Sample records for ponderosa pine landscape

  1. Cheesman Lake-a historical ponderosa pine landscape guiding restoration in the South Platte Watershed of the Colorado Front Range

    Treesearch

    Merrill R. Kaufmann; Paula J. Fornwalt; Laurie S. Huckaby; Jason M. Stoker

    2001-01-01

    An unlogged and ungrazed ponderosa pine/Douglas-fir landscape in the Colorado Front Range provides critical information for restoring forests in the South Platte watershed. A frame-based model was used to describe the relationship among the four primary patch conditions in the 35-km2 Cheesman Lake landscape: (1) openings, (2) ponderosa pine forest, (3) ponderosa pine/...

  2. Effect of prescribed burning on mortality of resettlement ponderosa pines in Grand Canyon National Park

    Treesearch

    G. Alan Kaufmann; W. Wallace Covington

    2001-01-01

    Ponderosa pine (Pinus ponderosa) trees established before Euro-American settlement are becoming rare on the landscape. Prescribed fire is the prime tool used to restore ponderosa pine ecosystems, but can cause high mortality in presettlement ponderosa pines. This study uses retrospective techniques to estimate mortality from prescribed burns within Grand Canyon...

  3. Ponderosa pine ecosystems

    Treesearch

    Russell T. Graham; Theresa B. Jain

    2005-01-01

    Ponderosa pine is a wide-ranging conifer occurring throughout the United States, southern Canada, and northern Mexico. Since the 1800s, ponderosa pine forests have fueled the economies of the West. In western North America, ponderosa pine grows predominantly in the moist and dry forests. In the Black Hills of South Dakota and the southern portion of its range, the...

  4. Chapter 5. Dynamics of ponderosa and Jeffrey pine forests

    Treesearch

    Penelope Morgan

    1994-01-01

    Ponderosa (Pinus ponderosa) and Jeffrey pine (Pinus jefferyi) forests are ecologically diverse ecosystems. The communities and landscapes in which these trees dominate are variable and often complex. Because of the economic value of resources, people have used these forests extensively.

  5. Changes in canopy fuels and fire behavior after ponderosa pine restoration treatments: A landscape perspective

    Treesearch

    J. P. Roccaforte; P. Z. Fule

    2008-01-01

    (Please note, this is an abstract only) We modeled crown fire behavior and assessed changes in canopy fuels before and after the implementation of restoration treatments in a ponderosa pine landscape at Mt. Trumbull, Arizona. We measured 117 permanent plots before (1996/1997) and after (2003) thinning and burning treatments. The plots are evenly distributed across the...

  6. Ponderosa pine in the Colorado Front Range: long historical fire and tree recruitment intervals and a case for landscape heterogeneity

    Treesearch

    M. R. Kaufmann; L. S. Huckaby; P. Gleason

    2000-01-01

    An unlogged forest landscape in the Colorado Front Range provides insight into historical characteristics of ponderosa pine/Douglas-fir landscapes where the past fire regime was mixed severity with mean fire intervals of 50 years or more. Natural fire and tree recruitment patterns resulted in considerable spatial and temporal heterogeneity, whereas nearby forest...

  7. Long-term, landscape patterns of past fire events in a montane ponderosa pine forest of central Colorado

    Treesearch

    Peter M. Brown; Merrill R. Kaufmann; Wayne D. Shepperd

    1999-01-01

    Parameters of fire regimes, including fire frequency, spatial extent of burned areas, fire severity, and season of fire occurrence, influence vegetation patterns over multiple scales. In this study, centuries-long patterns of fire events in a montane ponderosa pine - Douglas-fir forest landscape surrounding Cheesman Lake in central Colorado were reconstructed from fire...

  8. Vegetation response to stand structure and prescribed fire in an interior ponderosa pine ecosystem

    Treesearch

    Jianwei Zhang; Martin W. Ritchie; William W. Oliver

    2008-01-01

    A large-scale interior ponderosa pine (Pinus ponderosa Dougl. ex P. & C. Laws.) study was conducted at the Blacks Mountain Experimental Forest in northeastern California. The primary purpose of the study was to determine the influence of structural diversity on the dynamics of interior pine forests at the landscape scale. High structural...

  9. Ecology of southwestern ponderosa pine forests

    Treesearch

    William H. Moir; Brian W. Geils; Mary Ann Benoit; Dan Scurlock

    1997-01-01

    Ponderosa pine forests are important because of their wide distribution, commercial value, and because they provide habitat for many plants and animals. Ponderosa pine forests are noted for their variety of passerine birds resulting from variation in forest composition and structure modified by past and present human use. Subsequent chapters discuss how ponderosa pine...

  10. Western juniper and ponderosa pine ecotonal climate-growth relationships across landscape gradients in southern Oregon

    USGS Publications Warehouse

    Knutson, K.C.; Pyke, D.A.

    2008-01-01

    Forecasts of climate change for the Pacific northwestern United States predict warmer temperatures, increased winter precipitation, and drier summers. Prediction of forest growth responses to these climate fluctuations requires identification of climatic variables limiting tree growth, particularly at limits of free species distributions. We addressed this problem at the pine-woodland ecotone using tree-ring data for western juniper (Juniperus occidentalis var. occidentalis Hook.) and ponderosa pine (Pinus ponderosa Dougl. ex Loud.) from southern Oregon. Annual growth chronologies for 1950-2000 were developed for each species at 17 locations. Correlation and linear regression of climate-growth relationships revealed that radial growth in both species is highly dependent on October-June precipitation events that recharge growing season soil water. Mean annual radial growth for the nine driest years suggests that annual growth in both species is more sensitive to drought at lower elevations and sites with steeper slopes and sandy or rocky soils. Future increases in winter precipitation could increase productivity in both species at the pine-woodland ecotone. Growth responses, however, will also likely vary across landscape features, and our findings suggest that heightened sensitivity to future drought periods and increased temperatures in the two species will predominantly occur at lower elevation sites with poor water-holding capacities. ?? 2008 NRC.

  11. Should ponderosa pine be planted on lodgepole pine sites?

    Treesearch

    P.H. Cochran

    1984-01-01

    Repeated radiation frosts caused no apparent harm to the majority of lodgepole pine (Pinus contorta Dougl.) seedlings planted on a pumice flat in south-central Oregon. For most but not all of the ponderosa pine (Pinus ponderosa Dougl.) seedlings planted with the lodgepole pine, however, damage from radiation frost resulted in...

  12. Landscape-scale genetic variation in a forest outbreak species, the mountain pine beetle (Dendroctonus ponderosae)

    Treesearch

    K. E. Mock; B. J. Bentz; E. M. O' Neill; J. P. Chong; J. Orwin; M. E. Pfrender

    2007-01-01

    The mountain pine beetle Dendroctonus ponderosae is a native species currently experiencing large-scale outbreaks in western North American pine forests. We sought to describe the pattern of genetic variation across the range of this species, to determine whether there were detectable genetic differences between D. ponderosae...

  13. Financial analysis of pruning ponderosa pine.

    Treesearch

    Roger D. Fight; Natalie A. Bolon; James M. Cahill

    1992-01-01

    A recent lumber recovery study of pruned and unpruned ponderosa pine (Pinus ponderosa Dougl. ex Laws.) was used to project the financial return from pruning ponderosa pine in the Medford District of the Bureau of Land Management and in the Ochoco and Deschutes National Forests. The cost of pruning at which the investment would yield an expected 4-...

  14. Fire ecology of ponderosa pine and the rebuilding of fire-resilient ponderosa pine ecosystems

    Treesearch

    Stephen A. Fitzgerald

    2005-01-01

    The ponderosa pine ecosystems of the West have change dramatically since Euro-American settlement 140 years ago due to past land uses and the curtailment of natural fire. Today, ponderosa pine forests contain over abundance of fuel, and stand densities have increased from a range of 49-124 trees ha-1 (20-50 trees acre-1) to...

  15. AmeriFlux US-Vcp Valles Caldera Ponderosa Pine

    DOE Data Explorer

    Litvak, Marcy [University of New Mexico

    2016-01-01

    This is the AmeriFlux version of the carbon flux data for the site US-Vcp Valles Caldera Ponderosa Pine. Site Description - The Valles Caldera Ponderosa Pine site is located in the 1200km2 Jemez River basin of the Jemez Mountains in north-central New Mexico at the southern margin of the Rocky Mountain ecoregion. The Ponderosa Pine forest is the warmest and lowest (below 2700m) zone of the forests in the Valles Caldera National Preserve. Its vegetation is composed of a Ponderosa Pine (Pinus Ponderosa) overstory and a Gambel Oak (Quercus gambelii) understory.

  16. Management of ponderosa pine nutrition through fertilization

    Treesearch

    Mariann T. Garrison-Johnston; Terry M. Shaw; Peter G. Mika; Leonard R. Johnson

    2005-01-01

    The results of a series of replicated fertilization trials established throughout the Inland Northwest were reviewed for information specific to ponderosa pine (Pinus ponderosa P. and C. Lawson) nutrition. Ponderosa pine nitrogen (N) status was often better than the N-status of other Inland Northwest species, and therefore growth response to N...

  17. Ponderosa Pine reclamation at the Rosebud Mine

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Martin, P.R.

    1990-12-31

    The first operational plantings of ponderosa pine (Pinus ponderosa) were made on the Rosebud Mine near Colstrip, Montana in November of 1985. This paper discusses the five {open_quotes}R`s{close_quotes} of pine reclamation. These include the Reasons for planting ponderosa pine, ponderosa pine Research efforts and results, present Reclamation methods and materials, the Results of pine reclamation to date and the relationship of these results to the final bond Release criteria. Over 14,000 pine seedlings have been planted to date. They have been 1-0 to 3-0 bare root or 1-0 to 2-0 containerized stock. Plantings have been done by hand, with augersmore » and (primarily) with a modified Soil Conservation Service tree planter on {open_quotes}tree{close_quotes} and regular soils with and without animal damage protection in spring and fall. Percent survival has varied greatly from field to field influenced by record drought, wildlife, severe grasshopper depredation, cattle grazing and wildlife usage.« less

  18. A ponderosa pine-lodgepole pine spacing study in central Oregon: results after 20 years.

    Treesearch

    K.W. Seidel

    1989-01-01

    The growth response after 20 years from an initial spacing study established in a ponderosa pine (Pinus ponderosa Dougl. ex Laws.) and lodgepole pine (Pinus contorta Dougl. ex Loud.) plantation was measured in central Oregon. The study was designed to compare the growth rates of pure ponderosa pine, pure lodgepole pine, and a...

  19. Bark beetle-caused mortality in a drought-affected ponderosa pine landscape in Arizona, USA

    Treesearch

    Jose F. Negron; Joel D. McMillin; John A. Anhold; Dave Coulson

    2009-01-01

    Extensive ponderosa pine (Pinus ponderosa Dougl. ex Laws.) mortality associated with a widespread severe drought and increased bark beetle (Coleoptera: Curculionidae, Scolytinae) populations occurred in Arizona from 2001 to 2004. A complex of Ips beetles including: the Arizona fivespined ips, Ips lecontei Swaine...

  20. Songbird ecology in southwestern ponderosa pine forests: A literature review

    Treesearch

    William M. Block; Deborah M. Finch

    1997-01-01

    This publication reviews and synthesizes the literature about ponderosa pine forests of the Southwest, with emphasis on the biology, ecology, and conservation of songbirds. Critical bird-habitat management issues related to succession, snags, old growth, fire, logging, grazing, recreation, and landscape scale are addressed. Overviews of the ecology, current use, and...

  1. Ecology, silviculture, and management of Black Hills ponderosa pine

    Treesearch

    Wayne D. Shepperd; Michael A. Battaglia

    2002-01-01

    This paper presents a broad-based synthesis of the general ecology of the ponderosa pine ecosystem in the Black Hills. This synthesis contains information and results of research on ponderosa pine from numerous sources within the Black Hills ecosystem. We discuss the silvical characteristics of ponderosa pine, natural disturbances that govern ecosystem processes,...

  2. Using tree recruitment patterns and fire history to guide restoration of an unlogged ponderosa pine/Douglas-fir landscape in the southern Rocky Mountains after a century of fire suppression

    Treesearch

    Merrill R. Kaufmann; Laurie S. Huckaby; Paula J. Fornwalt; Jason M. Stoker; William H. Romme

    2003-01-01

    Tree age and fire history were studied in an unlogged ponderosa pine/Douglas-fir (Pinus ponderosa/Pseudotsuga menziesii) landscape in the Colorado Front Range mountains. These data were analysed to understand tree survival during fire and post-fire recruitment patterns after fire, as a basis for understanding the characteristics of, and restoration needs for, an...

  3. Using tree recruitment patterns and fire history to guide restoration of an unlogged ponderosa pine/Douglas‐fir landscape in the southern Rocky Mountains after a century of fire suppression

    USGS Publications Warehouse

    Kaufmann, M.R.; Huckaby, L.S.; Fornwalt, P.J.; Stoker, J.M.; Romme, W.H.

    2003-01-01

    Tree age and fire history were studied in an unlogged ponderosa pine/Douglas‐fir ( Pinus ponderosa/Pseudotsuga menziesii ) landscape in the Colorado Front Range mountains. These data were analysed to understand tree survival during fire and post‐fire recruitment patterns after fire, as a basis for understanding the characteristics of, and restoration needs for, an ecologically sustainable landscape. Comparisons of two independent tree age data sets indicated that sampling what subjectively appear to be the five oldest trees in a forest polygon could identify the oldest tree. Comparisons of the ages of the oldest trees in each data set with maps of fire history suggested that delays in establishment of trees, after stand‐replacing fire, ranged from a few years to more than a century. These data indicate that variable fire severity, including patches of stand replacement, and variable temporal patterns of tree recruitment into openings after fire were major causes of spatial heterogeneity of patch structure in the landscape. These effects suggest that restoring current dense and homogeneous ponderosa pine forests to an ecologically sustainable and dynamic condition should reflect the roles of fires and variable patterns of tree recruitment in regulating landscape structure.

  4. Diprionidae sawflies on lodgepole and ponderosa pines

    USDA-ARS?s Scientific Manuscript database

    Eight species of Diprionidae feed on lodgepole pine (Pinus contorta) and ponderosa pine (P. ponderosa) in western United States: Neodiprion burkei Middleton, N. annulus contortae Ross, N. autumnalis Smith, N. fulviceps (Cresson), N. gillettei (Rohwer), N. mundus Rohwer, N. ventralis Ross, and Zadi...

  5. Susceptibility of ponderosa pine, Pinus ponderosa (Dougl. Ex Laws.), to mountain pine beetle, Dendroctonus ponderosae Hopkins, attack in uneven-aged stands in the Black Hills of South Dakota and Wyoming USA

    Treesearch

    Jose F. Negron; Kurt Allen; Blaine Cook; John R. Withrow

    2008-01-01

    Mountain pine beetle, Dendroctonus ponderosae Hopkins can cause extensive tree mortality in ponderosa pine, Pinus ponderosa Dougl. ex Laws., forests in the Black Hills of South Dakota and Wyoming. Most studies that have examined stand susceptibility to mountain pine beetle have been conducted in even-aged stands. Land managers...

  6. Tall oil precursors in three western pines: ponderosa, lodgepole, and limber pine

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Conner, A.H.; Diehl, M.A.; Rowe, J.W.

    1980-01-01

    The nonvolatile diethyl ether extracts (NVEE) from ponderosa, lodgepole, and limber pines were analyzed to determine the amounts and chemical composition of the tall oil precursors (resin acids, fatty acids, and nonsaponifiables) and turpentine precursors available from these species. The results showed that crude tall oil compositions would be approximately as follows (% resin acids, % fatty acids, % nonsaponifiables); ponderosa pine - sapwood (15, 75, 10), heartwood (78, 7, 15); lodgepole pine - sapwood (24, 57, 19), heartwood (51, 26, 23); limber pine - sapwood (10, 82, 8), heartwood (23, 60, 17). The larger nonsaponifiables content, as compared tomore » southern pines, is the major factor in explaining the greater difficulty in the distillative refining of tall oil from these western species. Eight resin acids were found in ponderosa and lodgepole pine: palustric, isopimaric, abietic, dehydroabietic, and neoabietic acids predominated. Seven resin acids were identified from limber pine: anticopalic, isopimaric, abietic, and dehydroabietic acids predominated. The free and esterfied fatty acids from these species contained predominantly oleic and linoleic acids. In addition limber pine contained major amounts of 5, 9, 12-octadecatrienoic acid. The nonsaponifiables contained mostly diterpenes and the sterols, sitosterol and campesterol. The major turpentine components were: ponderosa pine - ..beta..-pinene and 3-carene; lodgepole pine - ..beta..-phellandrene; and limber pine - 3-carene, ..beta..-phellandrene, ..cap alpha..-piene, and ..beta..-pinene.« less

  7. Fuel loadings 5 years after a bark beetle outbreak in south-western USA ponderosa pine forests

    Treesearch

    Chad M. Hoffman; Carolyn Hull Sieg; Joel D. McMillin; Peter Z. Fule

    2012-01-01

    Landscape-level bark beetle (Coleoptera: Curculionidae, Scolytinae) outbreaks occurred in Arizona ponderosa pine (Pinus ponderosa Dougl. ex Law.) forests from 2001 to 2003 in response to severe drought and suitable forest conditions.We quantified surface fuel loadings and depths, and calculated canopy fuels based on forest structure attributes in 60 plots established 5...

  8. Genetic variation and hybridization of ponderosa pine

    Treesearch

    M. Thompson Conkle; William B. Critchfield

    1988-01-01

    Ponderosa pine's (Pinus ponderosa Laws.) geographic range is centered in the montane western United States, where it is the most widely adapted and ubiquitous conifer. A western variety, P. ponderosa var. ponderosa, extends from the mountains of southern California, northward, on the western and eastern...

  9. A test of grafting ponderosa pine.

    Treesearch

    Edwin L. Mowat; Roy R. Silen

    1957-01-01

    Cleft grafting ponderosa pine proved best of five methods of vegetative propagation tested for field use in the dry hot climate of eastern Oregon. Ponderosa pine has been successfully grafted under many conditions elsewhere, but first trials in this area during 1955 were a failure. This study was made in 1956 at the U.S. Forest Service nursery at Bend, Oreg., to find a...

  10. A density management diagram for even-aged ponderosa pine stands

    Treesearch

    James N. Long; John D. Shaw

    2005-01-01

    We developed a density management diagram (DMD) for ponderosa pine using Forest Inventory and Analysis (FIA) data. Analysis plots were drawn from all FIA plots in the western United States on which ponderosa pine occurred. A total of 766 plots met the criteria for analysis. Selection criteria were for purity, defined as ponderosa pine basal area 80% of plot basal area...

  11. Fire history in the ponderosa pine/Douglas-fir forests on the east slope of the Washington Cascades.

    Treesearch

    Richard L. Everett; Richard Schellhaas; Dave Keenum

    2000-01-01

    We collected 490 and 233 fire scars on two ponderosa pine (Pinus ponderosa)/Douglas-fir (Pseudotsuga menziesii) dominated landscapes on the east slope of the Washington Cascades that contained a record of 3901 and 2309 cross-dated fire events. During the pre-settlement period (1700/1750±1860), the Weibull median fire-free...

  12. Mountain pine beetle attack associated with low levels of 4-allylanisole in ponderosa pine.

    PubMed

    Emerick, Jay J; Snyder, Aaron I; Bower, Nathan W; Snyder, Marc A

    2008-08-01

    Mountain pine beetle (Dendroctonus ponderosae) is the most important insect pest in southern Rocky Mountain ponderosa pine (Pinus ponderosa) forests. Tree mortality is hastened by the various fungal pathogens that are symbiotic with the beetles. The phenylpropanoid 4-allylanisole is an antifungal and semiochemical for some pine beetle species. We analyzed 4-allylanisole and monoterpene profiles in the xylem oleoresin from a total of 107 trees at six sites from two chemotypes of ponderosa pine found in Colorado and New Mexico using gas chromatography-mass spectroscopy (GC-MS). Although monoterpene profiles were essentially the same in attacked and nonattacked trees, significantly lower levels of 4-allylanisole were found in attacked trees compared with trees that showed no evidence of attack for both chemotypes.

  13. Multiaged silviculture of ponderosa pine

    Treesearch

    Kevin L. O' Hara

    2005-01-01

    Ponderosa pine (Pinus ponderosa P. & C. Lawson) is highly suitable for management using multiaged systems. This suitability is primarily the result of a frequent, low severity disturbance regime, but also because it naturally occurs at low densities and has a long history of management to promote multiple age classes. Several different stocking...

  14. Large-scale thinning, ponderosa pine, and mountain pine beetle in the Black Hills, USA

    Treesearch

    Jose F. Negron; Kurt K. Allen; Angie Ambourn; Blaine Cook; Kenneth Marchand

    2017-01-01

    Mountain pine beetle (Dendroctonus ponderosae Hopkins) (MPB), can cause extensive ponderosa pine (Pinus ponderosa Dougl. ex Laws.) mortality in the Black Hills of South Dakota and Wyoming, USA. Lower tree densities have been associated with reduced MPB-caused tree mortality, but few studies have reported on large-scale thinning and most data come from small plots that...

  15. Predictions of fire behavior and resistance to control: for use with photo series for the ponderosa pine type, ponderosa pine and associated species type, and lodgepole pine type.

    Treesearch

    Franklin R. Ward; David V. Sandberg

    1981-01-01

    This publication presents tables on the behavior of fire and the resistance of fuels to control. The information is to be used with the publication, "Photo Series for Quantifying Forest Residues in the Ponderosa Pine Type, Ponderosa Pine and Associated Species Type, Lodgepole Pine Type" (Maxwell, Wayne G.; Ward, Franklin R. 1976. Gen. Tech. Rep. PNW-GTR-052....

  16. Phenology and density-dependent dispersal predict patterns of mountain pine beetle (Dendroctonus ponderosae) impact

    Treesearch

    James A. Powell; Barbara J. Bentz

    2014-01-01

    For species with irruptive population behavior, dispersal is an important component of outbreak dynamics. We developed and parameterized a mechanistic model describing mountain pine beetle (Dendroctonus ponderosae Hopkins) population demographics and dispersal across a landscape. Model components include temperature-dependent phenology, host tree colonization...

  17. Diseases as agents of disturbance in ponderosa pine

    Treesearch

    Gregory M. Filip

    2005-01-01

    Several diseases affect the growth and survival of ponderosa pine in the Pacific Northwest and serve as agents of disturbance. Probably the most widespread and damaging class of disease agents is dwarf mistletoe, which causes serious growth loss and mortality of ponderosa pine. Dwarf mistletoes (Arceuthobium spp.) are seed plants that can parasitize...

  18. Mountain Pine Beetle Dynamics and Reproductive Success in Post-Fire Lodgepole and Ponderosa Pine Forests in Northeastern Utah

    PubMed Central

    Lerch, Andrew P.; Pfammatter, Jesse A.

    2016-01-01

    Fire injury can increase tree susceptibility to some bark beetles (Curculionidae, Scolytinae), but whether wildfires can trigger outbreaks of species such as mountain pine beetle (Dendroctonus ponderosae Hopkins) is not well understood. We monitored 1173 lodgepole (Pinus contorta var. latifolia Doug.) and 599 ponderosa (Pinus ponderosa Doug. ex Law) pines for three years post-wildfire in the Uinta Mountains of northeastern Utah in an area with locally endemic mountain pine beetle. We examined how the degree and type of fire injury influenced beetle attacks, brood production, and subsequent tree mortality, and related these to beetle population changes over time. Mountain pine beetle population levels were high the first two post-fire years in lodgepole pine, and then declined. In ponderosa pine, populations declined each year after initial post-fire sampling. Compared to trees with strip or failed attacks, mass attacks occurred on trees with greater fire injury, in both species. Overall, a higher degree of damage to crowns and boles was associated with higher attack rates in ponderosa pines, but additional injury was more likely to decrease attack rates in lodgepole pines. In lodgepole pine, attacks were initially concentrated on fire-injured trees, but during subsequent years beetles attacked substantial numbers of uninjured trees. In ponderosa pine, attacks were primarily on injured trees each year, although these stands were more heavily burned and had few uninjured trees. In total, 46% of all lodgepole and 56% of ponderosa pines underwent some degree of attack. Adult brood emergence within caged bole sections decreased with increasing bole char in lodgepole pine but increased in ponderosa pine, however these relationships did not scale to whole trees. Mountain pine beetle populations in both tree species four years post-fire were substantially lower than the year after fire, and wildfire did not result in population outbreaks. PMID:27783632

  19. Mountain Pine Beetle Dynamics and Reproductive Success in Post-Fire Lodgepole and Ponderosa Pine Forests in Northeastern Utah.

    PubMed

    Lerch, Andrew P; Pfammatter, Jesse A; Bentz, Barbara J; Raffa, Kenneth F

    2016-01-01

    Fire injury can increase tree susceptibility to some bark beetles (Curculionidae, Scolytinae), but whether wildfires can trigger outbreaks of species such as mountain pine beetle (Dendroctonus ponderosae Hopkins) is not well understood. We monitored 1173 lodgepole (Pinus contorta var. latifolia Doug.) and 599 ponderosa (Pinus ponderosa Doug. ex Law) pines for three years post-wildfire in the Uinta Mountains of northeastern Utah in an area with locally endemic mountain pine beetle. We examined how the degree and type of fire injury influenced beetle attacks, brood production, and subsequent tree mortality, and related these to beetle population changes over time. Mountain pine beetle population levels were high the first two post-fire years in lodgepole pine, and then declined. In ponderosa pine, populations declined each year after initial post-fire sampling. Compared to trees with strip or failed attacks, mass attacks occurred on trees with greater fire injury, in both species. Overall, a higher degree of damage to crowns and boles was associated with higher attack rates in ponderosa pines, but additional injury was more likely to decrease attack rates in lodgepole pines. In lodgepole pine, attacks were initially concentrated on fire-injured trees, but during subsequent years beetles attacked substantial numbers of uninjured trees. In ponderosa pine, attacks were primarily on injured trees each year, although these stands were more heavily burned and had few uninjured trees. In total, 46% of all lodgepole and 56% of ponderosa pines underwent some degree of attack. Adult brood emergence within caged bole sections decreased with increasing bole char in lodgepole pine but increased in ponderosa pine, however these relationships did not scale to whole trees. Mountain pine beetle populations in both tree species four years post-fire were substantially lower than the year after fire, and wildfire did not result in population outbreaks.

  20. Growing stock levels in even-aged ponderosa pine

    Treesearch

    Clifford A. Myers

    1967-01-01

    Growth of the most widely distributed pine in North America is under joint study by the western Forest and Range Experiment Stations of the U. S. Forest Service. Young, even-aged ponderosa pine (Pinus ponderosa Laws.) stands are being examined over a wide range of tree sizes, stand densities, and site index. The single plan that co-...

  1. Prescribed burning in southwestern ponderosa pine

    Treesearch

    Stephen S Sackett; Sally M Haase; Michael G Harrington

    1996-01-01

    Prescribed burning is an effective way of restoring the fire process to ponderosa pine (Pinus ponderosa Dougl. ex Laws.) ecosystems of the Southwest. If used judiciously, fire can provide valuable effects for hazard reduction, natural regeneration, thinning, vegetation revitalization, and in general, better forest health. Relatively short burning...

  2. Differences in ponderosa pine isocupressic acid concentrations across space and time

    USDA-ARS?s Scientific Manuscript database

    Ponderosa Pine (Pinus ponderosa) is distributed throughout the western half of North America, where it is the most widely adapted and ubiquitous conifer. Ponderosa Pine contains isocupressic acid, a diterpene acid, which has been shown to be responsible for its abortifacient activity. The objectiv...

  3. Diagnostic phytoliths for a ponderosa pine-bunchgrass community near Flagstaff, Arizona

    Treesearch

    Becky K.: Kerns

    2001-01-01

    Phytolith analysis could play an important role in understanding vegetation dynamics in southwestern ponderosa pine (Pinus ponderosa) forests, which have been dramatically altered by fire suppression and other factors. My objectives were to develop a phytolith reference collection and classification system for a ponderosa pine-bunchgrass community...

  4. Modeling forest scenic beauty: Concepts and application to ponderosa pine

    Treesearch

    Thomas C. Brown; Terry C. Daniel

    1984-01-01

    Statistical models are presented which relate near-view scenic beauty of ponderosa pine stands in the Southwest to variables describing physical characteristics. The models suggest that herbage and large ponderosa pine contribute to scenic beauty, while numbers of small and intermediate-sized pine trees and downed wood, especially as slash, detract from scenic beauty....

  5. Recovery of ectomycorrhizal fungus communities fifteen years after fuels reduction treatments in ponderosa pine forests of the Blue Mountains, Oregon

    Treesearch

    Benjamin T.N. Hart; Jane E. Smith; Daniel L. Luoma; Jeff A. Hatten

    2018-01-01

    Managers use restorative fire and thinning for ecological benefits and to convert fuel-heavy forests to fuel-lean landscapes that lessen the threat of stand-replacing wildfire. In this study, we evaluated the long-term impact of thinning and prescribed fire on soil biochemistry and the mycorrhizal fungi associated with ponderosa pine (Pinus ponderosa...

  6. Testing Verbenone for reducing mountain pine beetle attacks in ponderosa pine in the Black Hills, South Dakota

    Treesearch

    Jose F. Negron; Kurt Allen; McMillin. Joel; Henry Burkwhat

    2006-01-01

    In 2000 and 2002, Verbenone, a compound with anti-aggregation properties for mountain pine beetle, Dendroctonus ponderosae, was tested for reducing attacks by the insect in Ponderosa pine, Pinus ponderosae forests. The verbenone was released to the environment with the use of permeable membranes; the first year with plastic...

  7. Fire-injured ponderosa pine provide a pulsed resource for bark beetles

    Treesearch

    Ryan S. Davis; Sharon Hood; Barbara J. Bentz

    2012-01-01

    Bark beetles can cause substantial mortality of trees that would otherwise survive fire injuries. Resin response of fire-injured northern Rocky Mountain ponderosa pine (Pinus ponderosa Douglas ex P. Lawson & C. Lawson) and specific injuries that contribute to increased bark beetle attack susceptibility and brood production are unknown. We monitored ponderosa pine...

  8. Ungulate ecology of ponderosa pine ecosystems in the northwest

    Treesearch

    Martin Vavra; Kenric Walburger; Timothy DelCurto

    2005-01-01

    Ponderosa pine ecosystems provide important foraging habitats for both wild and domestic ungulates. Livestock typically graze ponderosa pine ecosystems from May through October. Mule deer and elk may utilize these habitats on a yearlong basis in some areas. Stand density has a significant effect on understory production. Competition for soil moisture and nitrogen limit...

  9. Field guide to old ponderosa pines in the Colorado Front Range

    Treesearch

    Laurie Stroh Huckaby; Merrill R. Kaufmann; Paula J. Fornwalt; Jason M. Stoker; Chuck Dennis

    2003-01-01

    We describe the distinguishing physical characteristics of old ponderosa pine trees in the Front Range of Colorado and the ecological processes that tend to preserve them. Photographs illustrate identifying features of old ponderosa pines and show how to differentiate them from mature and young trees. The publication includes a photographic gallery of old ponderosa...

  10. Resilience of ponderosa and lodgepole pine forests to mountain pine beetle disturbance and limited regeneration

    USGS Publications Warehouse

    Briggs, Jenny S.; Hawbaker, Todd J.; Vandendriesche, Don

    2015-01-01

    After causing widespread mortality in lodgepole pine forests in North America, the mountain pine beetle (MPB) has recently also affected ponderosa pine, an alternate host species that may have different levels of resilience to this disturbance. We collected field data in ponderosa pine- and lodgepole pine-dominated forests attacked by MPB in Colorado and then simulated stand growth over 200 years using the Forest Vegetation Simulator. We compared scenarios of no disturbance with scenarios of MPB-caused mortality, both with and without regeneration. Results indicated that basal area and tree density recovered to predisturbance levels relatively rapidly (within 1‐8 decades) in both forest types. However, convergence of the disturbed conditions with simulated undisturbed conditions took longer (12‐20+ decades) and was delayed by the absence of regeneration. In MPB-affected ponderosa pine forests without regeneration, basal area did not converge with undisturbed conditions within 200 years, implying lower resilience in this ecosystem. Surface fuels accumulated rapidly in both forest types after MPB-induced mortality, remaining high for 3‐6 decades in simulations. Our results suggest that future patterns of succession, regeneration, fuel loading, climate, and disturbance interactions over long time periods should be considered in management strategies addressing MPB effects in either forest type, but particularly in ponderosa pine.

  11. Introduction of ponderosa pine and Douglas-fir to Argentina

    Treesearch

    Gerald E. Rehfeldt; Leonardo A. Gallo

    2001-01-01

    Patterns of shoot elongation of 2-yr seedlings from native North American populations of ponderosa pine and Douglas-fir were compared to those of Argentine land races originating from unknown provenances. The comparisons were conducted in Moscow, Idaho (USA), and suggested that the ponderosa pine land race was descended from a California provenance at low or middle...

  12. Short-term ecological consequences of collaborative restoration treatments in ponderosa pine forests of Colorado

    Treesearch

    Jennifer S. Briggs; Paula J. Fornwalt; Jonas A. Feinstein

    2017-01-01

    Ecological restoration treatments are being implemented at an increasing rate in ponderosa pine and other dry conifer forests across the western United States, via the USDA Forest Service’s Collaborative Forest Landscape Restoration (CFLR) program. In this program, collaborative stakeholder groups work with National Forests (NFs) to adaptively implement and monitor...

  13. Red Rot of Ponderosa Pine (FIDL)

    Treesearch

    Stuart R. Andrews

    1971-01-01

    Red rot caused by the fungus Polyporus anceps Peck is the most important heart rot of ponderosa pine (Pinus ponderosa Laws.) in the Southwest (in Arizona and New Mexico), the Black Hills of South Dakota, and some localities in Colorado, Montana, and Idaho. It causes only insignificant losses to this species elsewhere in the West. The red rot fungus rarely attacks other...

  14. Non-native plant invasions in managed and protected ponderosa pine/Douglas-fir forests of the Colorado Front Range

    Treesearch

    Paula J. Fornwalt; Merrill R. Kaufmann; Laurie S. Huckaby; Jason M. Stoker; Thomas J. Stohlgren

    2003-01-01

    We examined patterns of non-native plant diversity in protected and managed ponderosa pine/Douglas-fir forests of the Colorado Front Range. Cheesman Lake, a protected landscape, and Turkey Creek, a managed landscape, appear to have had similar natural disturbance histories prior to European settlement and fire protection during the last century. However, Turkey Creek...

  15. Thinning ponderosa pine (Pinus ponderosa) stands reduces mortality while maintaining stand productivity

    Treesearch

    Jianwei Zhang; Martin W. Ritchie; Douglas A. Maguire; William W. Oliver

    2013-01-01

    We analyzed 45-yr data collected from three ponderosa pine (Pinus ponderosa Lawson & C. Lawson) levels-of-growing-stock installations in Oregon (OR) and northern California (CA), USA, to determine the effect of stand density regimes on stand productivity and mortality. We found that periodic annual increment (PAI) of diameter, basal area (BA...

  16. Soil moisture and the distribution of lodgepole and ponderosa pine: a review of the literature.

    Treesearch

    Robert F. Tarrant

    1953-01-01

    Despite a number of published studies and observations of the factors affecting the distribution of lodgepole pine (Pinus contorta var, latifolia) and Ponderosa pine (Pinus ponderosa), some misunderstanding still exists as to the significance of the extensive stands of lodgepole pine in the ponderosa pine...

  17. Associations among breeding birds and gambel oak in Southwestern ponderosa pine forests

    Treesearch

    Stephanie Jentsch; R. William Mannan; Brett G. Dickson; William M. Block

    2008-01-01

    Ponderosa pine (Pinus ponderosa) forests with Gambel oak (Quercus gambelii) are associated with higher bird abundance and diversity than are ponderosa pine forests lacking Gambel oak. Little is known, however, about specific structural characteristics of Gambel oak trees, clumps, and stands that may be important to birds in...

  18. Examining Historical and Current Mixed-Severity Fire Regimes in Ponderosa Pine and Mixed-Conifer Forests of Western North America

    PubMed Central

    Odion, Dennis C.; Hanson, Chad T.; Arsenault, André; Baker, William L.; DellaSala, Dominick A.; Hutto, Richard L.; Klenner, Walt; Moritz, Max A.; Sherriff, Rosemary L.; Veblen, Thomas T.; Williams, Mark A.

    2014-01-01

    There is widespread concern that fire exclusion has led to an unprecedented threat of uncharacteristically severe fires in ponderosa pine (Pinus ponderosa Dougl. ex. Laws) and mixed-conifer forests of western North America. These extensive montane forests are considered to be adapted to a low/moderate-severity fire regime that maintained stands of relatively old trees. However, there is increasing recognition from landscape-scale assessments that, prior to any significant effects of fire exclusion, fires and forest structure were more variable in these forests. Biota in these forests are also dependent on the resources made available by higher-severity fire. A better understanding of historical fire regimes in the ponderosa pine and mixed-conifer forests of western North America is therefore needed to define reference conditions and help maintain characteristic ecological diversity of these systems. We compiled landscape-scale evidence of historical fire severity patterns in the ponderosa pine and mixed-conifer forests from published literature sources and stand ages available from the Forest Inventory and Analysis program in the USA. The consensus from this evidence is that the traditional reference conditions of low-severity fire regimes are inaccurate for most forests of western North America. Instead, most forests appear to have been characterized by mixed-severity fire that included ecologically significant amounts of weather-driven, high-severity fire. Diverse forests in different stages of succession, with a high proportion in relatively young stages, occurred prior to fire exclusion. Over the past century, successional diversity created by fire decreased. Our findings suggest that ecological management goals that incorporate successional diversity created by fire may support characteristic biodiversity, whereas current attempts to “restore” forests to open, low-severity fire conditions may not align with historical reference conditions in most ponderosa

  19. Examining historical and current mixed-severity fire regimes in ponderosa pine and mixed-conifer forests of western North America.

    PubMed

    Odion, Dennis C; Hanson, Chad T; Arsenault, André; Baker, William L; Dellasala, Dominick A; Hutto, Richard L; Klenner, Walt; Moritz, Max A; Sherriff, Rosemary L; Veblen, Thomas T; Williams, Mark A

    2014-01-01

    There is widespread concern that fire exclusion has led to an unprecedented threat of uncharacteristically severe fires in ponderosa pine (Pinus ponderosa Dougl. ex. Laws) and mixed-conifer forests of western North America. These extensive montane forests are considered to be adapted to a low/moderate-severity fire regime that maintained stands of relatively old trees. However, there is increasing recognition from landscape-scale assessments that, prior to any significant effects of fire exclusion, fires and forest structure were more variable in these forests. Biota in these forests are also dependent on the resources made available by higher-severity fire. A better understanding of historical fire regimes in the ponderosa pine and mixed-conifer forests of western North America is therefore needed to define reference conditions and help maintain characteristic ecological diversity of these systems. We compiled landscape-scale evidence of historical fire severity patterns in the ponderosa pine and mixed-conifer forests from published literature sources and stand ages available from the Forest Inventory and Analysis program in the USA. The consensus from this evidence is that the traditional reference conditions of low-severity fire regimes are inaccurate for most forests of western North America. Instead, most forests appear to have been characterized by mixed-severity fire that included ecologically significant amounts of weather-driven, high-severity fire. Diverse forests in different stages of succession, with a high proportion in relatively young stages, occurred prior to fire exclusion. Over the past century, successional diversity created by fire decreased. Our findings suggest that ecological management goals that incorporate successional diversity created by fire may support characteristic biodiversity, whereas current attempts to "restore" forests to open, low-severity fire conditions may not align with historical reference conditions in most ponderosa

  20. Latent resilience in ponderosa pine forest: effects of resumed frequent fire.

    PubMed

    Larson, Andrew J; Belote, R Travis; Cansler, C Alina; Parks, Sean A; Dietz, Matthew S

    2013-09-01

    Ecological systems often exhibit resilient states that are maintained through negative feedbacks. In ponderosa pine forests, fire historically represented the negative feedback mechanism that maintained ecosystem resilience; fire exclusion reduced that resilience, predisposing the transition to an alternative ecosystem state upon reintroduction of fire. We evaluated the effects of reintroduced frequent wildfire in unlogged, fire-excluded, ponderosa pine forest in the Bob Marshall Wilderness, Montana, USA. Initial reintroduction of fire in 2003 reduced tree density and consumed surface fuels, but also stimulated establishment of a dense cohort of lodgepole pine, maintaining a trajectory toward an alternative state. Resumption of a frequent fire regime by a second fire in 2011 restored a low-density forest dominated by large-diameter ponderosa pine by eliminating many regenerating lodgepole pines and by continuing to remove surface fuels and small-diameter lodgepole pine and Douglas-fir that established during the fire suppression era. Our data demonstrate that some unlogged, fire-excluded, ponderosa pine forests possess latent resilience to reintroduced fire. A passive model of simply allowing lightning-ignited fires to burn appears to be a viable approach to restoration of such forests.

  1. The concept: Restoring ecological structure and process in ponderosa pine forests

    Treesearch

    Stephen F. Arno

    1996-01-01

    Elimination of the historic pattern of frequent low-intensity fires in ponderosa pine and pine-mixed conifer forests has resulted in major ecological disruptions. Prior to 1900, open stands of large, long-lived, fire-resistant ponderosa pine were typical. These were accompanied in some areas by other fire-dependent species such as western larch. Today, as a result of...

  2. Feeding response of Ips paraconfusus to phloem and phloem metabolites of Heterobasidion annosum-inoculated ponderosa pine, Pinus ponderosa.

    PubMed

    McNee, William R; Bonello, Pierluigi; Storer, Andrew J; Wood, David L; Gordon, Thomas R

    2003-05-01

    In studies of feeding by the bark beetle, Ips paraconfusus, two pine stilbenes (pinosylvin and pinosylvin methyl ether), ferulic acid glucoside, and enantiomers of the four most common sugars present in ponderosa pine phloem (sucrose, glucose, fructose, and raffinose) did not stimulate or reduce male feeding when assayed on wet alpha-cellulose with or without stimulatory phloem extractives present. When allowed to feed on wet alpha-cellulose containing sequential extracts (hexane, methanol, and water) of ponderosa pine phloem, methanol and water extractives stimulated feeding, but hexane extractives did not. Males confined in wet alpha-cellulose containing aqueous or organic extracts of culture broths derived from phloem tissue and containing the root pathogen. Heterobasidion annosum, ingested less substrate than beetles confined to control preparations. In an assay using logs from uninoculated ponderosa pines, the mean lengths of phloem in the digestive tracts increased as time spent feeding increased. Males confined to the phloem of basal logs cut from ponderosa pines artificially inoculated with H. annosum ingested significantly less phloem than beetles in logs cut from trees that were (combined) mock-inoculated or uninoculated and did not contain the pathogen. However, individual pathogen-containing treatments were not significantly different from uninoculated controls. It was concluded that altered feeding rates are not a major factor which may explain why diseased ponderosa pines are colonized by I. paraconfusus.

  3. Mountain pine beetle dynamics and reproductive success in post-fire lodgepole and ponderosa pine forests in northeastern Utah

    Treesearch

    Andrew P. Lerch; Jesse A. Pfammatter; Barbara J. Bentz; Kenneth F. Raffa

    2016-01-01

    Fire injury can increase tree susceptibility to some bark beetles (Curculionidae, Scolytinae), but whether wildfires can trigger outbreaks of species such as mountain pine beetle (Dendroctonus ponderosae Hopkins) is not well understood. We monitored 1173 lodgepole (Pinus contorta var. latifolia Doug.) and 599 ponderosa (Pinus ponderosa Doug. ex Law) pines for three...

  4. Financial results of ponderosa pine forest restoration in southwestern Colorado

    Treesearch

    Dennis L. Lynch

    2001-01-01

    From 1996 to 1998, the Ponderosa Pine Partnership conducted an experimental forest restoration project on 493 acres of small diameter ponderosa pine in the San Juan National Forest, Montezuma County, Colorado. The ecological basis and the financial analysis for this project are discussed. Specific financial results of the project including products sold, revenues...

  5. Competing vegetation in ponderosa pine plantations: ecology and control

    Treesearch

    Philip M. McDonald; Gary O. Fiddler

    1989-01-01

    Planted ponderosa pine (Pinus ponderosa Dougl. ex Laws. var. ponderosa) seedlings in young plantations in California are at a disadvantage compared with competing shrubs, forbs, and grasses. In many instances, roots of competing plants begin expanding and exploiting the soil earlier and in greater numbers, thereby capturing the...

  6. Interior ponderosa pine in the Black Hills

    Treesearch

    Charles E. Boldt; Robert R. Alexander; Milo J. Larson

    1983-01-01

    The gross area of the Black Hills of South Dakota and associated Bear Lodge Mountains of eastern Wyoming is about 3.5 million acres (1.4 million ha). Roughly half the area supports forest or woodland cover. Essentially pure stands of climax Rocky Mountain ponderosa pine (Pinus ponderosa var. scopulorum Engelm.) predominate on about...

  7. ROLE OF CARBOHYDRATE SUPPLY IN WHITE AND BROWN ROOT RESPIRATION OF PONDEROSA PINE

    EPA Science Inventory

    Respiratory responses of fine ponderosa pine (Pinus ponderosa Laws) roots of differing morphology were measured to evaluate response to excision and to changes in the shoot light environment. Ponderosa pine seedlings were subject to either a 15:9 h light/dark environment over 24...

  8. Managing Gambel oak in southwestern ponderosa pine forests: the status of our knowledge

    Treesearch

    Scott R. Abella

    2008-01-01

    Gambel oak (Quercus gambelii) is a key deciduous species in southwestern ponderosa pine (Pinus ponderosa) forests and is important for wildlife habitat, soil processes, and human values. This report (1) summarizes Gambel oak's biological characteristics and importance in ponderosa pine forests, (2) synthesizes literature on...

  9. Geographic variation in ponderosa pine leader growth

    Treesearch

    James W. Hanover

    1963-01-01

    Growth of the shoot apices of 91 trees in a 45-year-old Pinus ponderosa Laws. provenance test was measured periodically with a transit. Analysis of the measurements led to the following conclusions: (1) 19 races of ponderosa pine planted near Priest River, Idaho, showed phenological, morphological, or physiological variation in six characters: date of beginning growth...

  10. Stand structure in eastside old-growth ponderosa pine forests of Oregon and northern California.

    Treesearch

    Andrew Youngblood; Timothy Max; Kent Coe

    2004-01-01

    Quantitative metrics of horizontal and vertical structural attributes in eastside old-growth ponderosa pine (Pinus ponderosa P. and C. Lawson var. ponderosa) forests were measured to guide the design of restoration prescriptions. The age, size structure, and the spatial patterns were investigated in old-growth ponderosa pine forests at three...

  11. Fire effects on Gambel oak in southwestern ponderosa pine-oak forests

    Treesearch

    Scott R. Abella; Peter Z. Fulé

    2008-01-01

    Gambel oak (Quercus gambelii) is ecologically and aesthetically valuable in southwestern ponderosa pine (Pinus ponderosa) forests. Fire effects on Gambel oak are important because fire may be used in pine-oak forests to manage oak directly or to accomplish other management objectives. We used published literature to: (1) ascertain...

  12. Contributions of silvicultural studies at Fort Valley to watershed management of Arizona's ponderosa pine forests

    Treesearch

    Gerald J. Gottfried; Peter F. Ffolliott; Daniel G. Neary

    2008-01-01

    Watershed management and water yield augmentation have been important objectives for chaparral, ponderosa pine, and mixed conifer management in Arizona and New Mexico. The ponderosa pine forests and other vegetation types generally occur in relatively high precipitation zones where the potential for increased water yields is great. The ponderosa pine forests have been...

  13. Intertree competition in uneven-aged ponderosa pine stands

    Treesearch

    C.W. Woodall; C.E. Fiedler; K.S. Milner

    2003-01-01

    Intertree competition indices and effects were examined in 14 uneven-aged ponderosa pine (Pinus ponderosa var. scopulorum Engelm.) stands in eastern Montana. Location, height, diameter at breast height (DBH), basal area increment, crown ratio, and sapwood area were determined for each tree (DBH >3.8 cm) on one stem-mapped plot...

  14. Conserving genetic diversity in Ponderosa Pine ecosystem restoration

    Treesearch

    L.E. DeWald

    2017-01-01

    Restoration treatments in the ponderosa pine (Pinus ponderosa P. & C. Lawson) ecosystems of the southwestern United States often include removing over 80 percent of post-EuroAmerican settlement-aged trees to create healthier forest structural conditions. These types of stand density reductions can have negative effects on genetic diversity. Allozyme analyses...

  15. Efficacy of verbenone for protecting ponderosa pine stands from western pine beetle (Coleoptera: Curculionidae, Scolytinae) attack in California

    Treesearch

    C.J. Fettig; S.R. McKelvey; R.R. Borys; C.P Dabney; S.M. Hamud; L.J. Nelson; S.J. Seybold

    2009-01-01

    The western pine beetle, Dendroctonus brevicomis LeConte (Coleoptera: Curculionidae: Scolytinae), is a major cause of ponderosa pine, Pinus ponderosa Dougl. ex Laws., mortality in much of western North America. Currently, techniques for managing D. brevicomis infestations are limited. Verbenone (4,6,6-...

  16. Spatial patterns of ponderosa pine regeneration in high-severity burn patches

    Treesearch

    Suzanne M. Owen; Carolyn H. Sieg; Andrew J. Sanchez. Meador; Peter Z. Fule; Jose M. Iniguez; L. Scott. Baggett; Paula J. Fornwalt; Michael A. Battaglia

    2017-01-01

    Contemporary wildfires in southwestern US ponderosa pine forests can leave uncharacteristically large patches of tree mortality, raising concerns about the lack of seed-producing trees, which can prevent or significantly delay ponderosa pine regeneration. We established 4-ha plots in high-severity burn patches in two Arizona wildfires, the 2000 Pumpkin and 2002 Rodeo-...

  17. Role of fire in restoration of a ponderosa pine forest, Washington

    Treesearch

    Richy J. Harrod; Richard W. Fonda; Mara K. McGrath

    2007-01-01

    Ponderosa pine forests in the Eastern Cascades of Washington support dense, overstocked stands in which crown fires are probable, owing to postsettlement sheep grazing, logging, and fire exclusion. In 1991, the Okanogan-Wenatchee National Forests began to apply long-term management techniques to reverse postsettlement changes in ponderosa pine forests. For 9 years, the...

  18. INTERACTION OF GRASS COMPETITION AND OZONE STRESS ON C/N RATIO IN PONDEROSA PINE

    EPA Science Inventory

    Individual ponderosa pine (Pinus ponderosa Dougl. ex Laws.) seedlings were grown with three levels of blue wild-rye grass (Elymus glaucus Buckl.) (0,32, or 88 plants m-2) to determine if the presence of a natural competitor altered ponderosa pine seedling response to ozone. Gras...

  19. Influence of elevation on bark beetle community structure in ponderosa pine stands of northern Arizona

    Treesearch

    Andrew Miller; Kelly Barton; Joel McMillin; Tom DeGomez; Karen Clancy; John Anhold

    2008-01-01

    (Please note, this is an abstract only) Bark beetles killed more than 20 million ponderosa pine trees in Arizona during 2002-2004. Historically, bark beetle populations remained endemic and ponderosa pine mortality was limited to localized areas in Arizona. Consequently, there is a lack of information on bark beetle community structure in ponderosa pine stands of...

  20. Bugs in the system: development of tools to minimize ponderosa pine losses from western pine beetle infestations

    Treesearch

    Christopher J. Fettig

    2005-01-01

    The western pine beetle, Dendroctonus brevicomis LeConte, is a major cause of ponderosa pine, Pinus ponderosa Dougl. Ex Laws., mortality in the western USA and particularly in California. Under certain conditions, the beetle can aggressively attack and kill apparently healthy trees of all ages and size classes. The average loss is...

  1. Tree mortality in drought-stressed mixed-conifer and ponderosa pine forests, Arizona, USA

    Treesearch

    Joseph L. Ganey; Scott C. Vojta

    2011-01-01

    We monitored tree mortality in northern Arizona (USA) mixed-conifer and ponderosa pine (Pinus ponderosa Dougl. ex Laws) forests from 1997 to 2007, a period of severe drought in this area. Mortality was pervasive, occurring on 100 and 98% of 53 mixed-conifer and 60 ponderosa pine plots (1-ha each), respectively. Most mortality was attributable to a suite of forest...

  2. Changes in Gambel oak densities in southwestern ponderosa pine forests since Euro-American settlement

    Treesearch

    Scott R. Abella; Peter Z. Fulé

    2008-01-01

    Densities of small-diameter ponderosa pine (Pinus ponderosa) trees have increased in southwestern ponderosa pine forests during a period of fire exclusion since Euro-American settlement in the late 1800s. However, less well known are potential changes in Gambel oak (Quercus gambelii) densities during this period in these forests....

  3. Lumber recovery from ponderosa pine in northern California.

    Treesearch

    Susan Ernst; Pong W.Y.

    1985-01-01

    Lumber recovery information from 942 logs from old- and young-growth ponderosa pine (Pinus ponderosa Dougl. ex Laws.) trees in northern California is presented. More than 58 percent of the lumber volume was found in 5/4 Shop, Moulding, and Select grades. About 25 percent of the total lumber volume was Moulding, and 24 percent was Standard and...

  4. Fungal endophytes in woody roots of Douglas-fir (Pseudotsuga menziesii) and ponderosa pine (Pinus ponderosa)

    Treesearch

    J. A. Hoff; Ned B. Klopfenstein; Geral I. McDonald; Jonalea R. Tonn; Mee-Sook Kim; Paul J. Zambino; Paul F. Hessburg; J. D. Rodgers; T. L. Peever; L. M. Carris

    2004-01-01

    The fungal community inhabiting large woody roots of healthy conifers has not been well documented. To provide more information about such communities, a survey was conducted using increment cores from the woody roots of symptomless Douglas-fir (Pseudotsuga menziesii) and ponderosa pine (Pinus ponderosa) growing in dry forests...

  5. Mountain pine beetle host selection between lodgepole and ponderosa pines in the southern Rocky Mountains

    USGS Publications Warehouse

    West, Daniel R.; Briggs, Jenny S.; Jacobi, William R.; Negron, Jose F.

    2016-01-01

    Recent evidence of range expansion and host transition by mountain pine beetle ( Dendroctonus ponderosae Hopkins; MPB) has suggested that MPB may not primarily breed in their natal host, but will switch hosts to an alternate tree species. As MPB populations expanded in lodgepole pine forests in the southern Rocky Mountains, we investigated the potential for movement into adjacent ponderosa pine forests. We conducted field and laboratory experiments to evaluate four aspects of MPB population dynamics and host selection behavior in the two hosts: emergence timing, sex ratios, host choice, and reproductive success. We found that peak MPB emergence from both hosts occurred simultaneously between late July and early August, and the sex ratio of emerging beetles did not differ between hosts. In two direct tests of MPB host selection, we identified a strong preference by MPB for ponderosa versus lodgepole pine. At field sites, we captured naturally emerging beetles from both natal hosts in choice arenas containing logs of both species. In the laboratory, we offered sections of bark and phloem from both species to individual insects in bioassays. In both tests, insects infested ponderosa over lodgepole pine at a ratio of almost 2:1, regardless of natal host species. Reproductive success (offspring/female) was similar in colonized logs of both hosts. Overall, our findings suggest that MPB may exhibit equally high rates of infestation and fecundity in an alternate host under favorable conditions.

  6. The Chilling Optimum of Idaho and Arizona Ponderosa Pine Buds

    Treesearch

    David L. Wenny; Daniel J. Swanson; R. Kasten Dumroese

    2002-01-01

    Ponderosa pine (Pinus ponderosa) seedlings from Idaho (var. ponderosa) and Arizona (var. scopulorum) grown in a container nursery received optimum chilling [2,010 hr (84 days) of temperatures below 5°C]. While seedlings were in the greenhouse, days required for 50% of the population to break bud were similar for both seed sources...

  7. Lessons learned from prescribed fire in ponderosa pine forests of the southern Sierra Nevada

    Treesearch

    Karen E. Bagne; Kathryn L. Purcell

    2009-01-01

    Prescribed fire is a commonly used management tool in fire-suppressed ponderosa pine (Pinus ponderosa) forests, but effects of these fires on birds are largely unstudied. We investigated both direct and indirect impacts on breeding birds in ponderosa pine forests of the southern Sierra Nevada where fires were applied in the spring. Following...

  8. Harvesting costs for management planning for ponderosa pine plantations.

    Treesearch

    Roger D. Fight; Alex Gicqueau; Bruce R. Hartsough

    1999-01-01

    The PPHARVST computer application is Windows-based, public-domain software used to estimate harvesting costs for management planning for ponderosa pine (Pinus ponderosa Dougl. ex Laws.) plantations. The equipment production rates were developed from existing studies. Equipment cost rates were based on 1996 prices for new...

  9. Insight into the hydraulics and resilience of Ponderosa pine seedlings using a mechanistic ecohydrologic model

    NASA Astrophysics Data System (ADS)

    Maneta, M. P.; Simeone, C.; Dobrowski, S.; Holden, Z.; Sapes, G.; Sala, A.; Begueria, S.

    2017-12-01

    In semiarid regions, drought-induced seedling mortality is considered to be caused by failure in the tree hydraulic column. Understanding the mechanisms that cause hydraulic failure and death in seedlings is important, among other things, to diagnose where some tree species may fail to regenerate, triggering demographic imbalances in the forest that could result in climate-driven shifts of tree species. Ponderosa pine is a common lower tree line species in the western US. Seedlings of ponderosa pine are often subject to low soil water potentials, which require lower water potentials in the xylem and leaves to maintain the negative pressure gradient that drives water upward. The resilience of the hydraulic column to hydraulic tension is species dependent, but from greenhouse experiments, we have identified general tension thresholds beyond which loss of xylem conductivity becomes critical, and mortality in Ponderosa pine seedlings start to occur. We describe this hydraulic behavior of plants using a mechanistic soil-vegetation-atmosphere transfer model. Before we use this models to understand water-stress induced seedling mortality at the landscape scale, we perform a modeling analysis of the dynamics of soil moisture, transpiration, leaf water potential and loss of plant water conductivity using detailed data from our green house experiments. The analysis is done using a spatially distributed model that simulates water fluxes, energy exchanges and water potentials in the soil-vegetation-atmosphere continuum. Plant hydraulic and physiological parameters of this model were calibrated using Monte Carlo methods against information on soil moisture, soil hydraulic potential, transpiration, leaf water potential and percent loss of conductivity in the xylem. This analysis permits us to construct a full portrait of the parameter space for Ponderosa pine seedling and generate posterior predictive distributions of tree response to understand the sensitivity of transpiration

  10. Classification tree and minimum-volume ellipsoid analyses of the distribution of ponderosa pine in the western USA

    USGS Publications Warehouse

    Norris, Jodi R.; Jackson, Stephen T.; Betancourt, Julio L.

    2006-01-01

    Aim? Ponderosa pine (Pinus ponderosa Douglas ex Lawson & C. Lawson) is an economically and ecologically important conifer that has a wide geographic range in the western USA, but is mostly absent from the geographic centre of its distribution - the Great Basin and adjoining mountain ranges. Much of its modern range was achieved by migration of geographically distinct Sierra Nevada (P. ponderosa var. ponderosa) and Rocky Mountain (P. ponderosa var. scopulorum) varieties in the last 10,000 years. Previous research has confirmed genetic differences between the two varieties, and measurable genetic exchange occurs where their ranges now overlap in western Montana. A variety of approaches in bioclimatic modelling is required to explore the ecological differences between these varieties and their implications for historical biogeography and impending changes in western landscapes. Location? Western USA. Methods? We used a classification tree analysis and a minimum-volume ellipsoid as models to explain the broad patterns of distribution of ponderosa pine in modern environments using climatic and edaphic variables. Most biogeographical modelling assumes that the target group represents a single, ecologically uniform taxonomic population. Classification tree analysis does not require this assumption because it allows the creation of pathways that predict multiple positive and negative outcomes. Thus, classification tree analysis can be used to test the ecological uniformity of the species. In addition, a multidimensional ellipsoid was constructed to describe the niche of each variety of ponderosa pine, and distances from the niche were calculated and mapped on a 4-km grid for each ecological variable. Results? The resulting classification tree identified three dominant pathways predicting ponderosa pine presence. Two of these three pathways correspond roughly to the distribution of var. ponderosa, and the third pathway generally corresponds to the distribution of var

  11. Fire hazard from precommercial thinning of ponderosa pine.

    Treesearch

    George R. Fahnestock

    1968-01-01

    Precommercial thinning lately has become a major feature in management of ponderosa pine (Pinus ponderosa Laws.) on the National Forests in Oregon and Washington. Nearly 47,000 acres were thinned in 1966, up from 9,196 in 1959; and the upward trend appears certain to continue. Current practice is to cut the trees with a powersaw about a foot above...

  12. Lumber recovery from ponderosa pine in western Montana.

    Treesearch

    Marlin E. Plank

    1982-01-01

    Lumber grade yields and recovery ratios are shown for a sample of ponderosa pine (Pinus ponderosa Dougl. ex Laws.) in western Montana. About 9 percent of the lumber produced was in Select grades, 48 percent in Shop grades, and 43 percent in Common grades. Information on log scale and yield is presented in tables by log grade and diameter class....

  13. Monoterpene emission from ponderosa pine

    NASA Technical Reports Server (NTRS)

    Lerdau, Manual; Dilts, Stephen B.; Westberg, Hal; Lamb, Brian K.; Allwine, Eugene J.

    1994-01-01

    We explore the variability in monoterpene emissions from ponderosa pine beyond that which can be explained by temperature alone. Specifically, we examine the roles that photosynthesis and needle monoterpene concentrations play in controlling emissions. We measure monoterpene concentrations and emissions, photosynthesis, temperature, and light availability in the late spring and late summer in a ponderosa pine forest in central Oregon. We use a combination of measurements from cuvettes and Teflon bag enclosures to show that photosynthesis is not correlated with emissions in the short term. We also show that needle monoterpene concentrations are highly correlated with emissions for two compounds, alpha-pinene and beta-pinene, but that Delta-carene concentrations are not correlated with emissions. We suggest that direct effects of light and photosynthesis do not need to be included in emission algorithms. Our results indicate that the role of needle concentration bears further investigation; our results for alpha-pinene and beta-pinene are explainable by a Raoult's law relationship, but we cannot yet explain the cause of our results with Delta-carene.

  14. Effect of pruning on growth of ponderosa pine.

    Treesearch

    Edwin L. Mowat

    1947-01-01

    A study of the influence of pruning various proportions of the crowns of young ponderosa pine trees upon growth, vigor, and mortality was started on the Pringle Falls Experimental Forest in 1941. After 5 years of tree growth, the study throws some light on how severely young pines should be pruned.

  15. Ponderosa pine reproduction in relation to seed supply at Challenge Experimental Forest

    Treesearch

    Edward S. Corbett

    1962-01-01

    Seed trees were selected in young-growth ponderosa pine to produce an estimated 25,000, 50,000, and 100,000 seeds per acre. A year after a good seed crop, ponderosa pine stocking in cutover plots was: 1,910, 4,020, and 4,820 seedlings per acre. Some additional regeneration of other species also occurred.

  16. Brush reduces growth of thinned ponderosa pine in northern California

    Treesearch

    William W. Oliver

    1984-01-01

    The effects of tree spacing and brush competition were evaluated on a ponderosa pine (Pinus ponderosa Dougl. ex Laws. var. ponderosa) site of low productivity in California's North Coast Range. Eleven-year-old saplings were thinned to square spacings of 2.1, 2.4, 3.0, and 4.3 m (7, 8, 10, and 14 ft), and all, half, and none of...

  17. Nuclear genetic variation across the range of ponderosa pine (Pinus ponderosa): Phylogeographic, taxonomic and conservation implications

    Treesearch

    Kevin M. Potter; Valerie D. Hipkins; Mary F. Mahalovich; Robert E. Means

    2015-01-01

    Ponderosa pine (Pinus ponderosa) is among the most broadly distributed conifer species of western North America, where it possesses considerable ecological, esthetic, and commercial value. It exhibits complicated patterns of morphological and genetic variation, suggesting that it may be in the process of differentiating into distinct regional...

  18. Two Case Histories For Using Prescribed Fire to Restore Ponderosa Pine Ecosystems in Northern Arizona

    Treesearch

    Stephen S. Sackett; Sally M. Haase

    1998-01-01

    Historic observations and research indicate that the ponderosa pine (Pinus ponderosa) ecosystem in the southwestern U.S. is now very different compared to pre-European settlement. Timber harvest, livestock grazing, and fire suppression have transformed an open ponderosa pine-bunch grass community into a dense forest overloaded with flammable...

  19. Mountain Pine Beetle Host Selection Between Lodgepole and Ponderosa Pines in the Southern Rocky Mountains.

    PubMed

    West, Daniel R; Briggs, Jennifer S; Jacobi, William R; Negrón, José F

    2016-02-01

    Recent evidence of range expansion and host transition by mountain pine beetle (Dendroctonus ponderosae Hopkins; MPB) has suggested that MPB may not primarily breed in their natal host, but will switch hosts to an alternate tree species. As MPB populations expanded in lodgepole pine forests in the southern Rocky Mountains, we investigated the potential for movement into adjacent ponderosa pine forests. We conducted field and laboratory experiments to evaluate four aspects of MPB population dynamics and host selection behavior in the two hosts: emergence timing, sex ratios, host choice, and reproductive success. We found that peak MPB emergence from both hosts occurred simultaneously between late July and early August, and the sex ratio of emerging beetles did not differ between hosts. In two direct tests of MPB host selection, we identified a strong preference by MPB for ponderosa versus lodgepole pine. At field sites, we captured naturally emerging beetles from both natal hosts in choice arenas containing logs of both species. In the laboratory, we offered sections of bark and phloem from both species to individual insects in bioassays. In both tests, insects infested ponderosa over lodgepole pine at a ratio of almost 2:1, regardless of natal host species. Reproductive success (offspring/female) was similar in colonized logs of both hosts. Overall, our findings suggest that MPB may exhibit equally high rates of infestation and fecundity in an alternate host under favorable conditions. © The Authors 2015. Published by Oxford University Press on behalf of Entomological Society of America. All rights reserved. For Permissions, please email: journals.permissions@oup.com.

  20. Mountain pine beetle, Dendroctonus ponderosae (Coleoptera: Curculionidae, Scolytinae)

    Treesearch

    Barbara Bentz

    2008-01-01

    The mountain pine beetle, Dendroctonus ponderosae Hopkins, is considered one of the most economically important insect species in coniferous forests of western North America. Adult beetles are capable of successfully reproducing in at least 12 North American species of Pinus (Pineacea) from southern British Columbia to northern Baja Mexico. Mountain pine beetle adults...

  1. Mountain pine beetle attack in ponderosa pine: Comparing methods for rating susceptibility

    Treesearch

    David C. Chojnacky; Barbara J. Bentz; Jesse A. Logan

    2000-01-01

    Two empirical methods for rating susceptibility of mountain pine beetle attack in ponderosa pine were evaluated. The methods were compared to stand data modeled to objectively rate each sampled stand for susceptibly to bark-beetle attack. Data on bark-beetle attacks, from a survey of 45 sites throughout the Colorado Plateau, were modeled using logistic regression to...

  2. Contributions of silvicultural studies at Fort Valley to watershed management of Arizona's ponderosa pine forests (P-53)

    Treesearch

    Gerald J. Gottfried; Peter F. Ffolliott; Daniel G. Neary

    2008-01-01

    Watershed management and water yield augmentation have been important objectives for chaparral, ponderosa pine, and mixed conifer management in Arizona and New Mexico. The ponderosa pine forests and other vegetation types generally occur in relatively high precipitation zones where the potential for increased water yields is great. The ponderosa pine forests have been...

  3. Stand density index in uneven-aged ponderosa pine stands

    Treesearch

    C.W. Woodall; C.E. Fiedler; K.S. Milner

    2003-01-01

    Stand density index (SDI) was developed to quantify relative stand density in even-aged stands. Application of SDI in uneven-aged stands has been described mathematically but not justified biologically. Diameter-class trends in SDI and sapwood area across 14 uneven-aged ponderosa pine (Pinus ponderosa Dougl. ex P. & C. Laws.) stands in eastern...

  4. Dwarf Mistletoe of Ponderosa Pine in the Southwest (FIDL)

    Treesearch

    Paul C. Lightle; Melvyn J. Weiss

    1974-01-01

    Southwestern dwarf mistletoe (Arceuthobuim vaginatum subsp. cryptopodum) occurs essentially throughout the range of ponderosa pine (Pinus ponderosa var. scopulorum) from northern Mexico through western Texas, Arizona, and New Mexico into Colorado and central Utah. In Arizona and New Mexico it is present on more than one-third of the commercial forest acreage and is...

  5. Blue wild-rye grass competition increases the effect of ozone on ponderosa pine seedlings.

    PubMed

    Andersen, C P; Hogsett, W E; Plocher, M; Rodecap, K; Lee, E H

    2001-03-01

    Individual ponderosa pine (Pinus ponderosa Dougl. ex Laws.) seedlings were grown in mesocosms with three densities of blue wild-rye grass (Elymus glaucus Buckl.) (equivalent to 0, 32 or 88 plants m-2) to determine if the presence of a natural competitor alters the response of ponderosa pine seedlings to ozone. After 3 years of ozone exposure, grass presence reduced total ponderosa pine dry mass by nearly 50%, whereas ozone alone had no significant effect on ponderosa pine growth. The combination of ozone and grass further reduced needle, stem and branch dry mass significantly below that induced by grass competition alone. Root:shoot ratios increased in response to the combined grass and ozone treatments. Grass competition significantly reduced soluble sugar concentrations in all ponderosa pine tissue components examined. Starch concentrations were highly variable but did not differ significantly between treatments. Ozone significantly reduced soluble sugar concentrations in fine roots and stems. In the absence of grass, ozone-treated seedlings tended to have higher tissue N concentrations than controls. In the presence of grass, ozone-treated seedlings had lower N concentrations than controls, resulting in a significant interaction between these two stresses in 1- and 2-year-old needles. Needle C:N ratios decreased in response to grass competition, as a result of increased N concentration and no change in C concentration. The opposite response was observed in ozone-treated seedlings as a result of decreased N concentrations, indicating that ozone-treated seedlings were unable to take up or retain as much nitrogen when grown in the presence of grass. We conclude that ponderosa pine seedlings are more susceptible to ozone when grown in competition with blue wild-rye grass.

  6. Management of ponderosa pine in the Southwest: As developed by research and experimental practice

    Treesearch

    G. A. Pearson

    1950-01-01

    Ponderosa pine (Pinus ponderosa) is the most widely distributed conifer in North America, and one of the most valuable. Commercial stands of the species are found in all of the 15 States which lie wholly or in part west of the 102d merinian, and in all but one it rank among the most important lumber producers. In the Southwest, ponderosa pine is of particular...

  7. Silvicultural systems for managing ponderosa pine

    Treesearch

    Andrew Youngblood

    2005-01-01

    Silviculturists have primarily relied on classical even-aged silvicultural systems (the planned series of treatments for tending, harvesting, and re-establishing a stand) for ponderosa pine, with uneven-aged systems used to a lesser degree. Current management practices involve greater innovation because of conflicting management objectives. Silvicultural systems used...

  8. Comparing growth of ponderosa pine in two growing media

    Treesearch

    R. Kasten Dumroese

    2009-01-01

    I compared growth of container ponderosa pine (Pinus ponderosa) seedlings grown in a 1:1 (v:v) Sphagnum peat moss:coarse vermiculite medium (P:V) and a 7:3 (v:v) Sphagnum peat moss:Douglas-fir sawdust medium (P:S) at three different irrigation regimes. By using exponential fertilization techniques, I was able to supply seedlings with similar amounts...

  9. Limited response of ponderosa pine bole defenses to wounding and fungi.

    PubMed

    Gaylord, Monica L; Hofstetter, Richard W; Kolb, Thomas E; Wagner, Michael R

    2011-04-01

    Tree defense against bark beetles (Curculionidae: Scolytinae) and their associated fungi generally comprises some combination of constitutive (primary) and induced (secondary) defenses. In pines, the primary constitutive defense against bark beetles consists of preformed resin stored in resin ducts. Induced defenses at the wound site (point of beetle entry) in pines may consist of an increase in resin flow and necrotic lesion formation. The quantity and quality of both induced and constitutive defenses can vary by species and season. The inducible defense response in ponderosa pine is not well understood. Our study examined the inducible defense response in ponderosa pine using traumatic mechanical wounding, and wounding with and without fungal inoculations with two different bark beetle-associated fungi (Ophiostoma minus and Grosmannia clavigera). Resin flow did not significantly increase in response to any treatment. In addition, necrotic lesion formation on the bole after fungal inoculation was minimal. Stand thinning, which has been shown to increase water availability, had no, or inconsistent, effects on inducible tree defense. Our results suggest that ponderosa pine bole defense against bark beetles and their associated fungi is primarily constitutive and not induced.

  10. Fall rates of prescribed fire-killed ponderosa pine. Forest Service research paper

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Harrington, M.G.

    1996-05-01

    Fall rates of prescribed fire-killed ponderosa pine were evaluated relative to tree and fire damage characteristics. High crown scorch and short survival time after fire injury were factors leading to a high probability of early tree fall. The role of chemical defense mechanisms is discussed. Results apply to prescribed-fire injured, second-growth ponderosa pine less than 16 inches diameter at breast height.

  11. Concentrations and deposition of nitrogenous air pollutants in a ponderosa/Jeffrey pine canopy

    Treesearch

    Andrzej Bytnerowicz; Mark E. Fenn; Michael J. Arbaugh

    1998-01-01

    Nitrogenous (N) air pollutant concentrations and surface deposition of nitrate (NO3-) and ammonium (NH4+) to branches of ponderosa pine (Pinus ponderosa Dougl. ex. Laws.) seedlings were measured on a vertical transect in a mature ponderosa/Jeffrey (...

  12. Lumber grade recovery from young ponderosa pine.

    Treesearch

    James E. Sowder

    1953-01-01

    Young ponderosa pine produces a good grade of common lumber, and close-grown trees produce better grades than those which were open-grown. This was shown by a study at the Pringle Falls Experimental Forest near Lapine in central Oregon.

  13. Identification and ecology of old ponderosa pine trees in the Colorado Front Range

    Treesearch

    Laurie Stroh Huckaby; Merrill R. Kaufmann; Paula J. Fornwalt; Jason M. Stoker; Chuck Dennis

    2003-01-01

    We describe the distinguishing physical characteristics of old ponderosa pine trees in the Front Range of Colorado, the processes that tend to preserve them, their past and present ecological significance, and their role in ecosystem restoration. Photographs illustrate identifying features of old ponderosa pines and show how to differentiate them from mature and young...

  14. Exploring climate niches of ponderosa pine (Pinus ponderosa Douglas ex Lawson) haplotypes in the western United States: implications for evolutionary history and conservation

    Treesearch

    Douglas J. Shinneman; Robert E. Means; Kevin M. Potter; Valerie D. Hipkins; Tzen-Yuh Chiang

    2016-01-01

    Ponderosa pine (Pinus ponderosa Douglas ex Lawson) occupies montane environments throughout western North America, where it is both an ecologically and economically important tree species. A recent study using mitochondrial DNA analysis demonstrated substantial genetic variation among ponderosa pine populations in the western U.S., identifying 10 haplotypes with unique...

  15. Long-term thinning alters ponderosa pine reproduction in northern Arizona

    Treesearch

    Kelsey N. Flathers; Thomas E. Kolb; John B. Bradford; Kristen M. Waring; W. Keith Moser

    2016-01-01

    The future of ponderosa pine (Pinus ponderosa var. scopulorum) forests in the southwestern United States is uncertain because climate-change-induced stresses are expected to increase tree mortality and place greater constraints on regeneration. Silvicultural treatments, which include thinning, are increasingly being used to address forest health concerns by...

  16. Ectomycorrhizal communities of ponderosa pine and lodgepole pine in the south-central Oregon pumice zone.

    PubMed

    Garcia, Maria O; Smith, Jane E; Luoma, Daniel L; Jones, Melanie D

    2016-05-01

    Forest ecosystems of the Pacific Northwest of the USA are changing as a result of climate change. Specifically, rise of global temperatures, decline of winter precipitation, earlier loss of snowpack, and increased summer drought are altering the range of Pinus contorta. Simultaneously, flux in environmental conditions within the historic P. contorta range may facilitate the encroachment of P. ponderosa into P. contorta territory. Furthermore, successful pine species migration may be constrained by the distribution or co-migration of ectomycorrhizal fungi (EMF). Knowledge of the linkages among soil fungal diversity, community structure, and environmental factors is critical to understanding the organization and stability of pine ecosystems. The objectives of this study were to establish a foundational knowledge of the EMF communities of P. ponderosa and P. contorta in the Deschutes National Forest, OR, USA, and to examine soil characteristics associated with community composition. We examined EMF root tips of P. ponderosa and P. contorta in soil cores and conducted soil chemistry analysis for P. ponderosa cores. Results indicate that Cenococcum geophilum, Rhizopogon salebrosus, and Inocybe flocculosa were dominant in both P. contorta and P. ponderosa soil cores. Rhizopogon spp. were ubiquitous in P. ponderosa cores. There was no significant difference in the species composition of EMF communities of P. ponderosa and P. contorta. Ordination analysis of P. ponderosa soils suggested that soil pH, plant-available phosphorus (Bray), total phosphorus (P), carbon (C), mineralizable nitrogen (N), ammonium (NH4), and nitrate (NO3) are driving EMF community composition in P. ponderosa stands. We found a significant linear relationship between EMF species richness and mineralizable N. In conclusion, P. ponderosa and P. contorta, within the Deschutes National Forest, share the same dominant EMF species, which implies that P. ponderosa may be able to successfully establish

  17. Long-term benefits to the growth of ponderosa pines from controlling southwestern pine tip moth (Lepidoptera: Tortricidae) and weeds.

    PubMed

    Wagner, Michael R; Chen, Zhong

    2004-12-01

    The southwestern pine tip moth, Rhyacionia neomexicana (Dyar) (Lepidoptera: Tortricidae), is a native forest pest that attacks seedlings and saplings of ponderosa pine, Pinus ponderosa Dougl. ex Laws, in the southwestern United States. Repeated attacks can cause severe deformation of host trees and significant long-term growth loss. Alternatively, effective control of R. neomexicana, vegetative competition, or both in young pine plantations may increase survival and growth of trees for many years after treatments are applied. We test the null hypothesis that 4 yr of R. neomexicana and weed control with insecticide, weeding, and insecticide plus weeding would not have any residual effect on survival and growth of trees in ponderosa pine plantation in northern Arizona 14 yr post-treatment, when the trees were 18 yr old. Both insecticide and weeding treatment increased tree growth and reduced the incidence of southwestern pine tip moth damage compared with the control. However, weeding alone also significantly increased tree survival, whereas insecticide alone did not. The insecticide plus weeding treatment had the greatest tree growth and survival, and the lowest rate of tip moth damage. Based on these results, we rejected our null hypothesis and concluded that there were detectable increases in the survival and growth of ponderosa pines 14 yr after treatments applied to control R. neomexicana and weeds.

  18. Historical and contemporary lessons from ponderosa pine genetic studies at the Fort Valley Experimental Forest, Arizona

    Treesearch

    Laura E. DeWald; Mary Frances Mahalovich

    2008-01-01

    Forest management will protect genetic integrity of tree species only if their genetic diversity is understood and considered in decision-making. Genetic knowledge is particularly important for species such as ponderosa pine (Pinus ponderosa Dougl. ex Laws.) that are distributed across wide geographic distances and types of climates. A ponderosa pine...

  19. Ponderosa pine lumber recovery in north-central Washington.

    Treesearch

    E.H. Clarke

    1961-01-01

    Prior to World War 11, the U. S. Forest Service (Region 6) adopted the policy of appraising ponderosa pine timber with a standardized set of lumber grade recovery data obtained from representative pine mills which are known to use average care in manufacturing and marketing. Such data were derived by combining the results of several mill studies made about 20 to 25...

  20. Optimum stand prescriptions for ponderosa pine

    Treesearch

    David W. Hann; J. Douglas Brodie; Kurt H. Riitters

    1983-01-01

    Two examples for a northern Arizona ponderosa pine stand illustrate the usefulness of dynamic programming in making silvicultural decisions. The first example analyzes the optimal planting density for bare land, while the second examines the optimal precommercial thinning intensity for a 43-year-old stand. Bot hexamples assume that the manager's primary objective...

  1. Soil compaction and initial height growth of planted ponderosa pine.

    Treesearch

    P. H. Cochran; Terry. Brock

    1985-01-01

    Early height growth of ponderosa pine (Pinus ponderosa Dougl. ex Laws.) seedlings planted in clearcuts in central Oregon was negatively correlated with increasing soil bulk density. Change in bulk density accounted for less than half the total variation in height growth. Although many other factors affect the development of seedlings, compaction...

  2. Estimating cubic volume of small diameter tree-length logs from ponderosa and lodgepole pine.

    Treesearch

    Marlin E. Plank; James M. Cahill

    1984-01-01

    A sample of 351 ponderosa pine (Pinus ponderosa Dougl. ex Laws.) and 509 lodgepole pine (Pinus contorta Dougl. ex Loud.) logs were used to evaluate the performance of three commonly used formulas for estimating cubic volume. Smalian's formula, Bruce's formula, and Huber's formula were tested to determine which...

  3. Habitat of birds in ponderosa pine and aspen/birch forest in the Black Hills, South Dakota

    Treesearch

    Todd R. Mills; Mark A. Rumble; Lester D. Flake

    2000-01-01

    Birds with both eastern and western distributions occur in the Black Hills of western South Dakota. This forest is mostly ponderosa pine (Pinus ponderosa) and is managed for timber. Logging alters forest characteristics and the bird community. We studied habitat relations of breeding songbirds at the stand- and site-level scales in ponderosa pine and...

  4. Fertilization and spacing effects on growth of planted ponderosa pine.

    Treesearch

    P.H. Cochran; R.P. Newman; James W. Barrett

    1991-01-01

    Fertilizer placed in the planting hole increased height growth of ponderosa pine (Pinus ponderosa Dougl. ex Laws.) early in the life of the plantation. Later broadcast applications of fertilizer may have had little effect on growth. Wider spacings produced larger trees but less volume per acre than narrower spacings after average tree height...

  5. Emergence of Buprestidae, Cerambycidae, and Scolytinae (Coleoptera) from mountain pine beetle-killed and fire-killed ponderosa pines in the Black Hills, South Dakota, USA

    Treesearch

    Sheryl L. Costello; William R. Jacobi; Jose F. Negron

    2013-01-01

    Wood borers (Coleoptera: Cerambycidae and Buprestidae) and bark beetles (Coleoptera: Curculionidae) infest ponderosa pines, Pinus ponderosa P. Lawson and C. Lawson, killed by mountain pine beetle (MPB), Dendroctonus ponderosae Hopkins, and fire. No data is available comparing wood borer and bark beetle densities or species guilds associated with MPB-killed or fire-...

  6. Growth and mortality of ponderosa pine poles thinned to various densities in the Blue Mountains of Oregon.

    Treesearch

    P.H. Cochran; James W. Barrett

    1995-01-01

    Growth and mortality in a ponderosa pine (Pinus ponderosa Dougl. ex Laws.) stand were investigated for 24 years. High mortality rates from mountain pine beetle (Dendroctonus ponderosae Hopkins) occurred on some plots where values for stand density index exceeded 140. Periodic annual increments for quadratic mean diameters...

  7. Guide to understory burning in ponderosa pine-larch-fir forests in the Intermountain West

    Treesearch

    Bruce M. Kilgore; George A. Curtis

    1987-01-01

    Summarizes the objectives, prescriptions, and techniques used in prescribed burning beneath the canopy of ponderosa pine stands, and stands of ponderosa pine mixed with western larch, Douglas-fir, and grand fir. Information was derived from 12 districts in two USDA Forest Service Regions and seven National Forests in Montana and Oregon.

  8. Tradeoffs in overstory and understory aboveground net primary productivity in southwestern ponderosa pine stands

    Treesearch

    Kyla E. Sabo; Stephen C. Hart; Carolyn Hull Sieg; John Duff Bailey

    2008-01-01

    Previous studies in ponderosa pine forests have quantified the relationship between overstory stand characteristics and understory production using tree measurements such as basal area. We built on these past studies by evaluating the tradeoff between overstory and understory aboveground net primary productivity (ANPP) in southwestern ponderosa pine forests at the...

  9. Mountain pine beetle-killed trees as snags in Black Hills ponderosa pine stands

    Treesearch

    J. M. Schmid; S. A. Mata; W. C. Schaupp

    2009-01-01

    Mountain pine beetle-killed ponderosa pine trees in three stands of different stocking levels near Bear Mountain in the Black Hills National Forest were surveyed over a 5-year period to determine how long they persisted as unbroken snags. Rate of breakage varied during the first 5 years after MPB infestation: only one tree broke during the first 2 years in the three...

  10. Stripcut-thinning of ponderosa pine stands: An Arizona case study

    Treesearch

    Peter F. Ffolliott; Malchus Baker

    2001-01-01

    Growth and structural changes in ponderosa pine (Pinus ponderosa) stands were studied over a 25-year posttreatment period to determine the impacts of a combined stripcut-thinning treatment. Trees on one-third of a watershed in north-central Arizona had been removed in clear-cut strips. Trees in the "leave" strips were thinned. Number of...

  11. Season of prescribed burn in ponderosa pine forests in eastern Oregon: impact on pine mortality.

    Treesearch

    Walter G. Thies; Douglas J. Westlind; Mark Loewen

    2005-01-01

    A study of the effects of season of prescribed burn on tree mortality was established in mixed-age ponderosa pine (Pinus ponderosa Dougl. ex Laws.) at the south end of the Blue Mountains near Burns, Oregon. Each of six previously thinned stands was subdivided into three experimental units and one of three treatments was randomly assigned to each:...

  12. Nonhost angiosperm volatiles and verbenone protect individual ponderosa pines from attack by western pine beetle and red turpentine beetle (Coleoptera: Curculionidae, Scolytinae)

    Treesearch

    Christopher J. Fettig; Christopher P. Dabney; Stepehen R. McKelvey; Dezene P.W. Huber

    2008-01-01

    Nonhost angiosperm volatiles (NAV) and verbenone were tested for their ability to protect individual ponderosa pines, Pinus ponderosa Dougl. ex. Laws., from attack by western pine beetle (WPB), Dendroctonus brevicomis LeConte, and red turpentine beetle (RTB), Dendroctonus valens LeConte (Coleoptera: Curculionidae, Scolytinae). A combination of (

  13. Performance of Ponderosa Pine on Bituminous Mine Spoils in Pennsylvania

    Treesearch

    Walter H. Davidson

    1977-01-01

    Seedlings from 40 seed sources of ponderosa pine (Pinus ponderosa Laws.) were planted on a strip-mine spoil in central Pennsylvania in 1969. Survival of seedlings from different sources ranged from 23 to 90 percent after six growing seasons. The average height of the seedlings ranged from 67 to 140 cm for the same period. Eight sources produced...

  14. Self-fertility of a central Oregon source of ponderosa pine.

    Treesearch

    Frank C. Sorensen

    1970-01-01

    This report will describe the effect of self-, cross-, and open- or wind-pollination on seed and seedling production of 19 ponderosa pine (Pinus ponderosa Laws.) trees in the eastern foothills of the Cascade Mountains south of Bend, Oreg. The study is part of a continuing investigation of self-fertility in several conifers growing in the Pacific...

  15. Heavy thinning of ponderosa pine stands: An Arizona case study

    Treesearch

    Peter F. Ffolliott; Jr. Baker; Gerald J. Gottfried

    2000-01-01

    Growth and structural changes in a mosaic of even-aged ponderosa pine (Pinus ponderosa) stands were studied for 25 years to determine the long-term impacts of a heavy thinning treatment to a basal-area level of 25 ft2/acre. Basal area and volume growth of these stands has increased since thinning and likely will continue to...

  16. Silvicultural recommendations for the management of ponderosa pine forest

    Treesearch

    Martin Alfonso Mendoza Briseno; Mary Ann Fajvan; Juan Manuel Chacon Sotelo; Alejandro Velazquez Martinez; Antonio Quinonez. Silva

    2014-01-01

    Ponderosa pines are the most important timber producing species in Mexico, and they also represent a major portion of the Usa and Canada timber production. These pines form near pure stands with simple and stable stand structure. They suffer only occasional disturbances, and they sustain a limited capacity to hold biodiversity and other senvironmental services. The...

  17. Rapid Increases in Forest Understory Diversity and Productivity following a Mountain Pine Beetle (Dendroctonus ponderosae) Outbreak in Pine Forests

    PubMed Central

    Pec, Gregory J.; Karst, Justine; Sywenky, Alexandra N.; Cigan, Paul W.; Erbilgin, Nadir; Simard, Suzanne W.; Cahill, James F.

    2015-01-01

    The current unprecedented outbreak of mountain pine beetle (Dendroctonus ponderosae) in lodgepole pine (Pinus contorta) forests of western Canada has resulted in a landscape consisting of a mosaic of forest stands at different stages of mortality. Within forest stands, understory communities are the reservoir of the majority of plant species diversity and influence the composition of future forests in response to disturbance. Although changes to stand composition following beetle outbreaks are well documented, information on immediate responses of forest understory plant communities is limited. The objective of this study was to examine the effects of D. ponderosae-induced tree mortality on initial changes in diversity and productivity of understory plant communities. We established a total of 110 1-m2 plots across eleven mature lodgepole pine forests to measure changes in understory diversity and productivity as a function of tree mortality and below ground resource availability across multiple years. Overall, understory community diversity and productivity increased across the gradient of increased tree mortality. Richness of herbaceous perennials increased with tree mortality as well as soil moisture and nutrient levels. In contrast, the diversity of woody perennials did not change across the gradient of tree mortality. Understory vegetation, namely herbaceous perennials, showed an immediate response to improved growing conditions caused by increases in tree mortality. How this increased pulse in understory richness and productivity affects future forest trajectories in a novel system is unknown. PMID:25859663

  18. Rapid Increases in forest understory diversity and productivity following a mountain pine beetle (Dendroctonus ponderosae) outbreak in pine forests.

    PubMed

    Pec, Gregory J; Karst, Justine; Sywenky, Alexandra N; Cigan, Paul W; Erbilgin, Nadir; Simard, Suzanne W; Cahill, James F

    2015-01-01

    The current unprecedented outbreak of mountain pine beetle (Dendroctonus ponderosae) in lodgepole pine (Pinus contorta) forests of western Canada has resulted in a landscape consisting of a mosaic of forest stands at different stages of mortality. Within forest stands, understory communities are the reservoir of the majority of plant species diversity and influence the composition of future forests in response to disturbance. Although changes to stand composition following beetle outbreaks are well documented, information on immediate responses of forest understory plant communities is limited. The objective of this study was to examine the effects of D. ponderosae-induced tree mortality on initial changes in diversity and productivity of understory plant communities. We established a total of 110 1-m2 plots across eleven mature lodgepole pine forests to measure changes in understory diversity and productivity as a function of tree mortality and below ground resource availability across multiple years. Overall, understory community diversity and productivity increased across the gradient of increased tree mortality. Richness of herbaceous perennials increased with tree mortality as well as soil moisture and nutrient levels. In contrast, the diversity of woody perennials did not change across the gradient of tree mortality. Understory vegetation, namely herbaceous perennials, showed an immediate response to improved growing conditions caused by increases in tree mortality. How this increased pulse in understory richness and productivity affects future forest trajectories in a novel system is unknown.

  19. Solar treatments for reducing survival of mountain pine beetle in infested ponderosa and lodgepole pine logs

    Treesearch

    Jose F. Negron; Wayne A. Shepperd; Steve A. Mata; John B. Popp; Lance A. Asherin; Anna W. Schoettle; John M. Schmid; David A. Leatherman

    2001-01-01

    Three experiments were conducted to evaluate the use of solar radiation for reducing survival of mountain pine beetle populations in infested logs. Ponderosa pine logs were used in experiments 1 and 2 and lodgepole pine logs were used in experiment 3. Experiment 1 comprised three treatments: (1) one-layer solar treatment without plastic sheeting and logs rotated one-...

  20. Efficacy of verbenone for protecting ponderosa pine stands from western pine beetle (Coleoptera: Curculionidae: Scolytinae) attack in California.

    PubMed

    Fettig, Christopher J; McKelvey, Stephen R; Borys, Robert R; Dabney, Christopher P; Hamud, Shakeeb M; Nelson, Lori J; Seybold, Steven J

    2009-10-01

    The western pine beetle, Dendroctonus brevicomis LeConte (Coleoptera: Curculionidae: Scolytinae), is a major cause of ponderosa pine, Pinus ponderosa Dougl. ex Laws., mortality in much of western North America. Currently, techniques for managing D. brevicomis infestations are limited. Verbenone (4,6,6-trimethylbicyclo [3.1.1] hept-3-en-2-one) is an antiaggregation pheromone of several Dendroctonus spp., including D. brevicomis, and it has been registered as a biopesticide for control of mountain pine beetle, Dendroctonus ponderosae Hopkins, and southern pine beetle, Dendroctonus frontalis Zimmermann. We evaluated the efficacy of a 5-g verbenone pouch [82%-(-); 50 mg/d] applied at 125 Ulha for protecting P. ponderosa stands (2 ha) from D. brevicomis attack over a 3-yr period. No significant differences in levels of D. brevicomis-caused tree mortality or the percentage of unsuccessfully attacked trees were found between verbenone-treated and untreated plots during each year or cumulatively over the 3-yr period. Laboratory analyses of release rates and chemical composition of volatiles emanating from verbenone pouches after field exposure found no deterioration of the active ingredient or physical malfunction of the release device. The mean release rate of pouches from all locations and exposure periods was 44.5 mg/d. In a trapping bioassay, the range of inhibition of the 5-g verbenone pouch was determined to be statistically constant 2 m from the release device. We discuss the implications of these and other results to the development of verbenone as a semiochemical-based tool for management of D. brevicomis infestations in P. ponderosa stands.

  1. How long do ponderosa pine snags stand?

    Treesearch

    Walter G. Dahms

    1949-01-01

    How long will the average ponderosa pine snag remain standing and thus contribute to greater rate of spread and resistance to control of forest fires? Are there any readily discernible characteristics that will enable us to predict which will fall soon and which will stand for a long time?

  2. A ponderosa pine-grand fir spacing study in central Oregon: results after 10 years.

    Treesearch

    K.W. Seidel

    1985-01-01

    The 10-year growth response from an initial spacing study established in a ponderosa pine (Pinus ponderosa Dougl, ex Laws.) and grand fir (Abies grandis (Dougl. ex D. Don) Lindl.) plantation was measured in central Oregon. The study was designed to compare the growth rates of pure pine, pure fir, and a 50-percent mixture of...

  3. Fire history in interior ponderosa pine communities of the Black Hills, South Dakota, USA

    Treesearch

    Peter M. Brown; Carolyn Hull Sieg

    1996-01-01

    Chronologies of fire events were reconstructed from crossdated fire-scarred ponderosa pine trees for four sites in the south-central Black Hills. Compared to other ponderosa pine forests in the southwest US or southern Rocky Mountains, these communities burned less frequently. For all sites combined, and using all fires detected, the mean fire interval (MFI), or number...

  4. Snow bending of sugar pine and ponderosa seedlings ... injury not permanent.

    Treesearch

    William W. Oliver

    1970-01-01

    Sugar pine and ponderosa pine seedlings in the Sierra Nevada, California, with stems bent by heavy snow loads were photographed the next summer and 10 years later. The photographs show that all trees recovered, leaving no permanent stem crook.

  5. Long-suppressed ponderosa pine seedlings respond to release.

    Treesearch

    Walter G. Dahms

    1960-01-01

    Long-suppressed ponderosa pine seedlings and saplings have a remarkable ability to recover and resume normal growth when released. This fact is strikingly demonstrated by a study begun in 1934 on the Pringle Falls Experimental Forest, near Bend, Oregon.

  6. Insects associated with ponderosa pine in Colorado

    Treesearch

    Robert E. Stevens; J. Wayne Brewer; David A. Leatherman

    1980-01-01

    Ponderosa pine serves as a host for a wide variety of insects. Many of these, including all the particularly destructive ones in Colorado, are discussed in this report. Included are a key to the major insect groups, an annotated list of the major groups, a glossary, and a list of references.

  7. Stomata open at night in pole-sized and mature ponderosa pine: implications for O3 exposure metrics

    Treesearch

    Nancy Grulke; R. Alonso; T. Nguyen; C. Cascio; W. Dobrowolski

    2004-01-01

    Ponderosa pine (Pinus ponderosa Dougl. exLaws.) is widely distributed in the western USA.We report the lack of stomatal closure at night in early summer for ponderosa pine at two of three sites investigated. Trees at a third site with lower nitrogen dioxide and nitric acid exposure, but greater drought stress, had slightly open stomata at night in...

  8. Effect of restoration thinning on mycorrhizal fungal propagules in a northern Arizona ponderosa pine forest: Preliminary results

    Treesearch

    Julie E. Korb; Nancy C. Johnson; W. W. Covington

    2001-01-01

    The inoculum potential for arbuscular mycorrhizal (AM) and ectomycorrhizal (EM) fungi were investigated in thinned and uncut control stands in a northern Arizona ponderosa pine forest. A corn bioassay was used to determine the relative amount of infective propagules of AM fungi, and a ponderosa pine (Pinus ponderosa) bioassay was used to determine the relative amount...

  9. Rapid Induction of Multiple Terpenoid Groups by Ponderosa Pine in Response to Bark Beetle-Associated Fungi.

    PubMed

    Keefover-Ring, Ken; Trowbridge, Amy; Mason, Charles J; Raffa, Kenneth F

    2016-01-01

    Ponderosa pine (Pinus ponderosa) is a major and widely distributed component of conifer biomes in western North America and provides substantial ecological and economic benefits. This tree is exposed to several tree-killing bark beetle-microbial complexes, including the mountain pine beetle (Dendroctonus ponderosae) and the phytopathogenic fungus Grosmannia clavigera that it vectors, which are among the most important. Induced responses play a crucial role in conifer defenses, yet these have not been reported in ponderosa pine. We compared concentrations of terpenes and a phenylpropanoid, two phytochemical classes with strong effects against bark beetles and their symbionts, in constitutive phloem tissue and in tissue following mechanical wounding or simulated D. ponderosae attack (mechanical wounding plus inoculation with G. clavigera). We also tested whether potential induced responses were localized or systemic. Ponderosa pines showed pronounced induced defenses to inoculation, increasing their total phloem concentrations of monoterpenes 22.3-fold, sesquiterpenes 56.7-fold, and diterpenes 34.8-fold within 17 days. In contrast, responses to mechanical wounding alone were only 5.2, 11.3, and 7.7-fold, respectively. Likewise, the phenylpropanoid estragole (4-allyanisole) rose to 19.1-fold constitutive levels after simulated attack but only 4.4-fold after mechanical wounding. Overall, we found no evidence of systemic induction after 17 days, which spans most of this herbivore's narrow peak attack period, as significant quantitative and compositional changes within and between terpenoid groups were localized to the wound site. Implications to the less frequent exploitation of ponderosa than lodgepole pine by D. ponderosae, and potential advantages of rapid localized over long-term systemic responses in this system, are discussed.

  10. Estimating probabilities of infestation and extent of damage by the roundheaded pine beetle in ponderosa pine in the Sacramento Mountains, New Mexico

    Treesearch

    Jose Negron

    1997-01-01

    Classification trees and linear regression analysis were used to build models to predict probabilities of infestation and amount of tree mortality in terms of basal area resulting from roundheaded pine beetle, Dendroctonus adjunctus Blandford, activity in ponderosa pine, Pinus ponderosa Laws., in the Sacramento Mountains, New Mexico. Classification trees were built for...

  11. Bark temperature patterns in ponderosa pine stands and their possible effects on mountain pine beetle behavior

    Treesearch

    J. M. Schmid; S. A. Mata; R. A. Schmidt

    1991-01-01

    Bark temperatures on the north and south sides of five ponderosa pines (Pinus ponderosa Laws.) in each of four growing stock levels in two areas in the Black Hills of South Dakota were monitored periodically from May through August 1989. Temperatures were significantly different among growing stock levels and between sides of the tree. The magnitude of differences...

  12. Silvicultural applications: Restoring ecological structure and process in ponderosa pine forests

    Treesearch

    Carl E. Fiedler

    1996-01-01

    A primary goal of restoration treatments in ponderosa pine (Pinus ponderosa)/fir forests is to create more open stand structures, thereby improving tree vigor and reducing vulnerability to insects, disease, and severe fire. An additional goal in some stands is to manipulate existing species composition and site conditions to favor regeneration of...

  13. Site classification of ponderosa pine stands under stocking control in California

    Treesearch

    Robert F. Powers; William W. Oliver

    1978-01-01

    Existing systems for estimating site index of ponderosa pine (Pinus ponderosa Laws.) do not apply well to California stands where stocking is controlled. A more suitable system has been developed using trends in natural height growth, derived from stem analysis of dominant trees in California. This site index system produces polymorphic patterns of...

  14. The Fort Valley Experimental Forest, ponderosa pine, and wildlife habitat research

    Treesearch

    David R. Patton

    2008-01-01

    Wildlife research at the Fort Valley Experimental Forest began with studies to determine how to control damage by wildlife and livestock to ponderosa pine (Pinus ponderosa) reproduction and tree growth. Studies on birds, small mammals, and mule deer (Odocoileus hemionus) browsing were initiated in the early 1930s and 1940s but...

  15. Mature ponderosa pine nutrient use and allocation responses to air pollution

    Treesearch

    Mark A. Poth; Mark E. Fenn

    1998-01-01

    Current-year needles from mature ponderosa pine (Pinus ponderosa Dougl. ex. Laws.) were sampled at four sites across the air pollution gradient in the San Bernardino Mountains in southern California. The sites, in order of decreasing air pollution exposure, included: Sky Forest (SF), Conference Center (CC), Camp Angelus (CA) and Heart Bar (HB)....

  16. Ponderosa pine resin defenses and growth: metrics matter.

    PubMed

    Hood, Sharon; Sala, Anna

    2015-11-01

    Bark beetles (Coleoptera: Curculionidae, Scolytinae) cause widespread tree mortality in coniferous forests worldwide. Constitutive and induced host defenses are important factors in an individual tree's ability to survive an attack and in bottom-up regulation of bark beetle population dynamics, yet quantifying defense levels is often difficult. For example, in Pinus spp., resin flow is important for resistance to bark beetles but is extremely variable among individuals and within a season. While resin is produced and stored in resin ducts, the specific resin duct metrics that best correlate with resin flow remain unclear. The ability and timing of some pine species to produce induced resin is also not well understood. We investigated (i) the relationships between ponderosa pine (Pinus ponderosa Lawson & C. Lawson) resin flow and axial resin duct characteristics, tree growth and physiological variables, and (ii) if mechanical wounding induces ponderosa pine resin flow and resin ducts in the absence of bark beetles. Resin flow increased later in the growing season under moderate water stress and was highest in faster growing trees. The best predictors of resin flow were nonstandardized measures of resin ducts, resin duct size and total resin duct area, both of which increased with tree growth. However, while faster growing trees tended to produce more resin, models of resin flow using only tree growth were not statistically significant. Further, the standardized measures of resin ducts, density and duct area relative to xylem area, decreased with tree growth rate, indicating that slower growing trees invested more in resin duct defenses per unit area of radial growth, despite a tendency to produce less resin overall. We also found that mechanical wounding induced ponderosa pine defenses, but this response was slow. Resin flow increased after 28 days, and resin duct production did not increase until the following year. These slow induced responses may allow

  17. Understanding ponderosa pine forest-grassland vegetation dynamics at Fort Valley Experimental Forest using phytolith analysis

    Treesearch

    Becky K. Kerns; Margaret M. Moore; Stephen C. Hart

    2008-01-01

    In the last century, ponderosa pine forests in the Southwest have changed from more open park-like stands of older trees to denser stands of younger, small-diameter trees. Considerable information exists regarding ponderosa pine forest fire history and recent shifts in stand structure and composition, yet quantitative studies investigating understory reference...

  18. Phoretic symbionts of the mountain pine beetle (Dendroctonus ponderosae Hopkins)

    Treesearch

    Javier E. Mercado; Richard W. Hofstetter; Danielle M. Reboletti; Jose F. Negron

    2014-01-01

    During its life cycle, the tree-killing mountain pine beetle Dendroctonus ponderosae Hopkins interacts with phoretic organisms such as mites, nematodes, fungi, and bacteria. The types of associations these organisms establish with the mountain pine beetle (MPB) vary from mutualistic to antagonistic. The most studied of these interactions are those between beetle and...

  19. Modeling cold tolerance in the mountain pine beetle, Dendroctonus ponderosae

    Treesearch

    Jacques Regniere; Barbara Bentz

    2007-01-01

    Cold-induced mortality is a key factor driving mountain pine beetle, Dendroctonus ponderosae, population dynamics. In this species, the supercooling point (SCP) is representative of mortality induced by acute cold exposure. Mountain pine beetle SCP and associated cold-induced mortality fluctuate throughout a generation, with the highest SCPs prior to and following...

  20. Black stain root disease studies on ponderosa pine parameters and disturbance treatments affecting infection and mortality

    Treesearch

    W.J. Otrosina; J.T. Kliejunas; S. Smith; D.R. Cluck; S.S. Sung; C.D. Cook

    2007-01-01

    Black stain root disease of ponderosa pine (Pinus ponderosa Doug. Ex Laws.), caused by Leptographium wageneri var. ponderosum (Harrington & Cobb) Harrington & Cobb, is increasing on many eastside Sierra Nevada pine stands in northeastern California. The disease is spread from tree to tree via root...

  1. Mountain pine beetle host selection between lodgepole and ponderosa pines in the southern Rocky Mountains

    Treesearch

    Daniel R. West; Jennifer S. Briggs; William R. Jacobi; Jose F. Negron

    2016-01-01

    Recent evidence of range expansion and host transition by mountain pine beetle (Dendroctonus ponderosae Hopkins; MPB) has suggested that MPB may not primarily breed in their natal host, but will switch hosts to an alternate tree species. As MPB populations expanded in lodgepole pine forests in the southern Rocky Mountains, we investigated the potential for...

  2. Patterns of conifer regeneration following high severity wildfire in ponderosa pine - dominated forests of the Colorado Front Range

    Treesearch

    Marin E. Chambers; Paula J. Fornwalt; Sparkle L. Malone; Michael Battaglia

    2016-01-01

    Many recent wildfires in ponderosa pine (Pinus ponderosa Lawson & C. Lawson) - dominated forests of the western United States have burned more severely than historical ones, generating concern about forest resilience. This concern stems from uncertainty about the ability of ponderosa pine and other co-occurring conifers to regenerate in areas where no...

  3. Snag distributions in relation to human access in ponderosa pine forests

    Treesearch

    Jeff P. Hollenbeck; Lisa J. Bate; Victoria A. Saab; John F. Lehmkuhl

    2013-01-01

    Ponderosa pine (Pinus ponderosa) forests in western North America provide habitat for numerous cavity-using wildlife species that often select large-diameter snags for nesting and roosting. Yet large snags are often removed for their commercial and firewood values. Consequently we evaluated effects of human access on snag densities and diameter-class distributions at...

  4. Historical and contemporary lessons from ponderosa pine genetic studies at the Fort Valley Experimental Forest, Arizona (P-53)

    Treesearch

    Laura E. DeWald; Mary Frances Mahalovich

    2008-01-01

    Forest management will protect genetic integrity of tree species only if their genetic diversity is understood and considered in decision-making. Genetic knowledge is particularly important for species such as ponderosa pine (Pinus ponderosa Dougl. ex Laws.) that are distributed across wide geographic distances and types of climates. A ponderosa pine study initiated in...

  5. Tree canopy types constrain plant distributions in ponderosa pine-Gambel oak forests, northern Arizona

    Treesearch

    Scott R. Abella

    2009-01-01

    Trees in many forests affect the soils and plants below their canopies. In current high-density southwestern ponderosa pine (Pinus ponderosa) forests, managers have opportunities to enhance multiple ecosystem values by manipulating tree density, distribution, and canopy cover through tree thinning. I performed a study in northern Arizona ponderosa...

  6. Emissions and Photochemistry of BVOCs in a Ponderosa Pine woodland

    NASA Astrophysics Data System (ADS)

    Kim, S.; Karl, T.; Rasmussen, R.; Apel, E.; Harley, P.; Waldo, S.; Roberts, S.; Guenther, A.

    2008-12-01

    We deployed two proton-transfer-reaction mass spectrometry instruments (PTR-MS, IONICON ANALYTIK) for ambient and branch enclosure measurements at the Manitou Experimental Forest, located in the Southern Rocky Mountain area as a part of the Bio-hydro-atmosphere interactions of Energy, Aerosols, Carbon, H2O, Organics and Nitrogen (BEACHON) field campaign in 2008. Vegetation at the field site is dominated by Ponderosa Pine. BVOC emissions from Ponderosa Pine along with temperature, photosynthetic photon flux density (ppfd), relative humidity, and CO2 uptake were measured from two branch-enclosures (shade and sun). Diurnal variations and the emission response to environmental conditions are described and compared to existing models. In addition, we analyzed the speciation of BVOCs from enclosures by GC-MS. We will present quantitative and qualitative characteristics of BVOC emissions from Ponderosa Pine and analytical characteristics of PTR-MS such as fragmentation patterns of semi-volatile compounds (sesquiterpene, bornyl acetate etc) that we identified as major emissions from the enclosures. BVOC emissions observed in the enclosures will be quantitatively compared to BVOC distributions in ambient air. We explore the presence of possibly unidentified BVOCs in the forest canopy by examining PTR-MS mass spectra of enclosure and ambient air samples based on mass scans between 40 - 210 amu.

  7. Unthinned slow-growing ponderosa pine (Pinus ponderosa) trees contain muted isotopic signals in tree rings as compared to thinned trees

    EPA Science Inventory

    We analysed the oxygen isotopic values of wood (δ18Ow) of 12 ponderosa pine (Pinus ponderosa) trees from control, moderately, and heavily thinned stands and compared them with existing wood-based estimates of carbon isotope discrimination (∆13C), basal area increment (BAI), and g...

  8. Water potential in ponderosa pine stands of different growing-stock levels

    Treesearch

    J. M. Schmid; S. A. Mata; R. K. Watkins; M. R. Kaufmann

    1991-01-01

    Water potential was measured in five ponderosa pine (Pinus ponderosa Laws.) in each of four stands of different growing-stock levels at two locations in the Black Hills of South Dakota. Mean water potentials at dawn and midday varied significantly among growing-stock levels at one location, but differences were not consistent. Mean dawn and midday water potentials...

  9. Lumber recovery from dead ponderosa pine in the Colorado front range.

    Treesearch

    Susan Willits; Richard O. Woodfin; Thomas A. Snellgrove

    1990-01-01

    Lumber recovery information from live and beetle-killed ponderosa pine (Pinus ponderosa Dougl. ex Laws) in the Colorado Front Range is presented. No significant difference in lumber volume was found among the samples. Significant differences were found in lumber value among the live, 1-year-dead, and 3- to 5-year-dead samples. About 10 percent of...

  10. Evaluation of insecticides for protecting southwestern ponderosa pines from attack by engraver beetles (Coleoptera: Curculionidae: Scolytinae)

    Treesearch

    Tom E. DeGomez; Christopher J. Hayes; John A. Anhold; Joel D. McMillin; Karen M. Clancy; Paul P. Bosu

    2006-01-01

    Insecticides that might protect pine trees from attack by engraver beetles (Ips spp.) have not been rigorously tested in the southwestern United States. We conducted two field experiments to evaluate the efficacy of several currently and potentially labeled preventative insecticides for protecting high-value ponderosa pine, Pinus ponderosa...

  11. Exploring Climate Niches of Ponderosa Pine (Pinus ponderosa Douglas ex Lawson) Haplotypes in the Western United States: Implications for Evolutionary History and Conservation

    PubMed Central

    Shinneman, Douglas J.; Potter, Kevin M.; Hipkins, Valerie D.

    2016-01-01

    Ponderosa pine (Pinus ponderosa Douglas ex Lawson) occupies montane environments throughout western North America, where it is both an ecologically and economically important tree species. A recent study using mitochondrial DNA analysis demonstrated substantial genetic variation among ponderosa pine populations in the western U.S., identifying 10 haplotypes with unique evolutionary lineages that generally correspond spatially with distributions of the Pacific (P. p. var. ponderosa) and Rocky Mountain (P. p. var. scopulorum) varieties. To elucidate the role of climate in shaping the phylogeographic history of ponderosa pine, we used nonparametric multiplicative regression to develop predictive climate niche models for two varieties and 10 haplotypes and to hindcast potential distribution of the varieties during the last glacial maximum (LGM), ~22,000 yr BP. Our climate niche models performed well for the varieties, but haplotype models were constrained in some cases by small datasets and unmeasured microclimate influences. The models suggest strong relationships between genetic lineages and climate. Particularly evident was the role of seasonal precipitation balance in most models, with winter- and summer-dominated precipitation regimes strongly associated with P. p. vars. ponderosa and scopulorum, respectively. Indeed, where present-day climate niches overlap between the varieties, introgression of two haplotypes also occurs along a steep clinal divide in western Montana. Reconstructed climate niches for the LGM suggest potentially suitable climate existed for the Pacific variety in the California Floristic province, the Great Basin, and Arizona highlands, while suitable climate for the Rocky Mountain variety may have existed across the southwestern interior highlands. These findings underscore potentially unique phylogeographic origins of modern ponderosa pine evolutionary lineages, including potential adaptations to Pleistocene climates associated with discrete

  12. Exploring Climate Niches of Ponderosa Pine (Pinus ponderosa Douglas ex Lawson) Haplotypes in the Western United States: Implications for Evolutionary History and Conservation.

    PubMed

    Shinneman, Douglas J; Means, Robert E; Potter, Kevin M; Hipkins, Valerie D

    2016-01-01

    Ponderosa pine (Pinus ponderosa Douglas ex Lawson) occupies montane environments throughout western North America, where it is both an ecologically and economically important tree species. A recent study using mitochondrial DNA analysis demonstrated substantial genetic variation among ponderosa pine populations in the western U.S., identifying 10 haplotypes with unique evolutionary lineages that generally correspond spatially with distributions of the Pacific (P. p. var. ponderosa) and Rocky Mountain (P. p. var. scopulorum) varieties. To elucidate the role of climate in shaping the phylogeographic history of ponderosa pine, we used nonparametric multiplicative regression to develop predictive climate niche models for two varieties and 10 haplotypes and to hindcast potential distribution of the varieties during the last glacial maximum (LGM), ~22,000 yr BP. Our climate niche models performed well for the varieties, but haplotype models were constrained in some cases by small datasets and unmeasured microclimate influences. The models suggest strong relationships between genetic lineages and climate. Particularly evident was the role of seasonal precipitation balance in most models, with winter- and summer-dominated precipitation regimes strongly associated with P. p. vars. ponderosa and scopulorum, respectively. Indeed, where present-day climate niches overlap between the varieties, introgression of two haplotypes also occurs along a steep clinal divide in western Montana. Reconstructed climate niches for the LGM suggest potentially suitable climate existed for the Pacific variety in the California Floristic province, the Great Basin, and Arizona highlands, while suitable climate for the Rocky Mountain variety may have existed across the southwestern interior highlands. These findings underscore potentially unique phylogeographic origins of modern ponderosa pine evolutionary lineages, including potential adaptations to Pleistocene climates associated with discrete

  13. Effectiveness of Esfenvalerate, Cyfluthrin, and Carbaryl in Protecting Individual Lodgepole Pines and Ponderosa Pines from Attack by Dendroctonus spp.

    Treesearch

    Michael I. Haverty; Patrick J. Shea; James T. Hoffman; John M. Wenz; Kenneth E. Gibson

    1998-01-01

    The effectiveness of registered and experimental application rates of insecticides esfenvalerate (Asana XL), cyfluthrin (Tempo WP and Tempo 2), and carbaryl (Sevimol and Sevin SL) was assessed for protection of individual high-value lodgepole pines from mountain pine beetles in Montana and ponderosa pines from western pine beetles in Idaho and California. This field...

  14. Changes in physiological attributes of ponderosa pine from seedling to mature tree

    Treesearch

    Nancy E. Grulke; William A. Retzlaff

    2001-01-01

    Plant physiological models are generally parameterized from many different sources of data, including chamber experiments and plantations, from seedlings to mature trees. We obtained a comprehensive data set for a natural stand of ponderosa pine (Pinus ponderosa Laws.) and used these data to parameterize the physiologically based model, TREGRO....

  15. Effects of photochemical oxidant injury of ponderosa and Jeffrey pines on susceptibility of sapwood and freshly cut stumps to Fomes annosus. [Pinus ponderosa; Pinus jeffreyi; Fomes annosus; Trichoderma spp. ; Polyporus versicolor; Poria Monticola

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    James, R.L.; Cobb, F.W. Jr.; Wilcox, W.W.

    1980-01-01

    Ponderosa and Jeffrey pine sapwood samples and freshly cut stumps from trees with different amounts of oxidant injury were inoculated with Fomes annosus. With stumps, percentage of surface cross-section area infected and extent of vertical colonization were determined 1 mo and 6-10 mo after inoculation, respectively. Increase in surface area infection with increased oxidant injury, expressed as upper-crown needle retention, was statistically significant for ponderosa pine (P=0.01), but was not for Jeffrey pine. Also, the rate of vertical colonization was greater in stumps from severely oxidant-injured trees than in those from slightly injured trees. The relationship between injury and colonizationmore » was significant for Jeffrey pine (P = 0.05) and for ponderosa pine at one site (P = 0.03), but nonsignificant (P = 0.18) for ponderosa pine at a second site. Increased susceptibility of stumps to F. annosus appeared to be associated with decreased colonization by other fungi (especially Trichoderma spp. and blue stain fungi). Laboratory tests indicated that decay susceptibility of excised sapwood to F. annosus apparently was not affected by oxidant injury with Jeffrey pine, but weight loss of ponderosa pine sapwood was correlated with decreased injury (greater needle retention). On the other hand, weight losses of Jeffrey pine caused by Polyporus versicolor and of ponderosa pine caused by Poria monticola were correlated with increased injury (increased needle chlorosis). 27 references, 2 figures, 3 tables.« less

  16. Suitability of live and fire-killed small-diameter ponderosa and lodgepole pine trees for manufacturing a new structural wood composite.

    PubMed

    Linton, J M; Barnes, H M; Seale, R D; Jones, P D; Lowell, E C; Hummel, S S

    2010-08-01

    Finding alternative uses for raw material from small-diameter trees is a critical problem throughout the United States. In western states, a lack of markets for small-diameter ponderosa pine (Pinus ponderosa) and lodgepole pine (Pinus contorta) can contribute to problems associated with overstocking. To test the feasibility of producing structural composite lumber (SCL) beams from these two western species, we used a new technology called steam-pressed scrim lumber (SPSL) based on scrimming technology developed in Australia. Both standing green and fire-killed ponderosa and lodgepole pine logs were used in an initial test. Fire-killed logs of both species were found to be unsuitable for producing SPSL but green logs were suitable for producing SPSL. For SPSL from green material, ponderosa pine had significantly higher modulus of rupture and work-to-maximum load values than did SPSL from lodgepole pine. Modulus of elasticity was higher for lodgepole pine. The presence of blows was greater with lodgepole pine than with ponderosa. Blows had a negative effect on the mechanical properties of ponderosa pine but no significant effect on the mechanical properties of SPSL from lodgepole pine. An evaluation of non-destructive testing methods showed that X-ray could be used to determine low density areas in parent beams. The use of a sonic compression wave tester for NDE evaluation of modulus of rupture showed some promise with SPSL but requires further research. (c) 2010 Elsevier Ltd. All rights reserved.

  17. Lumber recovery from ponderosa pine in the Black Hills, South Dakota.

    Treesearch

    Marlin E. Plank

    1985-01-01

    A sample of 400 ponderosa pine (Pinus ponderosa Dougl. ex Laws.) trees was selected from each of two sale areas in the Black Hills National Forest, South Dakota. The logs were processed through two sawmills into 1-inch-thick boards. Estimates of volume and value recovery based on cubic volume and board foot volume are shown in tables and figures....

  18. Understory-overstory relationships in ponderosa pine forests, Black Hills, South Dakota

    Treesearch

    Daniel W. Uresk; Kieth E. Severson

    1989-01-01

    Under-story-overstory relationships were examined over 7 different growing stock levels(GSLs) of 2 size classes(saplings,8-10 cm d.b.h. and poles, 15-18 cm d.b.h.) of ponderosa pine (Pinus ponderosa) in the Black Hills, South Dakota. Generally, production of graminoids, forbs, and shrubs was similar between sapling and pole stands. Trends among GSLs were also similar...

  19. Ponderosa pine managed-yield simulator: PPSIM users guide.

    Treesearch

    D.J. DeMars; J.W. Barrett

    1987-01-01

    PPSIM simulates yields of natural and managed ponderosa pine stands. Management practices and effects that can be simulated include commercial thinning, fertilization, genetic improvement, and presence of dwarf mistletoe. An option is available that adjusts growth to simulate local conditions. Equations used in PPSIM describe growth of natural...

  20. Maturity selection versus improvement selection: lessons from a mid-20th century controversy in the silviculture of ponderosa pine

    Treesearch

    Kevin L. O' Hara; Andy Youngblood; Kristen M. Waring

    2010-01-01

    Two rival silvicultural systems for promoting multiaged ponderosa pine stands emerged in the 1930s and 1940s. Maturity selection was developed to move the vast ponderosa pine (Pinus ponderosa) acreage in eastern Oregon and Washington into regulation to limit bark beetle losses. In the Southwest, improvement selection was designed to improve...

  1. Intraspecific niche models for ponderosa pine (Pinus ponderosa) suggest potential variability in population-level response to climate change.

    PubMed

    Maguire, Kaitlin C; Shinneman, Douglas J; Potter, Kevin M; Hipkins, Valerie D

    2018-03-14

    Unique responses to climate change can occur across intraspecific levels, resulting in individualistic adaptation or movement patterns among populations within a given species. Thus, the need to model potential responses among genetically distinct populations within a species is increasingly recognized. However, predictive models of future distributions are regularly fit at the species level, often because intraspecific variation is unknown or is identified only within limited sample locations. In this study, we considered the role of intraspecific variation to shape the geographic distribution of ponderosa pine (Pinus ponderosa), an ecologically and economically important tree species in North America. Morphological and genetic variation across the distribution of ponderosa pine suggest the need to model intraspecific populations: the two varieties (var. ponderosa and var. scopulorum) and several haplotype groups within each variety have been shown to occupy unique climatic niches, suggesting populations have distinct evolutionary lineages adapted to different environmental conditions. We utilized a recently-available, geographically-widespread dataset of intraspecific variation (haplotypes) for ponderosa pine and a recently-devised lineage distance modeling approach to derive additional, likely intraspecific occurrence locations. We confirmed the relative uniqueness of each haplotype-climate relationship using a niche-overlap analysis, and developed ecological niche models (ENMs) to project the distribution for two varieties and eight haplotypes under future climate forecasts. Future projections of haplotype niche distributions generally revealed greater potential range loss than predicted for the varieties. This difference may reflect intraspecific responses of distinct evolutionary lineages. However, directional trends are generally consistent across intraspecific levels, and include a loss of distributional area and an upward shift in elevation. Our results

  2. Intraspecific niche models for ponderosa pine (Pinus ponderosa) suggest potential variability in population-level response to climate change

    USGS Publications Warehouse

    Maguire, Kaitlin C.; Shinneman, Douglas; Potter, Kevin M.; Hipkins, Valerie D.

    2018-01-01

    Unique responses to climate change can occur across intraspecific levels, resulting in individualistic adaptation or movement patterns among populations within a given species. Thus, the need to model potential responses among genetically distinct populations within a species is increasingly recognized. However, predictive models of future distributions are regularly fit at the species level, often because intraspecific variation is unknown or is identified only within limited sample locations. In this study, we considered the role of intraspecific variation to shape the geographic distribution of ponderosa pine (Pinus ponderosa), an ecologically and economically important tree species in North America. Morphological and genetic variation across the distribution of ponderosa pine suggest the need to model intraspecific populations: the two varieties (var. ponderosa and var. scopulorum) and several haplotype groups within each variety have been shown to occupy unique climatic niches, suggesting populations have distinct evolutionary lineages adapted to different environmental conditions. We utilized a recently-available, geographically-widespread dataset of intraspecific variation (haplotypes) for ponderosa pine and a recently-devised lineage distance modeling approach to derive additional, likely intraspecific occurrence locations. We confirmed the relative uniqueness of each haplotype-climate relationship using a niche-overlap analysis, and developed ecological niche models (ENMs) to project the distribution for two varieties and eight haplotypes under future climate forecasts. Future projections of haplotype niche distributions generally revealed greater potential range loss than predicted for the varieties. This difference may reflect intraspecific responses of distinct evolutionary lineages. However, directional trends are generally consistent across intraspecific levels, and include a loss of distributional area and an upward shift in elevation. Our results

  3. Historical wildfire impacts on ponderosa pine tree overstories: An Arizona case study

    Treesearch

    Peter F. Ffolliott; Cody L. Stropki; Daniel G. Neary

    2008-01-01

    The Rodeo-Chediski Wildfire--the largest in Arizona's history--damaged or destroyed ecosystem resources and disrupted ecosystem functioning in a largely mosaic pattern throughout the ponderosa pine (Pinus ponderosa) forests exposed to the burn. Impacts of this wildfire on tree overstories were studied for 5 years (2002 to 2007) on two watersheds...

  4. Biomass and nutrient distributions in central Oregon second-growth ponderosa pine ecosystems.

    Treesearch

    Susan N. Little; Lauri J. Shainsky

    1995-01-01

    We investigated the distribution of biomass and nutrients in second-growth ponderosa pine (Pinus ponderosa Dougl. ex Laws.) ecosystems in central Oregon. Destructive sampling of aboveground and belowground tree biomass was carried out at six sites in the Deschutes National Forest; three of these sites also were intensively sampled for biomass and...

  5. Resin distribution in second-growth ponderosa pine

    Treesearch

    B.H. Paul

    1955-01-01

    In a study of specific gravity of second-growth ponderosa pine, there was visible evidence of resin in a part of the specific gravity specimens. Each specimen contained 10 annual growth rings in cross sections taken at 4 heights in the merchantable length of the trees. Since the presence of resin introduced an uncertain amount of error in the specific gravity values,...

  6. Restoration of southwestern ponderosa pine ecosystems with fire

    Treesearch

    Stephen S. Sackett; Sally M. Haase; Michael G. Harrington

    1994-01-01

    Heavy grazing and timbering during settlement by Europeans, and a policy of fire exclusion shortly after caused extensive structural and compositional changes to the southwestern ponderosa pine ecosystem. These changes have resulted in forest health problems, such as increased insect and disease epidemics, reduced wildlife habitat, and a serious wildfire hazard....

  7. Induction of Listeria monocytogenes infection by the consumption of ponderosa pine needles.

    PubMed Central

    Adams, C J; Neff, T E; Jackson, L L

    1979-01-01

    An infectious microorganism, identified as Listeria monocytogenes, has been isolated from the bloodstream of pregnant mice fed a diet containing Pinus ponderosa needles. When the isolate was injected into pregnant mice, reproductive dysfunction and other changes, including speckled livers, spleen atrophy, and hemorrhagic intestines, appeared to mimic the signs of the disease in pregnant mice fed pine needles. Moreover, these pathological changes are similar to those observed in cattle and other mammals experiencing abortions or toxemia, or both, attributed to the ingestion of P. ponderosa needles, suggesting that L. monocytogenes may be a part of the etiology of "pine needle abortion." PMID:113341

  8. Effects of long-term ozone exposure and drought on the photosynthetic capacity of ponderosa pine (Pinus ponderosa Laws.) New Phytol

    Treesearch

    Jan L. Beyers; George H. Riechers; Patrick J. Temple

    1992-01-01

    Seedlings of ponderosa pine (Pinus ponderosa Laws.) were grown for three years under three atmospheric ozone concentrations- clean air (CF), ambient ozone (NF), and 1-5 times ambient ozone (NF150) - at a moderatelypolluted site in the Sierra Nevada, under either well-watered or drought-stressed conditions. When the trees were 5 years old,...

  9. Effect of Periodic Burning on Soil Nitrogen Concentrations in Ponderosa Pine

    Treesearch

    W. W. Covington; S. S. Sackett

    1986-01-01

    To determine the effects of different burning intervals on soil N status in substands of sapling-, pole-, and sawtimber-sized ponderosa pine (Pinus ponderosa Laws.) we sampled plots burned at 1-, 2-, and 4-yr intervals by three strata at two depths (0-5 and 5-15 cm). Generally, NH4 +; and NO3 - concentrations were higher on plots repeatedly burned than on unburned...

  10. Effects of wildfire severity on small mammals in northern Arizona ponderosa pine forests

    Treesearch

    Sean C. Kyle; William M. Block

    2000-01-01

    We examined effects of a varied-severity wildfire on the community structure of small mammals and populations of the 2 most abundant species, the deer mouse (Peromyscus maniculatus) and the gray-collared chipmunk (Tamias cinereicollis), in northern Arizona ponderosa pine (Pinus ponderosa) forests. We examined 2...

  11. Frequent fire alters nitrogen transformations in ponderosa pine stands of the inland Northwest

    Treesearch

    Thomas H. DeLuca; Sala Anna

    2006-01-01

    Recurrent, low-severity fire in ponderosa pine (Pinus ponderosa)/interior Douglas-fir (Pseudotsuga menziesii var. glauca) forests is thought to have directly influenced nitrogen (N) cycling and availability. However, no studies to date have investigated the influence of natural fire intervals on soil processes in undisturbed...

  12. Ponderosa pine forest reconstruction: Comparisons with historical data

    Treesearch

    David W. Huffman; Margaret M. Moore; W. Wallace Covington; Joseph E. Crouse; Peter Z. Fule

    2001-01-01

    Dendroecological forest reconstruction techniques are used to estimate presettlement structure of northern Arizona ponderosa pine forests. To test the accuracy of these techniques, we remeasured 10 of the oldest forest plots in Arizona, a subset of 51 historical plots established throughout the region from 1909 to 1913, and compared reconstruction outputs to historical...

  13. Understanding ponderosa pine forest-grassland vegetation dynamics at Fort Valley Experimental Forest using phytolith analysis (P-53)

    Treesearch

    Becky K. Kerns; Margaret M. Moore; Stephen C. Hart

    2008-01-01

    In the last century, ponderosa pine forests in the Southwest have changed from more open park-like stands of older trees to denser stands of younger, smalldiameter trees. Considerable information exists regarding ponderosa pine forest fire history and recent shifts in stand structure and composition, yet quantitative studies investigating understory reference...

  14. Foliar nitrogen content and tree growth after prescribed fire in ponderosa pine.

    Treesearch

    J.D. Landsberg; P.H. Cochran; M.M. Finck; R.E. Martin

    1984-01-01

    This initial study of prescribed burning in ponderosa pine (Pinus ponderosa Doug. ex Laws.) stands in central Oregon showed that all periodic annual growth increments were reduced for trees alive four growing seasons later. Height growth was reduced 8 percent in areas burned by fires with moderate fuel consumption and 18 percent in areas with high...

  15. Spacing and shrub competition influence 20-year development of planted ponderosa pine

    Treesearch

    William W. Oliver

    1990-01-01

    Growth and stand development of ponderosa pine (Pinus ponderosa) were monitored for 20 years after planting at five different square spacings (6, 9, 12, 15, and 18 ft) in the presence or absence of competing shrubs on the westside Sierra Nevada. Mean tree size was positively correlated and stand values negatively correlated with spacing in the...

  16. Estimating value and volume of ponderosa pine trees by equations.

    Treesearch

    Martin E. Plank

    1981-01-01

    Equations for estimating the selling value and tally volume for ponderosa pine lumber from the standing trees are described. Only five characteristics are required for the equations. Development and application of the system are described.

  17. Exploring climate niches of ponderosa pine (Pinus ponderosa Douglas ex Lawson) haplotypes in the western United States: Implications for evolutionary history and conservation

    USGS Publications Warehouse

    Shinneman, Douglas; Means, Robert E.; Potter, Kevin M.; Hipkins, Valerie D.

    2016-01-01

    Ponderosa pine (Pinus ponderosa Douglas ex Lawson) occupies montane environments throughout western North America, where it is both an ecologically and economically important tree species. A recent study using mitochondrial DNA analysis demonstrated substantial genetic variation among ponderosa pine populations in the western U.S., identifying 10 haplotypes with unique evolutionary lineages that generally correspond spatially with distributions of the Pacific (P. p. var. ponderosa) and Rocky Mountain (P. p. var. scopulorum) varieties. To elucidate the role of climate in shaping the phylogeographic history of ponderosa pine, we used nonparametric multiplicative regression to develop predictive climate niche models for two varieties and 10 haplotypes and to hindcast potential distribution of the varieties during the last glacial maximum (LGM), ~22,000 yr BP. Our climate niche models performed well for the varieties, but haplotype models were constrained in some cases by small datasets and unmeasured microclimate influences. The models suggest strong relationships between genetic lineages and climate. Particularly evident was the role of seasonal precipitation balance in most models, with winter- and summer-dominated precipitation regimes strongly associated with P. p. vars. ponderosa and scopulorum, respectively. Indeed, where present-day climate niches overlap between the varieties, introgression of two haplotypes also occurs along a steep clinal divide in western Montana. Reconstructed climate niches for the LGM suggest potentially suitable climate existed for the Pacific variety in the California Floristic province, the Great Basin, and Arizona highlands, while suitable climate for the Rocky Mountain variety may have existed across the southwestern interior highlands. These findings underscore potentially unique phylogeographic origins of modern ponderosa pine evolutionary lineages, including potential adaptations to Pleistocene climates associated with

  18. Best predictors for postfire mortality of ponderosa pine trees in the Intermountain West

    Treesearch

    Carolyn Hull Sieg; Joel D. McMillin; James F. Fowler; Kurt K. Allen; Jose F. Negron; Linda L. Wadleigh; John A. Anhold; Ken E. Gibson

    2006-01-01

    Numerous wildfires in recent years have highlighted managers' needs for reliable tools to predict postfire mortality of ponderosa pine (Pinus ponderosa Dougl. ex Laws.) trees. General applicability of existing mortality models is uncertain, as researchers have used different sets of variables. We quantified tree attributes, crown and bole fire...

  19. FINE ROOT TURNOVER IN PONDEROSA PINE STANDS OF DIFFERENT AGES: FIRST-YEAR RESULTS

    EPA Science Inventory

    Root minirhizotron tubs were installed in two ponderosa pine (Pinus ponderosa Laws.) Stands of different ages to examine patterns of root growth and death. The old-growth site (OS) consists of a mixture of old (>250 years) and young trees (ca.45 yrs)< and is located near clamp S...

  20. The effect of manzanita and snowbrush competition on ponderosa pine reproduction.

    Treesearch

    Walter G. Dahms

    1950-01-01

    What effect does a dense cover of manzanita and snowbrush have on the establishment and growth of ponderosa pine reproduction? Observations on this point on the Pringle Falls Experimental Forest resulted in the following conclusions: (1) the brush did not significantly affect the early establishment of pine seedlings, but (2) it sharply reduced the growth of...

  1. Effects of stratification and temperature on seed germination speed and uniformity in central Oregon ponderosa pine (Pinus ponderosa Dougl. ex Laws.).

    Treesearch

    John C. Weber; Frank C. Sorensen

    1990-01-01

    Effects of stratification period and incubation temperature on seed germination speed and uniformity were investigated in a bulked seed lot of 200 ponderosa pine trees (Pinus ponderosa Dougl. ex Laws.) sampled from 149 locations in central Oregon. Mean rate of embryo development towards germination (l/days to 50 percent germination) and standard...

  2. Uptake and Distribution of Nitrogen from Acidic Fog within a Ponderosa Pine (Pinus ponderosa Laws.)/Litter/Soil System

    Treesearch

    Mark E. Fenn; Theodor D. Leininger

    1995-01-01

    The magnitude and importance of wet deposition of N in forests of the South Coast (Los Angeles) Air Basin have not been well characterized. We exposed 3-yr-old ponderosa pine (Pinus ponderos Laws.) seedlings growing in native forest soil to acidic fog treatments (pH 3.1) simulating fog chemistry from a pine forest near Los Angeles, California. Fog solutions contained...

  3. Growth models for ponderosa pine: I. Yield of unthinned plantations in northern California.

    Treesearch

    William W. Oliver; Robert F. Powers

    1978-01-01

    Yields for high-survival, unthinned ponderosa pine (Pinus ponderosa Laws.) plantations in northern California are estimated. Stems of 367 trees in 12 plantations were analyzed to produce a growth model simulating stand yields. Diameter, basal area, and net cubic volume yields by Site Indices50 40 through 120 are tabulated for...

  4. Growth of a 45-year-old ponderosa pine plantation: An Arizona case study

    Treesearch

    Peter F. Ffolliott; Gerald J. Gottfried; Cody L. Stropki; L. J. Heidmann

    2008-01-01

    Information on the growth of forest plantations is necessary for planning of ecosystem-based management of the plantations. This information is also useful in validating or refining computer simulators that estimate plantation growth into the future. Such growth information has been obtained from a 45-year-old ponderosa pine (Pinus ponderosa)...

  5. Pelletized ponderosa pine bark for adsorption of toxic heavy metals from water

    Treesearch

    Miyoung Oh; Mandla A. Tshabalala

    2007-01-01

    Bark flour from ponderosa pine (Pinus ponderosa) was consolidated into pellets using citric acid as cross-linking agent. The pellets were evaluated for removal of toxic heavy metals from synthetic aqueous solutions. When soaked in water, pellets did not leach tannins, and they showed high adsorption capacity for Cu(ll), Zn(ll), Cd(ll). and Ni(ll) under both equilibrium...

  6. Use of Hardwood Tree Species by Birds Nesting in Ponderosa Pine Forests

    Treesearch

    Kathryn L. Purcell; Douglas A. Drynan

    2008-01-01

    We examined the use of hardwood tree species for nesting by bird species breeding in ponderosa pine (Pinus ponderosa) forests in the Sierra National Forest, California. From 1995 through 2002, we located 668 nests of 36 bird species nesting in trees and snags on four 60-ha study sites. Two-thirds of all species nesting in trees or snags used...

  7. Contemporary human use of southwestern ponderosa pine forests

    Treesearch

    Carol Raish; Wang Yong; John M. Marzluff

    1997-01-01

    The ponderosa pine forests of the Southwest provide land, resources, products, and recreational opportunities for both urban and rural communities of the region and the nation. These human uses and activities affect resident and migratory bird populations in both negative and positive ways. This brief review focuses on three major kinds of human use that have the...

  8. Can prescribed fire be used to maintain fuel treatment effectiveness over time in Black Hills ponderosa pine forests?

    Treesearch

    Mike A. Battaglia; Frederick W. Smith; Wayne D. Shepperd

    2008-01-01

    We determine the time frame after initial fuel treatment when prescribed fire will be likely to produce high enough mortality rates in ponderosa pine (Pinus ponderosa var. scopulorum Dougl. ex Laws.) regeneration to be successful in maintaining treatment effectiveness in the Black Hills of South Dakota. We measured pine...

  9. High yields from young-growth ponderosa pine.

    Treesearch

    Edwin L. Mowat

    1947-01-01

    A ponderosa pine stand growing at a net rate of 618 board feet per acre per year may be rather amazing to foresters accustomed to the proverbial slow growth of this species in the virgin forest. Yet that is the average increment for the last 6 years of a 102-year-old even-aged stand on Lookout Mountain in the Pringle Falls Experimental Forest in central Oregon. During...

  10. Radial and stand-level thinning treatments: 15-year growth response of legacy ponderosa and Jeffrey pine trees

    Treesearch

    Sharon M. Hood; Daniel R. Cluck; Bobette E. Jones; Sean Pinnell

    2017-01-01

    Restoration efforts to improve vigor of large, old trees and decrease risk to high-intensity wildland fire and drought-mediated insect mortality often include reductions in stand density. We examined 15-year growth response of old ponderosa pine (Pinus ponderosa) and Jeffrey pine (Pinus jeffreyi) trees in northeastern California, U.S.A. to two levels of thinning...

  11. Dwarfmistletoe removal and reinvasion in jeffrey and ponderosa pine, northeastern California

    Treesearch

    Willis W. Wagener

    1965-01-01

    Nine years of removal of infections and prevention of seeding of dwarfmistletoe from a young mixed Jeffrey-ponderosa pine stand failed to eliminate the parasite. Spread of the parasite back into the experimental plot from infected trees outside averaged about 0.9 feet per year. One tree in 23 appeared to show some resistance to infection. Jeffrey pine had almost twice...

  12. Physiological responses of ponderosa pine in western Montana to thinning, prescribed fire and burning season.

    PubMed

    Sala, Anna; Peters, Gregory D; McIntyre, Lorna R; Harrington, Michael G

    2005-03-01

    Low-elevation ponderosa pine (Pinus ponderosa Dougl. ex. Laws.) forests of the northern Rocky Mountains historically experienced frequent low-intensity fires that maintained open uneven-aged stands. A century of fire exclusion has contributed to denser ponderosa pine forests with greater competition for resources, higher tree stress and greater risk of insect attack and stand-destroying fire. Active management intended to restore a semblance of the more sustainable historic stand structure and composition includes selective thinning and prescribed fire. However, little is known about the relative effects of these management practices on the physiological performance of ponderosa pine. We measured soil water and nitrogen availability, physiological performance and wood radial increment of second growth ponderosa pine trees at the Lick Creek Experimental Site in the Bitterroot National Forest, Montana, 8 and 9 years after the application of four treatments: thinning only; thinning followed by prescribed fire in the spring; thinning followed by prescribed fire in the fall; and untreated controls. Volumetric soil water content and resin capsule ammonium did not differ among treatments. Resin capsule nitrate in the control treatment was similar to that in all other treatments, although burned treatments had lower nitrate relative to the thinned-only treatment. Trees of similar size and canopy condition in the three thinned treatments (with and without fire) displayed higher leaf-area-based photosynthetic rate, stomatal conductance and mid-morning leaf water potential in June and July, and higher wood radial increment relative to trees in control units. Specific leaf area, mass-based leaf nitrogen content and carbon isotope discrimination did not vary among treatments. Our results suggest that, despite minimal differences in soil resource availability, trees in managed units where basal area was reduced had improved gas exchange and growth compared with trees in

  13. Ponderosa pine seed-tree removal reduces stocking only slightly

    Treesearch

    Philip M. McDonald

    1969-01-01

    After ponderosa pine seed trees were removed on the Challenge Experimental Forest, California, seedling stocking fell by 3.8 percent or about 212 seedlings per acre. This loss is slightly less than that incurred from natural mortality, and one that did not reduce regeneration levels below the minimum standard.

  14. Development of vegetation in a young ponderosa pine plantation: effect of treatment duration and time since disturbance

    Treesearch

    Philip M. McDonald; Gary O. Fiddler

    2007-01-01

    The density and development of deerbrush (Ceanothus integerrimus Hook. & Arn.), other shrubs, forbs, graminoids, and ponderosa pine (Pinus ponderosa Dougl. ex Laws. var. ponderosa) seedlings were evaluated in a young plantation in northern California from 1988 through 1997. Treatment regimes consisted of...

  15. Thinning stagnated ponderosa and Jeffrey pine stands in northeastern California: 30-year effects

    Treesearch

    Robert J. Lilieholm; Dennis E. Teeguarden; Donald T. Gordon

    1989-01-01

    Response to precommercial thinning in stagnated 55-year-old ponderosa (Pinus ponderosa Dougl. ex Laws.) and Jeffrey pine (P. jeffreyi Grev. and Balf.) stands in northeastern alifornia was rapid and long-lasting. During the first 5 years after thinning, average annual diameter at breast height (d.b.h.) and height growth of trees on...

  16. Coarse woody debris assay in northern Arizona mixed-conifer and ponderosa pine forests

    Treesearch

    Joseph L. Ganey; Scott C. Vojta

    2010-01-01

    Coarse woody debris (CWD) provides important ecosystem services in forests and affects fire behavior, yet information on amounts and types of CWD typically is limited. To provide such information, we sampled logs and stumps in mixed-conifer and ponderosa pine (Pinus ponderosa) forests in north-central Arizona. Spatial variability was prominent for all CWD parameters....

  17. A height increment equation for young ponderosa pine plantations using precipitation and soil factors

    Treesearch

    Fabian C.C. Uzoh

    2001-01-01

    A height increment equation was used to determine the effects of site quality and competing herbaceous vegetation on the development of ponderosa pine seedlings (Pinus ponderosa var. scopulorum Engelm.). Study areas were established in 36 plantations across northwest and west-central Montana on Champion International Corporation's timberland (...

  18. An overview of key silvicultural information for ponderosa pine

    Treesearch

    John Fiske; John Tappeiner

    2005-01-01

    This paper provides a selected list of classical references for the important silvicultural findings for ponderosa pine, and categorizes some of the key current literature, as well as some of the older, lesser known but important literature. The paper also provides some history of scientific developments, and sources of further information.

  19. The geographic selection mosaic for ponderosa pine and crossbills: a tale of two squirrels.

    PubMed

    Parchman, Thomas L; Benkman, Craig W

    2008-02-01

    Recent research demonstrates how the occurrence of a preemptive competitor (Tamiasciurus) gives rise to a geographic mosaic of coevolution for crossbills (Loxia) and conifers. We extend these studies by examining ponderosa pine (Pinus ponderosa), which produces more variable annual seed crops than the conifers in previous studies and often cooccurs with tree squirrels in the genus Sciurus that are less specialized than Tamiasciurus on conifer seed. We found no evidence of seed defenses evolving in response to selection exerted by S. aberti, which was apparently overwhelmed by selection resulting from inner bark feeding that caused many developing cones to be destroyed. In the absence of S. aberti, defenses directed at crossbills increased, favoring larger-billed crossbills and causing stronger reciprocal selection between crossbills and ponderosa pine. However, crossbill nomadism in response to cone crop fluctuations prevents localized reciprocal adaptation by crossbills. In contrast, evolution in response to S. griseus has incidentally defended cones against crossbills, limiting the geographic range of the interaction between crossbills and ponderosa pine. Our results suggest that annual resource variation does not prevent competitors from shaping selection mosaics, although such fluctuations likely prevent fine-scale geographic differentiation in predators that are nomadic in response to resource variability.

  20. The Fort Valley Experimental Forest, ponderosa pine, and wildlife habitat research (P-53)

    Treesearch

    David R. Patton

    2008-01-01

    Wildlife research at the Fort Valley Experimental Forest began with studies to determine how to control damage by wildlife and livestock to ponderosa pine (Pinus ponderosa) reproduction and tree growth. Studies on birds, small mammals, and mule deer (Odocoileus hemionus) browsing were initiated in the early 1930s and 1940s but these were short term efforts to develop...

  1. Identifying changes in tree form for harvested ponderosa pine in the Black Hills

    Treesearch

    Michael S. Williams; Raymond L. Czaplewski; Don L. Martinez

    1996-01-01

    Recent underestimates of total volume for timber sales in the Black Hills National Forest prompted analysis of two felled ponderosa pine (Pinus ponderosa Laws.) data sets that were collected approximately 10 years apart. Though neither data set collected was a representative sample of the Black Hills, both were similar in terms of diameter at breast height and total...

  2. Effects of fire exclusion on forest structure and composition in unlogged ponderosa pine/Douglas-fir forests

    Treesearch

    Eric G. Keeling; Anna Sala; Thomas H. DeLuca

    2006-01-01

    Research to date on effects of fire exclusion in ponderosa pine (Pinus ponderosa) forests has been limited by narrow geographical focus, by confounding effects due to prior logging at research sites, and by uncertainty from using reconstructions of past conditions to infer changes. For the work presented here, reference stands in unlogged ponderosa...

  3. Physiological adjustment of two full-sib families of ponderosa pine to elevated CO2

    Treesearch

    Nancy Grulke; J.L. Hom; S.W. Roberts

    1993-01-01

    Seeds from two full-sib families of ponderosa pine (Pinus ponderosa) with known differences in growth rates were germinated and grown in an ambient (350 µl l-1) or elevated (700 µl I-1) CO2 concentration. Gas exchange at both ambient and elevated CO2...

  4. Decadal Recruitment and Mortality of Ponderosa pine Predicted for the 21st Century Under five Downscaled Climate Change Scenarios

    NASA Astrophysics Data System (ADS)

    Ironside, K. E.; Cole, K. L.; Eischeid, J. K.; Garfin, G. M.; Shaw, J. D.; Cobb, N. S.

    2008-12-01

    Ponderosa pine (Pinus ponderosa var. scopulorum) is the dominant conifer in higher elevation regions of the southwestern United States. Because this species is so prominent, southwestern montane ecosystems will be significantly altered if this species is strongly affected by future climate changes. These changes could be highly challenging for land management agencies. In order to model the consequences of future climates, 20th Century recruitment events and mortality for ponderosa pine were characterized using measures of seasonal water balance (precipitation - potential evapotranspiration). These relationships, assuming they will remain unchanged, were then used to predict 21st Century changes in ponderosa pine occurrence in the southwest. Twenty-one AR4 IPCC General Circulation Model (GCM) A1B simulation results were ranked on their ability to simulate the later 20th Century (1950-2000 AD) precipitation seasonality, spatial patterns, and quantity in the western United States. Among the top ranked GCMs, five were selected for downscaling to a 4 km grid that represented a range in predictions in terms of changes in water balance. Predicted decadal changes in southwestern ponderosa pine for the 21st Century for these five climate change scenarios were calculated using a multiple quadratic logistic regression model. Similar models of other western tree species (Pinus edulis, Yucca brevifolia) predicted severe contractions, especially in the southern half of their ranges. However, the results for Ponderosa pine suggested future expansions throughout its range to both higher and lower elevations, as well as very significant expansions northward.

  5. Competitive Effects of Various Grasses and Forbs on Ponderosa Pine Seedlings

    Treesearch

    Katherine J. Elliott; Alan S. White

    1987-01-01

    Competition between ponderosa pine seedlings and various grasses and forbs was studied on :I site in northern Arizona burned in 1982 by a wildfire. Two-year-old pine seedlings were planted in 3.05 x 3.05 m plots in April 1983, followed by the sowing of grass and forb seeds on the same plots in July 1983 after summer rains had begun. Predawn xylem water potential...

  6. Quantifying post-fire ponderosa pine snags using GIS techniques on scanned aerial photographs

    NASA Astrophysics Data System (ADS)

    Kent, Kevin

    Snags are an important component of forest ecosystems because of their utility in forest-nutrient cycling and provision of critical wildlife habitat, as well as associated fuel management concerns relating to coarse woody debris (CWD). Knowledge of snag and CWD trajectories are needed for land managers to plan for long-term ecosystem change in post-fire regimes. This need will likely be exacerbated by increasingly warm and dry climatic conditions projected for the U.S. Southwest. One of the best prospects for studying fire-induced landscape change beyond the plot scale, but still at a resolution sufficient to resolve individual snags, is to utilize the available aerial photography record. Previous field-based studies of snag and CWD loads in the Southwest have relied on regional chronosequences to judge the recovery dynamic of ponderosa pine (Pinus ponderosa) burns. This previous research has been spatially and temporally restricted because of field survey extent limitations and uncertainty associated with the chronosequence approach (i.e., space-for-time substitution), which does not consider differences between specific site conditions and histories. This study develops highly automated methods for remotely quantifying and characterizing the spatial and temporal distribution of large snags associated with severe forest fires from very high resolution (VHR) landscape imagery I obtained from scans of aerial photos. Associated algorithms utilize the sharp edges, shape, shadow, and contrast characteristics of snags to enable feature recognition. Additionally, using snag shadow length, image acquisition time, and location information, heights were estimated for each identified snag. Furthermore, a novel solution was developed for extracting individual snags from areas of high snag density by overlaying parallel lines in the direction of the snag shadows and extracting local maxima lines contained by each snag polygon. Field survey data coincident to imagery coverage

  7. Early brush control promotes growth of ponderosa pine planted on bulldozed site

    Treesearch

    Jay R. Bentley; Stanley B. Carpenter; David A. Blakeman

    1971-01-01

    Test plots in a brushfield near Mount Shasta, California, were cleared by bulldozing in 1961, and planted with ponderosa pine seedlings in 1962. Brush regrowth was subjected to varying levels of control by spraying with herbicides. In the first 5 years, brush control definitely promoted the growth of pine seedlings. And this early control also promises to reduce the...

  8. Seven chemicals fail to protect Ponderosa pine from Armillaria root disease in central Washington.

    Treesearch

    Gregory M. Filip; Lewis F. Roth

    1987-01-01

    Chemicals were applied once to the root collars of small-diameter ponderosa pine (Pinus ponderosa Dougl. ex Laws.) to prevent mortality caused by Armillaria obscura (Pers.) Herink Roll-Harisen (A. meilea sensu lato). After 10 years, none of the 15 treatments appeared to reduce mortality in treated trees vs. untreated trees....

  9. Spatial distribution of ponderosa pine seedlings along environmental gradients within burned areas in the Black Hills, South Dakota

    Treesearch

    V. H. Bonnet; A. W. Schoettle; W. D. Shepperd

    2004-01-01

    In 2000, the Jasper fire in the Black Hills, SD, created a mosaic of burned and unburned patches of different sizes within the contiguous ponderosa pine forest. To study the spatial regeneration of ponderosa pine seedlings and the ecological gradients existing between burned and unburned areas two years after fire, we used a transect approach. We demonstrated that...

  10. Ponderosa pine lumber recovery in Lakeview, Oregon area.

    Treesearch

    E.E. Matson

    1955-01-01

    During January 1955, a study was made at the American Forest Products Corporation mill in Lakeview, Oregon, to determine the grades of lumber that can be expected from ponderosa pine timber grown in the Lakeview area of Oregon. All logs cut for a full week at this mill were scaled and graded and all the lumber produced was graded and tallied on the green chain by...

  11. Growth of ponderosa pine by keen tree class.

    Treesearch

    Philip A. Briegleb

    1943-01-01

    Every forester who works in the ponderosa pine woods is impressed by the tremendous range in size, quality, age, and thrift of the trees found in the virgin forest. So great is this variation from tree to tree that stand averages mean little to the timber marker trying to select trees of high value and insect risk for cutting, and at the same time reserve for future...

  12. SEASONAL PATTERNS OF FINE ROOT PRODUCTION AND TURNOVER IN PONDEROSA PINE STANDS OF DIFFERENT AGES

    EPA Science Inventory

    Root minirhizotron tubes were installed in two ponderosa pine (Pinus ponderosa Laws.) stands around three different tree age classes (16, 45, and > 250 yr old) to examine root spatial distribution in relation to canopy size and tree distribution, and to determine if rates of fine...

  13. Living artifacts: The ancient ponderosa pines of the West

    Treesearch

    Stephen F. Arno; Lars Östlund; Robert E. Keane

    2008-01-01

    Until late in the nineteenth century, magnificent ponderosa pine forests blanketed much of the inland West. They covered perhaps 30 million acres, an area the size of New York state, spreading across the mountains of New Mexico, Arizona, and California and flourishing throughout the eastern Cascades, the intermountain Pacific Northwest, and the Rocky Mountains...

  14. Influence of ponderosa pine overstory on forage quality in the Black Hills, South Dakota

    Treesearch

    Kieth E. Severson; Daniel W. Uresk

    1988-01-01

    Forage quality was assessed in pole and sapling ponderosa pine (Pinus ponderosa) stands growing at five stocking levels - 0, 5. 14, 23, and unthinned (which approximated 40 m2/ha basal area)-in the Black Hills of South Dakota. Crude protein, acid detergent fiber, acid detergent lignin, ash, calcium, and phosphorus were...

  15. Thinning decreases mortality and increases growth of Ponderosa pine in northeastern California

    Treesearch

    Gary O. Fiddler; Troy A. Fiddler; Dennis R. Hart; Philip M. McDonald

    1989-01-01

    Overstocked 70- to 90-year-old stands of ponderosa pine on medium- to low-quality sites were thinned in 1980 to 40, 55, and 70 percent of normal basal area and compared to an unthinned control. Mortality, diameter, and height in these northern California stands were measured annually from 1980 to 1987. After 8 years, mortality, primarily from mountain pine beetle (

  16. Reintroducing fire into a ponderosa pine forest with and without cattle grazing: understory vegetation response

    Treesearch

    Becky K. Kerns; Michelle Buonopane; Walter G. Thies; Christine. Niwa

    2011-01-01

    Reestablishing historical fire regimes is a high priority for North American coniferous forests, particularly ponderosa pine (Pinus ponderosa) ecosystems. These forests are also used extensively for cattle (Bos spp.) grazing. Prescribed fires are being applied on or planned for millions of hectares of these forests to reduce...

  17. The 2002 Rodeo-Chediski Wildfire's impacts on southwestern ponderosa pine ecosystems, hydrology, and fuels

    Treesearch

    Peter F. Ffolliott; Cody L. Stropki; Hui Chen; Daniel G. Neary

    2011-01-01

    The Rodeo-Chediski Wildfire burned nearly 462,600 acres in north-central Arizona in the summer of 2002. The wildfire damaged or destroyed ecosystem resources and disrupted the hydrologic functioning within the impacted ponderosa pine (Pinus ponderosa) forests in a largely mosaic pattern. Impacts of the wildfire on ecosystem resources, factors important to hydrologic...

  18. Biomass estimators for thinned second-growth ponderosa pine trees.

    Treesearch

    P.H. Cochran; J.W. Jennings; C.T. Youngberg

    1984-01-01

    Usable estimates of the mass of live foliage and limbs of sapling and pole-sized ponderosa pine in managed stands in central Oregon can be obtained with equations using the logarithm of diameter as the only independent variable. These equations produce only slightly higher root mean square deviations than equations that include additional independent variables. A...

  19. Postive seedling-shrub relationships in natural regeneration of ponderosa pine

    Treesearch

    Christopher R. Keyes; Douglas A. Maguire

    2005-01-01

    An understanding of natural regeneration processes, and the stand structural features that influence those processes, is vital to attaining goals associated with natural regeneration. This paper discusses natural regeneration concepts and the interactions that occur between shrubs and natural regeneration of ponderosa pine. The interactions observed recently in a...

  20. An analysis of genetic architecture in populations of Ponderosa Pine

    Treesearch

    Yan B. Linhart; Jeffry B. Mitton; Kareen B. Sturgeon; Martha L. Davis

    1981-01-01

    Patterns of genetic variation were studied in three populations of ponderosa pine in Colorado by using electrophoretically variable protein loci. Significant genetic differences were found between separate clusters of trees and between age classes within populations. In addition, data indicate that differential cone production and differential animal damage have...

  1. Old Black Hills ponderosa pines tell a story

    Treesearch

    Matthew J. Bunkers; L. Ronald Johnson; James R. Miller; Carolyn Hull Sieg

    1999-01-01

    A single ponderosa pine tree found in the central Black Hills of SouthDakota revealed its age of more than 700 years by its tree rings taken from coring in 1992. The purpose of this study was to examine historic climatic patterns from the 13th century through most of the 20th century as inferred from ring widths of this and other nearby trees. The steep, rocky site...

  2. Suitability of live and fire-killed small-diameter ponderosa and lodgepole pine trees for manufacturing a new structural wood composite

    Treesearch

    J.M. Linton; H.M. Barnes; R.D. Seale; P.D. Jones; E. Lowell; S.S. Hummel

    2010-01-01

    Finding alternative uses for raw material from small-diameter trees is a critical problem throughout the United States. In western states, a lack of markets for small-diameter ponderosa pine (Pinus ponderosa) and lodgepole pine (Pinus contorta ) can contribute to problems associated with overstocking. To test the feasibility of...

  3. Observations of bird numbers and species following a historic wildfire in Arizona ponderosa pine forests

    Treesearch

    Peter F. Ffolliott; Cody L. Stropki; Hui Chen; Daniel G. Neary

    2009-01-01

    The Rodeo-Chediski Wildfire, the largest in Arizona's history, damaged or destroyed ecosystem resources or disrupted ecosystem functioning in a mostly mosaic pattern throughout the ponderosa pine (Pinus ponderosa) forests exposed to the burn. Impacts of the wildfire on the occurrence of birds and their diversities were studied on...

  4. Height intercept for estimating site index in young ponderosa pine plantations and natural stands

    Treesearch

    William W. Oliver

    1972-01-01

    Site index is difficult to estimate with any reliability in ponderosa pine (Pinus ponderosa Laws.) stands below 20 yeas old. A method of estimating site index based on 4-year height intercepts (total length of the first four internodes above breast height) is described. Equations based on two sets of published site-index curves were developed. They...

  5. Isozyme markers associated with O(3) tolerance indicate shift in genetic structure of ponderosa and Jeffrey pine in Sequoia National Park, California.

    PubMed

    Staszak, J; Grulke, N E; Marrett, M J; Prus-Glowacki, W

    2007-10-01

    Effects of canopy ozone (O(3)) exposure and signatures of genetic structure using isozyme markers associated with O(3) tolerance were analyzed in approximately 20-, approximately 80-, and >200-yr-old ponderosa (Pinus ponderosa Dougl. ex Laws.) and Jeffrey pine (Pinus jeffreyi Grev. & Balf.) in Sequoia National Park, California. For both species, the number of alleles and genotypes per loci was higher in parental trees relative to saplings. In ponderosa pine, the heterozygosity value increased, and the fixation index indicated reduction of homozygosity with increasing tree age class. The opposite tendencies were observed for Jeffrey pine. Utilizing canopy attributes known to be responsive to O(3) exposure, ponderosa pine was more symptomatic than Jeffrey pine, and saplings were more symptomatic than old growth trees. We suggest that these trends are related to differing sensitivity of the two species to O(3) exposure, and to higher O(3) exposures and drought stress that younger trees may have experienced during germination and establishment.

  6. Fire effects on ponderosa pine soils and their management implications

    Treesearch

    W.W. Covington; S.S. Sackett

    1990-01-01

    Fire in southwestern ponderosa pine induces changes in soil properties including decreasing the amount of nutrients stored in fuels (forest floor, woody litter, and understory vegetation) increasing the amount of nutrients on the soil surface (the "ashbed effect"), and increasing the inorganic nitrogen and moisture content in the mineral soil. Soil...

  7. Seasonal changes in above- and belowground carbohydrate concentrations of ponderosa pine along a pollution gradient

    Treesearch

    Nancy E. Grulke; Chris P. Andersen; William E. Hogsett

    2001-01-01

    Seasonal patterns of carbohydrate concentration in coarse and fine roots, stem or bole, and foliage of ponderosa pine (Pinus ponderosa Laws) were described across five treeage classes from seedlings to mature trees at an atmospherically clean site. Relative to all other tree-age classes, seedlings exhibited greater tissue carbohydrate concentration...

  8. Prescribed fire effects on bark beetle activity and tree mortality in southwestern ponderosa pine forests

    Treesearch

    C.R. Breece; T.E. Kolb; B.G. Dickson; J.D. McMillin; K.M. Clancey

    2008-01-01

    Prescribed fire is an important tool in the management of ponderosa pine (Pinus ponderosa Dougl. ex Laws.) forests, yet effects on bark beetle (Coleoptera: Curculionidae, Scolytinae) activity and tree mortality are poorly understood in the southwestern U.S. We compared bark beetle attacks and tree mortality between paired prescribed-burned and...

  9. Fuel and stand characteristics in ponderosa pine infested with mountain pine beetle, Ips spp., and southwestern dwarf mistletoe in Colorado's northern Front Range

    Treesearch

    Jennifer Gene Klutsch

    2008-01-01

    The effect of forest disturbances, such as bark beetles and dwarf mistletoes, on fuel dynamics is important for understanding forest dynamics and heterogeneity. Fuel loads and other fuel parameters were assessed in areas of ponderosa pine (Pinus ponderosa Laws.) infested with southwestern dwarf mistletoe (Arceuthobium vaginatum...

  10. Pezizalean mycorrhizas and sporocarps in ponderosa pine (Pinus ponderosa) after prescribed fires in eastern Oregon, USA.

    Treesearch

    K.E. Fujimura; J.E. Smith; T.R. Horton; N.S. Weber; J.W. Spatafora

    2005-01-01

    Post-fire Pezizales fruit commonly in many forest types after fire. The objectives of this study were to determine which Pezizales appeared as sporocarps after a prescribed fire in the Blue Mountains of eastern Oregon, and whether species of Pezizales formed mycorrhizas on ponderosa pine, whether or not they were detected from sporocarps. Forty-two sporocarp...

  11. Photosynthetic response, survival, and growth of three ponderosa pine stocktypes under water stress enhanced by vegetative competition

    Treesearch

    Jeremiah R. Pinto; John D. Marshall; R. Kasten Dumroese; Anthony S. Davis; Douglas R. Cobos

    2012-01-01

    Selecting the proper stock type for reforestation on dry sites can be critical for the long-term survival and growth of seedlings. In this study, we use a novel approach to understand stock type selection on a site where drought was induced with vegetative competition. Three ponderosa pine (Pinus ponderosa Lawson & C. Lawson var. ponderosa C. Lawson) seedling stock...

  12. Presence of carbaryl in the smoke of treated lodgepole and ponderosa pine bark

    Treesearch

    Chris J. Peterson; Sheryl L. Costello

    2013-01-01

    Lodgepole and ponderosa pine trees were treated with a 2% carbaryl solution at recreational areas near Fort Collins, CO, in June 2010 as a prophylactic bole spray against the mountain pine beetle. Bark samples from treated and untreated trees were collected one day following application and at 4-month intervals for one year. The residual amount of carbaryl was...

  13. Recovery of young ponderosa pines damaged by herbicide spraying

    Treesearch

    Jay R. Bentley; David A. Blakeman; Stanley B. Carpenter

    1971-01-01

    Foliage injury and over-all tree damage to ponderosa pine plantations from aerial sprays of 2,4,5-T in 1965 were evaluated. Damage in 1966 was compared to tree growth from 1966 to 1968. The herbicide treatment caused above-normal damage to young trees when applied in September in a year with above-average summer precipitation and freezing weather soon after treatment....

  14. Ten-year observations on pruned ponderosa and jeffrey pine

    Treesearch

    Donald T. Gordon

    1959-01-01

    In 1956 Hallinl reported some observed effects of pruning ponderosa and Jeffrey pine based on the response after 5 years. This report is an extension of those observations, after a 10-year reexamination of the experiment. Hallin suggested that half the live crown, or six-tenths of total height of the tree could be pruned without adverse effects on growth. Pruning out...

  15. Evaluation of small-diameter ponderosa pine logs in bending

    Treesearch

    Debra Larson; Richard Mirth; Ronald Wolfe

    2004-01-01

    Ninety-nine roundwood bending specimens were tested over the course of seven months beginning in early 2002 at the College of Engineering and Technology at Northern Arizona University in Flagstaff, Arizona. These specimens were taken from 5.0- to 12.7-inch diameter at breast height ponderosa pine trees cut during the summer of 2001 from Unit 16 of the Fort Valley...

  16. Twenty-year growth of ponderosa pine saplings thinned to five spacings in central Oregon.

    Treesearch

    Barrett James W.

    1982-01-01

    Diameter, height, and volume growth and yield are given for plots thinned to 1000, 500, 250, 125, and 62 trees per acre in a 40- to 70-year-old stand of suppressed ponderosa pine (Pinus ponderosa Dougl. ex Laws.) saplings in central Oregon. Trees averaged about 1-inch in diameter and 8 feet in height at the time of thinning. Considerations for...

  17. Trends in snag populations in drought-stressed mixed-conifer and ponderosa pine forests (1997-2007)

    Treesearch

    Joseph L. Ganey; Scott C. Vojta

    2012-01-01

    Snags provide important biological legacies, resources for numerous species of native wildlife, and contribute to decay dynamics and ecological processes in forested ecosystems. We monitored trends in snag populations from 1997 to 2007 in drought-stressed mixed-conifer and ponderosa pine (Pinus ponderosa Dougl. ex Laws) forests, northern Arizona. Median snag density...

  18. Translocation of 14-C in ponderosa pine seedlings

    Treesearch

    Robert R. Ziemer

    1971-01-01

    The movement of 14-C from the old needles to the roots, and later to the new needles, was measured in 2-year-old ponderosa pine seedlings. The seedlings were in one of three growth stages at the time of the feeding of 14-CO-2: 9 days before spring bud break with no root activity; 7 days before spring bud break with high root activity; and 7 days after spring bud break...

  19. Silviculture of southwestern ponderosa pine: The status of our knowledge

    Treesearch

    Gilbert H. Schubert

    1974-01-01

    Describes the status of our knowledge of ponderosa pine silviculture in the southwestern States of Arizona, Colorado, New Mexico, and Utah. Economic value, impact on other uses, and the timber resource are discussed first, followed by ecological background, site quality, growth and yield, and silviculture and management. Relevant literature is discussed along with...

  20. Variations in the monoterpene composition of ponderosa pine wood oleoresin

    Treesearch

    Richard H. Smith

    1964-01-01

    A wide range in quantitative composition of the wood oleoresin monoterpenes was found among 64 ponderosa pines in the central Sierra Nevada by gas chromatographic analysis. An inverse relationship was found in the amount of β-pinene and Δ3-carene. Practically no difference in composition could be associated with (a) type of...

  1. Alternative ponderosa pine restoration treatments in the western United States

    Treesearch

    James McIver; Phillip Weatherspoon; Carl Edminster

    2001-01-01

    Compared to presettlement times, many ponderosa pine forests of the United States are now more dense and have greater quantities of fuels. Widespread treatments are needed in these forests to restore ecological integrity and to reduce the risk of uncharacteristically severe fires. Among possible restorative treatments, however, the appropriate balance among cuttings,...

  2. Climate change will restrict ponderosa pine forest regeneration in the 21st century in absence of disturbance

    NASA Astrophysics Data System (ADS)

    Petrie, M. D.; Bradford, J. B.; Hubbard, R. M.; Lauenroth, W. K.; Andrews, C.

    2016-12-01

    The persistence of ponderosa pine forests and the ability for these forests to colonize new habitats in the 21st century will be influenced by how climate change supports ponderosa pine regeneration through the demographic processes of seed production, germination and survival. Yet, the way that climate change may support or restrict the frequency of successful regeneration is unclear. We developed a quantitative, criteria-based framework to estimate ponderosa pine regeneration potential (RP: a metric from 0-1) in response to climate forcings and environmental conditions. We used the SOILWAT ecosystem water balance model to simulate drivers of air and soil temperature, evaporation and soil moisture availability for 47 ponderosa pine sites across the western United States, using meteorological data from 1910-2014, and projections from nine General Circulation Models and the RCP 8.5 emissions scenario for 2020-2099. Climate change simulations increased the success of early developmental stages of seed production and germination, and supported 49.7% higher RP in 2020-2059 compared to averages from 1910-2014. As temperatures increased in 2060-2099, survival scores decreased, and RP was reduced by 50.3% compared to 1910-2014. Although the frequency of years with high RP did not change in 2060-2099 (12% of years), the frequency of years with very low RP increased from 25% to 58% of years. Thus, climate change will initially support higher RP and more favorable years in 2020-2059, yet will reduce average RP and the frequency of years with moderate regeneration support in 2060-2099. Forest regeneration is complex and not fully-understood, but our results suggest it is likely that climate change alone will instigate restrictions to the persistence and expansion of ponderosa pine in the 21st century.

  3. CARBON STORAGE AND FLUXES IN PONDEROSA PINE AT DIFFERENT SUCCESSIONAL STAGES

    EPA Science Inventory

    We compared carbon storage and fluxes in young and old ponderosa pine stands in Oregon, including plant and soil storage, net primary productivity, respiration fluxes, and eddy flux estimates of net ecosystem exchange. The young site (Y site) was previously an old-growth pondero...

  4. Pheromone interruption of pine engraver, Ips pini, by pheromones of mountain pine beetle, Dendroctonus ponderosae (Coleoptera: Scolytidae)

    Treesearch

    Daniel R. Miller; John H. Borden

    2000-01-01

    The effect of pheromones of Dendroctonus ponderosae Hopkins on the attraction of Ips pini (Say) to its pheromone, ipsdienol, was investigated in stands of lodgepole pine. The mixture of cis- and trans-verbenol significantly reduced catches of I. pini in traps baited with...

  5. O3 uptake and drought stress effects on carbon acquisition of ponderosa pine in natural stands

    Treesearch

    N.E. Grulke; H.K. Preisler; C. Rose; J. Kirsch; L. Balduman

    2002-01-01

    • The effect of O3 exposure or uptake on carbon acquisition (net assimilation (A) or gross photosynthesis (Pg)), with and without drought stress, is reported here in 40-yr-old-ponderosa pine (Pinus ponderosa) trees. • Maximum daily gas exchange was...

  6. EFFECTS OF ELEVATED CO2 AND N-FERTILIZATION ON SURVIVAL OF PONDEROSA PINE FINE ROOTS

    EPA Science Inventory

    We used minihizaotrons to assess the effects of elevated CO2N and season on the life-span of ponderosa pine (Pinus ponderosa Dougl. Ex Laws.) fine roots. CO2 levels were ambient air (A), ambient air + 175 ?mol mol-1 (A+175) and ambient air + 350 ?mol mol-1 (A+350). N treatments ...

  7. Response of understory species to changes in ponderosa pine stocking levels in the Black Hills

    Treesearch

    Daniel W. Uresk; Kieth E. Severson

    1998-01-01

    The objective of this study was to test the hypothesis that there are no differences in understory production, by species, due to stocking levels of Pinus ponderosa (ponderosa pine). Understory production was estimated by species, on 3 replicates each of 8 growing stock levels, ranging from clearcuts to unthinned stands, in both sapling- and pole-...

  8. Crossdated fire histories (1650-1900) from ponderosa pine-dominated forests of Idaho and western Montana

    Treesearch

    Emily K. Heyerdahl; Penelope Morgan; James P. Riser

    2008-01-01

    For a broader study of the climate drivers of regional-fire years in the Northern Rockies, we reconstructed a history of surface fires at 21 sites in Idaho and western Montana. We targeted sites that historically sustained frequent surface fires and were dominated or codominated by ponderosa pine (Pinus ponderosa P. & C. Lawson). Our...

  9. Growth of a 45-year-old ponderosa pine plantation: An Arizona case study (P-53)

    Treesearch

    Peter F. Ffolliott; Gerald J. Gottfried; Cody L. Stropki; L. J. Heidmann

    2008-01-01

    Information on the growth of forest plantations is necessary for planning of ecosystem-based management of the plantations. This information is also useful in validating or refining computer simulators that estimate plantation growth into the future. Such growth information has been obtained from a 45-year-old ponderosa pine (Pinus ponderosa) plantation in the Hart...

  10. Estimation of crown cover in interior ponderosa pine stands: Effects of thinning and prescribed fire

    Treesearch

    Nicholas Vaughn; Martin W. Ritchie

    2005-01-01

    We evaluated the relationship between crown cover measured with a vertical sight tube and stand basal area per acre in treated (thinned, burned, and thinned and burned) and untreated interior ponderosa pine (Pinus ponderosa P. & C. Lawson) stands in northeastern California. Crown cover was significantly related to basal area at the plot level and...

  11. Cervid forage utilization in noncommercially thinned ponderosa pine forests

    USGS Publications Warehouse

    Gibbs, M.C.; Jenks, J.A.; Deperno, C.S.; Sowell, B.F.; Jenkins, Kurt J.

    2004-01-01

    To evaluate effects of noncommercial thinning, utilization of forages consumed by elk (Cervus elaphus L.), mule deer (Odocoileus hemionus Raf.), and white-tailed deer (Odocoileus virginianus Raf.) was measured in ponderosa pine (Pinus ponderosa P. & C. Lawson) stands in Custer State Park, S. D. Treatments consisted of unthinned (control; 22 to 32 m2/ha basal area), moderately thinned (12 to 22 m2/ha basal area), and heavily thinned (3 to 13 m2/ha basal area) stands of ponderosa pine. During June, July, and August, 1991 and 1992, about 7,000 individual plants were marked along permanent transects and percent-weight-removed by grazing was ocularly estimated. Sample plots were established along transects and plants within plots were clipped to estimate standing biomass. Pellet groups were counted throughout the study area to determine summer habitat use of elk and deer. Diet composition was evaluated using microhistological analysis of fecal samples. Average percent-weight-removed from all marked plants and percent-plants-grazed were used to evaluate forage utilization. Standing biomass of graminoids, shrubs, and forbs increased (P 0.05) across treatments. Forb use averaged less than 5% within sampling periods when measured as percent-weight-removed and percent-of-plants grazed and did not differ among treatments. Results of pellet group surveys indicated that cervids were primarily using meadow habitats. When averaged over the 2 years, forbs were the major forage class in deer diets, whereas graminoids were the major forage class in diets of elk.

  12. Influences of thinning, prescribed burning, and wildfire on soil processes and properties in southwestern ponderosa pine forests: A retrospective study

    Treesearch

    Kevin C. Grady; Stephen C. Hart

    2006-01-01

    Following Euro-American settlement in the late 1800s, fire suppression and livestock grazing in ponderosa pine-bunchgrass ecosystems of the southwestern US resulted in the replacement of grass openings with dense stands of ponderosa pine. This, in turn, has led to apparent decreases in decomposition, net N mineralization, and soil respiration (i.e., net soil CO2 efflux...

  13. Green foliage losses from ponderosa pines induced by Abert squirrels and snowstorms: A comparison. [Sciurus aberti; Pinus pondersosa

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Allred, W.S.; Gaud, W.S.

    1993-01-01

    Abert squirrels (Sciurus aberti) are obligate herbivores on ponderosa pine (Pinus ponderosa). The inner bark of pine shoots is considered one of the predominant food resources obtained by foraging squirrels. As squirrels forage for this resource they induce green needle losses from chosen feed trees. Amounts of induced green needle losses appear to vary according to the availability of alternative foods and squirrel population densities. Weather also induces green needle losses to ponderosa pines. Results of this study indicate that, at least in some years, heavy snowstorms can induce greater amounts of green needle losses than squirrels. Squirrel herbivory wasmore » not indicated as a factor in any tree mortality. However, losses due to snowstorms are more severe since they may cause the actual depletion of trees in the forest because of the tree mortality they inflict.« less

  14. Geographic variation in speed of seed germination in central Oregon ponderosa pine (Pinus ponderosa Dougl. ex Laws.).

    Treesearch

    John C. Weber; Frank C. Sorensen

    1992-01-01

    Variation in speed of seed germination was investigated among ponderosa pine trees representing 225 locations in central Oregon. Results suggested that at least some of the geographic variation is related to the severity of summer drought. In general, germination speed was greater in locations with shod, drought-limited growing seasons. Levels of geographic variation...

  15. Estimating past diameters of ponderosa pine in northern California

    Treesearch

    Robert F. Powers

    1969-01-01

    Past outside bark diameters (d.o.b.) of ponderosa pine in northern California can be estimated from the relation of bark thickness to stem diameter. An equation for estimating past d.o.b.'s has been developed. The relationship appears to be linear between 1 and 20 inches d.o.b. and differs from ratios reported for the species in other areas. Site quality and crown...

  16. Selection cutting reduces ponderosa pine losses at Pringle Falls.

    Treesearch

    Edwin L. Mowat

    1948-01-01

    One of the aims of selection cutting in ponderosa pine forests is to reduce the heavy mortality, from bark beetle attack and other causes that is common in most virgin stands. How much the loss can be reduced is shown by a 10-year record at the Pringle Falls Experimental Forest in central Oregon. Here selection cutting has reduced loss on cutting areas to only about...

  17. Restoration of the ponderosa pine ecosystem and its understory

    Treesearch

    Lee E. Hughes

    2008-01-01

    Restoration of the Mt. Logan ponderosa pine ecosystem has been on-going since 1995. This effort included tree thinning to a density based on what the tree density was in 1870. The desired plant community objectives from the Mt. Trumbull Resource Conservation Area Plan had a forest objective as 50% trees to be in old-growth - i.e., a diameter class of 20-31.9+ inch...

  18. Wood and understory production under a range of ponderosa pine stocking levels, Black Hills, South Dakota

    Treesearch

    Daniel W. Uresk; Carleton B. Edminster; Kieth E. Severson

    2000-01-01

    Stemwood and understory production (kg ha-1) were estimated during 3 nonconsecutive years on 5 growing stock levels of ponderosa pine including clearcuts and unthinned stands. Stemwood production was consistently greater at mid- and higher pine stocking levels, and understory production was greater in stands with less pine; however, there were no...

  19. Latent resilience in ponderosa pine forest: effects of resumed frequent fire

    Treesearch

    Andrew J. Larson; R. Travis Belote; C. Alina Cansler; Sean A. Parks; Matthew S. Dietz

    2013-01-01

    Ecological systems often exhibit resilient states that are maintained through negative feedbacks. In ponderosa pine forests, fire historically represented the negative feedback mechanism that maintained ecosystem resilience; fire exclusion reduced that resilience, predisposing the transition to an alternative ecosystem state upon reintroduction of fire. We evaluated...

  20. INTERACTIVE EFFECTS OF CO2 AND O3 ON A PONDEROSA PINE PLANT/LITTER/SOIL MESOCOSM

    EPA Science Inventory

    To study individual and combined impacts of two important atmospheric trace gases, CO2 and O3, on C and N cycling in forest ecosystems; a four-year experiment using a small-scale ponderosa pine (Pinus ponderosa Laws.) seedling/soil/litter system was initiated in April, 1998. Th...

  1. Postfire mortality of ponderosa pine and Douglas-fir: a review of methods to predict tree death

    Treesearch

    James F. Fowler; Carolyn Hull Sieg

    2004-01-01

    This review focused on the primary literature that described, modeled, or predicted the probability of postfire mortality in ponderosa pine (Pinus ponderosa) and Douglas-fir (Pseudotsuga menziesii). The methods and measurements that were used to predict postfire tree death tended to fall into two general categories: those focusing...

  2. Changes in snag populations in northern Arizona mixed-conifer and ponderosa pine forests, 1997-2002

    Treesearch

    Joseph L. Ganey; Scott C. Vojta

    2005-01-01

    Snags (standing dead trees) are important components of forests that contribute to ecological processes and provide habitat for many life forms. We monitored dynamics of snag populations on 1-ha plots in southwestern mixed-conifer (n = 53 plots) and ponderosa pine (Pinus ponderosa, n = 60 plots) forests in north-central Arizona from 1997 to 2002. Of...

  3. Density of large snags and logs in northern Arizona mixed-conifer and ponderosa pine forests

    Treesearch

    Joseph L. Ganey; Benjamin J. Bird; L. Scott Baggett; Jeffrey S. Jenness

    2015-01-01

    Large snags and logs provide important biological legacies and resources for native wildlife, yet data on populations of large snags and logs and factors influencing those populations are sparse. We monitored populations of large snags and logs in mixed-conifer and ponderosa pine (Pinus ponderosa) forests in northern Arizona from 1997 through 2012. We modeled density...

  4. Effectiveness of litter removal to prevent cambial kill-caused mortality in northern Arizona ponderosa pine

    Treesearch

    James F. Fowler; Carolyn Hull Sieg; Linda L. Wadleigh

    2010-01-01

    Removal of deep litter and duff from the base of mature southwestern ponderosa pine (Pinus ponderosa Laws.) is commonly recommended to reduce mortality after prescribed burns, but experimental studies that quantify the effectiveness of such practices in reducing mortality are lacking. After a pilot study on each of four sites in northern Arizona, we monitored 15-16...

  5. Characteristics of snags containing excavated cavities in northern Arizona mixed-conifer and ponderosa pine forests

    Treesearch

    Joseph L. Ganey; Scott C. Vojta

    2004-01-01

    Snags provide an important resource for a rich assemblage of cavity-nesting birds in the southwestern United States. To expand our knowledge of snag use by cavity-nesting birds in this region, we documented characteristics of snags with and without excavated cavities in mixed-conifer and ponderosa pine (Pinus ponderosa Dougl. ex Laws) forest in north...

  6. Trends in snag populations in Northern Arizona mixed-conifer and ponderosa pine forests, 1997-2012

    Treesearch

    J. L. Ganey; S. C. Vojta

    2014-01-01

    We monitored snag populations in drought-stressed mixed-conifer and ponderosa pine (Pinus ponderosa) forests, northern Arizona, at 5-yr intervals from 1997-2012. Snag density increased from 1997-2007 in both forest types, with accelerated change due to drought-related tree mortality during the period 2002-2007. Snag density declined non-significantly from 2007-2012,...

  7. Improved utilization of small-diameter ponderosa pine in glulam timber

    Treesearch

    Roland Hernandez; David W. Green; David E. Kretschmann; Steven P. Verrill

    2005-01-01

    This study involved the evaluation of ponderosa pine glulam made from lumber that was sawn from a small-diameter timber resource. Two different glulam beam depths were evaluated: 8 and 13 laminations. A comprehensive glulam test program was conducted to evaluate edgewise and flat-wise bending, shear, and tensile strength. Beam deflection was measured and a variety of...

  8. Early height growth of ponderosa pine forecasts dominance in plantations

    Treesearch

    William W. Oliver; Robert F. Powers

    1971-01-01

    Future crown class may be determined well in advance of intertree competition in plantation grown ponderosa pine. Regardless of site quality, dominant trees in 10 California plantations reached breast height ½ year sooner than codominants and 1-½ years sooner than intermediates. Dominant trees on poor sites reached breast height several years earlier than has been...

  9. Lumber recovery from small-diameter ponderosa pine from Flagstaff, Arizona

    Treesearch

    Eini C. Lowell; David W. Green

    2001-01-01

    Thousands of acres of densely stocked ponderosa pine forests surround Flagstaff, AZ. These stands are at high risk of fire, insect, and disease outbreak. Stand density management activity can be expensive, but product recovery from the thinned material could help defray removal costs. This project evaluated the yield and economic return of lumber recovered from small-...

  10. Ponderosa pine progenies: differential response to ultramafic and granitic soils

    Treesearch

    James L. Jenkinson

    1974-01-01

    Progenies of nine ponderosa pines native to one granitic and several ultramafic soils in the northern Sierra Nevada were grown on both soil types in a greenhouse. The progenies differed markedly in first-year growth on infertile ultramafic soils, but not on a fertile granitic soil. Growth differences between progenies were primarily related to differences in calcium...

  11. Area-wide application of verbenone-releasing flakes reduces mortality of whitebark pine Pinus albicaulis caused by the mountain pine beetle Dendroctonus ponderosae

    Treesearch

    Nancy E. Gillette; E. Matthew Hansen; Constance J. Mehmel; Sylvia R. Mori; Jeffrey N. Webster; Nadir Erbilgin; David L. Wood

    2012-01-01

    DISRUPT Micro-Flake Verbenone Bark Beetle Anti-Aggregant flakes (Hercon Environmental, Inc., Emigsville, Pennsylvania) were applied in two large-scale tests to assess their efficacy for protecting whitebark pine Pinus albicaulis Engelm. from attack by mountain pine beetle Dendroctonus ponderosae Hopkins (Coleoptera: Scolytinae) (MPB). At two locations, five...

  12. Pezizalean mycorrhizas and sporocarps in ponderosa pine (Pinus ponderosa) after prescribed fires in eastern Oregon, USA.

    PubMed

    Fujimura, K E; Smith, J E; Horton, T R; Weber, N S; Spatafora, J W

    2005-03-01

    Post-fire Pezizales fruit commonly in many forest types after fire. The objectives of this study were to determine which Pezizales appeared as sporocarps after a prescribed fire in the Blue Mountains of eastern Oregon, and whether species of Pezizales formed mycorrhizas on ponderosa pine, whether or not they were detected from sporocarps. Forty-two sporocarp collections in five genera (Anthracobia, Morchella, Peziza, Scutellinia, Tricharina) of post-fire Pezizales produced ten restriction fragment length polymorphism (RFLP) types. We found no root tips colonized by species of post-fire Pezizales fruiting at our site. However, 15% (6/39) of the RFLP types obtained from mycorrhizal roots within 32 soil cores were ascomycetes. Phylogenetic analyses of the 18S nuclear ribosomal DNA gene indicated that four of the six RFLP types clustered with two genera of the Pezizales, Wilcoxina and Geopora. Subsequent analyses indicated that two of these mycobionts were probably Wilcoxina rehmii, one Geopora cooperi, and one Geopora sp. The identities of two types were not successfully determined with PCR-based methods. Results contribute knowledge about the above- and below-ground ascomycete community in a ponderosa pine forest after a low intensity fire.

  13. Stability of the large tree component in treated and untreated late-seral interior ponderosa pine stands

    Treesearch

    Martin W. Ritchie; Brian M. Wing; Todd A. Hamilton

    2008-01-01

    Ponderosa pine (Pinus ponderosa Dougl. ex P. & C. Laws.) stands with late-seral features are found infrequently, owing to past management activities throughout western North America. Thus, management objectives often focus on maintaining existing late-seral stands. Observations over a 65 year period, of stands with not past history of harvest,...

  14. Recommendations for snag retention in southwestern mixed-conifer and ponderosa pine forests: History and current status

    Treesearch

    Joseph L. Ganey

    2016-01-01

    Snags provide habitat for numerous species of wildlife. Several authors have provided recommendations for snag retention in southwestern mixed-conifer and ponderosa pine (Pinus ponderosa) forests. Most recommendations were presented in terms of minimum snag density and/or size. I summarized the history of recommendations for snag retention in these forest...

  15. Woodpecker use and fall rates of snags created by killing ponderosa pine infected with dwarf mistletoe.

    Treesearch

    Catherine G. Parks; David A. Conklin; Larry Bednar; Helen. Maffei

    1999-01-01

    Ponderosa pine (Pinus ponderosa Dougl. ex Laws.) killed as part of a forest management project to reduce dwarf mistletoe (Arceuthobium sp.) in the Gila National Forest, New Mexico, were evaluated for wildlife value. One hundred and two dwarf mistletoe-infected trees were killed by basal burning, basal girdling, or by a...

  16. Evaluating the role of cutting treatments, fire and soil seed banks in an experimental framework in ponderosa pine forests of the Black Hills, South Dakota

    Treesearch

    Cody L. Wienk; Carolyn Hull Sieg; Guy R. McPherson

    2004-01-01

    Pinus ponderosa Laws. (ponderosa pine) forests have changed considerably during the past century, partly because recurrent fires have been absent for a century or more. A number of studies have explored the influence of timber harvest or burning on understory production in ponderosa pine forests, but study designs incorporating cutting and prescribed...

  17. AmeriFlux US-Me2 Metolius-intermediate aged ponderosa pine

    DOE Data Explorer

    Law, Bev [Oregon State University

    2016-01-01

    This is the AmeriFlux version of the carbon flux data for the site US-Me2 Metolius-intermediate aged ponderosa pine. Site Description - The mean stand age is 64 years old and the stand age of the oldest trees is about 100 years old. This site is one of the Metolius cluster sites with different age and disturbance classes and part of the AmeriFlux network (http://ameriflux.ornl.gov/fullsiteinfo.php?sid=88). The overstory is almost exclusively composed of ponderosa pine trees (Pinus ponderosa Doug. Ex P. Laws) with a few scattered incense cedars (Calocedrus decurrens (Torr.) Florin) and has a peak leaf area index (LAI) of 2.8 m2 m-2. Tree height is relatively homogeneous at about 16 m, and the mean tree density is approximately 325 trees ha-1 (Irvine et al., 2008). The understory is sparse with an LAI of 0.2 m2 m-2 and primarily composed of bitterbrush (Purshia tridentate (Push) DC.) and Manzanita (Arctostaphylos patula Greene). Soils at the site are sandy (69%/24%/7% sand/silt/clay at 0–0.2 m depth and 66%/27%/7% at 0.2–0.5 m depth, and 54%/ 35%/11% at 0.5–1.0 m depth), freely draining with a soil depth of approximately 1.5 m (Irvine et al., 2008; Law et al., 2001b; Schwarz et al., 2004).

  18. Frequencies of Null Alleles at Enzyme Loci in Natural Populations of Ponderosa and Red Pine

    PubMed Central

    Allendorf, Fred W.; Knudsen, Kathy L.; Blake, George M.

    1982-01-01

    Pinus ponderosa and P. resinosa population samples have mean frequencies of enzymatically inactive alleles of 0.0031 and 0.0028 at 29 and 27 enzyme loci, respectively. Such alleles are rare and are apparently maintained by selection-mutation balance. Ponderosa pine have much higher amounts of allozymic and polygenic phenotypic variation than red pine, yet both species have similar frequencies of null alleles. Thus, null alleles apparently do not contribute to polygenic variation, as has been suggested. The concordance between allozymic and polygenic variation adds support to the view that allozyme studies may be valuable in predicting the relative amount of polygenic variation in populations. PMID:17246067

  19. A comparison of bird species composition and abundance between late- and mid-seral ponderosa pine forests

    Treesearch

    T. Luke George; Steve Zack; William F. Jr. Laudenslayer

    2005-01-01

    We compared the relative abundance of bird species between two ponderosa pine (Pinus ponderosa) forests in northeastern California: one with a canopy of large old-growth trees present (Blacks Mountain Experimental Forest, BMEF) and the other with large trees essentially absent (Goosenest Adaptive Management Area, GAMA). We surveyed 24 units at BMEF...

  20. DDT spray for control of the ponderosa pine tip moth (Rhyacionia zozana [Kearfott])

    Treesearch

    Robert E. Stevens

    1965-01-01

    A water emulsion spray of DDT applied by hand sprayer to young trees infested with eggs and early-instar larvae of the ponderosa pine tip moth halted further larval activity and effectively prevented all damage.

  1. Response of dwarf mistletoe-infested ponderosa pine to thinning: 1. Sapling growth.

    Treesearch

    James W. Barrett; Lewis F. Roth

    1985-01-01

    Observations of thinned ponderosa pine infested with dwarf mistletoe over a 17-year period suggests that on average or better Sites most infested stands can be managed to produce usable wood products in reasonable time, if trends found in juvenile stands continue.

  2. Long-term post-wildfire correlates with avian community dynamics in ponderosa pine forests [Chapter J

    Treesearch

    Jamie S. Sanderlin; William M. Block; Brenda E. Strohmeyer

    2016-01-01

    We used a 10-year data set to illustrate the long-term correlates of wildfire on avian species richness in the ponderosa pine (Pinus ponderosa) forests of northern Arizona. This study was conducted in the vicinity of the Horseshoe and Hochderffer Fires, which occurred in 1996, and sampling began 1 year after the fires. Using point-count data from breeding...

  3. Estimating past diameters of several species in the ponderosa pine subregion of Oregon and Washington.

    Treesearch

    Benjamin. Spada

    1960-01-01

    In certain kinds of forest growth studies, an estimate of diameter outside bark is required for trees as of some previous date. Previously, papers were published by this Station regarding estimation of past diameters of Douglas-fir and ponderosa pine. Since they were issued, new and more complete data have been collected for most of the major species in the ponderosa...

  4. Resistance of ponderosa pine to western dwarf mistletoe in central Oregon

    Treesearch

    Robert F. Scharpf; Lewis F. Roth

    1992-01-01

    Ponderosa pines with little or no dwarf mistletoe in infested stands on the Deschutes, Ochoco, and Rogue River National Forests in Oregon were tested for resistance to dwarf mistletoe (Arceuthobium campylopodum). Small trees produced by grafting scions from the resistant and susceptible candidates onto seedling rootstock were planted in 1967-69...

  5. Economics of thinning stagnated ponderosa pine sapling stands in the pine-grass areas of central Washington.

    Treesearch

    Robert W. Sassaman; James W. Barrett; Justin G. Smith

    1972-01-01

    Present net worth values earned by investments in precommercial thinning of stagnated ponderosa pine sapling stands are reported for three stocking levels. Thirteen timber management regimes are ranked by their returns from timber only, and 22 regimes are ranked according to their returns from timber and forage, with and without the allowable cut effect.

  6. Acoustic analysis of warp potential of green ponderosa pine lumber

    Treesearch

    Xiping Wang; William T. Simpson

    2005-01-01

    This study evaluated the potential of acoustic analysis as presorting criteria to identify warp-prone boards before kiln drying. Dimension lumber, 38 by 89 mm (nominal 2 by 4 in.) and 2.44 m (8 ft) long, sawn from open-grown small-diameter ponderosa pine trees, was acoustically tested lengthwise at green condition. Three acoustic properties (acoustic speed, rate of...

  7. Structural lumber from suppressed-growth ponderosa pine from northern Arizona.

    Treesearch

    Thomas M. Gorman; David W. Green; Aldo G. Cisternas; Roland Hernandez; Eini C. Lowell

    2007-01-01

    Lumber was sawn from 150 suppressed-growth ponderosa pine trees, 6 to 16 inches in diameter, harvested near Flagstaff, Arizona. This paper presents grade recover and properties for dry 2 by 4s sawn from the logs and graded by a variety of structural grading systems. Flexural properties met or exceeded those listed in the National Design Specification. When graded as...

  8. A model of ponderosa pine growth response to prescribed burning

    Treesearch

    Elaine Kennedy Sutherland; W. Wallace Covington; Steve Andariese

    1991-01-01

    Our objective was to model the radial growth response of individual ponderosa pines to prescribed burning in northern Arizona. We sampled 188 trees from two study areas, which were burned in 1976. Within each study area, control and burned trees were of similar age, vigor, height, and competition index. At Chimney Spring, trees were older, less vigorous, and taller and...

  9. Product recovery of ponderosa pine in Arizona and New Mexico.

    Treesearch

    Thomas D. Fahey; Janet K. Ayer. Sachet

    1993-01-01

    A mill recovery study of ponderosa pine in Arizona and New Mexico showed wide variation in quality within the resource. Lumber grade ranged widely by log grade and diameter, with a major difference within grade 5 logs between old growth and young growth. Old growth produced mostly Shop and Selects grades of lumber while young growth produced mostly Dimension grades of...

  10. Mortality predictions of fire-injured large Douglas-fir and ponderosa pine in Oregon and Washington, USA

    Treesearch

    Lisa M. Ganio; Robert A. Progar

    2017-01-01

    Wild and prescribed fire-induced injury to forest trees can produce immediate or delayed tree mortality but fire-injured trees can also survive. Land managers use logistic regression models that incorporate tree-injury variables to discriminate between fatally injured trees and those that will survive. We used data from 4024 ponderosa pine (Pinus ponderosa...

  11. AUTOSAW simulations of lumber recovery for small-diameter Douglas-fir and ponderosa pine from southwestern Oregon.

    Treesearch

    R. James Barbour; Dean L. Parry; John Punches; John Forsman; Robert Ross

    2003-01-01

    Small-diameter (5- to 10-inch diameter at breast height) Douglas-fi r (Pseudotsuga menziesii (Mirb.) Franco) and ponderosa pine (Pinus ponderosa Dougl. ex Laws) trees were assessed for product potential by diagramming the location, size, and type of knots visible on the wood surface (inside bark) and using the AUTOSAW sawing simulator to evaluate...

  12. Predicting mortality of ponderosa pine regeneration after prescribed fire in the Black Hills, South Dakota, USA

    Treesearch

    Mike Battaglia; Frederick W. Smith; Wayne D. Shepperd

    2009-01-01

    Reduction of crown fire hazard in Pinus ponderosa forests in the Black Hills, SD, often focuses on the removal of overstorey trees to reduce crown bulk density. Dense ponderosa pine regeneration establishes several years after treatment and eventually increases crown fire risk if allowed to grow. Using prescribed fire to control this regeneration is...

  13. Prediction of delayed mortality of fire-damaged ponderosa pine following prescribed fires in eastern Oregon, USA.

    Treesearch

    Walter G. Thies; Douglas J. Westlina; Mark Loewen; Greg Brenner

    2006-01-01

    Prescribed burning is a management tool used to reduce fuel loads in western interior forests. Following a burn, managers need the ability to predict the mortality of individual trees based on easily observed characteristics. A study was established in six stands of mixed-age ponderosa pine (Pinus ponderosa Dougl. ex Laws.) with scattered western...

  14. Restoration of southwestern ponderosa pine forests: Implications and opportunities for wildlife

    Treesearch

    Catherine S. Wightman; Steven S. Rosenstock

    2008-01-01

    (Please note, this is an abstract only) After a century of fire suppression, livestock grazing, and even-aged timber harvest practices, forest managers in the Southwest face an enormous challenge. Millions of acres of ponderosa pine forest are extremely susceptible to uncharacteristic, high intensity wildfires, the consequences of which were amply demonstrated by...

  15. Lessons learned from fire use for restoring southwestern ponderosa pine ecosystems

    Treesearch

    Stephen S. Sackett; Sally M. Haase; Michael G. Harrington

    1996-01-01

    Since European settlement, the southwestern ponderosa pine ecosystem has experienced large scale alterations brought about by heavy grazing and timbering and a policy of attempted fire exclusion. These alterations are most evident as large increases in tree numbers and in forest floor organic matter. These changes have resulted in forest health problems, such...

  16. Seed origin and size of ponderosa pine planting stock grown at several California nurseries

    Treesearch

    Frank J. Baron; Gilbert H. Schubert

    1963-01-01

    Ponderosa pine planting stock (1-0 and 2-0) grown from five different seed collection zones in the California pine region differed noticeably in size. On the west side of the Sierra Nevada, seeds from zones above 4,000 feet yielded smaller seedlings than those from lower zones, but larger seedlings than those from east-side sources. Average dimensions (seedling weight...

  17. A comparison of the metabolism of the abortifacient compounds from Ponderosa pine needles in conditioned versus naive cattle.

    PubMed

    Welch, K D; Gardner, D R; Pfister, J A; Panter, K E; Zieglar, J; Hall, J O

    2012-12-01

    Isocupressic acid (ICA) is the abortifacient compound in ponderosa pine (Pinus ponderosa L.) needles, which can cause late-term abortions in cattle (Bos taurus). However, cattle rapidly metabolize ICA to agathic acid (AGA) and subsequent metabolites. When pine needles are dosed orally to cattle, no ICA is detected in their serum, whereas AGA is readily detected. Recent research has demonstrated that AGA is also an abortifacient compound in cattle. The observation has been made that when cattle are dosed with labdane acids for an extended time, the concentration of AGA in serum increases for 1 to 2 d but then decreases to baseline after 5 to 6 d even though they are still being dosed twice daily. Therefore, in this study we investigated whether cattle conditioned to pine needles metabolize ICA, and its metabolites, faster than naïve cattle. Agathic acid was readily detected in the serum of naïve cattle fed ponderosa pine needles, whereas very little AGA was detected in the serum of cattle conditioned to pine needles. We also compared the metabolism of ICA in vitro using rumen cultures from pine-needle-conditioned and naïve cattle. In the rumen cultures from conditioned cattle, AGA concentrations were dramatically less than rumen cultures from naïve cattle. Thus, an adaptation occurs to cattle conditioned to pine needles such that the metabolism AGA by the rumen microflora is altered.

  18. Twenty-year growth of thinned and unthinned ponderosa pine in the Methow Valley of northern Washington.

    Treesearch

    James W. Barrett

    1981-01-01

    Diameter, height and volume growth, and yield of thinned and unthinned plots are given for a suppressed, 47-year-old stand of ponderosa pine (Pinus ponderosa Dougl. ex Laws.) in the Methow Valley of northern Washington that averaged about 3 inches in diameter and 23 feet tall before thinning. Considerations are discussed for choosing tree spacing...

  19. Response of ponderosa pine plantations to competing vegetation control in Northern California, USA: A meta- analysis

    Treesearch

    Jianwei Zhang; Robert Powers; William Oliver; Young David

    2013-01-01

    A meta-analysis was performed to determine response of stand basal area growth to competing vegetation control (CVC) in ponderosa pine (Pinus ponderosa) plantations grown at 29 sites across northern California. These studies were installed during the last 50 years on site indices from 11 to 35 m at 50 years and often included other treatments...

  20. Efficacy of “Verbenone Plus” for protecting ponderosa pine trees and stands from Dendroctonus brevicomis (Coleoptera: Curculionidae) attack in British Columbia and California

    Treesearch

    Christopher J. Fettig; Stephen R. McKelvey; Christopher P. Dabney; Dezene P.W. Huber; Cameron C. Lait; Donald L Fowler; John H. Borden

    2012-01-01

    The western pine beetle, Dendroctonus brevicomis LeConte (Coleoptera: Curculionidae, Scolytinae), is a major cause of ponderosa pine, Pinus ponderosa Douglas ex Lawson, mortality in much of western North America. We review several years of research that led to the identification of Verbenone Plus, a novel four-component...

  1. Native root xylem embolism and stomatal closure in stands of Douglas-fir and ponderosa pine: mitigation by hydraulic redistribution.

    PubMed

    Domec, J-C; Warren, J M; Meinzer, F C; Brooks, J R; Coulombe, R

    2004-09-01

    Hydraulic redistribution (HR), the passive movement of water via roots from moist to drier portions of the soil, occurs in many ecosystems, influencing both plant and ecosystem-water use. We examined the effects of HR on root hydraulic functioning during drought in young and old-growth Douglas-fir [ Pseudotsuga menziesii (Mirb.) Franco] and ponderosa pine ( Pinus ponderosa Dougl. Ex Laws) trees growing in four sites. During the 2002 growing season, in situ xylem embolism, water deficit and xylem vulnerability to embolism were measured on medium roots (2-4-mm diameter) collected at 20-30 cm depth. Soil water content and water potentials were monitored concurrently to determine the extent of HR. Additionally, the water potential and stomatal conductance ( g(s)) of upper canopy leaves were measured throughout the growing season. In the site with young Douglas-fir trees, root embolism increased from 20 to 55 percent loss of conductivity (PLC) as the dry season progressed. In young ponderosa pine, root embolism increased from 45 to 75 PLC. In contrast, roots of old-growth Douglas-fir and ponderosa pine trees never experienced more than 30 and 40 PLC, respectively. HR kept soil water potential at 20-30 cm depth above -0.5 MPa in the old-growth Douglas-fir site and -1.8 MPa in the old-growth ponderosa pine site, which significantly reduced loss of shallow root function. In the young ponderosa pine stand, where little HR occurred, the water potential in the upper soil layers fell to about -2.8 MPa, which severely impaired root functioning and limited recovery when the fall rains returned. In both species, daily maximum g(s) decreased linearly with increasing root PLC, suggesting that root xylem embolism acted in concert with stomata to limit water loss, thereby maintaining minimum leaf water potential above critical values. HR appears to be an important mechanism for maintaining shallow root function during drought and preventing total stomatal closure.

  2. Recovering lost ground: effects of soil burn intensity on nutrients and ectomycorrhiza communities of ponderosa pine seedlings

    Treesearch

    Ariel D. Cowan; Jane E. Smith; Stephen A. Fitzgerald

    2016-01-01

    Fuel accumulation and climate shifts are predicted to increase the frequency of high-severity fires in ponderosa pine (Pinus ponderosa) forests of central Oregon. The combustion of fuels containing large downed wood can result in intense soil heating, alteration of soil properties, and mortality of microbes. Previous studies show ectomycorrhizal...

  3. Service life of treated and untreated Black Hills ponderosa pine fenceposts

    Treesearch

    Donald C. Markstrom; Lee R. Gjovik

    1992-01-01

    Service-life tests indicate that ponderosa pine fenceposts treated with preservatives performed well after field exposure of 30 years. Treating plants in the Black Hills area used commercial methods to treat the posts with creosote, pentachlorophenol, and waterborne arsenicals. Test sites were in the northern Great Plains-one in the semiarid western portion near Scenic...

  4. Is self-thinning in ponderosa pine ruled by Dendroctonus bark beetles?

    Treesearch

    William W. Oliver

    1995-01-01

    Stand density of even-aged stands of ponderosa pine in California seems to be ruled by Dendroctonus bark beetles, rather than the suppressioninduced mortality common for other tree species. Size-density trajectories were plotted for 155 permanent plots in both plantations and natural stands. Bark beetle kills created a limiting Stand Density Index of...

  5. Ponderosa pine needle length: an early indicator of release treatment effectiveness

    Treesearch

    Philip M. McDonald; Carl N. Skinner; Gary O. Fiddler

    1992-01-01

    Growth responses of ponderosa pine seedlings range from fast to slow after release and often demonstrate the effectiveness of the prescribed treatments. Although several morphological parameters have have identified as being sensitive to competition, no link to future growth and treatment effectiveness has been made. Shrubs and grasses in four 1- to 3-year-old...

  6. A decadal glimpse on climate and burn severity influences on ponderosa pine post-fire recovery

    NASA Astrophysics Data System (ADS)

    Newingham, B. A.; Hudak, A. T.; Bright, B. C.; Smith, A.; Khalyani, A. H.

    2016-12-01

    Climate change is predicted to affect plants at the margins of their distribution. Thus, ecosystem recovery after fire is likely to vary with climate and may be slowest in drier and hotter areas. However, fire regime characteristics, including burn severity, may also affect vegetation recovery. We assessed vegetation recovery one and 9-15 years post-fire in North American ponderosa pine ecosystems distributed across climate and burn severity gradients. Using climate predictors derived from downscaled 1993-2011 climate normals, we predicted vegetation recovery as indicated by Normalized Burn Ratio derived from 1984-2012 Landsat time series imagery. Additionally, we collected field vegetation measurements to examine local topographic controls on burn severity and post-fire vegetation recovery. At a regional scale, we hypothesized a positive relationship between precipitation and recovery time and a negative relationship between temperature and recovery time. At the local scale, we hypothesized southern aspects to recovery slower than northern aspects. We also predicted higher burn severity to slow recovery. Field data found attenuated ponderosa pine recovery in hotter and drier regions across all burn severity classes. We concluded that downscaled climate data and Landsat imagery collected at commensurate scales may provide insight into climate effects on post-fire vegetation recovery relevant to ponderosa pine forest managers.

  7. Comparative trends in log populations in northern Arizona mixed-conifer and ponderosa pine forests following severe drought

    Treesearch

    Joseph L. Ganey; Scott C. Vojta

    2017-01-01

    Logs provide an important form of coarse woody debris in forest systems, contributing to numerous ecological processes and affecting wildlife habitat and fuel complexes. Despite this, little information is available on the dynamics of log populations in southwestern ponderosa pine (Pinus ponderosa) and especially mixed-conifer forests. A recent episode of elevated tree...

  8. A review of precipitation and temperature control on seedling emergence and establishment for ponderosa and lodgepole pine forest regeneration

    USGS Publications Warehouse

    Petrie, Matthew; Wildeman, A.M.; Bradford, John B.; Hubbard, R.M.; Lauenroth, W.K.

    2016-01-01

    The persistence of ponderosa pine and lodgepole pine forests in the 21st century depends to a large extent on how seedling emergence and establishment are influenced by driving climate and environmental variables, which largely govern forest regeneration. We surveyed the literature, and identified 96 publications that reported data on dependent variables of seedling emergence and/or establishment and one or more independent variables of air temperature, soil temperature, precipitation and moisture availability. Our review suggests that seedling emergence and establishment for both species is highest at intermediate temperatures (20 to 25 °C), and higher precipitation and higher moisture availability support a higher percentage of seedling emergence and establishment at daily, monthly and annual timescales. We found that ponderosa pine seedlings may be more sensitive to temperature fluctuations whereas lodgepole pine seedlings may be more sensitive to moisture fluctuations. In a changing climate, increasing temperatures and declining moisture availability may hinder forest persistence by limiting seedling processes. Yet, only 23 studies in our review investigated the effects of driving climate and environmental variables directly. Furthermore, 74 studies occurred in a laboratory or greenhouse, which do not often replicate the conditions experienced by tree seedlings in a field setting. It is therefore difficult to provide strong conclusions on how sensitive emergence and establishment in ponderosa and lodgepole pine are to these specific driving variables, or to investigate their potential aggregate effects. Thus, the effects of many driving variables on seedling processes remain largely inconclusive. Our review stresses the need for additional field and laboratory studies to better elucidate the effects of driving climate and environmental variables on seedling emergence and establishment for ponderosa and lodgepole pine.

  9. An improved synthetic attractant for the mountain pine beetle, Dendroctonus ponderosae Hopkins (Coleoptera: Curculionidae: Scolytinae), in northeastern California

    Treesearch

    Brian Strom; Sheri Smith; D.A. Wakarchuk

    2008-01-01

    The mountain pine beetle, Dendroctonus ponderosae Hopkins 1902, is found in pine forests throughout the western U.S., north to northern British Columbia and Alberta, Canada and south to Mexico. It causes high levels of pine mortality throughout its range. Hosts include many species of Pinus (Pinaceae); in northern California,

  10. Site-index curves for young-growth ponderosa pine in northern Arizona

    Treesearch

    Charles O. Minor

    1964-01-01

    The productive capacity or site quality of an area enters into nearly every phase of forest management from regeneration to final harvest. No standards or measures of site quality have been developed specifically for ponderosa pine in the Southwest, which handicaps the forest manager. The major objective of the present study was to develop the basic site-index curves...

  11. Ponderosa pine needle blight in eastern Oregon during 1955 and 1956.

    Treesearch

    John Hunt; T.W. Childs

    1957-01-01

    A survey of ponderosa pine needle blight during 1955 and 1956, principally in eastern Oregon (where the disease is much more abundant than in Washington), provided information on extent of damage and on association of infection with certain environmental factors. Observations were largely confined to localities where the disease was known to be more than ordinarily...

  12. Eighty-eight years of change in a managed ponderosa pine forest

    Treesearch

    Helen Y. Smith; Stephen F. Arno

    1999-01-01

    This publication gives an overview of structural and other ecological changes associated with forest management and fire suppression since the early 1900's in a ponderosa pine forest, the most widespread forest type in the Western United States. Three sources of information are presented: (1) changes seen in a series of repeat photographs taken between 1909 and...

  13. Effects of stand density on top height estimation for ponderosa pine

    Treesearch

    Martin Ritchie; Jianwei Zhang; Todd Hamilton

    2012-01-01

    Site index, estimated as a function of dominant-tree height and age, is often used as an expression of site quality. This expression is assumed to be effectively independent of stand density. Observation of dominant height at two different ponderosa pine levels-of-growing-stock studies revealed that top height stability with respect to stand density depends on the...

  14. Economics of replacing young-growth ponderosa pine stands . . . a case study

    Treesearch

    Dennis E. Teeguarden

    1968-01-01

    Compares the expected capital value growth of five ponderosa pine stands (70 to 80 years old) on the Challenge Experimental Forest, Yuba County, Calif., with the cost of delaying harvest (defined as sum of stock-holding and land-holding costs). Suggests that replacement of all five stands would be financially desirable under constant stumpage prices. Recommends...

  15. Stocking levels and underlying assumptions for uneven-aged Ponderosa Pine stands.

    Treesearch

    P.H. Cochran

    1992-01-01

    Potential Problems With Q-Values Many ponderosa pine stands have a limited number of size classes, and it may be desirable to carry very large trees through several cutting cycles. Large numbers of trees below commercial size are not needed to provide adequate numbers of future replacement trees. Under these conditions, application of stand density index (SDI) can have...

  16. COMBINED EFFECTS OF CO2 AND O3 ON ANTIOXIDATIVE AND PHOTOPROTECTIVE DEFENSE SYSTEMS IN NEEDLES OF PONDEROSA PINE

    EPA Science Inventory

    To determine interactive effects of important environmental stresses on biochemical defense mechanisms of tree seedlings, we studied responses to elevated O3 and elevated atmospheric CO2 on antioxidative and photoprotective systems in needles of ponderosa pine (Pinus ponderosa Do...

  17. Early response of ponderosa pine to spacing and brush: observations on a 12-year-old plantation

    Treesearch

    William W. Oliver

    1979-01-01

    Ponderosa pine (Pinus ponderosa Laws.) was planted at five different spacings, from 6 by 6 to 18 by 18 feet, on a productive site in northern California. Spacing and brush effects on tree growth were evaluated both on plots where brush was allowed to develop and on plots kept free of brush. Competition between trees in brush-free plots began during...

  18. Effects of prescribed fire on winter assemblages of birds in ponderosa pine forests of northern Arizona

    Treesearch

    Theresa L. Pope; William M. Block

    2010-01-01

    Studies of birds in winter are rare in wildlife ecology despite winter being a critical time for birds. We examined winter assemblages of birds in ponderosa pine (Pinus ponderosa) forests of northern Arizona following prescribed fire. We conducted point counts on two study sites in northern Arizona from mid-October to mid-March 2004-2006. Each site had one unit treated...

  19. Radial growth rate increases in naturally occurring ponderosa pine trees: a late-20th century CO2 fertilization effect?

    PubMed

    Soulé, Peter T; Knapp, Paul A

    2006-01-01

    The primary objective of this study was to determine if gradually increasing levels of atmospheric CO2, as opposed to 'step' increases commonly employed in controlled studies, have a positive impact on radial growth rates of ponderosa pine (Pinus ponderosa) in natural environments, and to determine the spatial extent and variability of this growth enhancement. We developed a series of tree-ring chronologies from minimally disturbed sites across a spectrum of environmental conditions. A series of difference of means tests were used to compare radial growth post-1950, when the impacts of rising atmospheric CO2 are best expressed, with that pre-1950. Spearman's correlation was used to relate site stress to growth-rate changes. Significant increases in radial growth rates occurred post-1950, especially during drought years, with the greatest increases generally found at the most water-limited sites. Site harshness is positively related to enhanced radial growth rates. Atmospheric CO2 fertilization is probably operative, having a positive effect on radial growth rates of ponderosa pine through increasing water-use efficiency. A CO2-driven growth enhancement may affect ponderosa pine growing under both natural and controlled conditions.

  20. A ponderosa pine natural area reveals its secrets

    USGS Publications Warehouse

    Allen, Craig D.; Mac, Michael J.; Opler, Paul A.; Puckett Haecker, Catherine E.; Doran, Peter D.

    1998-01-01

    Monument Canyon Research Natural Area preserves an unlogged 259-hectare stand of old-growth ponderosa pine in the Jemez Mountains of New Mexico. This preserve, established in 1932, is the oldest research natural area in the state. This two-tiered forest displays an old-growth density of 100 stems per hectare (Muldavin et al. 1995), with an understory thicket of stagnant saplings and poles that raises the total stand density to an average of 5,954 stems per hectare, with concentrations as high as 21,617 stems per hectare (Fig. 1).

  1. Changes in the rumen bacterial microbiome of cattle exposed to ponderosa pine needles.

    PubMed

    Welch, K D; Stonecipher, C A; Gardner, D R; Cook, D; Pfister, J A

    2017-05-01

    Consumption of ponderosa pine needles, as well as needles and bark from a number of other trees, can cause abortions in cattle. The abortifacient compounds in these trees are labdane resin acids, including isocupressic acid and agathic acid. Previous research has demonstrated that cattle conditioned to pine needles metabolize the labdane resin acids more quickly than naïve cattle. The results from that study indicated that changes had occurred in the rumen of conditioned cattle. Therefore, in this study, the changes that occurred in the rumen bacterial microflora of cattle during exposure to ponderosa pine needles were evaluated. Cattle were dosed with ground pine needles twice daily for 7 d. Rumen samples were collected on d 0, 3, 7, and 14 (7 d after treatment stopped) and ruminal bacterial microbiome analyses were performed. There were 372 different genera of bacteria identified in the rumen samples. Principal coordinate analysis indicated that there was a significant difference in the rumen bacterial composition between the time points. There were 18 genera that increased in abundance from d 0 to d 7. Twenty three genera decreased in abundance from d 0 to d 7. The results from this study demonstrated that exposure of cattle to pine needles caused a clear shift in the rumen microbiome composition. In general, this shift lasted less than 1 wk post exposure, which indicates that any prophylactic treatment to manipulate the ruminal metabolism of the abortifacient compounds in pine needles would need to be continuously administered to maintain the necessary microbial composition in the rumen.

  2. Defoliation of interior Douglas-fir elicits carbon transfer and stress signalling to ponderosa pine neighbors through ectomycorrhizal networks.

    PubMed

    Song, Yuan Yuan; Simard, Suzanne W; Carroll, Allan; Mohn, William W; Zeng, Ren Sen

    2015-02-16

    Extensive regions of interior Douglas-fir (Pseudotsuga menziesii var. glauca, IDF) forests in North America are being damaged by drought and western spruce budworm (Choristoneura occidentalis). This damage is resulting from warmer and drier summers associated with climate change. To test whether defoliated IDF can directly transfer resources to ponderosa pine (Pinus ponderosae) regenerating nearby, thus aiding in forest recovery, we examined photosynthetic carbon transfer and defense enzyme response. We grew pairs of ectomycorrhizal IDF 'donor' and ponderosa pine 'receiver' seedlings in pots and isolated transfer pathways by comparing 35 μm, 0.5 μm and no mesh treatments; we then stressed IDF donors either through manual defoliation or infestation by the budworm. We found that manual defoliation of IDF donors led to transfer of photosynthetic carbon to neighboring receivers through mycorrhizal networks, but not through soil or root pathways. Both manual and insect defoliation of donors led to increased activity of peroxidase, polyphenol oxidase and superoxide dismutase in the ponderosa pine receivers, via a mechanism primarily dependent on the mycorrhizal network. These findings indicate that IDF can transfer resources and stress signals to interspecific neighbors, suggesting ectomycorrhizal networks can serve as agents of interspecific communication facilitating recovery and succession of forests after disturbance.

  3. Defoliation of interior Douglas-fir elicits carbon transfer and stress signalling to ponderosa pine neighbors through ectomycorrhizal networks

    PubMed Central

    Song, Yuan Yuan; Simard, Suzanne W.; Carroll, Allan; Mohn, William W.; Zeng, Ren Sen

    2015-01-01

    Extensive regions of interior Douglas-fir (Pseudotsuga menziesii var. glauca, IDF) forests in North America are being damaged by drought and western spruce budworm (Choristoneura occidentalis). This damage is resulting from warmer and drier summers associated with climate change. To test whether defoliated IDF can directly transfer resources to ponderosa pine (Pinus ponderosae) regenerating nearby, thus aiding in forest recovery, we examined photosynthetic carbon transfer and defense enzyme response. We grew pairs of ectomycorrhizal IDF ‘donor’ and ponderosa pine ‘receiver’ seedlings in pots and isolated transfer pathways by comparing 35 μm, 0.5 μm and no mesh treatments; we then stressed IDF donors either through manual defoliation or infestation by the budworm. We found that manual defoliation of IDF donors led to transfer of photosynthetic carbon to neighboring receivers through mycorrhizal networks, but not through soil or root pathways. Both manual and insect defoliation of donors led to increased activity of peroxidase, polyphenol oxidase and superoxide dismutase in the ponderosa pine receivers, via a mechanism primarily dependent on the mycorrhizal network. These findings indicate that IDF can transfer resources and stress signals to interspecific neighbors, suggesting ectomycorrhizal networks can serve as agents of interspecific communication facilitating recovery and succession of forests after disturbance. PMID:25683155

  4. The complexity of managing fire-dependent ecosystems in wilderness: relict ponderosa pine in the Bob Marshall Wilderness

    Treesearch

    Robert E. Keane; Stephen Arno; Laura J. Dickinson

    2006-01-01

    Isolated wilderness ecosystems with a history of frequent, low-severity fires have been altered due to many decades of fire exclusion and, as a result, are difficult to restore for philosophical and logistical reasons. In this paper, we describe the successional conditions of ponderosa pine (Pinus ponderosa) communities along the South Fork of the...

  5. Effects of wildfire on densities of secondary cavity-nesting birds in ponderosa pine forests of northern Arizona

    Treesearch

    Jill K. Dwyer; William M. Block

    2000-01-01

    Many catastrophic wildfires burned throughout forests in Arizona during the spring and summer of 1996 owing to severely dry conditions. One result of these fires was a loss of preexisting tree cavities for reproduction. In ponderosa pine (Pinus ponderosa) forests most cavities are found in dead trees; therefore, snags are a very important habitat...

  6. The mountain pine beetle: causes and consequences of an unprecedented outbreak

    Treesearch

    Allan L. Carroll

    2011-01-01

    The mountain pine beetle (Dendroctonus ponderosae) is native to the pine forests of western North America where it normally exists at very low densities, infesting only weakened or damaged trees. Under conditions conducive to survival, populations may erupt and spread over extensive landscapes, killing large numbers of healthy trees.

  7. 3-PG simulations of young ponderosa pine plantations under varied management intensity: why do they grow so differently?

    Treesearch

    Liang Wei; Marshall John; Jianwei Zhang; Hang Zhou; Robert Powers

    2014-01-01

    Models can be powerful tools for estimating forest productivity and guiding forest management, but their credibility and complexity are often an issue for forest managers. We parameterized a process-based forest growth model, 3-PG (Physiological Principles Predicting Growth), to simulate growth of ponderosa pine (Pinus ponderosa) plantations in...

  8. Genetically improved ponderosa pine seedlings outgrow nursery-run seedlings with and without competition -- Early findings

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    McDonald, P.M.; Fiddler, G.O.; Kitzmiller, J.H.

    1994-04-01

    Three classes of ponderosa pine (Pinus ponderosa) seedlings (nursery-run, wind-pollinated, control-pollinated) were evaluated for stem height and diameter at the USDA Forest Service's Placerville Nursery and the Georgetown Range District in northern California. Pines in all three classes were grown with competing vegetation or maintained in a free-to-grow condition. Control-pollinated seedlings were statistically taller than nursery-run counterparts when outplanted, and after 1 and 2 growing seasons in the field with and without competition. They also had significantly larger diameters when outplanted and after 2 growing seasons in the field when free to grow. Wind-pollinated seedlings grew taller than nursery-run seedlingsmore » when free to grow. A large amount of competing vegetation [bearclover (Chamaebatia foliolosa)--29,490 plants per acre; herbaceous vegetation--11,500; hardwood sprouts--233; and whiteleaf manzanita (Arctostaphylos viscida) seedlings--100] ensure that future pine development will be tested rigorously.« less

  9. Inhibiting effect of ponderosa pine seed trees on seedling growth

    Treesearch

    Philip M. McDonald

    1976-01-01

    Ponderosa pine seed trees, numbering 4, 8, and 12 per acre, were left standing for 9 years after harvest cutting on the Challenge Experimental Forest, Calif. Seedling heights were measured at ages 5, 9, and 14, and for all ages were poorest if within 20 feet of a seed tree. Seedlings 20 feet or less from a seed tree at the ages given lost the equivalent in years of...

  10. Impact of spring or fall repeated prescribed fire on growth of ponderosa pine in eastern Oregon, USA

    Treesearch

    Walter G. Thies; Douglas J. Westlind; Mark. Loewen

    2013-01-01

    Prescribed burning is used to reduce fuel loads and to return fire to its historic disturbance role in western forests. Managers need to know the effects of prescribed fire on tree growth. Growth of residual ponderosa pine (Pinus ponderosa Dougl. ex Laws.) was measured in an existing long-term study of the effects of season-of-prescribed burn in...

  11. Predicting the effects of tropospheric ozone on regional productivity of ponderosa pine and white fir.

    Treesearch

    D.A. Weinstein; J.A. Laurence; W.A. Retzlaff; J.S. Kern; E.H. Lee; W.E. Hogsett; J. Weber

    2005-01-01

    We simulated forest dynamics of the regional ponderosa pine-white fir conifer forest of the San Bernadino and Sierra Nevada mountains of California to determine the effects of high ozone concentrations over the next century and to compare the responses to our similar study for loblolly pine forests of the southeast. As in the earlier study, we linked two models, TREGRO...

  12. Guide to the Blacks Mountain Experimental Forest - A sustained yield experiment in ponderosa pine in northeastern California

    Treesearch

    E.I. Kotok

    1938-01-01

    Experimental forests, watersheds, and ranges are the field laboratories in the research structure of the Forest Service. The California Forest and Range Experiment Station maintains four experimental forests representing the more important timber types in the Pine Region.The Blacks Mountain Experimental Forest represents the ponderosa pine...

  13. Brush competition retards early stand development of planted ponderosa pine: update on a 24-year study

    Treesearch

    William W. Oliver

    1990-01-01

    Growth of ponderosa pine (Pinus ponderosa) was monitored for 24 years after planting at five different square spacings (6, 9. 12, 15, and 18 ft) in the presence or absence of competing brush on the westside Sierra Nevada. Spacing strongly influenced both mean dbh and basal area/ac. In plots maintained free of brush, diameters ranged from 5.1 in. at the 6-ft spacing to...

  14. Changes in the rumen bacterial microbiome of cattle exposed to ponderosa pine needles

    USDA-ARS?s Scientific Manuscript database

    Consumption of ponderosa pine needles, as well as needles and bark from a number of other trees, can cause abortions in cattle. The abortifacient compounds in these trees are labdane resin acids, including isocupressic acid and agathic acid. Previous research has demonstrated that cattle conditioned...

  15. ROOT GROWTH AND TURNOVER IN DIFFERENT AGED PONDEROSA PINE STANDS IN OREGON, USA

    EPA Science Inventory

    The impacts of pollution and climate change on soil carbon dynamics are poorly understood, in part due to a lack of information regarding root production and turnover in natural ecosystems. In order to examine how root dynamics change with stand age in ponderosa pine forests (...

  16. Seasonal influences on ozone uptake and foliar injury to ponderosa and Jeffrey pines at a southern California site

    Treesearch

    Patrick J. Temple; Paul R. Miller

    1998-01-01

    Ambient ozone was monitored from 1992 to 1994 near a forested site dominated by mature Jeffrey and ponderosa pines (Pinus jeffrey Grev. & Balf. and Pinus ponderosa Dougl. ex Laws.) at 2,000 m in the San Bernardino Mountains of southern California. Ozone injury symptoms, including percent chlorotic mottle and foliage retention,...

  17. Deciduous conifers: high n deposition and o3 exposure effects on growth and biomass allocation in ponderosa pine

    Treesearch

    Nancy Grulke; L. Balduman

    1999-01-01

    Ponderosa pines (Pinus ponderosa Dougl. ex. Laws) 21 to 60 yr old were used to assess the relative importance of environmental stressors (O3, drought) versus an enhancer (N deposition) on foliar retention, components of aboveground growth, and whole tree biomass allocation. Sites were chosen across a well-described gradient...

  18. Reducing fire hazard in ponderosa pine thinning slash by mechanical crushing

    Treesearch

    John R. Dell; Franklin R. Ward

    1969-01-01

    Precommercial thinning in ponderosa pine stands in the Western United States is a growing practice. Thinning slash can, however, be a serious fire hazard in dry areas. Crushing and compacting this slash may be one way of reducing the hazard. Three types of mechanical crushers were tested on the Deschutes National Forest, Oregon. Results indicate that at least one of...

  19. Development of post-fire crown damage mortality thresholds in ponderosa pine

    Treesearch

    James F. Fowler; Carolyn Hull Sieg; Joel McMillin; Kurt K. Allen; Jose F. Negron; Linda L. Wadleigh; John A. Anhold; Ken E. Gibson

    2010-01-01

    Previous research has shown that crown scorch volume and crown consumption volume are the major predictors of post-fire mortality in ponderosa pine. In this study, we use piecewise logistic regression models of crown scorch data from 6633 trees in five wildfires from the Intermountain West to locate a mortality threshold at 88% scorch by volume for trees with no crown...

  20. Growth of ten regional races of ponderosa pine in six plantations.

    Treesearch

    Thornton T. Munger

    1947-01-01

    Ponderosa pine occurs over a wide range--from the Great Plains to the Pacific and from British Columbia to Mexico. Within this range there is great variation in the form and life habits of this tree. It has been long recognized that "regional races" or strains are present. Some of these regional races have been recognized by some botanists as varieties, such...

  1. Importance of resin ducts in reducing ponderosa pine mortality from bark beetle attack.

    PubMed

    Kane, Jeffrey M; Kolb, Thomas E

    2010-11-01

    The relative importance of growth and defense to tree mortality during drought and bark beetle attacks is poorly understood. We addressed this issue by comparing growth and defense characteristics between 25 pairs of ponderosa pine (Pinus ponderosa) trees that survived and trees that died from drought-associated bark beetle attacks in forests of northern Arizona, USA. The three major findings of our research were: (1) xylem resin ducts in live trees were >10% larger (diameter), >25% denser (no. of resin ducts mm(-2)), and composed >50% more area per unit ring growth than dead trees; (2) measures of defense, such as resin duct production (no. of resin ducts year(-1)) and the proportion of xylem ring area to resin ducts, not growth, were the best model parameters of ponderosa pine mortality; and (3) most correlations between annual variation in growth and resin duct characteristics were positive suggesting that conditions conducive to growth also increase resin duct production. Our results suggest that trees that survive drought and subsequent bark beetle attacks invest more carbon in resin defense than trees that die, and that carbon allocation to resin ducts is a more important determinant of tree mortality than allocation to radial growth.

  2. Summary (Songbird ecology in southwestern ponderosa pine forests: A literature review)

    Treesearch

    William M. Block; Deborah M. Finch; Joseph L. Ganey; William H. Moir

    1997-01-01

    Most ornithological studies in Southwestern ponderosa pine forests have yielded results that are applicable only to the specific location and particular conditions of the study areas (for example, Green 1979 and Hurlbert 1984). In addition, varying interpretation of similar study results by investigators has limited our ability to extend or synthesize research results...

  3. Survival and sanitation of dwarf mistletoe-infected ponderosa pine following prescribed underburning

    Treesearch

    David A. Conklin; Brian W. Geils

    2008-01-01

    We present results on survival of ponderosa pine and reduction in dwarf mistletoe (Arceuthobium) infection after six operational prescribed underburns in New Mexico. Survival 3 years postburn for 1,585 trees fit a logistic relationship with crown scorch, bole char, and mistletoe. The scorch effect was best represented by classes as <90, 90,...

  4. Effects of Climate Variability and Accelerated Forest Thinning on Watershed-Scale Runoff in Southwestern USA Ponderosa Pine Forests

    PubMed Central

    Robles, Marcos D.; Marshall, Robert M.; O'Donnell, Frances; Smith, Edward B.; Haney, Jeanmarie A.; Gori, David F.

    2014-01-01

    The recent mortality of up to 20% of forests and woodlands in the southwestern United States, along with declining stream flows and projected future water shortages, heightens the need to understand how management practices can enhance forest resilience and functioning under unprecedented scales of drought and wildfire. To address this challenge, a combination of mechanical thinning and fire treatments are planned for 238,000 hectares (588,000 acres) of ponderosa pine (Pinus ponderosa) forests across central Arizona, USA. Mechanical thinning can increase runoff at fine scales, as well as reduce fire risk and tree water stress during drought, but the effects of this practice have not been studied at scales commensurate with recent forest disturbances or under a highly variable climate. Modifying a historical runoff model, we constructed scenarios to estimate increases in runoff from thinning ponderosa pine at the landscape and watershed scales based on driving variables: pace, extent and intensity of forest treatments and variability in winter precipitation. We found that runoff on thinned forests was about 20% greater than unthinned forests, regardless of whether treatments occurred in a drought or pluvial period. The magnitude of this increase is similar to observed declines in snowpack for the region, suggesting that accelerated thinning may lessen runoff losses due to warming effects. Gains in runoff were temporary (six years after treatment) and modest when compared to mean annual runoff from the study watersheds (0–3%). Nonetheless gains observed during drought periods could play a role in augmenting river flows on a seasonal basis, improving conditions for water-dependent natural resources, as well as benefit water supplies for downstream communities. Results of this study and others suggest that accelerated forest thinning at large scales could improve the water balance and resilience of forests and sustain the ecosystem services they provide. PMID

  5. Effects of climate variability and accelerated forest thinning on watershed-scale runoff in southwestern USA ponderosa pine forests.

    PubMed

    Robles, Marcos D; Marshall, Robert M; O'Donnell, Frances; Smith, Edward B; Haney, Jeanmarie A; Gori, David F

    2014-01-01

    The recent mortality of up to 20% of forests and woodlands in the southwestern United States, along with declining stream flows and projected future water shortages, heightens the need to understand how management practices can enhance forest resilience and functioning under unprecedented scales of drought and wildfire. To address this challenge, a combination of mechanical thinning and fire treatments are planned for 238,000 hectares (588,000 acres) of ponderosa pine (Pinus ponderosa) forests across central Arizona, USA. Mechanical thinning can increase runoff at fine scales, as well as reduce fire risk and tree water stress during drought, but the effects of this practice have not been studied at scales commensurate with recent forest disturbances or under a highly variable climate. Modifying a historical runoff model, we constructed scenarios to estimate increases in runoff from thinning ponderosa pine at the landscape and watershed scales based on driving variables: pace, extent and intensity of forest treatments and variability in winter precipitation. We found that runoff on thinned forests was about 20% greater than unthinned forests, regardless of whether treatments occurred in a drought or pluvial period. The magnitude of this increase is similar to observed declines in snowpack for the region, suggesting that accelerated thinning may lessen runoff losses due to warming effects. Gains in runoff were temporary (six years after treatment) and modest when compared to mean annual runoff from the study watersheds (0-3%). Nonetheless gains observed during drought periods could play a role in augmenting river flows on a seasonal basis, improving conditions for water-dependent natural resources, as well as benefit water supplies for downstream communities. Results of this study and others suggest that accelerated forest thinning at large scales could improve the water balance and resilience of forests and sustain the ecosystem services they provide.

  6. Habitat associations of the sagebrush lizard (Sceloporus graciosus): Potential responses of an ectotherm to ponderosa pine forest restoration treatments

    Treesearch

    Shawn C. Knox; Carol Chambers; Stephen S. Germaine

    2001-01-01

    Little is known about the response of ectotherms to ponderosa pine (Pinus ponderosa) restoration treatments. The ambient body temperature of an ectotherm affects its physiology, development, and behavior. Microhabitat availability and heterogeneity are critical factors in determining which thermoregulation choices are available to a terrestrial ectotherm (Stevenson...

  7. Understory vegetation response after 30 years of interval prescribed burning in two ponderosa pine sites in northern Arizona, USA

    Treesearch

    Catherine A. Scudieri; Carolyn Hull Sieg; Sally M. Haase; Andrea E. Thode; Stephen S. Sackett

    2010-01-01

    Southwestern USA ponderosa pine (Pinus ponderosa C. Lawson var. scopulorum Engelm.) forests evolved with frequent surface fires and have changed dramatically over the last century. Overstory tree density has sharply increased while abundance of understory vegetation has declined primarily due to the near cessation of fires. We...

  8. PARTITIONING OF WATER FLUX IN A SIERRA NEVADA PONDEROSA PINE PLANTATION. (R826601)

    EPA Science Inventory

    The weather patterns of the west side of the Sierra Nevada Mountains (cold, wet winters and hot, dry summers) strongly influence how water is partitioned between transpiration and evaporation and result in a specific strategy of water use by ponderosa pine trees (Pinus pond...

  9. Comparing timber and lumber from plantation and natural stands of ponderosa pine

    Treesearch

    Eini C. Lowell; Christine L. Todoroki; Ed. Thomas

    2009-01-01

    Data derived from empirical studies, coupled with modeling and simulation techniques, were used to compare tree and product quality from two stands of small-diameter ponderosa pine trees growing in northern California: one plantation, the other natural. The plantation had no management following establishment, and the natural stand had no active management. Fifty trees...

  10. Phytotoxic grass residues reduce germination and initial root growth of ponderosa pine

    Treesearch

    W. J. Rietveld

    1975-01-01

    Extracts of green foliage of Arizona fescue and mountain muhly significantly reduced germination of ponderosa pine seeds, and retarded speed of elongation and mean radicle length. Three possible routes of release of the inhibitor were investigated: (1) leaching from live foliage, (2) root exudation, and (3) overwinter leaching from dead residues. The principal route...

  11. Landscape models: helping land managers think big

    Treesearch

    Rachel White; Rhonda Mazza

    2011-01-01

    In a sun-baked, grassy clearing on the east side of the Cascade Range in central Washington, Pacific Northwest (PNW) Research Station landscape ecologist Miles Hemstrom and a group of ecologists and land managers from the Washington Department of Natural Resources (DNR) gather in the shade of a ponderosa pine. Hundreds of years old, this ancient pine has withstood...

  12. Ozone injury responses of ponderosa and Jeffrey pine in the Sierra Nevada and San Bernardino Mountains in California

    Treesearch

    Paul Miller; Raleigh Guthrey; Susan Schilling; John Carroll

    1998-01-01

    Ozone injury was monitored on foliage of ponderosa (Pinus ponderosa Dougl. ex Laws.) and Jeffrey (Pinus jeffreyi Grev. & Balf.) pines at 11 locations in the Sierra Nevada and 1 site in the San Bernardino Mountains of southern California. Ozone injury on all age cohorts of needles on about 1,600 trees was surveyed annually from...

  13. Resiliency of an interior ponderosa pine forest to bark beetle infestations following fuel-reduction and forest-restoration treatments

    Treesearch

    Christopher J Fettig; Stephen R. McKelvey

    2014-01-01

    Mechanical thinning and the application of prescribed fire are commonly used to restore fire-adapted forest ecosystems in the western United States. During a 10-year period, we monitored the effects of fuel-reduction and forest-restoration treatments on levels of tree mortality in an interior ponderosa pine, Pinus ponderosa Dougl. ex Laws., forest...

  14. Blacks Mountain Experimental Forest: bark beetle responses to differences in forest structure and the application of prescribed fire in interior ponderosa pine

    Treesearch

    Christopher J. Fettig; Robert R. Borys; Stephen R. McKelvey; Christopher P. Dabney

    2008-01-01

    Mechanical thinning and the application of prescribed fire are commonly used tools in the restoration of fire-adapted forest ecosystems. However, few studies have explored their effects on subsequent amounts of bark beetle caused tree mortality in interior ponderosa pine, Pinus ponderosa Dougl. ex P. & C. Laws. var. ponderosa. In...

  15. Seasoning degrade in kiln drying ponderosa pine in south central Oregon.

    Treesearch

    A.C. Knauss

    1957-01-01

    This report presents the results of a study to determine the loss in volume and value of lumber when kiln drying and surfacing the production from ponderosa pine logs. The study measured (1) the reduction in volume due to trimming and culling dry lumber after surfacing, (2) the reduction in grade due to seasoning defects, (3) the reduction in grade due to failure of...

  16. Glyoxysomes in Megagamethophyte of Germinating Ponderosa Pine Seeds 12

    PubMed Central

    Ching, Te May

    1970-01-01

    Decoated ponderosa pine (Pinus ponderosa Laws) seeds contained 40% lipids, which were mainly stored in megagametophytic tissue and were utilized or converted to sugars via the glyoxylate cycle during germination. Mitochondria and glyoxysomes were isolated from the tissue by sucrose density gradient centrifugation at different stages of germination. It was found that isocitrate lyase, malate synthase, and catalase were mainly bound in glyoxysomes. Aconitase and fumarase were chiefly localized in mitochondria, whereas citrate synthase was common for both. Both organelles increased in quantity and specific activity of their respective marker enzymes with the advancement of germination. When the megagametophyte was exhausted at the end of germination, the quantity of these organelles and the activity of their marker enzymes decreased abruptly. At the stage of highest lipolysis, the isolated mitochondria and glyoxysomes were able to synthesize protein from labeled amino acids. Both organellar fractions contained RNA and DNA. Some degree of autonomy in glyoxysomes is indicated. Images PMID:16657489

  17. The influence of partial cutting on mountain pine beetle-caused tree mortality in Black Hills ponderosa pine stands

    Treesearch

    J.M. Schmid; S.A. Mata; R.R. Kessler; J.B. Popp

    2007-01-01

    Ponderosa pine stands were partially cut to various stocking levels at five locations, periodically surveyed, and remeasured during the 20 years after installation. Mean diameter generally increased 2 inches over the 20-year period on most partially cut plots and less than 2 inches on unmanaged controls. Average diameter growth for diameter classes in partially cut...

  18. Local and regional variation in the monoterpenes of ponderosa pine wood oleoresin

    Treesearch

    R.H. Smith; R.L. Peloquin; P.C. Passof

    1969-01-01

    A gas chromatographic analysis of the mono-terpenes of 927 ponderosa pines, representing to some degree a major portion of the species' range, showed considerable local and regional diversity in composition. Five major monoterpenes— α-pinene, β-pinene, 3-carene, myrcene, and limonene—were analyzed. There is some evidence to support the...

  19. Ponderosa pine response to fertilization: influence of brush removal and soil type

    Treesearch

    Robert F. Powers; Grant D. Jackson

    1978-01-01

    First-year results of fertilization in a young ponderosa pine plantation on two contrasting soils were analyzed. Trees testing low in foliar nitrogen responded strongly to fertilization where brush had been removed, but failed to respond if brush remained. Height growth was doubled by certain treatment combinations on the less fertile Mariposa soil, but was not...

  20. Mechanical and Chemical Release in a 12-Year-Old Ponderosa Pine Plantation

    Treesearch

    Gary O. Fiddler; Philip M. McDonald

    1997-01-01

    A 12-year-old ponderosa pine plantation on the Tahoe National Forest in northern California was mechanically treated with a Hydro-Ax in an attempt to increase the survival and growth of the planted seedlings. Other release methods were not feasible because the shrubs in the mixed-shrub community (greenleaf manzanita, mountain whitethorn, bittercherry, coffeeberry) were...

  1. Genetic variation and seed transfer guidelines for ponderosa pine in central Oregon.

    Treesearch

    Frank C. Sorensen

    1994-01-01

    Adaptive genetic variation in seed and seedling traits for ponderosa pine from the east slopes of the Cascade Range in Oregon was analyzed by using 307 families from 227 locations. Factor scores from three principal components based on seed and seedling traits were related by multiple regression to latitude, distance from the Cascade crest, elevation, slope, and...

  2. Dynamic programming for optimization of timber production and grazing in ponderosa pine

    Treesearch

    Kurt H. Riitters; J. Douglas Brodie; David W. Hann

    1982-01-01

    Dynamic programming procedures are presented for optimizing thinning and rotation of even-aged ponderosa pine by using the four descriptors: age, basal area, number of trees, and time since thinning. Because both timber yield and grazing yield are functions of stand density, the two outputs-forage and timber-can both be optimized. The soil expectation values for single...

  3. Partitioning of water flux in a Sierra Nevada ponderosa pine plantation

    USGS Publications Warehouse

    Kurpius, M.R.; Panek, J.A.; Nikolov, N.T.; McKay, M.; Goldstein, Allen H.

    2003-01-01

    The weather patterns of the west side of the Sierra Nevada Mountains (cold, wet winters and hot, dry summers) strongly influence how water is partitioned between transpiration and evaporation and result in a specific strategy of water use by ponderosa pine trees (Pinus ponderosa) in this region. To investigate how year-round water fluxes were partitioned in a young ponderosa pine ecosystem in the Sierra Nevada Mountains, water fluxes were continually measured from June 2000 to May 2001 using a combination of sap flow and eddy covariance techniques (above- and below-canopy). Water fluxes were modeled at our study site using a biophysical model, FORFLUX. During summer and fall water fluxes were equally partitioned between transpiration and soil evaporation while transpiration dominated the water fluxes in winter and spring. The trees had high rates of canopy conductance and transpiration in the early morning and mid-late afternoon and a mid-day depression during the dry season. We used a diurnal centroid analysis to show that the timing of high canopy conductance and transpiration relative to high vapor pressure deficit (D) shifted with soil moisture: during periods of low soil moisture canopy conductance and transpiration peaked early in the day when D was low. Conversely, during periods of high soil moisture canopy conductance and transpiration peaked at the same time or later in the day than D. Our observations suggest a general strategy by the pine trees in which they maximize stomatal conductance, and therefore carbon fixation, throughout the day on warm sunny days with high soil moisture (i.e. warm periods in winter and late spring) and maximize stomatal conductance and carbon fixation in the morning through the dry periods. FORFLUX model estimates of evaporation and transpiration were close to measured/calculated values during the dry period, including the drought, but underestimated transpiration and overestimated evaporation during the wet period. ?? 2003

  4. Six-year post-fire mortality and health of relict ponderosa pines in the Bob Marshall Wilderness Area, Montana

    Treesearch

    Signe B. Leirfallom; Robert E. Keane

    2011-01-01

    In 2003, lightning-caused fires burned through relict ponderosa pine (Pinus ponderosa) stands in the Bob Marshall Wilderness, Montana, after decades of fire exclusion. Since many trees in these stands had Native American bark-peeling scars, concern arose about the adverse fire effects on this cultural and ecological resource. In 2004, Keane and others (2006) began a...

  5. Effects of low intensity prescribed fires on ponderosa pine forests in wilderness areas of Zion National Park, Utah

    Treesearch

    Henry V. Bastian

    2001-01-01

    Vegetation and fuel loading plots were monitored and sampled in wilderness areas treated with prescribed fire. Changes in ponderosa pine (Pinus ponderosa) forest structure tree species and fuel loading are presented. Plots were randomly stratified and established in burn units in 1995. Preliminary analysis of nine plots 2 years after burning show litter was reduced 54....

  6. Fuel reduction in residential and scenic forests: A comparison of three treatments in a western Montana ponderosa pine stand

    Treesearch

    Joe H. Scott

    1998-01-01

    Three contrasting thinning treatments to reduce fire hazard were implemented in a 100-year-old ponderosa pine/Douglas-fir (Pinus ponderosa/Pseudotsuga menzesii) stand on the Lolo National Forest, MT. All treatments included a commercial thinning designed to reduce crown fuels and provide revenue to offset costs. The treatments are...

  7. Changes in forest structure since 1860 in ponderosa pine dominated forests in the Colorado and Wyoming Front Range, USA

    Treesearch

    Mike A. Battaglia; Benjamin Gannon; Peter M. Brown; Paula J. Fornwalt; Antony S. Cheng; Laurie S. Huckaby

    2018-01-01

    Management practices since the late 19th century, including fire exclusion and harvesting, have altered the structure of ponderosa pine (Pinus ponderosa Douglas ex P. Lawson & C. Lawson) dominated forests across the western United States. These structural changes have the potential to contribute to uncharacteristic wildfire behavior and effects. Locally-...

  8. A comparison of the metabolism of the abortifacient compounds from Ponderosa Pine needles in conditioned versus naive cattle

    USDA-ARS?s Scientific Manuscript database

    Isocupressic acid (ICA) is the abortifacient compound in ponderosa pine needles, which can cause late term abortions in cattle. However, cattle rapidly metabolize ICA to agathic acid and subsequent metabolites. When pine needles are dosed orally to cattle, no ICA is detected in their serum while a...

  9. Effect of dietary protein level and quebracho tannin on consumption of pine needles (Pinus ponderosa) by beef cows

    USDA-ARS?s Scientific Manuscript database

    Ponderosa pine trees occupy over 15 million hectares of rangeland in western North America. Pregnant cows often consume pine needles (PN), and subsequently abort. The protein-to-energy ratio may be important in the ability of cattle to tolerate dietary terpenes. Tannins often co-occur with terpenes ...

  10. Genetic variation in the Ponderosae of the Southwest

    Treesearch

    Gerald E. Rehfeldt

    1993-01-01

    Ninety-five seedling populations of southwestern ponderosa pine (Pinus ponderosa var. scopulorum) along with single populations of Pinus engelmannii and Pinus arizonica were compared in four environmentally disparate common gardens. Differentiation among ponderosa pine populations was detected for a diverse assortment of variables that included patterns of shoot...

  11. Silviculture of ponderosa pine in the Black Hills: The status of our knowledge

    Treesearch

    Charles E. Boldt; James L. Van Deusen

    1974-01-01

    This Paper, intended as a guide for professional foresters, describes major silvicultural conditions likely to be encountered in the Black Hills, reasonable treatment options, and probable results and implications of these treatments. It also describes silvical characteristics and behavior of Black Hills ponderosa pine, and a variety of proven silvicultural tools....

  12. Response of ponderosa pine stands to pre-commercial thinning on Nez Perce and Spokane Tribal forests in the Inland Northwest, USA

    Treesearch

    Dennis E. Ferguson; John C. Byrne; William R. Wykoff; Brian Kummet; Ted Hensold

    2011-01-01

    Stands of dense, natural ponderosa pine (Pinus ponderosa var. ponderosa) regeneration were operationally, precommercially thinned at seven sites - four on Nez Perce Tribal lands in northern Idaho and three on Spokane Tribal lands in eastern Washington. Five spacing treatments were studied - control (no thinning), 5x5 ft, 7x7 ft, 10x10 ft, and 14x14 ft. Sample trees...

  13. Silvical characteristics of Jeffrey pine

    Treesearch

    William E. Hallin

    1957-01-01

    The most noteworthy feature of Jeffrey pine (Pinus jeffreyi Grev. & Balf. ) is its similarity in appearance and behavior to ponderosa pine (Pinus ponderosa Laws.), a much more widespread and better known species. At one time Jeffrey pine was considered to be a variety of ponderosa pine, and lumber markets make no...

  14. An experimental demonstration of stem damage as a predictor of fire-caused mortality for ponderosa pine

    USGS Publications Warehouse

    van Mantgem, P.; Schwartz, M.

    2004-01-01

    We subjected 159 small ponderosa pine (Pinus ponderosa Dougl. ex P. & C. Laws.) to treatments designed to test the relative importance of stem damage as a predictor of postfire mortality. The treatments consisted of a group with the basal bark artificially thinned, a second group with fuels removed from the base of the stem, and an untreated control. Following prescribed burning, crown scorch severity was equivalent among the groups. Postfire mortality was significantly less frequent in the fuels removal group than in the bark removal and control groups. No model of mortality for the fuels removal group was possible, because dead trees constituted <4% of subject trees. Mortality in the bark removal group was best predicted by crown scorch and stem scorch severity, whereas death in the control group was predicted by crown scorch severity and bark thickness. The relative lack of mortality in the fuels removal group and the increased sensitivity to stem damage in the bark removal group suggest that stem damage is a critical determinant of postfire mortality for small ponderosa pine.

  15. Influence of elevation on bark beetle (Coleoptera: Curculionidae, Scolytinae) community structure and flight periodicity in ponderosa pine forests of Arizona

    Treesearch

    Kelly K. Williams; Joel D. McMillin; Tom E. DeGomez; Karen M. Clancy; Andy Miller

    2008-01-01

    We examined abundance and flight periodicity of five Ips and six Dendroctonus species (Coleoptera: Curculionidae, Scolytinae) among three different elevation bands in ponderosa pine (Pinus ponderosa Douglas ex. Lawson) forests of northcentral Arizona. Bark beetle populations were monitored at 10 sites in each of three elevation...

  16. Conceptual framework for studying the effects of fuels treatments on avian communities in ponderosa pine forests of northern Arizona

    Treesearch

    Brett G. Dickson; William M. Block; Thomas D. Sisk

    2004-01-01

    Many ponderosa pine (Pinus ponderosa) forests in the western US are dense and contain excessive accumulations of ground and ladder fuels, resulting in forests at high risk of catastrophic fire. Prescribed fire and thinning are two potential tools used in the reduction of forest fuels, although the ecological and economic consequences of applying...

  17. A field guide to predict delayed mortality of fire-damaged ponderosa pine: application and validation of the Malheur model.

    Treesearch

    Walter G. Thies; Douglas J. Westlind; Mark Loewen; Greg Brenner

    2008-01-01

    The Malheur model for fire-caused delayed mortality is presented as an easily interpreted graph (mortality-probability calculator) as part of a one-page field guide that allows the user to determine postfire probability of mortality for ponderosa pine (Pinus ponderosa Dougl. ex Laws.). Following both prescribed burns and wildfires, managers need...

  18. Stand characteristics and downed woody debris accumulations associated with a mountain pine beetle (Dendroctonus ponderosae Hopkins) outbreak in Colorado

    Treesearch

    Jennifer G. Klutsch; Jose F. Negron; Sheryl L. Costello; Charles C. Rhoades; Daniel R. West; John Popp; Rick Caissie

    2009-01-01

    Lodgepole pine (Pinus contorta Dougl. ex Loud.)-dominated ecosystems in north-central Colorado are undergoing rapid and drastic changes associated with overstory tree mortality from a currentmountain pine beetle (Dendroctonus ponderosae Hopkins) outbreak. To characterize stand characteristics and downed woody debris loads during...

  19. Wildfire effects on a ponderosa pine ecosystem: An Arizona case study

    Treesearch

    R. E. Campbell; Jr. Baker; P. F. Ffolliott; F. R. Larson; C. C. Avery

    1977-01-01

    A wildfire of variable severity swept through 717 acres (290 ha) of ponderosa pine forest in north-central Arizona in May 1972. Where the fire was intense it killed 90% of the small trees and 50% of the sawtimber, burned 2.6 in (6.5 cm) of forest floor to the mineral soil, and induced a water-repellent layer in the sandier soils. The reduced infiltration rates, which...

  20. Postfire environmental conditions influence the spatial pattern of regeneration for Pinus ponderosa

    Treesearch

    V. H. Bonnet; Anna Schoettle; W. D. Shepperd

    2005-01-01

    Regeneration of ponderosa pine after fire depends on the patterns of seed availability and the environmental conditions that define safe sites for seedling establishment. A transect approach was applied in 2002 to determine the spatial distribution of regeneration from unburned to burned areas within the landscape impacted by the Jasper Fire of 2000 in the...

  1. Seasoning and surfacing degrade in kiln-drying ponderosa pine in eastern Washington.

    Treesearch

    A.C. Knauss; E.H. Clarke

    1961-01-01

    This report presents results of a study to determine the degrade (loss in volume and value) of ponderosa pine lumber when cut, kiln-dried, and surfaced in accordance with commercial practice. The study measured (1) loss in volume due to culling and trimming surfaced dry lumber because of sawing, seasoning, and surfacing defects; (2) reduction in grade due to seasoning...

  2. Thinning ponderosa pine in the Pacific Northwest: a summary of present information.

    Treesearch

    Edwin L. Mowat

    1953-01-01

    Proper thinning of dense young stands of ponderosa pine increases the growth rate of remaining trees and thereby shortens the time required to grow timber of any desired size. The extent of improvement in growth is shown by thinning studies in the Pacific Northwest covering periods of 10 to 24 years. These studies, together with general experience and research in other...

  3. User guide for HCR Estimator 2.0: software to calculate cost and revenue thresholds for harvesting small-diameter ponderosa pine.

    Treesearch

    Dennis R. Becker; Debra Larson; Eini C. Lowell; Robert B. Rummer

    2008-01-01

    The HCR (Harvest Cost-Revenue) Estimator is engineering and financial analysis software used to evaluate stand-level financial thresholds for harvesting small-diameter ponderosa pine (Pinus ponderosa Dougl. ex Laws.) in the Southwest United States. The Windows-based program helps contractors and planners to identify costs associated with tree...

  4. Ectomycorrhizal fungi associated with ponderosa pine and Douglas-fir: a comparison of species richness in native western North American forests and Patagonian plantations from Argentina.

    PubMed

    Barroetaveña, C; Cázares, E; Rajchenberg, M

    2007-07-01

    The putative ectomycorrhizal fungal species registered from sporocarps associated with ponderosa pine and Douglas-fir forests in their natural range distribution (i.e., western Canada, USA, and Mexico) and from plantations in south Argentina and other parts of the world are listed. One hundred and fifty seven taxa are reported for native ponderosa pine forests and 514 taxa for native Douglas-fir forests based on available literature and databases. A small group of genera comprises a high proportion of the species richness for native Douglas-fir (i.e., Cortinarius, Inocybe, and Russula), whereas in native ponderosa pine, the species richness is more evenly distributed among several genera. The comparison between ectomycorrhizal species richness associated with both trees in native forests and in Patagonia (Argentina) shows far fewer species in the latter, with 18 taxa for the ponderosa pine and 15 for the Douglas-fir. Epigeous species richness is clearly dominant in native Douglas-fir, whereas a more balanced relation epigeous/hypogeous richness is observed for native ponderosa pine; a similar trend was observed for Patagonian plantations. Most fungi in Patagonian Douglas-fir plantations have not been recorded in plantations elsewhere, except Suillus lakei and Thelephora terrestris, and only 56% of the fungal taxa recorded in Douglas-fir plantations around the world are known from native forests, the other taxa being new associations for this host, suggesting that new tree + ectomycorrhizal fungal taxa associations are favored in artificial situations as plantations.

  5. Five-year results of a ponderosa pine provenance study in the Black Hills

    Treesearch

    James L. Van Deusen

    1974-01-01

    Survival and height growth data were collected after five field growing seasons from ponderosa pine progeny representing 75 provenances of natural stands in the Great Plains and Northern Rockies. Results showed that trees from no other provenance survived significantly better or grew significantly taller than trees from the Black Hills. Trees from southern Colorado,...

  6. 2004 report on the health of Colorado's forests: Special issue: Ponderosa pine forests

    Treesearch

    Paige Lewis; Merrill R. Kaufmann; Laurie S. Huckaby; Dave Leatherman

    2005-01-01

    The 2004 Report on the Health of Colorado's Forests begins with an overview of significant incidents and trends in forest insect and disease activity across the state. The remainder of the Report provides an in-depth examination of the ecology, condition and management of Colorado's ponderosa pine forests. Unlike previous editions, which highlighted a range...

  7. Effect of redistributing windrowed topsoil on growth and development of ponderosa pine plantations

    Treesearch

    Jianwei Zhang; David H. Young; Kenneth R. Luckow

    2015-01-01

    Windrowing site preparation often displaces significant amounts of topsoil including nutrients and carbon into the strip-piles. Although short-term growth may increase due to the early control of competing vegetation, this practice can reduce long-term plantation productivity. Here, we report an experiment established in 1989 in a 28-year-old ponderosa pine (

  8. EFFECTS OF CO2 AND O3 IN PONDEROSA PINE PLANT/LITTER/SOIL MESOCOSMS

    EPA Science Inventory

    Forested ecosysems are subjected to interacting conditions whose joint impacts may be quite different from those from single factors. To understand the impacts of CO2 and O3 on forest ecosystems, in April 1998, we initiated a four-year study of a Ponderosa pine seedling/soil/lit...

  9. Reproductive isolation and environmental adaptation shape the phylogeography of mountain pine beetle (Dendroctonus ponderosa)

    Treesearch

    Eddy J. Dowle; Ryan R. Bracewell; Michael E. Pfrender; Karen E. Mock; Barbara J. Bentz; Gregory J. Ragland

    2017-01-01

    Chromosomal rearrangement can be an important mechanism driving population differentiation and incipient speciation. In the mountain pine beetle (MPB, Dendroctonus ponderosae), deletions on the Y chromosome that are polymorphic among populations are associated with reproductive incompatibility. Here, we used RAD sequencing across the entire MPB range in western North...

  10. Effects of nursery fertilizer and irrigation on ponderosa and lodgepole pine seedling size. Forest Service research note

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Sloan, J.P.

    1992-12-01

    Eight fertilizer treatments combined with three irrigation regimes were used when growing lodgepole and ponderosa pine seedlings on two soil types at Lucky Peak Nursery near Biose, ID. Seedlings of both species were larger on the sandy loam than the clay loam soil. Milorganite, an organic fertilizer derived from sewage sludge, reduced initial seedbeed densities but had no further effects. Ammonium nitrate increased seedling size on the clay loam, but not on the sandy loam soil. Increased irrigation was more effective in increasing seedling size on the sandy loam than on the clay loam soil. However, ponderosa pine receiving themore » least irrigation in the nursery grew the fastest for 3 years after being transplanted in the field, possibly because of drought conditioning.« less

  11. Frequent fire alters nitrogen transformations in ponderosa pine stands of the inland northwest.

    PubMed

    DeLuca, Thomas H; Sala, Anna

    2006-10-01

    Recurrent, low-severity fire in ponderosa pine (Pinus ponderosa)/interior Douglas-fir (Pseudotsuga menziesii var. glauca) forests is thought to have directly influenced nitrogen (N) cycling and availability. However, no studies to date have investigated the influence of natural fire intervals on soil processes in undisturbed forests, thereby limiting our ability to understand ecological processes and successional dynamics in this important ecosystem of the Rocky Mountain West. Here, we tested the standing hypothesis that recurrent fire in ponderosa pine/Douglas-fir forests of the Inland Northwest decreases total soil N, but increases N turnover and nutrient availability. We compared soils in stands unburned over the past 69-130 years vs. stands exposed to two or more fires over the last 130 years at seven distinct locations in two wilderness areas. Mineral soil samples were collected from each of the seven sites in June and July of 2003 and analyzed for pH, total C and N, potentially mineralizable N (PMN), and extractable NH4+, NO3-, PO4(-3), Ca+2, Mg+2, and K+. Nitrogen transformations were assessed at five sites by installing ionic resin capsules in the mineral soil in August of 2003 and by conducting laboratory assays of nitrification potential and net nitrification in aerobic incubations. Total N and PMN decreased in stands subjected to multiple fires. This loss of total N and labile N was not reflected in concentrations of extractable NH4+ and NO3-. Rather, multiple fires caused an increase in NO3 sorbed on ionic resins, nitrification potential, and net nitrification in spite of the burned stands not having been exposed to fire for at least 12-17 years. Charcoal collected from a recent fire site and added to unburned soils increased nitrification potential, suggesting that the decrease of charcoal in the absence of fire may play an important role in N transformations in fire-dependent ecosystems in the long term. Interestingly, we found no consistent effect of

  12. Stand-replacing wildfires increase nitrification for decades in southwestern ponderosa pine forests.

    PubMed

    Kurth, Valerie J; Hart, Stephen C; Ross, Christopher S; Kaye, Jason P; Fulé, Peter Z

    2014-05-01

    Stand-replacing wildfires are a novel disturbance within ponderosa pine (Pinus ponderosa) forests of the southwestern United States, and they can convert forests to grasslands or shrublands for decades. While most research shows that soil inorganic N pools and fluxes return to pre-fire levels within a few years, we wondered if vegetation conversion (ponderosa pine to bunchgrass) following stand-replacing fires might be accompanied by a long-term shift in N cycling processes. Using a 34-year stand-replacing wildfire chronosequence with paired, adjacent unburned patches, we examined the long-term dynamics of net and gross nitrogen (N) transformations. We hypothesized that N availability in burned patches would become more similar to those in unburned patches over time after fire as these areas become re-vegetated. Burned patches had higher net and gross nitrification rates than unburned patches (P < 0.01 for both), and nitrification accounted for a greater proportion of N mineralization in burned patches for both net (P < 0.01) and gross (P < 0.04) N transformation measurements. However, trends with time-after-fire were not observed for any other variables. Our findings contrast with previous work, which suggested that high nitrification rates are a short-term response to disturbance. Furthermore, high nitrification rates at our site were not simply correlated with the presence of herbaceous vegetation. Instead, we suggest that stand-replacing wildfire triggers a shift in N cycling that is maintained for at least three decades by various factors, including a shift from a woody to an herbaceous ecosystem and the presence of fire-deposited charcoal.

  13. CORRELATION BETWEEN OZONE EXPOSURE AND VISIBLE FOLIAR INJURY IN PONDEROSA AND JEFFREY PINES. (R825433)

    EPA Science Inventory

    Ozone exposure was related to ozone-induced visible foliar injury in ponderosa and Jeffrey pines growing on the western slopes of the Sierra Nevada Mountains of California. Measurements of ozone exposure, chlorotic mottle and fascicle retention were collected during the years ...

  14. Climate effects on fire regimes and tree recruitment in Black Hills ponderosa pine forests.

    PubMed

    Brown, Peter M

    2006-10-01

    Climate influences forest structure through effects on both species demography (recruitment and mortality) and disturbance regimes. Here, I compare multi-century chronologies of regional fire years and tree recruitment from ponderosa pine forests in the Black Hills of southwestern South Dakota and northeastern Wyoming to reconstructions of precipitation and global circulation indices. Regional fire years were affected by droughts and variations in both Pacific and Atlantic sea surface temperatures. Fires were synchronous with La Niñas, cool phases of the Pacific Decadal Oscillation (PDO), and warm phases of the Atlantic Multidecadal Oscillation (AMO). These quasi-periodic circulation features are associated with drought conditions over much of the western United States. The opposite pattern (El Niño, warm PDO, cool AMO) was associated with fewer fires than expected. Regional tree recruitment largely occurred during wet periods in precipitation reconstructions, with the most abundant recruitment coeval with an extended pluvial from the late 1700s to early 1800s. Widespread even-aged cohorts likely were not the result of large crown fires causing overstory mortality, but rather were caused by optimal climate conditions that contributed to synchronous regional recruitment and longer intervals between surface fires. Synchronous recruitment driven by climate is an example of the Moran effect. The presence of abundant fire-scarred trees in multi-aged stands supports a prevailing historical model for ponderosa pine forests in which recurrent surface fires affected heterogenous forest structure, although the Black Hills apparently had a greater range of fire behavior and resulting forest structure over multi-decadal time scales than ponderosa pine forests of the Southwest that burned more often.

  15. Stemwood production patterns in ponderosa pine: effects of stand dynamics and other factors

    Treesearch

    Michael J. Arbaugh; David L. Peterson

    1993-01-01

    Growth patterns of vertical stems in nine ponderosa pines from a stand in the southern Sierra Nevada were analyzed for recent changes due to stand dominance position, age, climate, and ozone exposure. Large positive correlations were found between increments in volume growth and basal area at d.b.h. The results indicated that patterns of wood distribution along the...

  16. Response of dwarf mistletoe-infested ponderosa pine to thinning: 2. Dwarf mistletoe propagation.

    Treesearch

    Lewis F. Roth; James W. Barrett

    1985-01-01

    Propagation of dwarf mistletoe in ponderosa pine saplings is little influenced by thinning overly dense stands to 250 trees per acre. Numerous plants that appear soon after thinning develop from formerly latent plants in the suppressed under-story. Subsequently, dwarf mistletoe propagates nearly as fast as tree crowns enlarge but the rate differs widely among trees....

  17. Effects of elevation and seed source on tracheid length in young ponderosa pine

    Treesearch

    R. M. Echols

    1973-01-01

    Tracheid lengths in 30-year-old ponderosa pine progeny test plantations in the central Sierra Nevada of California were analyzed for effects of (a) elevation of seed parents and (b) elevation and location of test sites. The influence of elevation of seed parents on progeny tracheid length was not significant. Plantation location was significant, but interaction between...

  18. Estimating air-drying times of small-diameter ponderosa pine and Douglas-fir logs

    Treesearch

    William T. Simpson; Xiping Wang

    2004-01-01

    One potential use for small-diameter ponderosa pine and Douglas-fir timber is in log form. Many potential uses of logs require some degree of drying. Even though these small diameters may be considered small in the forestry context, their size when compared to typical lumber thickness dimensions is large. These logs, however, may require uneconomically long kiln-drying...

  19. Growth after thinning ponderosa and Jeffrey pine pole stands in northeastern California

    Treesearch

    William W. Oliver

    1972-01-01

    Thinning ponderosa and Jeffrey pine pole stands (6 to 8 inches d.b.h.) on Meyer Site Classes IV and V land (site index 65 to 80) stimulates growth in diameter and height. This was concluded from data on 12 thinned plots scattered over northeastern California, in natural stands and in a plantation. Basal areas immediately after thinning ranged from 13 to 149 square feet...

  20. Isozyme markers associated with O3 tolerance indicate shift in genetic structure of ponderosa and Jeffrey pine in Sequoia National Park, California

    Treesearch

    J. Staszak; Nancy Grulke; M.J. Marrett; W. Prus-Glowacki

    2007-01-01

    Effects of canopy ozone (O3) exposure and signatures of genetic structure using isozyme markers associated with O3 tolerance were analyzed in ~20-, ~80-, and >200-yr-old ponderosa (Pinus ponderosa Dougl. ex Laws.) and Jeffrey pine (Pinus jeffreyi Grev. & Balf.) in...

  1. Are high elevation pines equally vulnerable to climate change-induced mountain pine beetle attack?

    Treesearch

    Barbara J. Bentz; Erika L. Eidson

    2016-01-01

    Mountain pine beetle (Dendroctonus ponderosae) (MPB), a native insect to western North America, caused extensive tree mortality in pine ecosystems during a recent warm and dry period. More than 24 million acres were affected, including in the relatively low elevation lodgepole (Pinus contorta) and ponderosa (P. ponderosa) pines, and the high-elevation whitebark (P....

  2. Mountain Pine Beetle

    Treesearch

    Gene D. Amman; Mark D. McGregor; Robert E. Jr. Dolph

    1989-01-01

    The mountain pine beetle, Dendroctonus ponderosae Hopkins, is a member of a group of beetles known as bark beetles: Except when adults emerge and attack new trees, the mountain pine beetle completes its life cycle under the bark. The beetle attacks and kills lodgepole, ponderosa, sugar, and western white pines. Outbreaks frequently develop in lodgepole pine stands that...

  3. Effect of body condition on consumption of pine needles (Pinus ponderosa) by beef cows

    USDA-ARS?s Scientific Manuscript database

    We determined if cattle in low (LBC) or high body condition (HBC) would consume different amounts of green pine needles (Pinus ponderosa). Cattle (mature; open Hereford and Hereford x Angus) were fed an adequate basal diet (alfalfa pellets) for Trials 1 and 2; during Trials 3 and 4 cows were fed hig...

  4. Response of planted ponderosa pine seedlings to weed control by herbicide in western Montana

    Treesearch

    Fabian C.C. Uzoh

    1999-01-01

    The effects of competing herbaceous vegetation on the growth of ponderosa pine seedlings with and without herbicide Pronone were characterized in this 1987-1990 study. Study areas were established in 36 plantations across western Montana on Champion International Corporation's timberland (currently owned by Plum Creek Timber Company). The study sites were divided...

  5. Comparison of a -pinene and myrcene on attraction of mountain pine beetle, Dendroctonus ponderosae (Coleoptera: Scolytidae) to pheromones in stands of western white pine

    Treesearch

    Daniel R. Miller; B. Staffan Lindgren

    2000-01-01

    Multiple-funnel traps baited with exo-brevicomin and a mixture of cis- and trans-verbenol were used to test the relative attractiveness of myrcene and (-)-a -pinene to the mountain pine beetle, Dendroctonus ponderosae Hopkins, in a stand...

  6. Role of selection versus historical isolation in racial differentiation of ponderosa pine in southern Oregon: an investigation of alternative hypotheses.

    Treesearch

    Frank C. Sorensen; Nancy L. Mandel; Jan E. Aagaard

    2001-01-01

    Continuous populations identified as Pacific and North Plateau races of ponderosa pine (Pinus ponderosa P. Laws. ex C. Laws.) are parapatric along the crest of the Cascade Range in southern Oregon. A 3-year common-garden study of bud phenology and seedling vigor was performed to estimate the nature and magnitude of differentiation between races, to...

  7. Overstory and understory dynamics in a ponderosa pine plantation vary with stand density in the Sierra Nevada: 40-year results

    Treesearch

    Jianwei Zhang; William W. Oliver; Martin W. Ritchie; Donald L. Neal

    2013-01-01

    We periodically measured overstory ponderosa pine (Pinus ponderosa) growth and understory cover and abundance in a long-term study on the west slope of the Sierra Nevada, California, USA. The study was established in 1969 in a 20-year-old plantation, thinned to basal areas of 9, 16, 23, 30, and 37 m2 ha-1...

  8. Understory vegetation response to mechanical mastication and other fuels treatments in a ponderosa pine forest

    Treesearch

    Jeffrey M. Kane; J. Morgan Varner; Eric E. Knapp

    2010-01-01

    Questions: What influence does mechanical mastication and other fuel treatments have on: (1) canopy and forest floor response variables that influence understory plant development; (2) initial understory vegetation cover, diversity, and composition; and (3) shrub and non-native species density in a secondgrowth ponderosa pine forest....

  9. Methods of cutting ponderosa pine in the Southwest - Establishment report: Even-aged yield study, Plot 11

    Treesearch

    Gilbert H. Schubert

    1963-01-01

    The objectie of the study is to obtain information on the growth and yield of even-aged stands of different stocking levels. The plot will also serve as the start of a detailed growing stock study for the ponderosa pine region.

  10. Three studies on ponderosa pine management on the Warm Springs Indian Reservation: stocking control in uneven-aged stands, forest products from fire-damage trees, and fuels reduction

    Treesearch

    John V. Arena

    2005-01-01

    Over 60,000 acres of ponderosa pine (Pinus ponderosa P. and C. Lawson) forest on the Warm Springs Indian Reservation (WSIR) in Oregon are managed using an uneven-age system. Three on-going studies on WSIR address current issues in the management of pine forests: determining levels of growing stock for uneven-age management, fire effects on wood...

  11. Cost / effectiveness analysis of ponderosa pine ecosystem restoration in Flagstaff Arizona's wildland-urban interface

    Treesearch

    Guy Pinjuv; P. J. Daugherty; Bruce E. Fox

    2001-01-01

    Ponderosa pine ecosystem restoration in Fort Valley (located east of Flagstaff, Arizona) has been proposed as a method of restoring ecosystem health and lowering the risk of catastrophic wildfire in Flagstaff's wildland-urban interface. Three methods of harvest are being used to carry out restoration treatments: hand harvesting, cut-to-length harvesting, and whole...

  12. Restoring eastside ponderosa pine ecosystems at the Blacks Mountain Experimental Forest: a case study

    Treesearch

    Jianwei Zhang; Martin W. Ritchie

    2008-01-01

    The ecological research project of interior ponderosa pine forests at the Blacks Mountain Experimental Forest in northeastern California was initiated by an interdisciplinary team of scientists in the early 1990s. The objectives of this study were to determine the effect of stand structure, and prescribed fire on vegetation growth, resilience, and sustainability of...

  13. Prescribed burning in the interior ponderosa pine type of northeastern California . . . a preliminary study

    Treesearch

    Donald T. Gordon

    1967-01-01

    Three prescribed burns, made in 1959-62, in the interior ponderosa pine type of northeastern California are described in terms of meteorological factors, fire behavior, and effects of burning. Results were generally unsatisfactory. Factors that limit the usefulness of prescribed burning for reducing fuel hazard or for thinning in young-growth stands include stand...

  14. A review of precipitation and temperature control on seedling emergence and establishment for ponderosa and lodgepole pine forest regeneration

    Treesearch

    M. D. Petrie; A. M. Wildeman; J. B. Bradford; Robert Hubbard; W. K. Lauenroth

    2016-01-01

    The persistence of ponderosa pine and lodgepole pine forests in the 21st century depends to a large extent on how seedling emergence and establishment are influenced by driving climate and environmental variables, which largely govern forest regeneration. We surveyed the literature, and identified 96 publications that reported data on dependent variables of seedling...

  15. Growth after partial cutting of ponderosa pine on permanent sample plots in eastern Oregon.

    Treesearch

    Edwin L. Mowat

    1961-01-01

    Between the years 1913 and 1938, seven sets of permanent sample plots were established on the Whitman, Malheur, Rogue River, and Deschutes National Forests in eastern and central Oregon to study the results of various methods of selection cutting in old-growth ponderosa pine stands. This report briefly describes these studies and gives statistics on board-foot growth...

  16. Economic assessment of using a mobile Micromill® for processing small-diameter ponderosa pine.

    Treesearch

    Dennis R. Becker; Evan E. Hjerpe; Eini C. Lowell

    2004-01-01

    An economic assessment of an SLP5000 Diesel Micromill® was conducted to determine the maintenance and operation costs and the logistics of a mobile sawmill used to process small-diameter ponderosa pine. The Micromill® was first introduced in 1997 and has since received considerable attention. In 2003, the USDA Forest Service, Pacific Northwest Research Station...

  17. Development of a mixed shrub–ponderosa pine community in a natural and treated condition

    Treesearch

    Philip M. McDonald; Gary O. Fiddler

    1995-01-01

    On a medium site in northern California, a mostly shrub community was treated by two manual release techniques and by two herbicides, to study its development in both a natural (control) and treated condition. Survival and growth of planted ponderosa pine seedlings were quantified for 8 to 11 years after initial treatment applications. Treatments included manual...

  18. Modeling precipitation-runoff relationships to determine water yield from a ponderosa pine forest watershed

    Treesearch

    Assefa S. Desta

    2006-01-01

    A stochastic precipitation-runoff modeling is used to estimate a cold and warm-seasons water yield from a ponderosa pine forested watershed in the north-central Arizona. The model consists of two parts namely, simulation of the temporal and spatial distribution of precipitation using a stochastic, event-based approach and estimation of water yield from the watershed...

  19. Growth data for 29 elevational transect years from the California study of ponderosa pine

    Treesearch

    M. Thompson Conkle

    1973-01-01

    Ponderosa pine progenies from parents restricted in latitude but spanning 7.000 feet of elevation show significant growth differences in plantations at low-, mid-, and high-elevation test sites. At low- and mid-elevation sites tree heights and diameters of progenies from high-elevation parents were the smallest; those from the low-elevation parents, intermediate; and...

  20. Growth of Ponderosa pine poles thinned to different stocking levels in central Oregon.

    Treesearch

    James W. Barrett

    1983-01-01

    This paper presents 15-year results of one installation of a west-wide study of growing-stock levels in even-aged ponderosa pine. Growth was related to growing-stock level in a 65-year-old pole stand on an above average site. Periodic growth is presented for 10 years after the initial thinning and for 5 years after a second thinning to six assigned growing-stock levels...

  1. The west-wide ponderosa pine levels-of-growing-stock study at age 40

    Treesearch

    William W. Oliver

    2005-01-01

    In the 1960s a series of levels-of-growing-stock studies was established in young, even-aged stands throughout the range of ponderosa pine in the western United States. Using a common study plan, installations were begun in the Black Hills of South Dakota, eastern and central Oregon, the Coconino Plateau of Arizona and the west slope of the Sierra Nevada in California...

  2. Treatment duration and time since disturbance affect vegetation development in a young ponderosa pine plantation

    Treesearch

    Gary O. Fiddler; Philip M. McDonald

    1999-01-01

    The effect of early and delayed release treatments (designated as "Treat first 3 years" and "Treat second 3 years," respectively) on diameter, height, and foliar cover of ponderosa pine seedlings, and density, foliar cover, and height of competing vegetation was evaluated in a young northern California plantation. Manual grubbing created vegetation...

  3. Fine woody fuel particle diameters for improved planar intersect fuel loading estimates in Southern Rocky Mountain ponderosa pine forests

    Treesearch

    Emma Vakili; Chad M. Hoffman; Robert E. Keane

    2016-01-01

    Fuel loading estimates from planar intersect sampling protocols for fine dead down woody surface fuels require an approximation of the mean squared diameter (d2) of 1-h (0-0.63 cm), 10-h (0.63-2.54 cm), and 100-h (2.54-7.62 cm) timelag size classes. The objective of this study is to determine d2 in ponderosa pine (Pinus ponderosa) forests of New Mexico and Colorado,...

  4. Use of a bark thickness—tree diameter relationship for estimating past diameters of ponderosa pine trees.

    Treesearch

    Floyd A. Johnson

    1956-01-01

    Whenever past diameters of ponderosa pine trees are required for growth studies or for other purposes they can be estimated with these formulas: (1) trees 10 inches and over in diameter at breast height Dp=Dn - Wg (1.121) Where Dp...

  5. Snow breakage in a pole-sized ponderosa pine plantation ... more damage at high stand-densities

    Treesearch

    Robert F. Powers; William W. Oliver

    1970-01-01

    Damage by snow breakage to pole-sized ponderosa pine (Pinus pondvosa Laws.) increased as stand density increased. In a plantation on the west slope of California's Sierra Nevada, the tallest trees were most often broken. Thinning in the sapling stage is recommended as a preventative measure in dense plantations subject to heavy snowfall.

  6. A statistical approach to estimate O3 uptake of ponderosa pine in a mediterranean climate

    Treesearch

    N.E. Grulke; H.K. Preisler; C.C. Fan; W.A. Retzlaff

    2002-01-01

    In highly polluted sites, stomatal behavior is sluggish with respect to light, vapor pressure deficit, and internal CO2 concentration (Ci) and poorly described by existing models. Statistical models were developed to estimate stomatal conductance (gs) of 40-year-old ponderosa pine at three sites differing in pollutant exposure for the purpose of...

  7. Genetic variation in ponderosa pine: A 15-year test of provenances in the Great Plains

    Treesearch

    David F. Van Haverbeke

    1986-01-01

    Survival was highest and height growth greatest in ponderosa pine provenances from northcentral Nebraska, southwest South Dakota, and the High Plains region. Genotype x environment interaction was minimal in central and northern Great Plains plantations. Age/age correlations indicate provenances expressing superior height growth can be identified after 5 or 10 years....

  8. Crossability and relationships of Washoe pine

    Treesearch

    William B. Critchfield

    1984-01-01

    Washoe pine, related to ponderosa pine but occurring at higher elevations along the western edge of the Great Basin, crosses freely with the Rocky Mountain race of ponderosa despite evidence of long-term separation. Washoe is morphologically distinct from the parapatric Pacific race of ponderosa pine, and the two taxa are kept separate partly by genetic barriers that...

  9. Impact of mountain pine beetle outbreaks on forest albedo and radiative forcing, as derived from Moderate Resolution Imaging Spectroradiometer, Rocky Mountains, USA

    NASA Astrophysics Data System (ADS)

    Vanderhoof, M.; Williams, C. A.; Ghimire, B.; Rogan, J.

    2013-12-01

    pine beetle (Dendroctonus ponderosae) outbreaks in North America are widespread and have potentially large-scale impacts on albedo and associated radiative forcing. Mountain pine beetle outbreaks in Colorado and southern Wyoming have resulted in persistent and significant increases in both winter albedo (change peaked 10 years post outbreak at 0.06 ± 0.01 and 0.05 ± 0.01, in lodgepole pine (Pinus contorta) and ponderosa pine (Pinus ponderosa) stands, respectively) and spring albedo (change peaked 10 years post outbreak at 0.06 ± 0.01 and 0.04 ± 0.01, in lodgepole pine and ponderosa pine stands, respectively). Instantaneous top-of-atmosphere radiative forcing peaked for both lodgepole pine and ponderosa pine stands in winter at 10 years post outbreak at -1.7 ± 0.2 W m-2 and -1.4 ± 0.2 W m-2, respectively. The persistent increase in albedo with time since mountain pine beetle disturbance combined with the continued progression of the attack across the landscape from 1994-2011 resulted in an exponential increase in winter and annual radiative cooling (MW) over time. In 2011 the rate of radiative forcing within the study area reached -982.7 ± 139.0 MW, -269.8 ± 38.2 MW, -31.1 ± 4.4 MW, and -147.8 ± 20.9 MW in winter, spring, summer, and fall, respectively. An increase in radiative cooling has the potential to decrease sensible and/or latent heat flux by reducing available energy. Such changes could affect current mountain pine beetle outbreaks which are influenced by climatic conditions.

  10. Growth of ponderosa pine thinned to different stocking levels in the western United States

    Treesearch

    William W. Oliver; Carleton B. Edminster

    1988-01-01

    Growth of ponderosa pine was studied by the western Forest and Range Experiment Stations of the USDA Forest Service in response to increasing demands for better and more precise estimates of yields possible through intensive management. We summarized results of 15 to 20 years of growth after thinning each of five stands to a wide range of stocking levels. The stands-...

  11. Condition of live fire-scarred ponderosa pine eleven years after removing partial cross-sections

    Treesearch

    Emily K. Heyerdahl; Steven J. McKay

    2008-01-01

    Our objective is to report mortality rates for ponderosa pine trees in Oregon ten to eleven years after removing a fire-scarred partial cross-section from them, and five years after an initial survey of post-sampling mortality. We surveyed 138 live trees from which we removed fire-scarred partial crosssections in 1994/95 and 387 similarly sized, unsampled neighbor...

  12. Timing and duration of release affect vegetation development in a young ponderosa pine plantation

    Treesearch

    Philip M. McDonald; Gary O. Fiddler

    2001-01-01

    The density and development of greenleaf manzanita, other shrubs, and graminoids were evaluated in a young ponderosa pine plantation on a poor site in northern California from 1988 through 1997. Manual grubbing to a 5-foot radius created treatment regimes that lasted for 3 to 6 years and vegetation recovery times of 4 to 10 years. The duration and timing of the...

  13. Effects of CO2 and Nitrogen Fertilization on Growth and Nutrient Content of Juvenile Ponderosa Pine (NDP-061A)

    DOE Data Explorer

    Johnson, Dale W [Desert Research Inst. (DRI), Reno, NV (United States); University of Nevada, Reno, NV (United States); Ball, J. Timothy [Desert Research Inst. (DRI), Reno, NV (United States); Walker, Roger F [University of Nevada, Reno, NV (United States)

    1998-03-01

    This data set presents measured values of plant diameter and height, biomass of plant components, and nutrient (carbon, nitrogen, phosphorus, sulfur, potassium, calcium, magnesium, boron, copper, iron, manganese, and zinc) concentrations from a study of the effects of carbon dioxide and nitrogen fertilization on ponderosa pine (Pinus ponderosa Dougl. ex Laws.) conducted in open-top chambers in Placerville, California, from 1991 through 1996. This data set contains values from 1991 through 1993.

  14. Structural injury underlying mottling in ponderosa pine needles exposed to ambient ozone concentrations in the San Bernardino Mountains near Los Angeles, California

    Treesearch

    Pierre Vollenweider; Mark E. Fenn; Terry Menard; Madeleine Gunthardt-Goerg; Andrzej Bytnerowicz

    2013-01-01

    For several decades, southern California experienced the worst ozone pollution ever reported. Peak ozone concentrations have, however, declined steadily since 1980. In this study, the structural injuries underlying ozone symptoms in needles of ponderosa pine (Pinus ponderosa Dougl. ex Laws.) collected in summer 2006 from one of the most polluted sites in the San...

  15. Areas of Agreement and Disagreement Regarding Ponderosa Pine and Mixed Conifer Forest Fire Regimes: A Dialogue with Stevens et al.

    PubMed Central

    Odion, Dennis C.; Hanson, Chad T.; Baker, William L.; DellaSala, Dominick A.; Williams, Mark A.

    2016-01-01

    In a recent PLOS ONE paper, we conducted an evidence-based analysis of current versus historical fire regimes and concluded that traditionally defined reference conditions of low-severity fire regimes for ponderosa pine (Pinus ponderosa) and mixed-conifer forests were incomplete, missing considerable variability in forest structure and fire regimes. Stevens et al. (this issue) agree that high-severity fire was a component of these forests, but disagree that one of the several sources of evidence, stand age from a large number of forest inventory and analysis (FIA) plots across the western USA, support our findings that severe fire played more than a minor role ecologically in these forests. Here we highlight areas of agreement and disagreement about past fire, and analyze the methods Stevens et al. used to assess the FIA stand-age data. We found a major problem with a calculation they used to conclude that the FIA data were not useful for evaluating fire regimes. Their calculation, as well as a narrowing of the definition of high-severity fire from the one we used, leads to a large underestimate of conditions consistent with historical high-severity fire. The FIA stand age data do have limitations but they are consistent with other landscape-inference data sources in supporting a broader paradigm about historical variability of fire in ponderosa and mixed-conifer forests than had been traditionally recognized, as described in our previous PLOS paper. PMID:27195808

  16. Areas of Agreement and Disagreement Regarding Ponderosa Pine and Mixed Conifer Forest Fire Regimes: A Dialogue with Stevens et al.

    PubMed

    Odion, Dennis C; Hanson, Chad T; Baker, William L; DellaSala, Dominick A; Williams, Mark A

    2016-01-01

    In a recent PLOS ONE paper, we conducted an evidence-based analysis of current versus historical fire regimes and concluded that traditionally defined reference conditions of low-severity fire regimes for ponderosa pine (Pinus ponderosa) and mixed-conifer forests were incomplete, missing considerable variability in forest structure and fire regimes. Stevens et al. (this issue) agree that high-severity fire was a component of these forests, but disagree that one of the several sources of evidence, stand age from a large number of forest inventory and analysis (FIA) plots across the western USA, support our findings that severe fire played more than a minor role ecologically in these forests. Here we highlight areas of agreement and disagreement about past fire, and analyze the methods Stevens et al. used to assess the FIA stand-age data. We found a major problem with a calculation they used to conclude that the FIA data were not useful for evaluating fire regimes. Their calculation, as well as a narrowing of the definition of high-severity fire from the one we used, leads to a large underestimate of conditions consistent with historical high-severity fire. The FIA stand age data do have limitations but they are consistent with other landscape-inference data sources in supporting a broader paradigm about historical variability of fire in ponderosa and mixed-conifer forests than had been traditionally recognized, as described in our previous PLOS paper.

  17. Simulating historical disturbance regimes and stand structures in old-forest ponderosa pine/Douglas-fir forests

    Treesearch

    Mike Hillis; Vick Applegate; Steve Slaughter; Michael G. Harrington; Helen Smith

    2001-01-01

    Forest Service land managers, with the collaborative assistance from research, applied a disturbance based restoration strategy to rehabilitate a greatly-altered, high risk Northern Rocky Mountain old-forest ponderosa pine-Douglas-fir stand. Age-class structure and fire history for the site have been documented in two research papers (Arno and others 1995, 1997)....

  18. Reticulate evolution and incomplete lineage sorting among the ponderosa pines.

    PubMed

    Willyard, Ann; Cronn, Richard; Liston, Aaron

    2009-08-01

    Interspecific gene flow via hybridization may play a major role in evolution by creating reticulate rather than hierarchical lineages in plant species. Occasional diploid pine hybrids indicate the potential for introgression, but reticulation is hard to detect because ancestral polymorphism is still shared across many groups of pine species. Nucleotide sequences for 53 accessions from 17 species in subsection Ponderosae (Pinus) provide evidence for reticulate evolution. Two discordant patterns among independent low-copy nuclear gene trees and a chloroplast haplotype are better explained by introgression than incomplete lineage sorting or other causes of incongruence. Conflicting resolution of three monophyletic Pinus coulteri accessions is best explained by ancient introgression followed by a genetic bottleneck. More recent hybridization transferred a chloroplast from P. jeffreyi to a sympatric P. washoensis individual. We conclude that incomplete lineage sorting could account for other examples of non-monophyly, and caution against any analysis based on single-accession or single-locus sampling in Pinus.

  19. Fire-induced erosion and millennial-scale climate change in northern ponderosa pine forests.

    PubMed

    Pierce, Jennifer L; Meyer, Grant A; Jull, A J Timothy

    2004-11-04

    Western US ponderosa pine forests have recently suffered extensive stand-replacing fires followed by hillslope erosion and sedimentation. These fires are usually attributed to increased stand density as a result of fire suppression, grazing and other land use, and are often considered uncharacteristic or unprecedented. Tree-ring records from the past 500 years indicate that before Euro-American settlement, frequent, low-severity fires maintained open stands. However, the pre-settlement period between about ad 1500 and ad 1900 was also generally colder than present, raising the possibility that rapid twentieth-century warming promoted recent catastrophic fires. Here we date fire-related sediment deposits in alluvial fans in central Idaho to reconstruct Holocene fire history in xeric ponderosa pine forests and examine links to climate. We find that colder periods experienced frequent low-severity fires, probably fuelled by increased understory growth. Warmer periods experienced severe droughts, stand-replacing fires and large debris-flow events that comprise a large component of long-term erosion and coincide with similar events in sub-alpine forests of Yellowstone National Park. Our results suggest that given the powerful influence of climate, restoration of processes typical of pre-settlement times may be difficult in a warmer future that promotes severe fires.

  20. Vegetation Trends in a 31-Year-Old Ponderosa Pine Plantation: Effect of Different Shrub Densities

    Treesearch

    Philip M. McDonald; Celeste S. Abbott

    1997-01-01

    On a poor site in northern California, a brushfield community was treated in various ways which left initial densities of no shrubs, light, medium, and heavy shrubs. Density and development (height, foliar cover, crown volume) for three shrub species (alone and combined), one grass, and planted ponderosa pine in these categories were quantified from 1966 to 1992....

  1. Ponderosa pine growth response to soil strength in the volcanic ash soils of central Oregon.

    Treesearch

    R.T. Parker; D.A. Maguire; D.D. Marshall; P. Cochran

    2007-01-01

    Mechanical harvesting and associated logging activities have the capacity to compact soil across large portions of harvest units. Two thinning treatments (felled only versus felled and skidded) in 70- to 80-year-old ponderosa pine stands were replicated at three sites with volcanic soils in central Oregon. Growth in diameter, height, and volume of residual trees were...

  2. Estimating site index of ponderosa pine in Northern California...standard curves, soil series, stem analysis

    Treesearch

    Robert F. Powers

    1972-01-01

    Four sets of standard site index curves based on statewide or regionwide averages were compared with data on natural growth from nine young stands of ponderosa pine in northern California. The curves tested were by Meyer; Dunning; Dunning and Reineke; and Arvanitis, Lindquist, and Palley. The effects of soils on height growth were also studied. Among the curves tested...

  3. Effectiveness of litter removal in preventing mortality of yellow barked ponderosa pine in northern Arizona

    Treesearch

    James F. Fowler; Carolyn H. Sieg; Linda Wadleigh; Sally Haase

    2007-01-01

    Removal of deep litter and duff from the base of mature southwestern ponderosa pine trees is commonly recommended to reduce mortality following prescribed burns, but experimental studies that quantify the effectiveness of such practices in reducing mortality are lacking. Following a pilot study, on each of four sites in northern Arizona we monitored 15-16 sets of 8...

  4. Restoring forest structure and process stabilizes forest carbon in wildfire-prone southwestern ponderosa pine forests

    Treesearch

    Matthew D. Hurteau; Shuang Liang; Katherine L. Martin; Malcolm P. North; George W. Koch; Bruce A. Hungate

    2016-01-01

    Changing climate and a legacy of fire-exclusion have increased the probability of high-severity wildfire, leading to an increased risk of forest carbon loss in ponderosa pine forests in the southwestern USA. Efforts to reduce high-severity fire risk through forest thinning and prescribed burning require both the removal and emission of carbon from these forests, and...

  5. Mountain pine beetle-caused mortality over eight years in two pine hosts in mixed conifer stands of the southern Rocky Mountains

    USGS Publications Warehouse

    West, Daniel R.; Briggs, Jennifer S.; Jacobi, William R.; Negrón, José F.

    2014-01-01

    Eruptive mountain pine beetle (Dendroctonus ponderosae, MPB) populations have caused widespread mortality of pines throughout western North America since the late 1990s. Early work by A.D. Hopkins suggested that when alternate host species are available, MPB will prefer to breed in the host to which it has become adapted. In Colorado, epidemic MPB populations that originated in lodgepole pine expanded into mixed-conifer stands containing ponderosa pine, a related host. We evaluated the susceptibility of both hosts to successful MPB colonization in a survey of 19 sites in pine-dominated mixed-conifer stands spanning 140 km of the Front Range, CO, USA. In each of three 0.2-ha plots at each site, we (1) assessed trees in the annual flights of 2008–2011 to compare MPB-caused mortality between lodgepole and ponderosa pine; (2) recorded previous MPB-caused tree mortality from 2004–2007 to establish baseline mortality levels; and (3) measured characteristics of the stands (e.g. tree basal area) and sites (e.g. elevation, aspect) that might be correlated with MPB colonization. Uninfested average live basal area of lodgepole and ponderosa pine was 74% of total basal area before 2004. We found that for both species, annual percent basal area of attacked trees was greatest in one year (2009), and was lower in all other years (2004–2007, 2008, 2010, and 2011). Both pine species had similar average total mortality of 38–39% by 2011. Significant predictors of ponderosa pine mortality in a given year were basal area of uninfested ponderosa pine and the previous year’s mortality levels in both ponderosa and lodgepole pine. Lodgepole pine mortality was predicted by uninfested basal areas of both lodgepole and ponderosa pine, and the previous year’s lodgepole pine mortality. These results indicate host selection by MPB from lodgepole pine natal hosts into ponderosa pine the following year, but not the reverse. In both species, diameters of attacked trees within each year

  6. Inoculum reduction measures to manage Armillaria root disease in a severely infected stand of ponderosa pine in south-central Washington: 35-year results

    Treesearch

    Charles G. Shaw; D.W. Omdal; A. Ramsey-Kroll; L.F. Roth

    2012-01-01

    A stand of ponderosa pine (Pinus ponderosa) severely affected by Armillaria root disease was treated with five different levels of sanitation by root removal to reduce root disease losses in the regenerating stand. Treatments included the following: (1) all trees pushed over by machine, maximum removal of roots by machine ripping, and visible...

  7. Evaluation of insecticides for protecting Southwestern ponderosa pines from attack by engraver beetles (Coleoptera: Curculionidae: Scolytinae).

    PubMed

    DeGomez, Tom E; Hayes, Christopher J; Anhold, John A; McMillin, Joel D; Clancy, Karen M; Bosu, Paul P

    2006-04-01

    Insecticides that might protect pine trees from attack by engraver beetles (Ips spp.) have not been rigorously tested in the southwestern United States. We conducted two field experiments to evaluate the efficacy of several currently and potentially labeled preventative insecticides for protecting high-value ponderosa pine, Pinus ponderosa Dougl ex. Laws., from attack by engraver beetles. Preventative sprays (0.19% permethrin [Permethrin Plus C]; 0.03, 0.06, and 0.12% bifenthrin [Onyx]; and 1.0 and 2.0% carbaryl [Sevin SL] formulations) and systemic implants (0.875 g per capsule acephate [Acecap] and 0.650 g per capsule dinotefuran) were assessed on bolts (sections of logs) as a surrogate for live trees for a period of 13 mo posttreatment. The pine engraver, Ips pini (Say), was the most common bark beetle found attacking control and treated bolts, but sixspined ips, Ips calligraphus (Germar), and Ips lecontei Swain also were present. After approximately 13 mo posttreatment in one experiment, the spray treatments with 2.0% carbaryl, 0.19% permethrin, and 0.06 or 0.12% bifenthrin prevented Ips attack on the bolts at a protection level of > or = 70%. The acephate and dinotefuran systemic insecticides, and the 0.03% bifenthrin spray, provided inadequate (< or = 36%) protection in this experiment. For the other experiment, sprayed applications of 1.0% carbaryl, 0.19% permethrin, and 0.06% bifenthrin prevented beetle attack at protection levels of > or = 90, > or = 80, and > or = 70%, respectively, when bolts were exposed to Ips beetle attack for approximately 9-15 wk posttreatment. The sprays with 0.19% permethrin and 0.06% bifenthrin also provided > or = 90% protection when bolts were exposed for approximately 15-54 wk posttreatment. We concluded that under the conditions tested, 1.0 and 2.0% carbaryl, 0.19% permethrin, and 0.06 and 0.12% binfenthrin were acceptable preventative treatments for protecting ponderosa pine from successful engraver beetle attack for one

  8. Fifteen-year growth patterns after thinning a ponderosa-Jeffrey pine plantation in northeastern California

    Treesearch

    William W. Oliver

    1979-01-01

    Growth was analyzed after one thinning in a plantation of pole-size ponderosa and Jeffrey pines on land having a site index of 50 feet at 50 years. Periodic annual increment was determined for each of three 5-year periods. On this basis, increment in diameter and cubic volume were found to he related closely to stand basal area only. Basal area and height increment,...

  9. Southwestern Pine Tip Moth

    Treesearch

    Daniel T. Jennings; Robert E. Stevens

    1982-01-01

    The southwestern pine tip moth, Rhyacionia neomexicana (Dyar), injures young ponderosa pines (Pinus ponderosa Dougl. ex Laws) in the Southwest, central Rockies, and midwestern plains. Larvae feed on and destroy new, expanding shoots, often seriously reducing terminal growth of both naturally regenerated and planted pines. The tip moth is especially damaging to trees on...

  10. AmeriFlux US-CZ2 Sierra Critical Zone, Sierra Transect, Ponderosa Pine Forest, Soaproot Saddle

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Goulden, Michael

    This is the AmeriFlux version of the carbon flux data for the site US-CZ2 Sierra Critical Zone, Sierra Transect, Ponderosa Pine Forest, Soaproot Saddle. Site Description - Half hourly data are available at https://www.ess.uci.edu/~california/. This site is one of four Southern Sierra Critical Zone Observatory flux towers operated along an elevation gradient (sites are USCZ1, USCZ2, USCZ3 and USCZ4). This site is an oak/pine forest, with occasional thinning and wildfire, a prescribed understory burn ~2012, and severe drought and ~80% canopy mortality in 2011-15

  11. Height growth and site index curves for managed even-aged stands of ponderosa pine in the Pacific Northwest

    Treesearch

    James W. Barrett

    1978-01-01

    This paper presents height growth and site index curves and equations for even-aged, managed stands of ponderosa pine east of the Cascade Range in Oregon and Washington where height growth has not been suppressed by high density or related factors.

  12. Mark-recapture estimation of snag standing rates in northern Arizona mixed-conifer and ponderosa pine forests

    Treesearch

    Joseph L. Ganey; Gary C. White; Jeffrey S. Jenness; Scott C. Vojta

    2015-01-01

    Snags (standing dead trees) are important components of forests that provide resources for numerous species of wildlife and contribute to decay dynamics and other ecological processes. Managers charged with managing populations of snags need information about standing rates of snags and factors influencing those rates, yet such data are limited for ponderosa pine (...

  13. Forest restoration and fuels reduction in ponderosa pine and dry mixed conifer in the Southwest

    Treesearch

    Marlin Johnson

    2008-01-01

    (Please note, this is an abstract only) Most people agree that ponderosa pine and dry mixed conifer stands need to be thinned and burned to move the stands to within a normal range of variability. Unfortunately, people are in disagreement beyond that point. To some, restoration and fuels reduction means restoring stands to more open, pre-European (pre-1880) conditions...

  14. Distribution of fine roots of ponderosa pine and Douglas-fir in a central Idaho forest

    Treesearch

    Gabriel Dumm; Lauren Fins; Russell T. Graham; Theresa B. Jain

    2008-01-01

    This study describes soil horizon depth and fine root distribution in cores collected at two distances from the boles of Douglas-fir and ponderosa pine trees at a study site in a central Idaho forest. Concentration and content of fine roots extracted from soil cores were compared among species, soil horizons, tree size, and distance from bole. Approximately 80% of...

  15. Condition of live fire-scarred ponderosa pine trees six years after removing partial cross sections

    Treesearch

    Emily K. Heyerdahl; Steven J. McKay

    2001-01-01

    Our objective was to document the effect of fire-history sampling on the mortality of mature ponderosa pine trees in Oregon. We examined 138 trees from which fire-scarred partial cross sections had been removed five to six years earlier, and 386 similarly sized, unsampled neighbor trees, from 78 plots distributed over about 5,000 ha. Mortality was low for both groups....

  16. Developing resilient ponderosa pine forests with mechanical thinning and prescribed fire in central Oregon's pumice region

    Treesearch

    Matt D. Busse; P.H. Cochran; William E. Hopkins; William H. Johnson; Gregg M. Riegel; Gary O. Fiddler; Alice W. Ratcliff; Carol J. Shestak

    2009-01-01

    Thinning and prescribed burning are common management practices for reducing fuel buildup in ponderosa pine forests. However, it is not well understood if their combined use is required to lower wildfire risk and to help restore natural ecological function. We compared 16 treatment combinations of thinning, prescribed fire, and slash retention for two decades...

  17. Response of antelope bitterbrush to repeated prescribed burning in Central Oregon ponderosa pine forests

    Treesearch

    Matt D. Busse; Gregg M. Riegel

    2009-01-01

    Antelope  bitterbrush is a dominant shrub in many interior ponderosa pine forests  in the western United States. How it responds to prescribed fire is not  well understood, yet is of considerable concern to wildlife and fire  managers alike given its importance as a browse species and as a ladder  fuel in these fire-prone forests. We quantified bitterbrush...

  18. Quantum Yields in Mixed-Conifer Forests and Ponderosa Pine Plantations

    NASA Astrophysics Data System (ADS)

    Wei, L.; Marshall, J. D.; Zhang, J.

    2008-12-01

    Most process-based physiological models require canopy quantum yield of photosynthesis as a starting point to simulate carbon sequestration and subsequently gross primary production (GPP). The quantum yield is a measure of photosynthetic efficiency expressed in moles of CO2 assimilated per mole of photons absorbed; the process is influenced by environmental factors. In the summer 2008, we measured quantum yields on both sun and shade leaves for four conifer species at five sites within Mica Creek Experimental Watershed (MCEW) in northern Idaho and one conifer species at three sites in northern California. The MCEW forest is typical of mixed conifer stands dominated by grand fir (Abies grandis (Douglas ex D. Don) Lindl.). In northern California, the three sites with contrasting site qualities are ponderosa pine (Pinus ponderosa C. Lawson var. ponderosa) plantations that were experimentally treated with vegetation control, fertilization, and a combination of both. We found that quantum yields in MCEW ranged from ~0.045 to ~0.075 mol CO2 per mol incident photon. However, there were no significant differences between canopy positions, or among sites or tree species. In northern California, the mean value of quantum yield of three sites was 0.051 mol CO2/mol incident photon. No significant difference in quantum yield was found between canopy positions, or among treatments or sites. The results suggest that these conifer species maintain relatively consistent quantum yield in both MCEW and northern California. This consistency simplifies the use of a process-based model to accurately predict forest productivity in these areas.

  19. Roots and Their Rhizosphere of Fremont Cottonwood and Ponderosa Pine Substantially Stimulated Soil Organic Carbon Decomposition.

    NASA Astrophysics Data System (ADS)

    Dijkstra, F. A.; Cheng, W.

    2006-12-01

    There is increasing evidence that living plant roots can significantly alter soil microbial activity and soil organic carbon (SOC) decomposition. Most research on rhizosphere effects on SOC has been done in short-term experiments using annual plants. Here we test if rhizosphere processes of two woody perennial plant species, Fremont cottonwood (Populus fremontii) and Ponderosa pine (Pinus ponderosa), affect SOC decomposition in three different soil types in a 395-day greenhouse experiment. We continuously labeled plants with depleted 13C, which allowed us to separate plant-derived CO2-C from original soil-derived CO2-C in soil respiration measurements. Results show that after 100 days of planting both cottonwood (by 79%) and pine (by 108%) significantly increased soil carbon decomposition compared to soils without plants ("primed C"). We observed no differences in primed C among the three soil types, despite their differences in total and labile carbon and available nitrogen content. Instead, primed C was positively related to foliar biomass. Our results indicate that rhizosphere effects on SOC decomposition play an important role in the carbon cycle of forested ecosystems.

  20. Price projections for selected grades of Douglas-fir, coast hem-fir, inland hem-fir, and ponderosa pine lumber.

    Treesearch

    Richard W. Haynes; Roger D. Fight

    1992-01-01

    Grade-specific price projections were developed for Douglas-fir, coast hem-fir, inland hem-fir, and ponderosa pine lumber. These grade-specific price projections can be used in evaluating management practices that will affect the quality of saw logs produced under various management regimes.

  1. Late Holocene geomorphic record of fire in ponderosa pine and mixed-conifer forests, Kendrick Mountain, northern Arizona, USA

    USGS Publications Warehouse

    Jenkins, S.E.; Hull, Sieg C.; Anderson, D.E.; Kaufman, D.S.; Pearthree, P.A.

    2011-01-01

    Long-term fire history reconstructions enhance our understanding of fire behaviour and associated geomorphic hazards in forested ecosystems. We used 14C ages on charcoal from fire-induced debris-flow deposits to date prehistoric fires on Kendrick Mountain, northern Arizona, USA. Fire-related debris-flow sedimentation dominates Holocene fan deposition in the study area. Radiocarbon ages indicate that stand-replacing fire has been an important phenomenon in late Holocene ponderosa pine (Pinus ponderosa) and ponderosa pine-mixed conifer forests on steep slopes. Fires have occurred on centennial scales during this period, although temporal hiatuses between recorded fires vary widely and appear to have decreased during the past 2000 years. Steep slopes and complex terrain may be responsible for localised crown fire behaviour through preheating by vertical fuel arrangement and accumulation of excessive fuels. Holocene wildfire-induced debris flow events occurred without a clear relationship to regional climatic shifts (decadal to millennial), suggesting that interannual moisture variability may determine fire year. Fire-debris flow sequences are recorded when (1) sufficient time has passed (centuries) to accumulate fuels; and (2) stored sediment is available to support debris flows. The frequency of reconstructed debris flows should be considered a minimum for severe events in the study area, as fuel production may outpace sediment storage. ?? IAWF 2011.

  2. Stomata open at night in pole-sized and mature ponderosa pine: implications for O3 exposure metrics.

    PubMed

    Grulke, N E; Alonso, R; Nguyen, T; Cascio, C; Dobrowolski, W

    2004-09-01

    Ponderosa pine (Pinus ponderosa Dougl. ex Laws.) is widely distributed in the western USA. We report the lack of stomatal closure at night in early summer for ponderosa pine at two of three sites investigated. Trees at a third site with lower nitrogen dioxide and nitric acid exposure, but greater drought stress, had slightly open stomata at night in early summer but closed stomata at night for the rest of the summer. The three sites had similar background ozone exposure during the summer of measurement (2001). Nighttime stomatal conductance (gs) ranged from one tenth to one fifth that of maximum daytime values. In general, pole-sized trees (< 40 years old) had greater nighttime gs than mature trees (> 250 years old). In late summer, nighttime gs was low (< 3.0 mmol H2O m(-2) s(-1)) for both tree size classes at all sites. Measurable nighttime gs has also been reported in other conifers, but the values we observed were higher. In June, nighttime ozone (O3) uptake accounted for 9, 5 and 3% of the total daily O3 uptake of pole-sized trees from west to east across the San Bernardino Mountains. In late summer, O3 uptake at night was < 2% of diel uptake at all sites. Nocturnal O3 uptake may contribute to greater oxidant injury development, especially in pole-sized trees in early summer.

  3. Change in soil fungal community structure driven by a decline in ectomycorrhizal fungi following a mountain pine beetle (Dendroctonus ponderosae) outbreak.

    PubMed

    Pec, Gregory J; Karst, Justine; Taylor, D Lee; Cigan, Paul W; Erbilgin, Nadir; Cooke, Janice E K; Simard, Suzanne W; Cahill, James F

    2017-01-01

    Western North American landscapes are rapidly being transformed by forest die-off caused by mountain pine beetle (Dendroctonus ponderosae), with implications for plant and soil communities. The mechanisms that drive changes in soil community structure, particularly for the highly prevalent ectomycorrhizal fungi in pine forests, are complex and intertwined. Critical to enhancing understanding will be disentangling the relative importance of host tree mortality from changes in soil chemistry following tree death. Here, we used a recent bark beetle outbreak in lodgepole pine (Pinus contorta) forests of western Canada to test whether the effects of tree mortality altered the richness and composition of belowground fungal communities, including ectomycorrhizal and saprotrophic fungi. We also determined the effects of environmental factors (i.e. soil nutrients, moisture, and phenolics) and geographical distance, both of which can influence the richness and composition of soil fungi. The richness of both groups of soil fungi declined and the overall composition was altered by beetle-induced tree mortality. Soil nutrients, soil phenolics and geographical distance influenced the community structure of soil fungi; however, the relative importance of these factors differed between ectomycorrhizal and saprotrophic fungi. The independent effects of tree mortality, soil phenolics and geographical distance influenced the community composition of ectomycorrhizal fungi, while the community composition of saprotrophic fungi was weakly but significantly correlated with the geographical distance of plots. Taken together, our results indicate that both deterministic and stochastic processes structure soil fungal communities following landscape-scale insect outbreaks and reflect the independent roles tree mortality, soil chemistry and geographical distance play in regulating the community composition of soil fungi. © 2016 The Authors. New Phytologist © 2016 New Phytologist Trust.

  4. Controls on vegetation structure in Southwestern ponderosa pine forests, 1941 and 2004.

    PubMed

    Bakker, Jonathan D; Moore, Margaret M

    2007-09-01

    Long-term studies can broaden our ecological understanding and are particularly important when examining contingent effects that involve changes to dominance by long-lived species. Such a change occurred during the last century in Southwestern (USA) ponderosa pine (Pinus ponderosa) forests. We used five livestock grazing exclosures established in 1912 to quantify vegetation structure in 1941 and 2004. Our objectives were to (1) assess the effects of historical livestock grazing on overstory structure and age distribution, (2) assess the effects of recent livestock grazing and overstory on understory vegetation, and (3) quantify and explain changes in understory vegetation between 1941 and 2004. In 1941, canopy cover of tree regeneration was significantly higher inside exclosures. In 2004, total tree canopy cover was twice as high, density was three times higher, trees were smaller, and total basal area was 40% higher inside exclosures. Understory species density, herbaceous plant density, and herbaceous cover were negatively correlated with overstory vegetation in both years. Most understory variables did not differ between grazing treatments in 1941 but were lower inside exclosures in 2004. Differences between grazing treatments disappeared once overstory effects were accounted for, indicating that they were due to the differential overstory response to historical livestock grazing practices. Between 1941 and 2004, species density declined by 34%, herbaceous plant density by 37%, shrub cover by 69%, total herbaceous cover by 59%, graminoid cover by 39%, and forb cover by 82%. However, these variables did not differ between grazing treatments or years once overstory effects were accounted for, indicating that the declines were driven by the increased dominance of the overstory during this period. Our results demonstrate that historical livestock grazing practices are an aspect of land-use history that can affect ecosystem development. Grazing history must be

  5. Recovery of a bearclover (Chamaebatia foliolosa) plant community after site preparation and planting of ponderosa pine seedlings

    Treesearch

    Philip M. McDonald; Gary O. Fiddler

    1999-01-01

    Bearclover inhabits thousands of acres of forest land in northern and central California, but little quantification of its recovery after timber harvest, site preparation, and planting is available. And the species composition and development of the ensuing plant community is largely unknown. Density, foliar cover, and height of planted ponderosa pine seedlings,...

  6. Xylem vulnerability to cavitation in Pseudotsuga menziesii and Pinus ponderosa from contrasting habitats.

    PubMed

    Stout, Deborah H; Sala, Anna

    2003-01-01

    In the Rocky Mountains, ponderosa pine (Pinus ponderosa (ssp.) ponderosa Dougl. ex P. Laws. & C. Laws) often co-occurs with Douglas-fir (Pseudotsuga menziesii var. glauca (Mayr) Franco). Despite previous reports showing higher shoot vulnerability to water-stress-induced cavitation in ponderosa pine, this species extends into drier habitats than Douglas-fir. We examined: (1) whether roots and shoots of ponderosa pine in riparian and slope habitats are more vulnerable to water-stress-induced cavitation than those of Douglas-fir; (2) whether species-specific differences in vulnerability translate into differences in specific conductivity in the field; and (3) whether the ability of ponderosa pine to extend into drier sites is a result of (a) greater plasticity in hydraulic properties or (b) functional or structural adjustments. Roots and shoots of ponderosa pine were significantly more vulnerable to water-stress-induced cavitation (overall mean cavitation pressure, Psi(50%) +/- SE = -3.11 +/- 0.32 MPa for shoots and -0.99 +/- 0.16 MPa for roots) than those of Douglas-fir (Psi(50%) +/- SE = -4.83 +/- 0.40 MPa for shoots and -2.12 +/- 0.35 MPa for roots). However, shoot specific conductivity did not differ between species in the field. For both species, roots were more vulnerable to cavitation than shoots. Overall, changes in vulnerability from riparian to slope habitats were small for both species. Greater declines in stomatal conductance as the summer proceeded, combined with higher allocation to sapwood and greater sapwood water storage, appeared to contribute to the ability of ponderosa pine to thrive in dry habitats despite relatively high vulnerability to water-stress-induced cavitation.

  7. Using acoustic analysis to presort warp-prone ponderosa pine 2 by 4s before kiln-drying

    Treesearch

    Xiping Wang; William T. Simpson

    2006-01-01

    This study evaluated the potential of acoustic analysis as presorting criteria to identify warp-prone boards before kiln-drying. Dimension lumber, 38 by 89 mm (nominal 2 by 4 in.) and 2.44 m (8 it) long, sawn from open-grown small-diameter ponderosa pine trees, was acoustically tested lengthwise at green condition. Three acoustic properties (acoustic speed, rate of...

  8. Effect of cattle grazing, seeded grass, and an herbicide on ponderosa pine seedling survival and growth

    Treesearch

    Philip M. McDonald; Gary O. Fiddler

    1999-01-01

    On a site of above-average quality in northern California, an early shrub-forb-grass plant community was treated by artificially seeding two forage grass species at plantation age 3, cattle grazing with and without seeded grasses, and applying a soil-active chemical (Velpar). Planted ponderosa pines were part of this community. Results for a 10-year period (1988-1997)...

  9. Growth Response of Pinus ponderosa following a Mixed-Severity Wildfire in the Black Hills, South Dakota

    Treesearch

    Tara keyser; Fredrick Smith; Wayne Sheppard

    2010-01-01

    In late summer 2000 the Jasper Fire burned ~34,000 ha of ponderosa pine forest in the Black Hills of South Dakota. Although regarded as a catastrophic event, the Jasper Fire left a mosaic of fire severity across the landscape, with live trees present in areas burned under low and moderate fire severity. In October 2005, we cored 96 trees from unburned, low-severity,...

  10. Stand and fuel treatments for restoring old-growth ponderosa pine forests in the interior west (Boise Basin Experimental Forest)

    Treesearch

    Russell T. Graham; Theresa B. Jain

    2007-01-01

    Fire exclusion, especially in the dry forests (i.e. those dominated or potentially dominated by ponderosa pine) has most often altered tree and shrub composition and structure and, though often overlooked in many locales, the forest floor from conditions that occurred historically (pre-1900).

  11. Tree regeneration spatial patterns in ponderosa pine forests following stand-replacing fire: Influence of topography and neighbors

    Treesearch

    Justin P. Ziegler; Chad M. Hoffman; Paula J. Fornwalt; Carolyn H. Sieg; Michael A. Battaglia; Marin E. Chambers; Jose M. Iniguez

    2017-01-01

    Shifting fire regimes alter forest structure assembly in ponderosa pine forests and may produce structural heterogeneity following stand-replacing fire due, in part, to fine-scale variability in growing environments. We mapped tree regeneration in eighteen plots 11 to 15 years after stand-replacing fire in Colorado and South Dakota, USA. We used point pattern analyses...

  12. A survey of effects of intentional burning on fuels and timber stands of ponderosa pine in Arizona

    Treesearch

    A. W. Lindenmuth

    1960-01-01

    Limited intentional burning of ponderosa pine timberlands to achieve net benefits has been going on for years. Many statements of results have been published. Most statements have been based on small plot records or subjective observations. Hence, the statements have been questioned by some readers and have led to debate over the merits of intentionally burning...

  13. Fuel loads and simulated fire behavior in "old-stage" beetle-infested ponderosa pine of the Colorado Plateau

    Treesearch

    E. Matthew Hansen; Morris C. Johnson; Barbara J. Bentz; James C. Vandygriff; A. Steven Munson

    2015-01-01

    Recent bark beetle outbreaks in western North America have led to concerns regarding changes in fuel profiles and associated changes in fire behavior. Data are lacking for a range of infestation severities and time since outbreak, especially for relatively arid cover types. We surveyed fuel loads and simulated fire behavior for ponderosa pine stands of the...

  14. A modified tree classification for use in growth studies and timber marking in Black Hills ponderosa pine

    Treesearch

    E. M. Hornibrook

    1939-01-01

    A satisfactory silvicultural management of ponderosa pine stands requires a judicious selection of trees to be left in the reserve stand. The timber marker must know what type of tree has the greatest growth potentialities and what type of tree will respond but slightly upon being released. The silvicultural problem in marking therefore is one of recognizing the...

  15. Two-dimensional heat flow analysis applied to heat sterilization of ponderosa pine and Douglas-fir square timbers

    Treesearch

    William T. Simpson

    2004-01-01

    Equations for a two-dimensional finite difference heat flow analysis were developed and applied to ponderosa pine and Douglas-fir square timbers to calculate the time required to heat the center of the squares to target temperature. The squares were solid piled, which made their surfaces inaccessible to the heating air, and thus surface temperatures failed to attain...

  16. Recovery of ponderosa pine ecosystem carbon and water fluxes from thinning and stand-replacing fire.

    PubMed

    Dore, Sabina; Montes-Helu, Mario; Hart, Stephen C; Hungate, Bruce A; Koch, George W; Moon, John B; Finkral, Alex J; Kolb, Thomas E

    2012-10-01

    Carbon uptake by forests is a major sink in the global carbon cycle, helping buffer the rising concentration of CO 2 in the atmosphere, yet the potential for future carbon uptake by forests is uncertain. Climate warming and drought can reduce forest carbon uptake by reducing photosynthesis, increasing respiration, and by increasing the frequency and intensity of wildfires, leading to large releases of stored carbon. Five years of eddy covariance measurements in a ponderosa pine (Pinus ponderosa)-dominated ecosystem in northern Arizona showed that an intense wildfire that converted forest into sparse grassland shifted site carbon balance from sink to source for at least 15 years after burning. In contrast, recovery of carbon sink strength after thinning, a management practice used to reduce the likelihood of intense wildfires, was rapid. Comparisons between an undisturbed-control site and an experimentally thinned site showed that thinning reduced carbon sink strength only for the first two posttreatment years. In the third and fourth posttreatment years, annual carbon sink strength of the thinned site was higher than the undisturbed site because thinning reduced aridity and drought limitation to carbon uptake. As a result, annual maximum gross primary production occurred when temperature was 3 °C higher at the thinned site compared with the undisturbed site. The severe fire consistently reduced annual evapotranspiration (range of 12-30%), whereas effects of thinning were smaller and transient, and could not be detected in the fourth year after thinning. Our results show large and persistent effects of intense fire and minor and short-lived effects of thinning on southwestern ponderosa pine ecosystem carbon and water exchanges. © 2012 Blackwell Publishing Ltd.

  17. Landscape dynamics of mountain pine beetles

    Treesearch

    John E. Lundquist; Robin M. Reich

    2014-01-01

    The magnitude and urgency of current mountain pine beetle outbreaks in the western United States and Canada have resulted in numerous studies of the dynamics and impacts of these insects in forested ecosystems. This paper reviews some of the aspects of the spatial dynamics and landscape ecology of this bark beetle. Landscape heterogeneity influences dispersal patterns...

  18. Monoterpene emissions from a Ponderosa Pine forest. Does age matter?

    NASA Astrophysics Data System (ADS)

    Madronich, M. B.; Guenther, A. B.; Wessman, C. A.

    2011-12-01

    Determining the emissions rate of biogenic volatile organic carbon (BVOC) from plants is a challenge. Biological variability makes it difficult to assess accurately those emissions rates. It is known that photosynthetic active radiation (PAR), temperature, nutrients as well as the biology of the plant affect emissions. However, less is known about the variability of the emissions with respect to the life cycle of the plants. This study is focusing on the difference of monoterpene emission rates from mature Ponderosa Pine trees and saplings in the field. Preliminary calculations show that there is a significant difference between total monoterpene emissions in mature trees (0.24±0.04 μgC/gdwh) and saplings (0.37±0.02 μgC/gdwh).

  19. Interactive effects of CO2 and O3 on a ponderosa pine plant/litter/soil mesocosm.

    PubMed

    Olszyk, D M; Johnson, M G; Phillips, D L; Seidler, R J; Tingey, D T; Watrud, L S

    2001-01-01

    To study individual and combined impacts of two important atmospheric trace gases, CO2 and O3, on C and N cycling in forest ecosystems; a multi-year experiment using a small-scale ponderosa pine (Pinus ponderosa Laws.) seedling/soil/litter system was initiated in April 1998. The experiment was conducted in outdoor, sun-lit chambers where aboveground and belowground ecological processes could be studied in detail. This paper describes the approach and methodology used, and presents preliminary data for the first two growing seasons. CO2 treatments were ambient and elevated (ambient + 280 ppm). O3 treatments were elevated (hourly averages to 159 ppb, cumulative exposure > 60 ppb O3, SUM 06 approximately 10.37 ppm h), and a low control level (nearly all hourly averages <40 ppb. SUM 06 approximately 0.07 ppm h). Significant (P < 0.05) individual and interactive effects occurred with elevated CO2 and elevated O3. Elevated CO2 increased needle-level net photosynthetic rates over both seasons. Following the first season, the highest photosynthetic rates were for trees which had previously received elevated O3 in addition to elevated CO2. Elevated CO2 increased seedling stem diameters, with the greatest increase at low O3. Elevated CO2 decreased current year needle % N in the summer. For 1-year-old needles measured in the fall there was a decrease in % N with elevated CO2 at low O3, but an increase in % N with elevated CO2 at elevated O3. Nitrogen fixation (measured by acetylene reduction) was low in ponderosa pine litter and there were no significant CO2 or O3 effects. Neither elevated CO2 nor elevated O3 affected standing root biomass or root length density. Elevated O3 decreased the % N in coarse-fine (1-2 mm diameter) but not in fine (< 1 mm diameter) roots. Both elevated CO2 and elevated O3 tended to increase the number of fungal colony forming units (CFUs) in the AC soil horizon, and elevated O3 tended to decrease bacterial CFUs in the C soil horizon. Thus, after two

  20. Dry deposition of nitrogen and sulfur to ponderosa and jeffrey pine in the San Bernardino National Forest in southern California

    Treesearch

    Mark E. Fenn; Andrzej Bytnerowicz

    1993-01-01

    Little is known about the concentrations, deposition rates, and effects of nitrogenous and sulfurous compounds in photochemical smog in the San Bernardino National Forest (SBNF) in southern Calfirnia. Dry deposition of NO3- and NH4+ to foliage of ponderosa pine (Pinus...