugated isoflavone phytoalexins, were also involved in the AMF-mediated plant response to infection.
APA, Harvard, Vancouver, ISO, and other styles
24

Drakulic, Jassy, Caroline Gorton, Ana Perez-Sierra, Gerard Clover, and Liz Beal. "Associations Between Armillaria Species and Host Plants in U.K. Gardens." Plant Disease 101, no. 11 (November 2017): 1903–9. http://dx.doi.org/10.1094/pdis-04-17-0472-re.

Full text
Abstract:
Honey fungus (Armillaria spp.) root rot is the disease most frequently inquired about by U.K. gardeners to the Royal Horticultural Society. Armillaria epidemiology has been studied within forestry and agriculture, but data are lacking within gardens, which have greater host plant diversity than orchards and vineyards and greater disturbance than woodlands. Which Armillaria species are responsible for garden disease, and how the broad range of susceptible ornamentals are differentially affected is not known. To address this, isolates of Armillaria were obtained from dead and dying plants from across the U.K. over a 4-year period (2004 to 2007). Species were identified by PCR-RFLP for IGS, and further verified by species-specific PCR for EF-1 α. Of the seven species known in the U.K., three were identified: A. mellea (83.1%), A. gallica (15.8%), and A. ostoyae (1.1%). Armillaria was isolated from trees, shrubs, and nonwoody plants including bulbs and vegetables, with newly recorded hosts listed herein. A. mellea was associated with infections of multiple hosts, and with all infections of the most common host, Ligustrum. In sites where more than one Armillaria species was found, the combination was of A. mellea and A. gallica, raising questions regarding the interactions of these species in U.K. gardens.
APA, Harvard, Vancouver, ISO, and other styles
25

Hood, I. A., M. O. Kimberley, and J. F. Gardner. "Stock-Type Susceptibility and Delineation of Treatment Areas for a Cryptic Pinus radiata Root Disease." Phytopathology® 96, no. 6 (June 2006): 630–36. http://dx.doi.org/10.1094/phyto-96-0630.

Full text
Abstract:
Planting material with superior resistance to Armillaria root disease was identified in a field trial established to investigate variation in Armillaria infection among different Pinus radiata nursery stock types. At stand age 6.4 years, total infection incidence, mortality, and degree of root collar girdling by Armillaria spp. were all significantly lower among trees derived from both rooted stool bed cuttings (physiological age 1 to 3 years) and rooted field cuttings (physiological age 3 to 6 years) than among those grown from seedlings. Cutting types did not differ significantly from one another. No significant differences were found between stock types in stem diameter, but trees from stool bed cuttings were significantly taller than seedling trees. Whether these differences remain detectable later in the rotation, initial results suggest that it may be advantageous to plant robust stock, of either cuttings or seedlings, on Armillaria-infested sites. The incidence of infection in living, green-crowned trees was unevenly distributed across the trial site, and was greater nearer to trees killed by Armillaria spp. than further away (significant within a radius of 10 m). By mapping visible Armillaria-caused mortality prior to thinning, it may be possible to delineate areas with a higher incidence of concealed chronic infection, thus defining infested sites for postharvest treatment.
APA, Harvard, Vancouver, ISO, and other styles
26

Ferguson, B. A., T. A. Dreisbach, C. G. Parks, G. M. Filip, and C. L. Schmitt. "Coarse-scale population structure of pathogenic Armillaria species in a mixed-conifer forest in the Blue Mountains of northeast Oregon." Canadian Journal of Forest Research 33, no. 4 (April 1, 2003): 612–23. http://dx.doi.org/10.1139/x03-065.

Full text
Abstract:
The coarse-scale population structure of pathogenic Armillaria (Fr.) Staude species was determined on approximately 16 100 ha of relatively dry, mixed-conifer forest in the Blue Mountains of northeast Oregon. Sampling of recently dead or live, symptomatic conifers produced 112 isolates of Armillaria from six tree species. Armillaria species identifications done by using a polymerase chain reaction based diagnostic and diploid–diploid pairings produced identical results: 108 of the isolates were Armillaria ostoyae (Romagn.) Herink and four were North American Biological Species X (NABS X). Five genets of A. ostoyae and one of NABS X were identified through the use of somatic incompatibility pairings among the putatively diploid isolates. Armillaria ostoyae genet sizes were approximately 20, 95, 195, 260, and 965 ha; cumulative colonization of the study area was at least 9.5%. The maximum distance between isolates from the 965-ha A. ostoyae genet was approximately 3810 m, and use of three estimates of A. ostoyae spread rate in conifer forests resulted in age estimates for the genet ranging from 1900 to 8650 years. Results are discussed in relation to possible mechanisms that influenced the establishment, expansion, and expression of these genets; the genetic structure and stability of Armillaria; and the implications for disease management in this and similar forests.
APA, Harvard, Vancouver, ISO, and other styles
27

Keca, Nenad. "In vitro interactions between Armillaria species and potential biocontrol fungi." Bulletin of the Faculty of Forestry, no. 100 (2009): 129–42. http://dx.doi.org/10.2298/gsf0900129k.

Full text
Abstract:
Interaction between Armillaria species and seven other fungi were tested in vitro. Tree antagonistic (Trichoderma viride, Trichotecium roseum and Penicillium sp.) and four decaying (Hypholoma fasciculare? Hypholoma capnoides, Phlebiopsis gigantea, and Pleurotus ostreatus) fungi were chosen for this study. The best results were noted for Trichoderma viride, because fungus was able to kill both mycelia and rhizomorphs of Armillaria species, while Hypholoma spp. inhibited both growth of Armillaria colonies and rhizomorph production.
APA, Harvard, Vancouver, ISO, and other styles
28

Entry, James A., and Kermit Cromack Jr. "Effect of pH, aluminum, and sulfate on a high-elevation Armillaria isolate cultured in vitro." Canadian Journal of Forest Research 17, no. 3 (March 1, 1987): 260–62. http://dx.doi.org/10.1139/x87-043.

Full text
Abstract:
An Armillaria (Ft.) isolate compatible with Armillaria biological species II according to haploid tester challenge, was taken from a dying high-elevation red spruce (Picearubens Sarg.) tree in New York and cultured in unbuffered Melin–Norkrans medium containing 0, 50, 100, or 200 mg/L of Al3+ as Al2(SO4)3 at pH 3, 4, or 5. Total fungal weight and hyphal extension increased significantly as pH rose yet increased only slightly when aluminum was added to the medium. The pH of the medium decreased as fungal weight increased. Weight was linearly correlated with hyphal extension (r2 = 0.87) at all aluminum concentrations and pH values tested. Fungal weight and hyphal extension significantly decreased when Armillaria II was cultured in a duplicate experiment with 0, 395, 790, or 1580 mg/L of [Formula: see text] as MgSO4, indicating that increasing [Formula: see text] concentrations were harmful to fungal growth. Field studies demonstrating predisturbance distribution of Armillaria functioning as a stress pathogen of P. rubens are needed to help determine whether soil conditions created by acid deposition are inhibiting Armillaria growth.
APA, Harvard, Vancouver, ISO, and other styles
29

Brazee, Nicholas J., and Robert L. Wick. "Armillaria species distribution and site relationships in Pinus- and Tsuga-dominated forests in Massachusetts." Canadian Journal of Forest Research 41, no. 7 (July 2011): 1477–90. http://dx.doi.org/10.1139/x11-076.

Full text
Abstract:
The primary objective of this study was to determine the composition of Armillaria species in northeastern North American Pinus - and Tsuga -dominated forests. This was accomplished by sampling 32 plots at eight sites within pitch pine ( Pinus rigida Mill.), eastern white pine ( Pinus strobus L.), eastern white pine – mixed oak, and eastern hemlock ( Tsuga canadensis (L.) Carr.) forests. In total, 320 isolates were collected from 19 host tree species, with 207 of 320 (65%) of all isolations coming from Pinus and Tsuga. Armillaria solidipes Peck was the most abundant species, making up 188 of 320 (59%) of all isolations, which included 39 isolations from hardwoods. Meanwhile, Armillaria mellea (Vahl) P. Kumm. was collected a total of 27 times from eastern white and pitch pine. These two Armillaria species co-occurred at five of the eight sites sampled. Chi-square analyses showed that incidence of Armillaria species were significantly different by forest type. Pitch pine forests had a higher incidence of A. solidipes (p < 0.001), eastern white pine forests had a higher incidence of A. mellea (p = 0.001), and eastern hemlock forests had a higher incidence of Armillaria gallica Marxm. & Romagn. (p = 0.002) compared with expected values. The distribution of A. solidipes varied significantly by soil drainage and soil type, with a higher incidence on excessively drained (p < 0.001) and loamy sand (p < 0.001) soils.
APA, Harvard, Vancouver, ISO, and other styles
30

Whitney, R. D. "Armillaria Root Rot Damage in Softwood Plantations in Ontario." Forestry Chronicle 64, no. 4 (August 1, 1988): 345–51. http://dx.doi.org/10.5558/tfc64345-4.

Full text
Abstract:
Armillaria root rot. caused most likely by Armillaria obscura (Pers) Herink, killed 6-to 21-year-old white spruce, black spruce, jack pine and red pine saplings in each of 49 plantations examined in northern Ontario. Annual mortality in the four species over the last 2 to 6 years averaged 1.4%, 1.5%, 0.5% and 0.2%, respectively. In all but one of 25 white spruce and red pine plantations (43 to 58 years old) in eastern and southern Ontario. Armillaria root rot was associated with mortality. Accumulated mortality in white spruce and red pine (initially recorded in 1978) averaged 7.6% and 11.7%, respectively, as of 1986. Current annual mortality for all plantations ranged from 0% to 16%. Key words: root rot. Armillaria obscura, white spruce, black spruce, jack pine, red pine.
APA, Harvard, Vancouver, ISO, and other styles
31

Patil, Swanand R., and Shubham V. Yadav. "Photographic record of Armillaria mellea a bioluminescent fungi from Lonavala in Western Ghats, India." Journal of Threatened Taxa 14, no. 2 (February 26, 2022): 20692–94. http://dx.doi.org/10.11609/jott.7677.14.2.20692-20694.

Full text
Abstract:
Armillaria mellea is a bioluminescent fungus from the Basidiomycota division. The previous known record of the fungus is from Bhimashankar Wildlife Sanctuary. This paper provides data and photographic evidence of Armillaria mellea in Lonavala.
APA, Harvard, Vancouver, ISO, and other styles
32

Jankovský, L., P. Cudlín, and I. Moravec. "Root decays as a potential predisposition factor of a bark beetle disaster in the Šumava Mts." Journal of Forest Science 49, No. 3 (January 16, 2012): 125–32. http://dx.doi.org/10.17221/4687-jfs.

Full text
Abstract:
Root decay infection and potential relations to Ips typographus L. outbreaks in the &Scaron;umava Mts. (Bohemian Forest) were monitored in 3 permanent sample plots. As an originator of root decays honey fungus predominated, in particular cases Heterobasidion annosum (Fr.) Bref. was also recorded. As for honey fungus species, Armillaria ostoyae (Romagn.) Herink predominated, however, A. cepistipes Velenovsk&yacute; and A. borealis Marxm&uuml;ller et Korhonen were also determined. Other wood-destroying fungi were also recorded, e.g. Stereum sanguinolentum (ALB. &amp; SCHW.: FR.) FR. and Climacocystis borealis (FR.) KOTL. Although Armillaria foci were localized directly in a forest edge after bark beetle disaster, it is not possible to state definite relationships between Ips typographus L. invasion and root system infection by Armillaria. The found out rate of infection is, with respect to an altitude over 1,100 m, extremely high not corresponding to existing knowledge on the behaviour of Armillaria in the region of Central Europe. The extent of Norway spruce infection by Armillaria ostoyae (Romagn.) Herink can give evidence of the chronic stress load of spruce trees in the area. &nbsp; &nbsp;
APA, Harvard, Vancouver, ISO, and other styles
33

Cai, Jinlong, Bilian Chen, Wenchao Li, Peng Xu, Yongguo Di, Huini Xu, and Kunzhi Li. "Transcriptome analysis reveals the regulatory mode by which NAA promotes the growth of Armillaria gallica." PLOS ONE 17, no. 11 (November 21, 2022): e0277701. http://dx.doi.org/10.1371/journal.pone.0277701.

Full text
Abstract:
A symbiotic relationship is observed between Armillaria and the Chinese herbal medicine Gastrodia elata (G. elata). Armillaria is a nutrient source for the growth of G. elata, and its nutrient metabolism efficiency affects the growth and development of G. elata. Auxin has been reported to stimulate Armillaria species, but the molecular mechanism remains unknown. We found that naphthalene acetic acid (NAA) can also promote the growth of A. gallica. Moreover, we identified a total of 2071 differentially expressed genes (DEGs) by analyzing the transcriptome sequencing data of A. gallica at 5 and 10 hour of NAA treatment. Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) analyses showed that these unigenes were significantly enriched in the metabolism pathways of arginine, proline, propanoate, phenylalanine and tryptophan. The expression levels of the general amino acid permease (Gap), ammonium transporter (AMT), glutamate dehydrogenase (GDH), glutamine synthetase (GS), Zn(II) 2Cys6 and C2H2 transcription factor genes were upregulated. Our transcriptome analysis showed that the amino acid and nitrogen metabolism pathways in Armillaria were rapidly induced within hours after NAA treatment. These results provide valuable insights into the molecular mechanisms by which NAA promotes the growth of Armillaria species.
APA, Harvard, Vancouver, ISO, and other styles
34

Logan, D. P., and P. A. Alspach. "Negative association between chorus cicada Amphipsalta zelandica and armillaria root disease in kiwifruit." New Zealand Plant Protection 60 (August 1, 2007): 235–40. http://dx.doi.org/10.30843/nzpp.2007.60.4616.

Full text
Abstract:
Armillaria root disease is a serious fungal disease of kiwifruit that can ultimately end in the death of infected vines One commerciallyavailable treatment for infected vines is to inject compost teas into subsoil with compressed air Compost teas are applied to improve soil and hence plant health and their use is based on the generalisation that healthy plants have fewer pests Accordingly there is a suggestion that healthy kiwifruit vines free of armillaria root disease support fewer cicadas Exuviae of the dominant cicada species in kiwifruit Amphipsalta zelandica (Boisduval) were counted and symptoms of armillaria root disease scored for 300 contiguous vines in a commerciallymanaged block of kiwifruit More A zelandica exuviae were found on healthy vines than on vines with symptoms of armillaria root disease and in this instance the generalisation that healthy plants have fewer pests was not true
APA, Harvard, Vancouver, ISO, and other styles
35

Treštić, Tarik, Midhat Usčuplić, Osman Mujezinović, and Mirzeta Memišević. "ARMILLARIA GLJIVE U CENTRALNOJ BOSNI." Radovi Šumarskog fakulteta Univerziteta u Sarajevu 33, no. 1 (December 1, 2003): 41–46. http://dx.doi.org/10.54652/rsf.2003.v33.i1.213.

Full text
Abstract:
Armillaria je rod patogena i saprofita. Neke od ovih vrsta uzrokuju bolesti truleži korijena i ponekad veliku smrtnost stabala izloženih faktorima stresa, npr. požari, insekti, imela, zagađenje itd. Armillaria spp. su uobičajene u Bosni i Hercegovini, javljaju se kako u prirodnim šumama tako i na plantažama, ali nema saznanja o tome koja vrsta naseljava određena područja i koja je od njih značajan patogen. Prikazani rad čini prvi korak u istraživanju Armillaria u Bosni i Hercegovini kako bi se identifikovale uobičajene vrste i bolje razumjela njihova uloga u našim šumskim ekosistemima.
APA, Harvard, Vancouver, ISO, and other styles
36

Pankuch, J. M., P. V. Blenis, V. J. Lieffers, and K. I. Mallett. "Fungal colonization of aspen roots following mechanical site preparation." Canadian Journal of Forest Research 33, no. 12 (December 1, 2003): 2372–79. http://dx.doi.org/10.1139/x03-172.

Full text
Abstract:
Fungal colonization of aspen (Populus tremuloides Michx.) roots was examined in boreal mixedwood sites that were mechanically site prepared 8–10 years earlier for white spruce (Picea glauca (Moench) Voss) regeneration using disc trenchers or ripper plows. A survey of root wounds determined that Armillaria sinapina Bérubé & Dessureault and Armillaria ostoyae (Romagn.) Herink were both wound pathogens of aspen; however, A. sinapina was more frequently associated with wounds than was A. ostoyae. Armillaria ostoyae was more common on unwounded root tissues. Sixty percent of wounds infected by A. sinapina were not compartmentalized and the likelihood of an A. sinapina infection did not increase with increasing wound size. Pathogenic fungi other than Armillaria were rarely associated with root wounds. Sever wounds were associated with furrows; scrape wounds were located both along and between furrows irrespective of the site-preparation technique (ripper plow vs. disk trencher).
APA, Harvard, Vancouver, ISO, and other styles
37

Marçais, B., and P. M. Wargo. "Impact of liming on the abundance and vigor of Armillaria rhizomorphs in Allegheny hardwoods stands." Canadian Journal of Forest Research 30, no. 12 (December 1, 2000): 1847–57. http://dx.doi.org/10.1139/x00-107.

Full text
Abstract:
Abundance of rhizomorphs of Armillaria was characterized in 1995-1996 in 32 plots located in sugar maple (Acer saccharum Marsh.) stands in the Susquehannock State Forest (Pennsylvania, U.S.A.). All of the plots were thinned, and half of the plots were limed in 1985 when the plots were established. Frequency and abundance of Armillaria rhizomorphs in soil samples, on dead wood food bases (stumps, snags, fallen logs), and on the root collar of living sugar maples were determined in each plot. Rhizomorph vigor was evaluated by measuring their ability to colonize fresh striped maple (Acer pennsylvanicum L.) stem sections in the soil, or potato tubers in the laboratory. Isolates of Armillaria were obtained from rhizomorphs in the soil samples and species were determined by somatic incompatibility tests. Armillaria calvescens Bérubé & Dessureault was the major species present, representing about 66% of the isolates. Armillaria gemina Bérubé & Dessureault and Armillaria mellea (Vahl:Fr.) Kummer were also identified in the plots. Frequency of rhizomorphs in the soil, on food bases, abundance of rhizomorphs on root collars, as well as the proportion of rhizomorphs per plot that regenerated and (or) colonized fresh substrates were all correlated. However, abundance of ectotrophic rhizomorphs on the root collar was only weakly correlated with the other components of rhizomorph abundance and vigor. Frequency of the rhizomorphs as well as their ability to colonize fresh substrates were greater in plots either limed or with a high proportion of the basal area in sugar maple prior to thinning. By contrast, abundance of ectotrophic rhizomorphs on root collars was not affected by these factors.
APA, Harvard, Vancouver, ISO, and other styles
38

Dettman, Jeremy R., and Bart J. van der Kamp. "The population structure of Armillaria ostoyae and Armillaria sinapina in the central interior of British Columbia." Canadian Journal of Botany 79, no. 5 (May 1, 2001): 600–611. http://dx.doi.org/10.1139/b01-033.

Full text
Abstract:
The population structures of Armillaria ostoyae (Romagn.) Herink and Armillaria sinapina Bérubé & Dessureault were investigated at a study site near Williams Lake in the central interior of British Columbia. One hundred and twenty eight fungal isolates were examined from nine infection centers and individual genets were delineated using random amplified polymorphic DNA (RAPD) analysis and somatic incompatibility tests. Six A. ostoyae genets ranging in size from 0.70 to >15 ha were detected. The population structure of A. ostoyae was consistent with a clonal reproductive strategy, and infection centers were occupied by single A. ostoyae genets or ramets thereof. Eighteen relatively small A. sinapina genets were detected, with infection centers being occupied by multiple genets. Armillaria sinapina appears to be more pathogenic to coniferous hosts than previously reported in the region. Armillaria ostoyae appears to initiate new infections of available substrate via airborne basidiospores at a lower frequency than A. sinapina. However once established, A. ostoyae can spread quite aggressively and capture significant amounts of secondary resources, while A. sinapina is unable to do so. The results of somatic incompatibility tests used to differentiate genets corresponded with the results of RAPD analysis, with only one minor discrepancy.Key words: Armillaria, RAPD, population structure, genet, ramet, clone.
APA, Harvard, Vancouver, ISO, and other styles
39

Wright, Mark. "Update on Armillaria ectypa." Field Mycology 8, no. 2 (April 2007): 41–43. http://dx.doi.org/10.1016/s1468-1641(10)60450-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Juan Guo, Wen, and Shun Xing Guo. "Triterpene from Armillaria mellea." Chemistry of Natural Compounds 46, no. 6 (January 2011): 995–96. http://dx.doi.org/10.1007/s10600-011-9809-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Motta, Jerome J., Diane Cope Peabody, and Robert B. Peabody. "Quantitative Differences in Nuclear Dna Content Between Armillaria Mellea and Armillaria Bulbosa." Mycologia 78, no. 6 (November 1986): 963–65. http://dx.doi.org/10.1080/00275514.1986.12025357.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Motta, Jerome J., Diane Cope Peabody, and Robert B. Peabody. "Quantitative Differences in Nuclear DNA Content between Armillaria mellea and Armillaria bulbosa." Mycologia 78, no. 6 (November 1986): 963. http://dx.doi.org/10.2307/3807438.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Lochman, J., O. Šerý, L. Jankovský, and V. Mikes. "Discrimination of Czech Armillaria species based on PCR method and high performance liquid chromatography." Plant Protection Science 38, SI 1 - 6th Conf EFPP 2002 (January 1, 2002): S31—S34. http://dx.doi.org/10.17221/10316-pps.

Full text
Abstract:
The genus Armillaria belongs to basidiomycetes and has been known to induce root rot disease and to cause extensive economic losses to a forest crop. We analysed about 40 isolates of Armillaria collected in Czech Republic by PCR and restriction analysis using gel electrophoresis and ion-exchange HPLC. Restrictase Hinf I was able to discriminate all investigated Armillaria species. The sensitivity and resolution of HPLC method was better than that performed by gel electrophoresis. HPLC was able to detect some heterozygous. The results prove the similarity of the species A. borealis, A. cepistipes, A. gallica, A. ostoyae in difference of A. mellea and A. tabescens.
APA, Harvard, Vancouver, ISO, and other styles
44

Bruhn, J. N., J. D. Mihail, and T. R. Meyer. "Using spatial and temporal patterns of Armillaria root disease to formulate management recommendations for Ontario's black spruce (Piceamariana) seed orchards." Canadian Journal of Forest Research 26, no. 2 (February 1, 1996): 298–305. http://dx.doi.org/10.1139/x26-033.

Full text
Abstract:
Between 1982 and 1989, 22 black spruce (Piceamariana (Mill.) BSP) seed orchards were established on cleared jack pine (Pinusbanksiana Lamb.) forest land in northwest Ontario. These orchards were located on stressful sites for black spruce to hasten seed production. Mortality caused by Armillariaostoyae (Romagn.) Herink was observed in most of these orchards within 3 years of establishment. This study was initiated to quantitatively describe the temporal progress and spatial patterns of Armillaria root disease mortality in five representative orchards, to determine future operational management implications. In the four orchards where epidemics developed, temporal disease progress was nonlinear and was better described by the monomolecular function than by the Gompertz or logistic functions. Monomolecular rates of disease increase were 0.0062–0.0346. Applying these rates, we estimated that cumulative Armillaria root disease mortality will be 9–41% and 25–79%, at 20 and 50 years after planting, respectively. Armillaria root disease mortality was spatially aggregated in all four orchards. Trees adjacent to Armillaria-killed trees had an increased probability of mortality from Armillaria root disease. Successive epidemics may develop in these orchards; their timing and severity will be affected by orchard management practices. Measures of spruce family performance in these orchards are compromised by the aggregated distributions of different A. ostoyae genets and the root disease they cause.
APA, Harvard, Vancouver, ISO, and other styles
45

Coetzee, Martin, Brenda Wingfield, and Michael Wingfield. "Armillaria Root-Rot Pathogens: Species Boundaries and Global Distribution." Pathogens 7, no. 4 (October 24, 2018): 83. http://dx.doi.org/10.3390/pathogens7040083.

Full text
Abstract:
This review considers current knowledge surrounding species boundaries of the Armillaria root-rot pathogens and their distribution. In addition, a phylogenetic tree using translation elongation factor subunit 1-alpha (tef-1α) from isolates across the globe are used to present a global phylogenetic framework for the genus. Defining species boundaries based on DNA sequence-inferred phylogenies has been a central focus of contemporary mycology. The results of such studies have in many cases resolved the biogeographic history of species, mechanisms involved in dispersal, the taxonomy of species and how certain phenotypic characteristics have evolved throughout lineage diversification. Such advances have also occurred in the case of Armillaria spp. that include important causal agents of tree root rots. This commenced with the first phylogeny for Armillaria that was based on IGS-1 (intergenic spacer region one) DNA sequence data, published in 1992. Since then phylogenies were produced using alternative loci, either as single gene phylogenies or based on concatenated data. Collectively these phylogenies revealed species clusters in Armillaria linked to their geographic distributions and importantly species complexes that warrant further research.
APA, Harvard, Vancouver, ISO, and other styles
46

Motta, Jerome J., and Kari Korhonen. "A Note on Armillaria mellea and Armillaria bulbosa from the Middle Atlantic States." Mycologia 78, no. 3 (May 1986): 471. http://dx.doi.org/10.2307/3793052.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Holdenrieder, Von O. "Beobachtungen zum Vorkommen von Armillaria obscura und Armillaria cepistipes an Tanne in Südbayern." European Journal of Forest Pathology 16, no. 5-6 (October 1986): 375–79. http://dx.doi.org/10.1111/j.1439-0329.1986.tb00203.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Prospero, Simone, Esther Jung, Tetyana Tsykun, and Daniel Rigling. "Eight microsatellite markers for Armillaria cepistipes and their transferability to other Armillaria species." European Journal of Plant Pathology 127, no. 2 (March 7, 2010): 165–70. http://dx.doi.org/10.1007/s10658-010-9594-0.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Baumgartner, Kendra, and Amy E. Warnock. "A Soil Inoculant Inhibits Armillaria mellea In Vitro and Improves Productivity of Grapevines with Root Disease." Plant Disease 90, no. 4 (April 2006): 439–44. http://dx.doi.org/10.1094/pd-90-0439.

Full text
Abstract:
A soil inoculant, Vesta (Biologically Integrated Organics, Inc., Sonoma, CA), was tested for its ability to inhibit Armillaria mellea, causal agent of Armillaria root disease of grapevine (Vitis vinifera). Colony diameter of A. mellea was significantly inhibited by undiluted inoculant (P < 0.0001) and by bacterial isolates cultured from the inoculant (Bacillus subtilis, B. lentimorbus, Comamonas testosteroni, Pseudomonas aeruginosa, P. mendocina; P < 0.0001) relative to diameter of the nontreated control. Efficacy of the inoculant for postinfection control of Armillaria root disease of grapevine was examined in an A. mellea-infested vineyard in northern California. Inoculant was applied via drip-irrigation to vine rows in replicate blocks in 2003 and 2004. Yield, growth, mineral nutrition, and juice quality parameters of healthy and symptomatic vines were measured in treated and nontreated vine rows. Significantly decreased petiole P and K concentrations and significantly lower soluble solids content in fruit from symptomatic vines demonstrated that Armillaria root disease negatively affects vine mineral nutritional status and fruit quality, findings that have not been previously reported for an agronomic host of A. mellea. The inoculant significantly increased cluster weights of symptomatic vines (109.63 g/cluster), relative to those of symptomatic-nontreated vines (92.05 g/cluster), to levels comparable to those of healthy vines (122.09 g/cluster). However, the inoculant did not decrease the rate of symptom development or mortality of treated vines from 2002 to 2004. The results of our field experiment suggest that the inoculant may not prevent Armillaria root disease, but can provide therapeutic benefit by improving productivity of infected vines.
APA, Harvard, Vancouver, ISO, and other styles
50

Mallett, K. I. "Host range and geographic distribution of Armillaria root rot pathogens in the Canadian prairie provinces." Canadian Journal of Forest Research 20, no. 12 (December 1, 1990): 1859–63. http://dx.doi.org/10.1139/x90-249.

Full text
Abstract:
A survey to identify Armillaria root rot pathogens, their host range, and geographic distribution was conducted in the Canadian prairie provinces. Collections of basidiocarps and isolates from the wood of gymptomatic or dead trees were made. Armillaria species were identified by interfertility testing and by the L-DOPA method. Three Armillaria species, A. ostoyae (Romagn.) Herink, A. sinapina Bérubé & Dessureault, and A. calvescens Bérubé & Dessureault, were identified. Armillariaostoyae was the most common species in both the subalpine and boreal forests and was found on a wide variety of coniferous and deciduous host species. Armillariasinapina was in both the boreal and subalpine forests but occurred primarily on deciduous host species. Armillariacalvescens was rare and was found only in the boreal forest on both coniferous and deciduous host species.
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography