Academic literature on the topic 'Slash Pine (Pinus Elliotti)'

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

Select a source type:

Consult the lists of relevant articles, books, theses, conference reports, and other scholarly sources on the topic 'Slash Pine (Pinus Elliotti).'

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.

Journal articles on the topic "Slash Pine (Pinus Elliotti)"

1

Highsmith, Maxine T., John Frampton, David O'Malley, James Richmond, and Martesa Webb. "Susceptibility of parent and interspecific F1 hybrid pine trees to tip moth damage in a coastal North Carolina planting." Canadian Journal of Forest Research 31, no. 5 (May 1, 2001): 919–23. http://dx.doi.org/10.1139/x01-022.

Full text
Abstract:
Tip moth damage among families of parent pine species and their interspecific F1 hybrids was quantitatively assessed in a coastal planting in North Carolina. Three slash pine (Pinus elliotti var. elliotti Engelm.), two loblolly pine (Pinus taeda L.), and four interspecific F1 hybrid pine families were used. The F1 hybrids were as susceptible to damage by Nantucket pine tip moth (Rhyacionia frustrana (Comst.)), as was their susceptible loblolly pine parent. Their phenotypes support a dominant or partially dominant mode of inheritance for susceptibility. The phenotype of one slash pine family was not statistically different from the phenotypes of the loblolly and F1 hybrid pines. The high susceptibility of that one slash pine family appeared to be intrinsic, even though slash pine is considered resistant to tip moth damage. Tip moth damage on the two other slash pine families was significantly lower.
APA, Harvard, Vancouver, ISO, and other styles
2

Hunt, Ellis V., and J. David Lenhart. "Fusiform Rust Trends in East Texas." Southern Journal of Applied Forestry 10, no. 4 (November 1, 1986): 215–16. http://dx.doi.org/10.1093/sjaf/10.4.215.

Full text
Abstract:
Abstract Four surveys of pine plantations in East Texas between 1969 and 1984 indicate that fusiform rust (Cronartium quercuum (Berk.) Miyabe ex Shirai f. sp. fusiforme) infection rates are increasing on slash pine (Pinus elliottii Engelm. var. elliottii) and either decreasing or about constant on loblolly pine (Pinus taeda L.). Currently, stem infections occur on about 1 in 2 slash pines and 1 in 14 loblolly pines. South. J. Appl. For. 10:215-216, Nov. 1986.
APA, Harvard, Vancouver, ISO, and other styles
3

Ross, Darrell W., Göran Birgersson, Karl E. Espelie, and C. Wayne Berisford. "Monoterpene emissions and cuticular lipids of loblolly and slash pines: potential bases for oviposition preference of the Nantucket pine tip moth." Canadian Journal of Botany 73, no. 1 (January 1, 1995): 21–25. http://dx.doi.org/10.1139/b95-003.

Full text
Abstract:
Monoterpene emissions from intact 5- to 29-month-old loblolly and slash pine seedlings contained α-pinene, camphene, β-pinene, myrcene, sabinene, β-phellandrene, and limonene. α-Pinene made up > 50% of the volatiles from both species. β-Pinene was significantly more abundant in slash (35.6%) than in loblolly pine (15.3%), while myrcene was significantly more abundant in loblolly (10.9%) than in slash pine (3.4%). Cuticular lipids represented 0.11 and 0.06% of the dry weight biomass of shoots from loblolly and slash pines, respectively. Species differences in cuticular lipid composition were primarily in relative proportions of a group of unidentified compounds that appear to be saturated and unsaturated diols and (or) hydroxy alcohols with chain lengths of about 18 carbons. 10-Nonacosanol made up 16.2 and 14.1% of the total lipids recovered from loblolly and slash pines, respectively. The Nantucket pine tip moth, Rhyacionia frustrana (Comstock), may use these chemical differences to distinguish the susceptible loblolly pines from the resistant slash pines. Key words: Pinus taeda, Pinus elliottii, monoterpene emissions, cuticular lipids, Rhyacionia frustrana.
APA, Harvard, Vancouver, ISO, and other styles
4

Lenhart, J. David, Terry L. Hackett, Charlie J. Laman, Thomas J. Wiswell, and Jock A. Blackard. "Tree Content and Taper Functions for Loblolly and Slash Pine Trees Planted on Non-Old-Fields in East Texas." Southern Journal of Applied Forestry 11, no. 3 (August 1, 1987): 147–51. http://dx.doi.org/10.1093/sjaf/11.3.147.

Full text
Abstract:
Abstract Equations are presented to estimate total or partial stem content in cubic feet and pounds (green or dry) for loblolly pine (Pinus taeda L.) and slash pine (Pinus elliotti Engelm.) trees planted on non-old-fields in East Texas. Equations are included to estimate the content of the completetree (stem and branches). In addition, a set of compatible stem taper functions are described. South. J. Appl. For. 11(3):147-151.
APA, Harvard, Vancouver, ISO, and other styles
5

Hanberry, Brice B. "Transition from Fire-Dependent Open Forests: Alternative Ecosystem States in the Southeastern United States." Diversity 13, no. 9 (August 29, 2021): 411. http://dx.doi.org/10.3390/d13090411.

Full text
Abstract:
Land use and fire exclusion have influenced ecosystems worldwide, resulting in alternative ecosystem states. Here, I provide two examples from the southeastern United States of fire-dependent open pine and pine-oak forest loss and examine dynamics of the replacement forests, given continued long-term declines in foundation longleaf (Pinus palustris) and shortleaf (Pinus echinata) pines and recent increases in commercial loblolly (Pinus taeda) and slash (Pinus elliottii var. elliottii) pines. Shortleaf pine-oak forest historically may have been dominant on about 32 to 38 million ha, a provisional estimate based on historical composition of 75% of all trees, and has decreased to about 2.5 million ha currently; shortleaf pine now is 3% of all trees in the northern province. Longleaf pine forest decreased from about 30 million ha, totaling 75% of all trees, to 1.3 million ha and 3% of all trees in contemporary forests of the southern province. The initial transition from open pine ecosystems to closed forests, primarily comprised of broadleaf species, was countered by conversion to loblolly and slash pine plantations. Loblolly pine now accounts for 37% of all trees. Loss of fire-dependent ecosystems and their foundation tree species affect associated biodiversity, or the species that succeed under fire disturbance.
APA, Harvard, Vancouver, ISO, and other styles
6

Hooker, Jamie M., Brian P. Oswald, Jeremy P. Stovall, Yuhui Weng, Hans M. Williams, and Jason Grogan. "Third Year Survival, Growth, and Water Relations of West Gulf Coastal Plain Pines in East Texas." Forest Science 67, no. 3 (May 10, 2021): 347–55. http://dx.doi.org/10.1093/forsci/fxab005.

Full text
Abstract:
Abstract West Gulf Coastal Plain provenance loblolly (Pinus taeda L.), longleaf (Pinus palustris Mill.), shortleaf (Pinus echinata Mill.), and slash pines (Pinus elliottii Engelm.) were planted in December 2015 on three east Texas sites to compare initial growth and survival. Three years after planting, survival ranged from 26.4% to 76.4%. Damage by Texas leafcutter ants (Atta texana) caused significant mortality on one site, and feral hog (Sus scrofa) herbivory and uprooting greatly affected survival at two other sites. Tree heights were greater in loblolly and slash pine than in shortleaf and longleaf pine, whereas diameters were greater in loblolly than in slash, shortleaf, and longleaf. Height and survival rates were greater in Shelby County and were lowest in Cherokee County. Midday leaf-level water potentials were most negative for shortleaf and loblolly pines and varied across the three sites. Tree heights were significantly but weakly (R = −0.23) correlated to leaf-level water potentials.
APA, Harvard, Vancouver, ISO, and other styles
7

Merkel, Edward P., Carl W. Fatzinger, and Wayne N. Dixon. "Keys for Distinguishing Thrips (Thysanoptera) Commonly Found on Slash Pine in Florida." Journal of Entomological Science 29, no. 1 (January 1, 1994): 92–99. http://dx.doi.org/10.18474/0749-8004-29.1.92.

Full text
Abstract:
The slash pine flower thrips (SPFT), Gnophothrips fuscus (Morgan) commonly damages cone crops of slash pines, Pinus elliottii Engelmann var. elliottii. It occurs in the crowns of seed orchard trees in association with three other species of thrips — Leptothrips pini (Watson), Frankliniella bispinosa (Morgan), and Oxythrips pallidiventris Hood. Pest management strategies are being developed for southern pine seed orchards that rely on the rapid and accurate identification of different species of pests. Two laboratory keys are presented to distinguish the four species of thrips in the laboratory.
APA, Harvard, Vancouver, ISO, and other styles
8

Gonzalez-Benecke, Carlos A., Timothy A. Martin, and Wendell P. Cropper,. "Whole-tree water relations of co-occurring mature Pinus palustris and Pinus elliottii var. elliottii." Canadian Journal of Forest Research 41, no. 3 (March 2011): 509–23. http://dx.doi.org/10.1139/x10-230.

Full text
Abstract:
The natural range of longleaf pine ( Pinus palustris P. Mill.) and slash pine ( Pinus elliottii var. elliottii Engelm.) includes most of the southeastern US Coastal Plain, and there is now considerable interest in using these species for ecological forestry, restoration, and carbon sequestration. It is therefore surprising that little information is currently available concerning differences in their ecological water relations in natural stands. In this study, we compared water use, stomatal conductance at the crown scale (Gcrown), and whole-tree hydraulic conductance of mature pine trees growing in a naturally regenerated mixed stand on a flatwoods site in north-central Florida. We found remarkable similarities between longleaf and slash pine in stored water use, nocturnal transpiration, and whole-tree hydraulic conductance. Mean daily transpiration rate was higher for slash than for longleaf pine, averaging 39 and 26 L·tree–1, respectively. This difference was determined by variations in tree leaf area. Slash pine had 60% more leaf area per unit basal sapwood area than longleaf pine, but the larger plasticity of longleaf pine stomatal regulation partially compensated for leaf area differences: longleaf pine had higher Gcrown on days with high volumetric water content (θv) but this was reduced to similar or even lower values than for slash pine on days with low θv. There was no species difference in the sensitivity of Gcrown to increasing vapor pressure deficit.
APA, Harvard, Vancouver, ISO, and other styles
9

Eberhardt, Thomas L., Joseph Dahlen, and Laurence Schimleck. "Species comparison of the physical properties of loblolly and slash pine wood and bark." Canadian Journal of Forest Research 47, no. 11 (November 2017): 1495–505. http://dx.doi.org/10.1139/cjfr-2017-0091.

Full text
Abstract:
Composition of the southern pine forest is now predominated by two species, loblolly pine (Pinus taeda L.) and slash pine (Pinus elliottii Engelm.), owing to fire suppression activities, natural regeneration on abandoned agricultural lands, and extensive planting. Comparison of the wood and bark physical properties of these pines is of interest in terms of the yields of usable biomass and, for the bark, its ecological functionality on a living tree. Trees from a species comparison study were used to generate wood and bark property data, on a whole-tree basis, and for stem disks collected at breast height. Models were constructed to explain the effect of relative height on wood and bark properties. When comparing the whole-tree data, slash pine wood (0.523 versus 0.498) and bark (0.368 versus 0.311) specific gravity values were higher, both offset by lower moisture contents; slash pine also produced a higher percentage of bark on a dry-mass basis (17% versus 12.5%). Unlike wood properties, bark properties showed significant between-species differences when determined at breast height alone, the exception being moisture content. In terms of yield, harvests of a green tonne of loblolly pine and slash pine would give approximately the same dry mass of wood, but slash pine provides more bark.
APA, Harvard, Vancouver, ISO, and other styles
10

Sluder, Earl R. "Fusiform Rust in Crosses Among Resistant and Susceptible Loblolly and Slash Pines." Southern Journal of Applied Forestry 13, no. 4 (November 1, 1989): 174–77. http://dx.doi.org/10.1093/sjaf/13.4.174.

Full text
Abstract:
Abstract Progenies from a half-diallel cross among six loblolly pines (Pinus taeda L.) and another among six slash pines (P. elliottii Engelm. var. elliottii) were field-tested in central Georgia for fusiform-rust (Cronartium quercuum [Berk.] Miyabe ex Shirai f. sp. fusiforme) resistance. Threeof each set of six parents were rated resistant (R) and three susceptible (S) to the fungus relative to check lots in previous progeny tests. For both species, relative susceptibility of the three types of progenies at age 5 years was R x R < R x S < S x S. The R x R and R x S progeniesin slash pine had considerably less infection than did the same types of progenies in loblolly pines. All S x S progenies were heavily infected. Heritability estimates for percentage infection and average galls per tree indicated that in these progenies, family selection should be based onpercentage infection for slash pine and galls per tree for loblolly pine. South. J. Appl. For. 13(4):174-177.
APA, Harvard, Vancouver, ISO, and other styles

Dissertations / Theses on the topic "Slash Pine (Pinus Elliotti)"

1

Pswarayi, Idah Zviripayi. "Genetic parameters and selection indices for a population of Pinus elliottii Engelm. var. elliottii." Thesis, University of Oxford, 1993. http://ora.ox.ac.uk/objects/uuid:97fc4675-4dae-4b43-a15e-d3e9f52f6948.

Full text
Abstract:
P. elliottii Engelm. var. elliottii is an important exotic plantation species in Zimbabwe, where it is grown for saw-timber and resin production. Three progeny tests, originating from factorial matings between parents selected in plantations, were assessed at five, eight and 15 years. The objectives of the study were to characterise quantitative genetic variation in the population through the estimation of genetic parameters, and to use these parameters in combined indices to select for specified breeding objectives for P. elliottii in Zimbabwe. All traits of interest were under a reasonable degree of additive genetic control, and the magnitudes of nonadditive genetic variances were almost invariably much less than those of additive genetic variances. Narrow-sense heritabilities for growth traits, wood density and resin yield were moderate to high, ranging from 0.3 - 0.42; those for stem straightness and branching traits were lower, ranging between 0.10 and 0.25. Genetic correlations at each of the assessment ages were more variable; of most consequence for this study were the slight negative correlations between wood density and both stem diameter and volume, and the slight positive correlation between density and height. Age-age correlations for growth traits were high, indicating potential for early selection. Age-age correlations for other traits ranged from moderately negative to highly positive. Although statistically significant for many traits, genotype-environment interaction was judged by a number of criteria to be of little practical importance. No one site was the most efficient for selection across the range of traits for establishment at other sites; rather, a set of pooled parameters was estimated for application on sites typical of those on which commercial plantations of P. elliottii are established. Selection indices were constructed for four breeding objectives, representing differing assumptions about the relative importance of saw-timber and resin production. Indices for both direct and indirect selection were compared in terms of genetic gain, efficiency and accuracy, which were influenced by the differential weighting of traits in the breeding objective. The highest gains, efficiency and accuracies were for the breeding objective of saw-timber only; increasing the emphasis on resin production reduced each of these parameters, and also had a more adverse impact on wood density. For a breeding objective corresponding to or emphasizing saw-timber production, selection based on diameter or height at five years was best; selection on the latter has the advantage of maintaining wood density at around its present level. Should resin production also be important, resin yield or a correlated trait must be included in the index. Efficiencies of indirect selection were highest at five years, regardless of the breeding objective or selection criteria considered. The lack of economic information was a considerable hinderance in conducting these analyses. The construction of more complete indices, incorporating information from all siblings represented in the factorial mating design, was also investigated for the breeding objective of saw-timber production. These indices were compared in terms of gain and accuracy, and their effect on population structure in the subsequent generation. Selection based on the most complete index resulted in the greatest gain and accuracy, but also in the greatest reduction in additive genetic variance in the next generation. These results highlight the dilemma facing breeders charged both with achieving gains in the short term and maintaining diversity over the longer term. Breeding strategies which facilitate differential intensities of selection and breeding, and the maintenance of a large effective population size, are seen as the best means to resolve these conflicting demands; some implications for the breeding population of P. elliottii in Zimbabwe are discussed.
APA, Harvard, Vancouver, ISO, and other styles
2

Pagliarini, Maximiliano Kawahata. "Genotype by environment interaction in slash pine and methodologies comparison for radiata pine wood properties /." Ilha Solteira, 2016. http://hdl.handle.net/11449/141895.

Full text
Abstract:
Orientador: Ananda Virginia de Aguiar
Abstract: Exotic forest species have been introduced in Brazil in order to promote improvements in socioeconomic development and help to reduce the pressure caused to native forests. With growing demand for these species, research on genetic improvement has increased to find new, more productive germplasm and preferably in less time. Two species were used in the study: slash pine (Pinus elliottii Engelm. var. elliottii) and radiata pine (Pinus radiata D. Don). The first part of the study had the purpose to identify the stability, adaptability, productivity and genetic parameters, in addition to selection gain and genetic divergence in slash pine open pollinated second generation progenies considering phenotypic trait. Two tests were established, one in Ponta Grossa-PR with 24 progenies and one in Ribeirão Branco-SP with 44 progenies, both in Brazil, to identify the most productive genotypes for commercial planting areas in both sites. There was significant variation (p<0.01) among progenies for growth and form traits. The high coefficients of genetic variation for wood volume (14.31% to 16.24% - Ribeirão Branco-SP and 31.78% to 33.77% - Ponta Grossa-PR) and heritability (0.10 to 0.15 – Ribeirão Branco-SP and 0.36 to 0.48 – Ponta Grossa-PR) have shown low environmental influence on phenotypic variation, which is important for the prediction of genetic gain by selecting and confirming genetic potential in both places, especially Ponta Grossa. The effect of genotype x environment interact... (Complete abstract click electronic access below)
Resumo: Espécies exóticas de Pinus foram introduzidas no Brasil para promoverem o crescimento socioeconômico do país e ajudar na redução da pressão causada pelo uso de florestas nativas Com a crescente demanda por essas espécies, pesquisas em melhoramento genético tem aumentado na busca de novos germoplasma mais produtivos em menor tempo. Duas espécies foram utilizadas no presente trabalho: Pinus elliottii Engelm. var. elliottii e Pinus radiata D. Don. A primeira parte do trabalho teve a finalidade de identificar a estabilidade, a adaptabilidade, a produtividade e os parâmetros genéticos, além do ganho de seleção e diversidade genética em progênies de polinização aberta de segunda geração de P. elliottii var. elliottii considerando os caracteres fenotípicos. Foram estabelecidos dois testes, um em Ponta Grossa-PR com 24 progênies e outro em Ribeirão Branco-SP com 44 progênies visando identificar os genótipos mais produtivos para áreas de plantio comercial em ambos locais. Foi observada variação significativa (p<0,01) entre as progênies para os caracteres de crescimento e alguns caracteres de forma. Os altos coeficientes de variação genética para volume de madeira (14,31% a 16,24% - Ribeirão Branco e 31,78% a 33,77% - Ponta Grossa) e herdabilidade (0,10 a 0,15 – Ribeirão Branco e 0,36 a 0,48 – Ponta Grossa) mostraram baixa influência do ambiente na variação fenotípica, o que é importante para a predição do ganho genético mediante a seleção e confirmam potencial genético em ambos os loc... (Resumo completo, clicar acesso eletrônico abaixo)
Doutor
APA, Harvard, Vancouver, ISO, and other styles
3

Pagliarini, Maximiliano Kawahata [UNESP]. "Genotype by environment interaction in slash pine and methodologies comparison for radiata pine wood properties." Universidade Estadual Paulista (UNESP), 2016. http://hdl.handle.net/11449/141895.

Full text
Abstract:
Submitted by Maximiliano Kawahata Pagliarini null (max.pagliarini@gmail.com) on 2016-07-19T18:22:23Z No. of bitstreams: 1 Tese_Maximiliano_Pagliarini_PPGAgronomia_UNESP.pdf: 2759397 bytes, checksum: 252905ebd212e708c5070c9c9f59d79a (MD5)
Approved for entry into archive by Ana Paula Grisoto (grisotoana@reitoria.unesp.br) on 2016-07-20T18:24:06Z (GMT) No. of bitstreams: 1 pagliarini_mk_dr_ilha.pdf: 2759397 bytes, checksum: 252905ebd212e708c5070c9c9f59d79a (MD5)
Made available in DSpace on 2016-07-20T18:24:06Z (GMT). No. of bitstreams: 1 pagliarini_mk_dr_ilha.pdf: 2759397 bytes, checksum: 252905ebd212e708c5070c9c9f59d79a (MD5) Previous issue date: 2016-05-20
Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES)
Espécies exóticas de Pinus foram introduzidas no Brasil para promoverem o crescimento socioeconômico do país e ajudar na redução da pressão causada pelo uso de florestas nativas Com a crescente demanda por essas espécies, pesquisas em melhoramento genético tem aumentado na busca de novos germoplasma mais produtivos em menor tempo. Duas espécies foram utilizadas no presente trabalho: Pinus elliottii Engelm. var. elliottii e Pinus radiata D. Don. A primeira parte do trabalho teve a finalidade de identificar a estabilidade, a adaptabilidade, a produtividade e os parâmetros genéticos, além do ganho de seleção e diversidade genética em progênies de polinização aberta de segunda geração de P. elliottii var. elliottii considerando os caracteres fenotípicos. Foram estabelecidos dois testes, um em Ponta Grossa-PR com 24 progênies e outro em Ribeirão Branco-SP com 44 progênies visando identificar os genótipos mais produtivos para áreas de plantio comercial em ambos locais. Foi observada variação significativa (p<0,01) entre as progênies para os caracteres de crescimento e alguns caracteres de forma. Os altos coeficientes de variação genética para volume de madeira (14,31% a 16,24% - Ribeirão Branco e 31,78% a 33,77% - Ponta Grossa) e herdabilidade (0,10 a 0,15 – Ribeirão Branco e 0,36 a 0,48 – Ponta Grossa) mostraram baixa influência do ambiente na variação fenotípica, o que é importante para a predição do ganho genético mediante a seleção e confirmam potencial genético em ambos os locais, especialmente Ponta Grossa. O efeito da interação genótipo x ambiente é simples. As progênies plantadas em um local poderão também ser plantadas no outro. Dentre essas as C-197, C-189-1, C-084-2 e C-032-2 são indicadas para plantações tanto na região estudada do estado de São Paulo quanto do Paraná. Apesar de um número maior de progênies em Ribeirão Branco, constatou-se o mesmo número de agrupamentos de progênies pelo método UPGMA e de otimização de Tocher em ambos os testes. Existe diversidade genética entre as progênies de P. elliottii. Para programas de melhoramento, recomenda-se o cruzamento entre progênies de grupos divergentes para aumentar a variação genética, e consequentemente, o ganho genético nas gerações subsequentes, sem esquecer de se levar em consideração a performance do caráter de interesse. O objetivo do trabalho em P. radiata foi relacionar os resultados de características da madeira obtidas a partir de dois métodos Pilodyn e SilviScan visando validar uma metodologia eficiente para fenotipagem de um maior número de amostras. Um teste com 30 progênies de P. radiata foi estabelecido em Flynn na Austrália. As características avaliadas foram densidade da madeira, o ângulo microfibrilar e o módulo de elasticidade. A correlação genética e fenotípica entre os caracteres da madeira obtidas a partir dos dois métodos e a herdabilidade individual no sentido restrito foram estimadas. Os dados de Pilodyn apresentaram alta herdabilidade e alta correlação genética e fenotípica entre densidade de madeira e moderada com ângulo microfibrilar e módulo de elasticidade. Os resultados confirmam que o Pylodyn é um efetivo método indireto e rápido para avaliação de parâmetros genéticos para caracteres de qualidade madeira em P. radiata.
Exotic forest species have been introduced in Brazil in order to promote improvements in socioeconomic development and help to reduce the pressure caused to native forests. With growing demand for these species, research on genetic improvement has increased to find new, more productive germplasm and preferably in less time. Two species were used in the study: slash pine (Pinus elliottii Engelm. var. elliottii) and radiata pine (Pinus radiata D. Don). The first part of the study had the purpose to identify the stability, adaptability, productivity and genetic parameters, in addition to selection gain and genetic divergence in slash pine open pollinated second generation progenies considering phenotypic trait. Two tests were established, one in Ponta Grossa-PR with 24 progenies and one in Ribeirão Branco-SP with 44 progenies, both in Brazil, to identify the most productive genotypes for commercial planting areas in both sites. There was significant variation (p<0.01) among progenies for growth and form traits. The high coefficients of genetic variation for wood volume (14.31% to 16.24% - Ribeirão Branco-SP and 31.78% to 33.77% - Ponta Grossa-PR) and heritability (0.10 to 0.15 – Ribeirão Branco-SP and 0.36 to 0.48 – Ponta Grossa-PR) have shown low environmental influence on phenotypic variation, which is important for the prediction of genetic gain by selecting and confirming genetic potential in both places, especially Ponta Grossa. The effect of genotype x environment interaction is simple. Progenies planted in one site can also be planted in the other. Among these C-197, C-189-1, C-084-2 and C-032-2 progenies are suitable for plantations in both studied region of São Paulo and Paraná. Although larger number of progenies in Ribeirão Branco, it was found the same number of clusters through UPGMA and Tocher methods in both tests. There is genetic diversity among slash pine progenies. For breeding programs, it is recommended to cross progenies between different groups to increase genetic variation, and consequently the genetic gain in subsequent generations, not forgetting to take into account the performance of interest trait. The objective of the study in Radiata pine was relate wood quality traits obtained from two methods Pilodyn and SilviScan to validate an efficient phenotyping methodology for a greater number of samples. A test with 30 progenies of Radiata pine was established in Flynn Australia. The evaluated traits were wood density, microfibril ange and modulus of elasticity. Genetic and phenotypic correlation between traits of wood quality obtained from two methods and narrow-sense individual heritability were estimated. The Pilodyn data showed high heritability and high genetic and phenotypic correlation between wood density and moderate with microfibril angle and modulus of elasticity. The results confirm that the Pylodyn is an effective indirect and rapid method for evaluation of genetic parameters for wood quality traits in Radiata pine.
APA, Harvard, Vancouver, ISO, and other styles
4

Teague, Kara Elizabeth. "Environmental ramifications of the fire ecology of slash pine (Pinus elliottii) a study of population dynamics and dispersal following a fire event /." [Tampa, Fla. : s.n.], 2003. http://purl.fcla.edu/fcla/etd/SFE0000089.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Quinde, Abad Augusto. "Behaviour of the major resin- and fatty acids of slash pine (Pinus elliottii) during organosolv pulping." Thesis, University of British Columbia, 1990. http://hdl.handle.net/2429/30658.

Full text
Abstract:
A high extractive-content temperate conifer wood (Pinus elliottii) was examined as a pulpwood source by organosolv pulping. Particularly, the behavior of the resin- and fatty acids during the lignin solvolysis process was studied in detail. For this purpose the resin-and fatty acids were characterized in the wood, and after pulping trials in order to reveal their fate during pulping, using catalyzed 80% aqueous alcohol (methanol) as solvent. Wood extractives were removed by both methanolic cold maceration and Soxhlet extraction techniques. The resin-and fatty acid fractions thus collected were saponified and/or methylated and characterized by gas liquid chromatography (GC) and gas chromatography-mass spectrometry (GC-MS). No significant differences were found in regard to extraction efficiencies between the two types of cold extractions. Furthermore, there was no significant difference between these two types of cold extractions in comparison with the procedure described by TAPPI standard T 204 os-76. Pulping experiments were performed at 205°C for periods of 5, 20, 40, and 60 min. Lignins, which precipitated on cooling of the black liquor (Lignin fraction I), were set aside for further extractions and chemical analyses. The molecular weight distribution of these lignins was determined by size exclusion chromatography on an HPLC and their quantity was determined either gravimetrically or volumetrically. Precipitated Lignin Fraction I, suspected of containing some adsorbed extractives and some fiber fragments, was transferred to a tared crucible. The lignin and extractives were sequentially dissolved by using tetrahydrofuran (THF), acetone and methanol. This solution was evaporated, the residue redissolved in methanol-water (80:20) and the solution liquid-liquid extracted with diethyl ether in a separatory funnel followed by methylation prior to GC and GC-MS analysis. Quantification of the resin- and fatty acids in the wood and those recovered after organosolv pulping was performed using an internal standard (methyl heptadecanoate) added prior to the extraction steps. The extractives dissolved in the black liquor were isolated by a ternary liquid-liquid extraction scheme using diethyl ether, methylated with fresh diazomethane, and the resin- and fatty acids methyl esters characterized by GC and GC-MS. The extractives present in the pulp were isolated (removed) by a Soxhlet extraction procedure with methanol and" the resin- and fatty acids fractions characterized as above. Resin- and fatty acids surviving the high-temperature pulping process, were found mainly in the black liquor. After the 60 min cook, the black liquor contained 78.1% and 71.6% of resin- and fatty acids, respectively, while the pulp retained 11.7% and 8.2%, respectively of the extractives originally present in wood. "Lignin fraction I" adsorbed 10.2% and 20.2% of the resin- and fatty acids, respectively. Contrarily, if all of the lignin is precipitated (Lignin fraction II). prior to liquid/liquid extraction of the black liquor with diethyl ether, 98% and 60.4% of the resin- and fatty acids co-precipitate with the lignin and 2.0% and 39.6%, respectively, remain dissolved in the aqueous filtrate. Industrial organosolv lignin isolated after solvent pulping of pine was thus shown to contain most (98%) of the resin acids and 39.6% of the fatty acids normally found in pines. Although not tested, it is supposed that lignins isolated by precipitation from the black liquor after organosolv pulping of other species cannot be considered as "pristine lignins" as described hitherto in the technical literature, since such lignins are heavily contaminated by the extractives of the wood species. In light of these findings all data on chemical and physical characterization of organosolv lignins and their reactivity will have to be reexamined and reassessed to remove the effect of the extractives as contaminants.
Forestry, Faculty of
Graduate
APA, Harvard, Vancouver, ISO, and other styles
6

Harley, Grant L., Henri D. Grissino-Mayer, and Sally P. Horn. "The Dendrochronology Of Pinus Elliottii In The Lower Florida Keys: Chronology Development And Climate Response." Tree-Ring Society, 2011. http://hdl.handle.net/10150/622627.

Full text
Abstract:
South Florida slash pine (Pinus elliottii var. densa) is the southernmost pine species in the United States and the foundation species of the globally endangered pine rockland communities in south Florida. To test if slash pine produces annual growth rings in the Lower Florida Keys, we counted the number of rings on samples collected from the North Big Pine Key site (NBP), which contained a fire scar from a known wildfire and a known date for hurricane-induced tree mortality (2006 or 2007). In addition, a crossdated tree-ring chronology (1871–2009) was developed from living trees and remnant wood found at the site and compared to divisional climate data to determine how the regional climate regime influences radial growth. Our analyses demonstrated that slash pine forms anatomically distinct, annual growth rings with the consistent year-to-year variability necessary for rigorous dendrochronological studies. Response-function and correlation analysis showed that annual growth of slash pine at NBP is primarily influenced by water availability during the growing season. However, no significant correlations were found between tree growth and the Atlantic Multidecadal Oscillation or the El Niño-Southern Oscillation. Our study reveals the potential of producing high-quality dendrochronological data in southern Florida from slash pine, which should prove useful in further studies on fire history and tree phenology and for assessing the projected impacts of impending climate change on the fragile pine rockland community.
APA, Harvard, Vancouver, ISO, and other styles
7

Wyss, Lozano Hoyos Tania. "Pinus elliottii var. densa Seedling Performance Reflects Ectomycorrhizas, Soil Nutrient Availability and Root Competition." Scholarly Repository, 2010. http://scholarlyrepository.miami.edu/oa_dissertations/496.

Full text
Abstract:
Ectomycorrhizas generally improve seedling mineral nutrition and growth, so I hypothesized that decline of the Florida native pine variety Pinus elliottii var. densa Little & Dorman is related to deficiency of appropriate ectomycorrhizal (ECM) fungi in the pine's native flatwoods. At Archbold Biological Station I examined how quickly ECM fungi colonize P. elliottii var. densa seedlings and I compared the effect of local absence versus presence of adult pines on ECM colonization and pine seedling performance. Under controlled greenhouse conditions, I investigated how a wide range of ECM colonization and spread of extraradical mycelium throughout a large volume of relatively infertile, flatwoods soil enhance the mineral nutrition and growth of pine seedlings. In a field bioassay, I transplanted two-month-old pine seedlings to three flatwoods sites with low (4 pines/400 square m), medium (9 pines/400 square m), and high (19 pines/400 square m) adult pine densities. I subsequently excavated seedlings every two weeks for four-and-a-half months and determined their ECM colonization, response to shade, and response to surrounding grass density. Across all sites, pine seedlings in high shade had a higher mean chlorophyll concentration and lower stem dry weight than in full sun. Competition with grass reduced seedling survival and stem dry weight. Initial colonization was rapid and not different among sites, with 5.4 % of roots colonized 15 days after transplant. Pine seedlings had midpoint means of 29.5 %, 18.1 % and 21.3 % ECM root tips in low, medium and high adult pine density sites, respectively, suggesting that pine seedlings establishing in flatwoods encounter sufficient ECM fungi to support their growth, regardless of adult pine density. In a field experiment, I determined in the presence versus absence of adult pines if pine seedlings had higher ECM colonization and consequent improved survival, mineral nutrition, and growth. Within and beyond pine stands, I transplanted seedlings into intact or drilled, hyphae in-growth pipes buried in the ground. I placed autoclaved or fresh ECM root inoculum in two sets of intact pipes, and autoclaved inoculum in drilled pipes into which mycorrhizal hyphae could extend from the surrounding vegetation. Seven-and-a-half months after transplant, ECM hyphae had penetrated the drilled pipes and colonized pine seedlings, but roots from the surrounding vegetation also penetrated pipes. Extraneous roots reduced the survival of seedlings both within and beyond pine stands, but extraneous roots reduced seedling growth only beyond pine stands. Because percentage ECM root tips was higher in the presence (53 %) than in the absence (38.8%) of adult pines, pine stands might benefit the competitive ability of seedlings by increased ECM colonization and possibly by common mycorrhizal networks connecting seedlings to adults. Because beneficial effects of ECM in the field were small, I also examined ECM effects on pine seedlings in a greenhouse experiment. I manipulated ECM fungus colonization and the volume of flatwoods soil to which extraradical mycelium had access. In a small volume of soil (220 mL), fresh ECM root inoculum promoted the mycorrhizal colonization of seedlings versus those receiving autoclaved roots, but seedling growth and uptake of Mg, Ca, and Zn was lower with fresh than with autoclaved root inoculum. Growth and mineral nutrient uptake likely was enhanced by a pulse of nutrients from autoclaved roots, but for inoculated plants may have been reduced because of nutrient retention by saprotrophic microorganisms degrading fresh ECM roots and because of mineral nutrient retention by ECM fungi. Ectomycorrhizal seedlings with extraradical mycelium access to a large soil volume had higher mean chlorophyll concentration than those in a small soil volume. Weekly disturbance of the extraradical mycelium, however, reduced foliar contents of Mn, K, P, N, and Zn by one-third to one-half, and reduced needle dry weight of seedlings by one-third, demonstrating the importance of extraradical mycelium accessing a large volume of soil when it is nutrient-poor. My research demonstrates that ECM fungi are widespread in flatwoods and rapidly colonize pine seedlings. ECM fungus inocula are greater in the presence than in the absence of adult pines, and ECM or seedlings' connections to a common mycorrhizal network improve seedlings' belowground competitive ability. ECM especially enhance seedling mineral nutrition and growth when undisturbed, extraradical mycelium extends throughout a large volume of soil. Populations of Pinus elliottii var. densa might best regenerate in flatwoods if seedlings recruit near adult pines and where there is little competition for light, water, and mineral nutrients.
APA, Harvard, Vancouver, ISO, and other styles
8

Medina, Perez Alex Mauricio. "Survival and promotion of female and male strobili by topgrafting in a third-cycle slash pine (Pinus elliottii var. elliottii) breeding program." [Gainesville, Fla.] : University of Florida, 2005. http://purl.fcla.edu/fcla/etd/UFE0011827.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Lister, Andrew Joseph. "Spatial Patterns in a 40-year-old Slash Pine (Pinus elliottii Engelm.) Forest in the Coastal Plain of South Carolina." Thesis, Virginia Tech, 1998. http://hdl.handle.net/10919/36472.

Full text
Abstract:
A study was conducted at the Savannah River Site near Aiken, SC to: 1) characterize the spatial patterns of soil and forest floor variables (moisture, pH, soil phosphate, forest floor and soil carbon and nitrogen, and soil available nitrogen), 2) assess the spatial patterns of the plant community, and 3) investigate spatial relationships among the variables and between the variables and woody vegetation. Spatial soil and litter samples were collected on five 0.25 hectare plots, and relationships were explored using Pearson's correlation tests, canonical correlation analysis, variogram modeling and kriging. The average range of spatial autocorrelation for the forest floor variables was >45 m, while that for soil variables was 12 m. Woody stem basal area exhibited spatial autocorrelation at ranges of less than 12 m, and was only weakly correlated with forest floor and soil resource patterns. Few strong spatial correlations among the forest floor and soil variables were observed. The means and variances of the variables were low, and differences in resource levels probably had little impact on the spatial pattern of vegetation. Results indicate a weak, differential effect of species group on litter quality, a weak relationship between large pine trees and soil nitrogen patterns, and a general homogeneity of the stands.
Master of Science
APA, Harvard, Vancouver, ISO, and other styles
10

Teague, Kara Elizabeth. "Environmental Ramification of the Fire Ecology of Slash Pine (Pinus elliottii): A Study of Population Dynamics and Dispersal following a Fire Event." Scholar Commons, 2003. https://scholarcommons.usf.edu/etd/1491.

Full text
Abstract:
With increasing encroachment on natural communities by anthropogenic activity, it is important to understand the functions of natural ecosystems in an effort to conserve natural areas. A first-hand study of the population dynamics of South Florida Slash Pine (P. elliottii Engelm. var. densa) following a fire event provided insight to its recovery and dispersal following a fire. A natural fire (lightning-induced) occurred in the spring of 2000 at the T. Mabry Carlton, Jr. Reserve, Sarasota County, providing an opportunity to study aspects of slash pine in relation to fire. One objective of my research was to look at dispersal/recruitment conditions and slash pine dynamics in relation to fire. I looked at the varying degrees of tree mortality due to fire at different stands of slash pines. I also looked at the stands in terms of stand composition and spatial arrangement of surviving adults. Finally, I studied how variable seedling establishment and survival was between stands. Few inferences could be drawn between fire and these individual analyses; however, all analyses revealed that at the scale of this study, pine flatwoods are patchy. I also looked at the dispersal of slash pines following a fire event. I modeled my research after Ribbens et al. (1994) and Clark et al. (1998), who took a phenomenological approach to dispersal modeling. This approach involved using distances between adults and seeds/seedlings and fecundity of adults to create dispersal models based on maximum likelihood estimates (MLE). I found that, while I could predict a model within acceptable parameters for most of the stands, more data was needed to predict models that better fit the data. This finding, along with the fact that I recovered no seed data for analysis, suggests factors are contributing to dispersal and recruitment (e.g. cone-crop) that need to be accounted for in the future.
APA, Harvard, Vancouver, ISO, and other styles

Books on the topic "Slash Pine (Pinus Elliotti)"

1

Z, Rakotovao Ramasiarivelo. Essai de provenances de Pinus elliottii no 12/74 - EPR - Morarano: Résultats à 14 ans. [Antananarivo]: Repoblika Demokratika Malagasy, Ministère de la recherche scientifique et technologique pour le développement, Centre national de la recherche appliquée au développement rural, Département de recherches forestières et piscicoles, 1989.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
2

Sluder, Earl R. Further comparisons between infection of loblolly and slash pines by fusiform rust after artificial inoculation or planting. Asheville, NC: U.S. Dept. of Agriculture, Forest Service, Southeastern Forest Experiment Station, 1986.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
3

Sequeira-Vindas, Wilber. Canopy structure, light penetration and tree growth at different configurations in various 18-year-old slash pine (Pinus elliottii Engelm. var. elliottii) stands in Florida. 1991.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
4

McNabb, Kenneth Lee. The relationship of carbohydrate reserves to the quality of bare-root Pinus elliottii var. elliottii (Engelm.) seedlings produced in a northern Florida nursery. 1985.

Find full text
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "Slash Pine (Pinus Elliotti)"

1

Lesney, M. S. "Slash Pine (Pinus elliottii Engelm.)." In Biotechnology in Agriculture and Forestry, 288–303. Berlin, Heidelberg: Springer Berlin Heidelberg, 1991. http://dx.doi.org/10.1007/978-3-662-13231-9_18.

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

Newton, R. J., K. A. Marek-Swize, M. E. Magallanes-Cedeno, N. Dong, S. Sen, and S. M. Jain. "Somatic Embryogenesis in Slash Pine (Pinus Elliottii Engelm.)." In Somatic Embryogenesis in Woody Plants, 183–95. Dordrecht: Springer Netherlands, 1995. http://dx.doi.org/10.1007/978-94-011-0960-4_11.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Newton, R. J., N. Dong, S. Sen, K. Marek-Swize, and S. Chang. "Genetic Transformation in Pinus elliottii Engelm. (Slash Pine)." In Biotechnology in Agriculture and Forestry, 280–96. Berlin, Heidelberg: Springer Berlin Heidelberg, 1996. http://dx.doi.org/10.1007/978-3-662-09368-9_25.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Lesney, M. S., J. D. Johnson, Theresa Korhnak, and M. W. McCaffery. "In Vitro Manipulation of Slash Pine (Pinus Elliottii)." In Genetic Manipulation of Woody Plants, 43–55. Boston, MA: Springer US, 1988. http://dx.doi.org/10.1007/978-1-4613-1661-9_3.

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

Reports on the topic "Slash Pine (Pinus Elliotti)"

1

Boyle, Maxwell, and Elizabeth Rico. Terrestrial vegetation monitoring at Timucuan Ecological and Historic Preserve: 2019 data summary—Version 2.0. National Park Service, February 2022. http://dx.doi.org/10.36967/nrds-2290196.

Full text
Abstract:
The Southeast Coast Network (SECN) conducts long-term terrestrial vegetation monitoring as part of the nationwide Inventory and Monitoring Program of the National Park Service (NPS). The vegetation community vital sign is one of the primary-tier resources identified by SECN park managers, and it is currently conducted on 15 network parks (DeVivo et al. 2008). Monitoring plants and their associated communities over time allows for targeted understanding of ecosystems within the SECN geography, which provides managers information about the degree of change within their parks’ natural vegetation. 2019 marks the first year of conducting this monitoring effort on four SECN parks, including Timucuan Ecological and Historic Preserve (TIMU). A total of 23 vegetation plots were established in the park in May and June. Data collected in each plot include species richness across multiple spatial scales, species-specific cover and constancy, species-specific woody stem seedling/sapling counts and adult tree (greater than 10 centimeters [3.9 inches (in)]) diameter at breast height (DBH), overall tree health, landform, soil, observed disturbance, and woody biomass (i.e., fuel load) estimates. This report summarizes the baseline (year 1) terrestrial vegetation data collected at Timucuan Ecological and Historic Preserve in 2019. Data were stratified across three dominant broadly defined habitats within the park (Coastal Plain Nonalluvial Wetlands, Coastal Plain Open Uplands and Woodlands, and Maritime Upland Forests and Shrublands) and three land parcels (Cedar Point, Theodore Roosevelt, and Thomas Creek). Noteworthy findings include: A total of 157 vascular plant taxa (species or lower) were observed across 23 vegetation plots, including nine species not previously known from the park. Three plots were located in the footprint of the Yellow Bluff Fire, and were sampled only two weeks following the fire event. Muscadine (Muscadinia rotundifolia), cat greenbrier (Smilax glauca), water oak (Quercus nigra), and swamp tupelo (Nyssa biflora) were the most frequently encountered species in Coastal Plain Nonalluvial Wetland habitat; saw palmetto (Serenoa repens), slash pine (Pinus elliottii), and gallberry (Ilex glabra) were the most frequently encountered species in Coastal Plain Open Upland and Woodland habitat; and Darlington oak (Quercus hemisphaerica), Spanish moss (Tillandsia usenoides), and red bay (Persea borbonia) were the most frequently encountered species in Maritime Upland Forests and Shrublands. There were no exotic species of the Florida Exotic Pest Plant Council list of invasive plants (FLEPPC 2020) observed on any of these plots. Both red bay and swamp bay (Persea palustris) were largely absent from the tree stratum in these plots; however, they were present (occasionally in high abundance) in the seedling and sapling strata across all habitat types. Buckthorn bully (Sideroxylon lycioides)—listed as Endangered in the state of Florida by the Florida Department of Agriculture and Consumer Services (FDACS 2020)—was observed in three Maritime Upland Forest and Shrubland plots. The tree strata in each broadly defined habitat were dominated by the following species: Coastal Plain Nonalluvial Wetlands-loblolly bay (Gordonia lasianthus) Coastal Plain Open Uplands and Woodlands-longleaf pine (Pinus palustris) Maritime Upland Forests and Shrublands-oaks (Quercus sp.) Most stems within the tree strata exhibited healthy vigor and only moderate dieback across all habitat types. However, there was a large amount of standing dead trees in plots within Maritime Upland Forests and Shrublands. Downed woody biomass (fuel loads) were highest in the Cedar Point and Thomas Creek land parcels.
APA, Harvard, Vancouver, ISO, and other styles
2

Newton, Ronald, Joseph Riov, and John Cairney. Isolation and Functional Analysis of Drought-Induced Genes in Pinus. United States Department of Agriculture, September 1993. http://dx.doi.org/10.32747/1993.7568752.bard.

Full text
Abstract:
Drought is a common factor limiting timber production in the U.S. and Israel. Loblolly (Pinus taeda) and alleppo pine (Pinus halepensis) seedling survival is reduced when out planted, and growth and reproduction are often hindered by periodic droughts during later stages of tree development. Molecular and gene responses to drought stress have not been characterized. The objectives were to characterize drought-induced gene clones from these pines, to determine the effects of a growth regulator on drought tolerance, ABA levels, and drought-induced gene expression in alleppo pine, and to develop procedures for loblolly pine transformation. Nearly 20 cDNA clones influenced by gradual, prolonged drought stress have been isolated. Many of these have been shown to be induced by drought stress, whereas several others are down-regulated. These are the first drought-induced genes isolated from a pine species. Two genomic clones (lp5-1 and lp3-1) have been sequenced and characterized, and each has been found to be associated with a gene family. Clone lp5 appears to code for a cell wall protein, and clone lp3 codes for a nuclear protein. The former may be associated with changing the elastic properties of the cell wall, while the latter may be involved in signal transduction and/or protection from desiccation in the nucleus. Clone lp3 is similar to a drought-induced gene from tomato and is regulated by ABA. Several DNA sequences that are specific to induction during growth-retardation in alleppo pine by uniconazole have been identified. The active DNA species is now being identified. Promoters from genomic clones, lp3 and lp5, have been sequenced. Both are functional when fused with the gus reporter gene and transferred to other plant tissues as well as responding to a simulated drought stress. Through exodeletion analysis, it has been established that the promoter ABRE element of lp3 responds to ABA and that drought-induction of lp3 expression may also involve ABA. Stable tobacco transformants carrying either the lp5 or the lp3 promoter fused to a reporter gus gene have been obtained. The lp5lgus fusion was expressed at several stages of tobacco development and differentiation including the reproductive stage. There was no difference in phenotype between the transformants and the wild type. Embryogenesis procedures were developed for slash pine, but attempts to couple this process with gene transfer and plantlet transformation were not successful. Transformation of pine using Agrobacterium appears tractable, but molecular data supporting stable integration of the Agrobacterium-transferred gene are still inconclusive.
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