xv aw .dwt, mwwhfis \. . XI... 1‘ TU. : . w n . .r 3;. akvhuza I. «I 1... arm Km... 1!. n b- x 1 .. V I. n7)?1rna.“t . if .5 .flflfii l . : hug). 04”.;de“3 «5.9.5 1? hug. - .53 .5. .553 . .3 .. .. 1.. a . 3. 3 h .35 :2. 3 at: .5: ‘ s. 9-.- It LIBRARY Michigan State University This is to certify that the dissertation entitled DISTRIBUTION AND ECOPHYSIOLOGY OF THE PONDEROSAE IN THE SANTA CATALINAMOUNTAINS OF SOUTHERN ARIZONA presented by JASON SCOTT KILGORE has been accepted towards fulfillment of the requirements for the PhD. degree in Plant Biology and Ecology, Evolution, and Behavioral 4Biology Ari/Cw Major Prof/erssor’s Signature 3/, 2 50)" Date MSU is an affirmative-action, equal-opportunity employer PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/07 p:lClRC/DateDue.indd—p.1 DISTRIBUTION AND ECOPHYSIOLOGY OF THE PONDEROSAE IN THE SANTA CATALINA MOUNTAINS OF SOUTHERN ARIZONA By Jason Scott Kilgore A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Plant Biology Ecology. Evolution, and Behavioral Biology Program 2007 ABSTRACT DISTRIBUTION AND ECOPHYSIOLOGY OF THE PONDEROSAE IN THE SANTA CATALINA MOUNTAINS OF SOUTHERN ARIZONA By Jason Scott Kilgore I investigated the distribution and ecophysiological factors for establishment and survival for Ponderosae in the Santa Catalina Mountains of southern Arizona. The complex topography of this system produces variable microhabitats that obscure a simple elevational pattern of Arizona pine (Pinus arizonica Engelmann) lower than the Southwestern race of Rocky Mountain ponderosa pine (P. ponderosa var. scopulorum Engelmann). In addition, an intermediate needle morphotype, a putative hybrid between P. arizonica and P. ponderosa var. scopulorum, is recognized from this and other mountain islands in the American Southwest. Following extensive fires, Coronado National Forest identified the need to assess their distributional patterns for reforestation. Modeling of documented occurrences in relation to climatic variables indicated clear distinction in distribution and ecological niche space between P. arizonica and P. ponderosa var. scopulorum, while the putative hybrid displayed intermediate distribution and occupation of niche space. High intratree variance in needle number and the overlapping distribution and ecological niche space support the hybridization hypothesis. In controlled tests, germination success and time to initial germination were similar, but P. arizonica seeds germinated over a longer period of time. In the field, neither species germinated in the spring, while, following the monsoon, P. an‘zonica (25%) had lower germination at low and high elevation than did P. ponderosa var. scopulorum (52%). Controlled whole-plant freezing tests revealed greater cold tolerance in needle, bud, and cambial tissue by P. ponderosa var. scopulorum seedlings than P. arizonica seedlings over periods of simulated fall, winter, and spring. Seedlings of P. ponderosa var. scopulorum exposed to ambient light and freezing temperatures maintained higher photosynthetic quantum yield than did seedlings of P. an‘zonica and the putative hybrid. In the field, small P. ponderosa var. scopulorum trees maintained higher net photosynthesis and midday quantum yield than did similarly sized P. arizonica trees. When soil water availability was low and incident light levels were high during the winter, small P. ponderosa var. scopulorum trees maintained higher leaf water status yet similar net photosynthesis than P. an‘zonica trees. During the arid foresummer, diurnal variation in gas exchange was higher in P. ponderosa var. scopulorum than in P. arizonica, suggesting greater stomatal response to water-limiting conditions. However, integrated water-use efficiency and leaf nitrogen content did not differ between species. During severe dehydration, P. ponderosa var. scopulorum maintained higher photosynthetic quantum yield than did P. arizonica. However, xylem conductivity and vulnerability to cavitation did not differ between species. P. ponderosa var. scopulorum has greater cold tolerance, controls transpirational water loss, and maintains photosynthetic function across the seasons better than P. arizonica. Copyright by JASON SCOTT KILGORE 2007 ACKNOWLEDGEMENTS The research described in this dissertation could not have been accomplished without the support of my family and many friends. I especially appreciate the enduring support and love from my wife, Pam, and son, Oliver. My extended family, including Lucy Dueck, Justin Kilgore, Bob and Kathy Reynolds, Kim Reynolds, and Jason Tallant, provided unquestioning support, trust, and curiosity in what I was doing in Arizona or at the lab through the night. I thank my committee - Drs. Frank Telewski, Peter Murphy, Bryan Epperson, and Brian Maurer - for serving my research interests. Frank was not only my major advisor and mentor but also has become a good friend; I truly appreciate his companionship and contributions to field and labwork. Pete has been my prudent mentor and friend for over a decade. And I eagerly anticipate exploring the genetic complexities of the Ponderosae with Bryan in the future. I have been enormously fortunate to have many people volunteer their time and resources to me through this dissertation research. Friends who trudged the Santa Catalina Mountains with or for me have included Frank Telewski, Kris Kern, Stephanie Blumer, Gretchen Friedlander, Samantha Warwick, Clare and Bob Peterson, Jill Grimes, Tom and Annita Harlan, Holly Bechetti, and Joe and Christina Hoscheidt. Folks from the Coronado National Forest, especially Bill Hart but also Terri Austin, Jennifer Ruyle, and Josh Taiz, have been incredibly supportive and interested in this research; these are the managers of the forest who need this research. Barb Eisele (Trees for Mount Lemmon) was instrumental in connecting this research with reforestation interests for the Summerhaven community as well as in contributing seedlings to our research program; she was always interested in trading home-cooked meals for research updates. Jim Burns and Steve Leavitt at the Laboratory of Tree- Ring Research provided shipping destinations and nitrogen cylinders, while Carol Mack at Mount Lemmon General Store kindly arranged shipping departures. Bob Peterson and the gang at the Steward Observatory always made sure they had room for us at the top. Loma and Jim Griffith always had a place for us in Tucson and provided transportation, including a 4WD vehicle, whenever needed. Back in Michigan, I had ample assistance and support from friends. Folks who contributed to lab and greenhouse adventures include Alex Lindsey, Stephanie Blumer, Frank Telewski, Samantha Warwick, Jessica Reif, Tricia Mitchell, Casey Bartrem, Calvin Glaspie, Jameel AI-Haddad, Justin Savu, Pam Kilgore, Crystal Wallace, Will Paddock, Abbie Schrotenboer, Lynn Sage, Todd Robinson, Andrea Corpolongo, Scott Smith, Hope Rankin, Christopher Start, and Andrew Murphy. Anna Jacobsen is a valued friend, colleague, and contributor to this research. Although not serving on my committee, Drs. Lissa Leege, Ken Poff, Bert Cregg, Frank Ewers, Jim Flore, David Rothstein, Carolyn Malmstrom, Merritt Turetsky, Andrew McAdam, and Andy Jarosz provided invaluable advice, equipment, and/or office space. Jeff Wilson, Bob Goodwin, and Nate Siegert assisted me with computer and GIS issues. Lynn Sage, Mark Hammond, and Jim Klug arranged for cold conditions, while Dave Freville’s crew ensured warm (greenhouse) conditions. Mike Grillo and Heather Hallen were invaluable friends vi and colleagues. No person acts without their departmental staff; I appreciate the support and assistance by Tracey Barner, Kasey Baldwin, Jan McGowan, Jill Richey, and Holly Nieusma. Last, | wish to acknowledge those units that have financially contributed to this research. The Department of Plant Biology supported a considerable amount of travel, conference, and non-MSU course-related expenses through the Paul Taylor Fund, as well as supporting me through the experience of teaching assistantships. The College of Natural Science granted me the Dr. Marvin Hensley Endowed Fellowship and the Dissertation Completion Fellowship, while the Ecology, Evolution, and Behavioral Biology (EEBB) program granted me Travel and Summer Fellowships. Campus Planning and Administration has provided logistical support, while Frank has supplemented funding with his research account. I appreciate all that these units have contributed to my research. vii TABLE OF CONTENTS CHAPTER1 AN INTRODUCTION TO SOUTHWESTERN PONDEROSA PINE, ARIZONA PINE, AND THEIR ENVIRONMENT CHAPTER 2 DISTRIBUTION AND ECOLOGICAL NICHE DIFFERENTIATION OF SYMPATRIC PONDEROSAE TAXA IN THE SANTA CATALINA MOUNTAINS OF ARIZONA CHAPTER 3 INFLUENCE OF SEED GERMINATION ECOLOGY ON THE DISTRIBUTION OF PONDEROSAE IN THE SANTA CATALINA MOUNTAINS OF ARIZONA CHAPTER 4 DIFFERENTIAL COLD TOLERANCE SEGREGATES DISTRIBUTION OF PONDEROSAE IN THE SANTA CATALINA MOUNTAINS OF CHAPTER 5 THE ROLE OF DROUGHT TOLERANCE IN STRUCTURING PONDEROSAE DISTRIBUTION IN THE SANTA CATALINA MOUNTAINS OF ARIZONA CHAPTER 6 GENERAL CONCLUSIONS viii 23 82 116 160 ...220 ...226 244 LIST OF TABLES Table 2-1. Models and environmental variables used in the modeling of species distributions, or suitable habitat (Phillips et al. 2006a), using Maxent (Phillips et al. 2006b) Table 2-2. Summary statistics for regression of mean number of needles per fascicle per tree by morphotype and combined species against elevation at which the tree was located Table 2-3. Comparison of influence of spatial extent by shape of calibration area on model output Table 2-4. Performance characteristics and statistics for prediction models generated by Maxent for the rectangular study region... .. Table 2-5. Climatically suitable area estimates for the three morphotypes based on cumulative probabilities from the Lit and Lit.Thresh models... Table 2-6. Thresholds for the three morphotypes based on the Lit. Thresh model... Table 2-7. Ecological similarity matrix based on the ability of one taxon’s model (columns) to predict the documented occurrences of each of the taxa (rows) Table 3-1. Seeds inventoried from cones collected from the Santa Catalina Mountains in September-October 2005 Table 3-2. Type II tests of 2-factor ANOVA of percent germination (pergerm), time to first germination (initiate), and time to last realized germination (complete) for PIN_PON seed as a function of site (MTL and PAL), stratification (strat; 0-, 15-, and 30—day), and their interaction ...... Table 3-3. Type II tests of 2-factor ANOVA of percent germination (pergerm), time to first germination (initiate), and time to last realized germination (complete) for P|N_ARI seed as a function of site (PAL and RC/LIZ), stratification (strat; 0-, 15-, and 30-day), and their interaction... .. Table 3-4. Type II tests of 2-factor ANOVA of percent germination (pergerm), time to first germination (initiate), and time to last realized germination (complete) for the PAL site as a function of morphotype (PIN_ PON, Mixed, and PIN _ARI)W stratification ”(strat;. 0-, 15-, and 30-day), and their interaction. ...43 .51 ...55 .58 .62 ...64 .64 .88 ..... 99 100 ...102 Table 3-5. Type II tests of 2-factor ANOVA of natural-log-transformed percent realized germination (pergerm), time to first germination (initiate), and time to last germination (complete) for pooled sites as a function of morphotype (PIN_ PON, Mixed, and PIN :AHRI), stratification (strat;0-,15- and 30—day), and their interaction” ..............103 Table 3-6. Pairwise comparisons (Tukey’s HSD) of PIN_PON and PIN_ARI seeds for natural-log-transformed percent realized germination (pergerm), time to first germination (initiate), and time to last germination (complete) as a function of stratification treatment for pooled sites... .103 Table 3-7. ANOVA for germination success of PIN _PON and PIN _ARI seeds at the Mount Lemmon and Willow Canyon sites (adjusted R2=0.1625)... 104 Table 3-8. Multiple comparisons between site (Mount Lemmon - MTL and Willow "Canyon- WC), morphotype and interactions "by TukeysHSD... .. .. ................105 Table 4-1. Seedlings and source trees from the Santa Catalina Mountains used in the controlled whole-seedling experiment........................... .....125 Table 4-2. Conditions for hardening, acclimating, and deacclimating PIN___PON and PIN_ARI seedlings for the whole-seedling freezing test... ....126 Table 4-3. Temperature profiles for each of the freezing tests outlined in Table 4-2129 Table 4-4. ANOVA results for dark-acclimated chlorophyll fluorescence (DACF), relative conductivity (RC), and cambial (Camb) and bud (Bud) mortality for PIN _PON and PIN _ARI seedlings exposed to freezing treatments (Tx) during the acclimation phase... .. .. .137 Table 4-5. Cold tolerance (LTso) 0f PIN_PON and PIN__ARI estimated as a function of needle dark-acclimated chlorophyll fluorescence (ndl - DACF), needle relative conductivity (ndl - RC), and cambium and bud mortality across three stages of cold acclimation: Hardening, Winter, and Deacclimation..............................................................................144 Table 4-6. Type II tests of ANCOVA for PSII quantum yield (LACF) fitted to morphotype and the covariate air temperature (Temp) for 3- year-old potted seedlings with frozen soil water........................................146 Table 4-7. ANCOVA results for predawn dark-acclimated chlorophyll fluorescence (pDACF), midday light-acclimated chlorophyll fluorescence (mLACF), predawn (ppsi) and midday (mpsi) xylem pressure potential, and midday net photosynthesis (mphoto) for PIN _PON and PIN_ARI trees at three sites (MTL, PAL, and LIZ) In the Santa Catalina MountaIns In January 2006.. . .. . . ... Table 4-8. Type II ANOVA results for predawn dark-acclimated chlorophyll fluorescence (pDACF), midday light-acclimated chlorophyll fluorescence (mLACF), predawn (ppsi) and midday (mpsi) xylem pressure potential, and midday net photosynthesis (mphoto) for PIN_ PON and PIN _ARI trees at three sites (MTL“. PAL, and Lil) In the Santa Catalina Mountains In January 2006... Table 4-9. Mean responses by PIN_PON and PIN_ARI trees by site (MTL, PAL, and LIZ) and across sites (Combined) following removal of three outlying data values Table 4-10. Cold tolerance (LT50) of combined PIN_PON and PIN_ARI tissues from this study and from Burr et al. (1990) across three stages of cold acclimation: Hardening, Winter, and Deacclimation...... Table 5-1. Characteristics of 'small’ trees used for gas exchange and xylem pressure potential measurements in 2005-06 at the Mount Lemmon (MTL), Palisade Rock (PAL), and Lizard Rock (LIZ) sites... Table 5-2. Comparison of seasonal xylem water potential (Wp) of combined (both), PIN_PON, and PIN_ARI fascicles harvested at predawn, midday, and evening across an elevational gradient... Table 5-3. Nested ANOVA results xylem pressure potential (W,,) in needles from PIN_PON and PIN_ARI trees across an elevation gradient, 3 seasons (“‘Arid" is arid foresummer), and 2-3 daytime periods... Table 5-4 Univariate ANOVA results for carbon isotope (6130), nitrogen isotope (615N), carbon (%C), nitrogen (%N), and carbon-to- n-itrogen (C: N) content in needles for each year from PIN _PON and PIN _ARI trees at PAL. Table 5- 5. Nested ANOVA results for carbon isotope (613C), nitrogen isotope (615N), carbon (%C), nitrogen (%N), and carbon-to- -nitrogen (C: N) content' In needles from PIN_PON and PIN_ARI trees at PAL. xi ...148 ...150 152 153 166 190 .194 ...195 ...198 Table 5-6. ANCOVA results for fresh mass loss (pmass) and PSII excitation capture efficiency (Fv’lFm’) of dehydrating needles for PIN_PON and PIN_ARI stems from four sites in the Santa Catalina Mountains: MTL, PAL, RCL, and LI2203 Table 5-7. ANOVA results for xylem pressure potential at 50% loss of hydraulic conductivity (”’50) and xylem specific hydraulic conductivity (Ks) for PIN_PON and PIN_ARI stems collected from three sites in the Santa Catalina Mountains: MTL, PAL, and LlZ/RCL... ..207 Table A-1. Geographic coordinates and mean needle number for trees from the distribution study ("SCAT")... ..............226 Table A-2. Correlation matrix for mean (1971-2000; PRISM 2006) annual and monthly precipitation (Precip, top), minimum temperature (MinTemp, bottom), and maximum temperature (MaxTemp, next page) for the Santa Catalina Mountain study region...........................................241 xii LIST OF FIGURES Figure 1- 1. Vegetation of the Santa Catalina Mountains as drawn by Whittaker and Niering .....(1965) from 400 vegetatIon samples In a gradient analysis... Figure 2—1. General distribution of ponderosa pine (top, Pinus ponderosa, Thompson et al. 1999), and location of study region in the Santa Catalina Mountains in southeastern Arizona (bottom, Figure 2-2. Distribution of sampling points collected during the distribution study (“SCAT”) and other projects ("Non- SCAT").” in the study region of the Santa Catalina Mountains... Figure 2-3. Frequency of trees (n=671) by mean needle number per fascicle Figure 2-4. Variability in needle number within a tree as a function of its mean needle number Figure 2- 5. Mean annual needle number (solid, :80) and sample size (hollow) for the PIN _PON ”(A)”. Mixed ”(O)” and PIN _ARI ...(D). morphotypes (Peloquin 1984)... Figure 2-6. Distribution of occurrences for the three morphotypes in the region using the distribution study (“SCAT”) data Figure 2-7. Mean number of needles per fascicle as a function of elevation foreach morphotype Figure 2-8. Distribution of the three morphotypes in climatic space with regard to mean monthly minimum and maximum January temperatures... Figure 2-9. Distribution of the three morphotypes in climatic space with regard to mean monthly maximum June temperature and monthly July preCIpItatIon Figure 2-10. Distribution of cumulative probability of suitable habitat for PIN_PON relative to maximum June temperature and annual precipitation in two spatial extents (“polygon” and “rectangle") for Maxent model caIIbratIon xiii .....18 32 ...35 47 .48 ..48 .50 .51 .52 .54 ...56 Figure 2-11. Distribution of cumulative probability of suitable habitat for the three morphotypes in the region for the Lit model by Maxent... Figure 2- 12. Distribution of cumulative probability of suitable habitat for the three morphotypes In the region for the Lit. Threshold model by Maxent... Figure 2-13. Distribution of cumulative probability of suitable habitat for the three morphotypes in the region for the NED.Asp model by Maxent... Figure 3-1. Prepared seedbed within the predator-exclusion cage at the Willow Canyon SIte Figure 3-2. Predator-exclusion cage after placement over sown seeds at the Willow Canyon site (January 2006) Figure 3-3. Mean seed germination (top), time to first germination (middle), and elapsed time to the last seed germination (bottom) for the full data set . . . Figure 3-4. Germination as a function of stratification treatment for seeds from PIN_PON trees at MTL and PAL Figure 3-5. Germination as a function of stratification treatment for seeds from PIN_ARI trees at PAL and RC/LIZ Figure 3-6. Germination as a function of stratification treatment for combined morphotypes at the PAL Slte Figure 3-7. Germination of PIN_PON and PIN_ARI seeds planted at high (Mount Lemmon) and low (Willow Canyon) elevation... Figure 3-8. A proposed path of selection for conditional (secondary) dormancy of seeds produced by Arizona and Southwestern ponderosa pine in the mountain islands of the Southwest experiencing bimodal precipitation patterns . .. .. Figure 4-1. Dark-acclimated chlorophyll fluorescence (left) and relative conductivity (right) for PIN_PON and PIN_ARI after different freezing treatments across three stages of cold acclimation: Hardening (top), Acclimation (middle), and Deacclimation (bottom).................................... Figure 4-2. Cambial and bud mortality for PIN_PON and PIN_ARI after different freezing treatments across three stages of cold acclimation: Hardening (top). Winter (middle), and Deacclimation (bottom). .. xiv .61 ...63 ...70 .94 .94 ...97 .98 100 101 104 ...113 .140 141 Figure 4-3. Linear regression of dark-acclimated chlorophyll fluorescence, relative conductivity, and cambial and bud mortality against treatment temperature for PIN_PON (solid line) and PIN_ARI (dashed line) across three stages of cold acclimation: Hardening (thin line), Winter (normal line), and Deacclimation (thick line). .. ... ... ... ... ... .... Figure 4-4. Slope (thin line) and intercept (thick line) for regressions of dark-acclimated chlorophyll fluorescence, relative conductivity, and cambial and bud mortality against treatment temperature for PIN_PON (solid line) and PIN_ARI (dashed line) across weeks of cold acclimation... .. Figure 4-5. PSII quantum yield (LACF) as a function of air temperature when measured from 3-year—old potted seedlings with frozen soil water: PIN_PON (solid line), Mixed (dotted line), and PIN_ARI (dashed line). .. Figure 4-7. Variable chlorophyll fluorescence (Fv/Fm, solid; Fv’lFm’, empty) as a function of leaf temperature for PIN_PON and PIN_ARI trees combined across sites in January 2006 Figure 4- 8. Xylem pressure potential (919) from excised needles of PIN _PON and PIN _ARI trees at three sites (MTL” PAL, and LIZ) In the Santa Catalina Mountains In January 2006... Figure 4- 9. Midday net photosynthesis (A) results for PIN _PON and PIN _ARI trees at three sites (MTL, PAL, and HLIZ)’ In the Santa Catalina Mountains In January 2006... Figure 5-1. Diurnal variation in net photosynthetic rate (A) and transpiration (Tr) as a function of incident radiation (PAR) and temperature (7.03;) for paired PIN_ PON and PIN _ARI trees at PAL during winter (January 2006)... Figure 5-2. Diurnal variation in net photosynthetic rate (A) and transpiration (Tr) as a function of incident radiation (PAR) and temperature (T leaf) for paired PIN_PON and PIN_ARI trees at PAL during the arid foresummer (June 2005) Figure 5-3. Diurnal variation in net photosynthetic rate (A) and transpiration (Tr) as a function of incident radiation (PAR) and temperature (That) for paired PIN_PON and PIN_ARI trees at PAL during the monsoon season (August 2005) Figure 5-4. Diurnal by seasonal needle xylem pressure potential (WP) for PIN_ PON and PIN _ARI trees growing in the Santa Catalina Mountains... XV 142 144 146 .147 ...151 ...151 ...181 183 .186 ...190 Figure 5-5. Variation (standard deviation, SD) in diurnal needle xylem pressure potential (Wp) by season for PIN_PON and PIN_ARI trees growing in the Santa Catalina MountaIns Figure 5-6. Needle xylem pressure potential (91p) by winter (top,left), arid foresummer (bottom), and monsoon (top,n'ght) for PIN_PON and PIN_ARI trees growing along an elevational gradient in the Santa Catalina MountaIns Figure 5- 7. Annual variation in mean carbon (513C, solid lines) and nitrogen (615M, dashed lines) isotope composition for leaf tissue of PIN_PON and PIN_ARI trees at PAL... Figure 5-8. Annual variation in mean carbon (solid lines) and nitrogen (dashed lines) content for leaf tissue of PIN_PON and PIN_ARI trees atPAL Figure 5- 9. Carbon isotopic composition (613C) relative to carbon content in needles from PIN_ PON ”(solid line)“ and PIN _ARI m(dashed Iine)trees atPAL... Figure 5- 10. Carbon isotopic composition (613C) relative to nitrogen content In needles from PIN_ PON (solid line)”. and PIN _ARI “(dashed Iine)trees atPAL... Figure 5-11. Nitrogen isotopic composition (6‘5N) relative to nitrogen content in needles from PIN_ PON (solid line)”. and PIN _ARI ”(dashed Iine)trees atPAl.... . Figure 5-12. Nitrogen isotopic composition (6‘5N) relative to carbon- to-nitrogen content (ON) in needles from PIN_PON (solid line) and PIN_ARI (dashed line) trees at PAL Figure 5-13. Dehydration of fascicles for PIN_PON (solid lines) and PIN_ARI (dashed lines) stems from four sites in the Santa Catalina Mountains: MTL, PAL, RCL, and LIZ Figure 5-14. Mean dehydration of fascicles for PIN_PON (solid lines) and PIN_ARI (dashed lines) stems across four sites in the Santa Catalina MountaIns Figure 5-15. Photosynthetic efficiency as a function of needle dehydration for PIN_PON (solid lines) and PIN_ARI (dashed lines) stems from four sites in the Santa Catalina Mountains: MTL, PAL, RCL, and LIZ....... .192 193 ...195 ...197 ...199 ...199 ...200 .200 .202 .203 .. .205 Figure 5- 16. Photosynthetic efficiency as a function of needle dehydration for PIN _PON (solid lines) and PIN _ARI (dashed lines) stems across four sites in the Santa Catalina Mountains... Figure 5-17. Xylem vulnerability curves for PIN_PON (solid lines) and PIN_ARI (dashed lines) stems collected from three sites in the Santa Catalina Mountains: MTL, PAL, and LlZ/RCL Figure 5-18. Xylem specific hydraulic conductivity (K...) for PIN_PON and PIN_ARI stems collected from three sites in the Santa Catalina Mountains: MTL, PAL, and LIZ/RCL Figure 5- 19. Stomatal openings in PIN _PON “(left)”. and PIN _ARI "”(right) bySEM Figure 5-20. Chloroforrn-cleared stomatal openings in PIN _ARI of top and transverse views by SEM... ...205 207 208 ...211 ...212 CHAPTER 1 AN INTRODUCTION TO SOUTHWESTERN PONDEROSA PINE, ARIZONA PINE, AND THEIR ENVIRONMENT From the time of Theophrastus (1916) some 2300 years ago, humans have been fascinated with the distribution and variation of species. Vast human and land resources have been invested in the quest to understand the biogeography of life. This information is invaluable when considering future distributions of species and their responses to global climate change patterns and other anthropogenic influences shaping the biogeography of species. A number of important questions can be posed in this regard. What restricts geographic ranges, and why does a group of organisms vary within its range and yet may be reproductively compatible across the range? How has the heterogeneous environment selected for modified and plastic features? At what point are organisms from two morphologically different populations considered different species? What effect does interspecies gene flow have on the systematics and evolution of the clade? Overall, how have the synergy of geologic processes, climate, and time interacted with ecological and evolutionary processes to produce the diversity of life around us? Finally, how have human interactions affected, or will affect, biodiversity? The objective of this dissertation is to combine information from geographic range and ecophysiology to understand the distribution of two closely related conifers - Rocky Mountain ponderosa pine (P. ponderosa var. scopulorum Engelmann) and Arizona pine (P. arizonica Engelmann; formerly P. ponderosa var. arizonica (Engelmann) Shaw) — in the Santa Catalina Mountains of southern Arizona, USA. Within the isolated mountain islands of the American Southwest, the systematics of the ponderosa pines remains unresolved (Rehfeldt 1993, 1999). Shared ancestry, dynamic climate and topography, strong selection gradients, and hybridization have contributed to the difficulties in classifying populations into species. Focusing on one geographic system - the Santa Catalina Mountains — can elucidate processes occurring across other mountain islands and species. Furthermore, following the devastating Bullock (2002) and Aspen (2003) fires, the Coronado National Forest identified the need for distribution information for improved management of the forest’s genetic diversity. This introduction serves to set the stage for the subsequent chapters on the distribution, ecophysiology, and management implications of Arizona and ponderosa pine in the Santa Catalina Mountains. Following a summary of the putative origin and current status of the systematics of these species, I will describe the geographic system in which my research was conducted. The distribution and ecological modeling of the Ponderosae in the Santa Catalina Mountains will be given in Chapter 2, while the ecophysiological studies that shed light on the causes of their distributions will be addressed in Chapters 3-5. Each of these chapters is intended as an individual paper without reference to the other chapters or this introduction. Last, Chapter 6 will summarize the findings and implications of this dissertation. Origins of Ponderosa Pine The model species for this research, Arizona and ponderosa pine, are members of the Genus Pinus, Subgenus Pinus (Diploxylon or Hard Pines), Section Pinus, and Subsection Ponderosae. Fossil evidence suggests that the common ancestor for this genus emerged in the mid-latitudes of then-contiguous northeastern United States and western Europe during the late Jurassic (Axelrod 1986, Millar 1993). At the time, the unified Laurasia was relatively homogeneous in topography and climate. Tectonic activity was low, sea levels were high, average temperatures were 10-20° C warmer than present in the middle and high latitudes, and seasonality and latitudinal gradients in the Northern Hemisphere were reduced compared to the present (Millar 1993). Subsequently, the newly emerged Pinus began spreading east to west (Millar 1993) toward the Western Interior seaway (Baldridge 2004). By the end of the Cretaceous, Laurasia began to split, the Western Interior seaway began to recede (Baldridge 2004), the Laramide orogeny (tectonism) began creating new topography (e.g., elevation, dry slopes, rain shadows), and the mid-latitudes began to dry leaving the low and high latitudes more humid (Axelrod 1986, Millar 1993). This new geologic and climatic variability created the opportunity for adaptation to new environments (Axelrod and Raven 1985), and, facilitated by symbiotic associations with ectotrophic mycorrhizae, Pinus began speciating into the major subsections of the genus by the late Cretaceous (Axelrod 1983). Along with 5 other subsections in Pinus, Pondemsae likely originated in the Cordilleran region of Colorado south into Mexico (Axelrod 1983). Following on the heels of the western migration and subsectional speciation of Pinus were the newly emerging boreotropical angiosperms in the early Tertiary (Millar 1993). Increasingly warm and humid conditions in the mid- latitudes during the Paleocene favored the more competitive and faster growing angiosperms at the expense of pines (Millar 1993). Consequently, by the early Eocene, widespread extirpation of pines in the lower elevations of the mid- latitudes as well as the first major fragmentation of remaining pines into three refugia are observed in the fossil record (Millar 1993). Some pines most closely allied to Ponderosae retreated north to a circumpolar high—latitude zone (65-80° N) where sufficient photosynthesis was still possible due to carbon dioxide levels higher than at present (Millar 1993). Other Ponderosae lineages found refugium at high elevation within the mid-latitudes (e.g., Colorado), while other lineages sought the warmer and drier Eocene climate of low latitudes (e.g., Mexico and Central America) (Millar 1993). The fragmentation of subsection Ponderosae during the Eocene is thought to be the most significant event leading to the speciation of Pinus ponderosa in the north (Axelrod and Raven 1985), with fossil evidence near British Columbia (Axelrod 1986), and other distinct species (possibly P. arizonica) to the south, as well as generating new centers of pine diversity (Millar 1993). At the close of the Eocene, the evolutionary environment changed significantly. Average annual temperatures dropped 10-13° C with decreased summer rainfall and increased seasonality, while mean annual range of temperatures increased from 3-5° to 25° C (Millar 1993). Renewed tectonic activity produced extensive volcanism and uplift producing the Rocky Mountains and Mexican Sierra Madre Occidental and Oriental (Millar 1993, Baldridge 2004). The complex climatic and geologic changes resulted in widespread extirpation of the boreotropical angiosperms and concomitant expansion of pines back into the mid-latitudes in some places in a few million years by the early Oligocene (Millar 1993). As the climate warmed and summer aridity increased through the Oligocene and into the Miocene (Axelrod and Raven 1985), pines expanded their range and abundance across latitudes, including mountainous low latitudes (Millar 1993). The former refugia served as centers of pine diversity, radiating outward to colonize the diverse environment. With at least a dozen closely related extant species of Ponderosae endemic to the refugia in Mexico and Central America, the events of the Eocene can still be seen through active radiation and speciation within Ponderosae and other subsections of Pinus (Millar 1993). The impact of the Eocene in forcing large latitudinal and environmental shifts in distribution, and thereby dissecting the genus into subsections, significantly contributed to the diversity in Pinus observed today (Millar 1993). Climatic fluctuations following the Oligocene to the present were significantly less than those during the close of the Eocene. For example, the amplitude of average temperatures during the 2.4-million-year Pleistocene (5-10° C) was about the same as climatic cycles during the 20-million-year Eocene (Millar 1993). Pleistocene glacial events caused significant shifts in pine distributions generally from north to south or from high to low elevation. The alternating distribution expansions and contractions (“pulsing”) allowed gene flow between formerly disjunct taxa followed by isolation over dozens of warming and cooling cycles. Through local adaptation, hybridization, and isolation, the genetic structure of Southwestern flora, including Pinus, has changed, or pulsed, over the last 20 million years through the present. However, a discontinuity exists in the Quaternary fossil record for the presence of P. ponderosa in the Southwestern isolated mountain systems, which were generated long before by Eocene-Miocene crustal extension and tectonic uplift (Baldridge 2004). Packrat midden data from van Devender et al. (1984) extending back 40 ka did not show the presence of P. ponderosa in New Mexico’s mountains until the last glacial maximum (18 ka), while no such evidence appears in the Great Basin, Colorado Plateau, and current Southwest desert regions for the same period (Thompson and Anderson 2000). Yet pollen from the late Wisconsin (11.4 ka) has been detected in the Sonoran Desert (Martin and Mehringer 1965), perhaps from distant forests (Betancourt et al. 1990). Furthermore, packrat midden data from montane forests in Arizona (van Devender 1990) support the current belief that P. ponderosa was present at the end of the Wisconsin (14 ka) but did not extend into the Sonoran Desert nor form continuous cover between mountain systems or the high Colorado Plateau to the north (Betancourt et al. 1990). During some pre-Wisconsin glacial, P. ponderosa must have migrated south from the Colorado Plateau and Rocky Mountains. Likewise, P. arizonica, with its purported more southern origin, presumably migrated through the lowland conifers (e.g., subsection Cembroides) northward into southern Arizona’s mountains during a cooler but more humid glacial. Systematics of Ponderosa Pine Due to the inherent complexity of the topography, historical biogeography and evolution, as well as current evolutionary response to Quaternary climate changes (Axelrod 1986), the systematics of subsection Ponderosae is challenging and unresolved. In general, species closely related to ponderosa pine have a wide distribution ranging from British Columbia to Montana and south along the Cascade-Sierra Nevada, Rocky Mountains, isolated mountain systems of the Great Basin, Colorado Plateau, and disjunct mountain islands down into the Sierra Madre Occidental and Oriental of Mexico (Conkle and Critchfield 1988). Ponderosa pine is one of the most widely distributed Pinus in western North America (Oliver and Ryker 1990). Of the many classifications of Pinus, Shaw (1914, 1924) placed P. ponderosa (Douglas ex Lawson) within the Group Australes, although currently the more accepted placement is within the subsection Ponderosae in the classification by Little and Critchfield (1969). This placement is based on morphologic features and is supported by nuclear and chloroplast phylogenies (Gernandt 2005). In 1880, Engelmann described a Rocky Mountain form (var. scopulorum) of P. ponderosa as having smaller cones, shorter needles (leaves), and 2-needled fascicles as compared to populations found in Oregon and California (Farjon and Styles 1997). The relatively low (<3%) frequency of 2- needled fascicles (Haller 1965), significant isozyme differentiation (Conkle and Critchfield 1988), and differential growth response in range-wide provenance tests (Conkle and Critchfield 1988) of the more western populations (var. ponderosa) effectively segregate the Rocky Mountain populations. Later, two distinct races of var. scopulorum were identified: Rocky Mountain and Southwestern (Conkle and Critchfield 1988). Within var. scopulorum, the Southwestern trees produce higher frequency (>85% vs. 15—40%; Haller 1965) of 3-needled fascicles (average 2.7-3.2 needles/fascicle; Peloquin 1984), differential growth responses (eg, slower early-season growth and double growth flushes; Conkle and Critchfield 1988), and more d-pinene (50% vs. 5%; Smith 1977), a monoterpene found in the xylem resin, compared to the Rocky Mountain race. While the Southwestern race is clearly recognized, the transition to the Rocky Mountain race occurs gradually and broadly over the Great Basin and Colorado Plateau (Haller1965). Shaw’s ( 1914, 1924) classification of P. ponderosa included P. an'zonica, P. engelmannii, and P. jeffreyi, while Little and Critchfield (1969) and The Flora of North America (Kral 1993) considered P. arizonica to be a variety (var. arizonica Engelmann) of P. ponderosa. However, the consensus in this field is that Arizona pine is a distinct species (P. arizonica Engelmann) from Southwestern ponderosa pine (Peloquin 1984, Conkle and Critchfield 1988, Price et al. 1998) due to its more southern distribution in isolated mountains of Arizona, New Mexico, and Texas south into Mexico (Farjon and Styles 1997), production of 5- needled fascicles (4.6-5.4 needles/fascicle; Peloquin 1984), and generation of close to 100% d-pinene with no A3-carene (Peloquin 1971). Obfuscating contemporary American classification of the species, authors more familiar with the Mexican Pinus redescribed P. arizonica with usually fewer needles per fascicle and further classified the species into additional varieties. In Perry (1991), P. an'zonica occurs in the isolated mountains of southeastern Arizona through the Sierra Madre Occidental, produces 3 (occasionally 4-5) needles per fascicle, and has 6-10 (vs. 2-6 in P. ponderosa var. scopulorum) needle resin canals. He further describes a P. arizonica var. stonniae (Martinez) in scattered populations in west Texas and northeast Mexico with similar needle number, longer needles (20-30 cm vs. 12-22 cm), and fewer (3-8) resin canals than P. arizonica. Farjon and Styles (1997) split P. arizonica into three separate varieties: arizonica, stonniae, and cooperi. The first variety is distributed in southern Arizona and New Mexico through the Sierra Madre Occidental and is characterized by variable needle number (3-4, 5 closer to Sonora) and needle length (10-20 cm) (Farjon and Styles 1997). As by Perry (1991), Farjon and Styles (1997) characterized P. arizonica var. stormiae with a more eastern distribution, longer (14-25 cm) needles, and longer cones. Finally, although found in Durango and the central Sierra Madre Occidental, P. arizonica var. coopen' (Blanco) is characterized by producing 5 (4-5) needles per fascicle and shorter (6-10 cm) needles (Farjon and Styles 1997). The great difficulty in segregating the various morphological variations (morphotypes) within subsection Ponderosae is in part due to environmental clines producing geographic variation within the taxa but also their interfertility. Conkle and Critchfield (1988) summarized results of over 50 years of artificial crossing of Ponderosae at the Institute of Forest Genetics in Placerville, California. This work demonstrated moderate (5-40%) crossability of the Pacific race of ponderosa pine (P. ponderosa var. ponderosa) with the Rocky Mountain race, Arizona pine, Apache pine (P. engelmannii Carriere), and other Mexican species in the Ponderosae; the Rocky Mountain race of ponderosa pine and Arizona pine were not crossed. Within the Mexican species, crossability was high (>40%) (Conkle and Critchfield 1988). Based on 14 species-specific morphologic and monoterpenoid characteristics, Peloquin (1971, 1984) concluded that Southwestern ponderosa pine, Arizona pine, and Apache pine naturally hybridize in all combinations, including rare three-way hybridization. Like Arizona pine, Apache pine lacks A3—carene and has 100% d-pinene (Peloquin 1984) but produces 3 relatively long (25-35 cm; Farjon and Styles 1997) needles per fascicle. Biogeographically, these species tend to be elevationally parapatric with Apache pine replaced by Arizona pine which is then replaced by Southwestern ponderosa pine with increasing elevation (Peloquin 1984). When grown in common gardens, genetically and environmentally controlled characteristics of the three species, as well as putative hybrids produced mostly from Arizona pine and Southwestern ponderosa pine, demonstrate the influence of geographic clines (Rehfeldt 1993). In the Southwestern mountain islands, Rehfeldt et al. (1996) identified a “Taxon X” that combined the three-needled Southwestern 10 ponderosa pine with trees producing a mixed (3,4, and 5) number of needles within a given year. This hypothetical Taxon X also demonstrates a strong geographic cline (i.e., spatial structure) in needle morphology (Rehfeldt 1999, Epperson et al. 2001) and growth (Rehfeldt 1999), with little spatial structure in allozymes (Epperson et al. 2003). Despite significant spatial autocorrelation, geographic clines do not necessarily signify hybridization but rather environmental influence and/or genetic differentiation. For example, Epperson et al. (unpublished data) have documented significant spatial autocorrelation in chloroplast DNA single-sequence repeats (chNA SSRs) from sympatric Ponderosae and have identified two unique populations based on chNA corresponding to the Arizona pine and Taxon X morphotypes. However, chNA is paternally inherited via pollen in Pinus (Wagner 1992) and since pollen flow in ponderosa pine is considered unlimited (Latta et al. 1998), this genetic structure could also be interpreted as unidirectional hybridization of Arizona pine by Southwestern ponderosa pine pollen. In Japan, Watano et al. (1995) proposed unidirectional hybridization based on chNA haplotype patterns within a putative hybrid transition zone between Pinus pumila and P. parviflora var. pentaphylla. Additionally, comparison of multiple gene markers (Watana et al. 1996, Latta and Mitton 1997, Gugerli et al. 2001, Ribeiro et al. 2002, Burban and Petit 2003), including maternally inherited mitochondrial DNA and biparentally inherited nuclear DNA (Wagner 1992), is required to detect presence and direction of hybridization. For example, in a separate paper, Watano et al. (1996) found unidirectional introgression from mitochondrial DNA haplotype patterns but in an 11 equal and opposite direction as the chNA haplotypes in their earlier paper, thereby concluding that gene exchange through both pollen and seed dispersal contribute to the formation of the hybrid zone in P. pumila and P. parviflora var. pentaphylla. Demonstrated strong interfertility and geographic variation within subsection Ponderosae, like other subsections in Pinus (Price et al. 1998), indicate that introgression across sympatric populations of Ponderosae is possible and still requires more detailed investigation. The Santa Catalina Mountains of Southern Arizona The origin of the contemporary topography and geology of the Santa Catalina Mountains began 35 Ma with the final subduction of the Farallon Plate beneath the North American Plate (Baldridge 2004). For 140 Ma, the Farallon Plate had been slipping beneath the North American Plate, thus fueling the extensive Laramide orogeny across the West (Baldridge 2004). By 28 Ma, the trailing end of the Farallon dipped beneath the vast continent thus fueling silicic volcanism from central Mexico through southern Arizona, New Mexico, Texas, and Colorado and forming the Sierra Madre Occidentals and many of the isolated mountain islands of the Southwest (Baldridge 2004). Furthermore, changes in the gravitational potential energy of the lithosphere of the Southwest, due in part to the Pacific Plate grinding against the North American Plate and thinner, more buoyant mantle in the Southwest, generated an extensive force that effectively structured the contemporary Southwest landscape (Baldridge 2004). 12 This extensional orogeny and deformation process was important in producing mountains and drop fault zones throughout the Southwest, thus forming the Basin and Range province (Baldridge 2004). The Santa Catalina Mountains were formed through one type of orogenic extension known as a metamorphic core complex (Force 1997, Baldridge 2004). As upper-crust non- metamorphic (sedimentary) rock became strained by extension, low-angle shear zones formed above the core complex of middle-crust metamorphic (gneiss; Lucchitta 2001) and plutonic rocks, which were then exposed and forced upward in elevation (Force 1997, Baldridge 2004). As the sheared upper crust thickened along the fault line, it became denser and sank (“gravitational collapse”) lower in elevation than the newly exposed metamorphic rock (Baldridge 2004). The Santa Catalina Mountains are particularly complicated compared to other metamorphic core complexes clue to multidirectional extension and shearing forces, intrusions, and uplifts (Force 1997). In general, since crustal tension occurred in an east-west direction, linear mountain ranges like the Santa Catalina Mountains formed perpendicular to the strain direction with broad sediment-filled valleys, or basins (i.e., Basin and Range province), separating the ranges (Baldridge 2004). Roughly triangular in shape and extending over 510 kmz, the Santa Catalina Mountains rise from a basal elevation of less than 760 m in the Tucson Basin to 2,791 m (9,157 ft) on Mount Lemmon. The WNW — ESE ridgelines are separated by steep escarpments into the smaller forerange on the southwest side and the major Mount Bigelow (2608 m) and Mount Lemmon crest to the 13 northwest. The Samaniego, Red, and Oracle Ridges extend the crestline to the north. Shallow lithosols characterize the soil in this precipitous environment, with granitic-based soils in the south and a more complex interspersion of soils derived from various granites, argillite, quartzite, schist, andesite, slate, shale, and limestone in the northern 2/3 of the range (Whittaker and Niering 1965, 1968; Force 1997). Soil characteristics in the primarily forested zone (1830-2130 m elevation) show low fertility on diorite and limestone, respectively: bare rock 5- 20 and 20-60%; pH 5.5-7.0 and 8.0; organic matter content 4.4 and 5.1%; nitrogen content 0.12 and 0.21%; carbon:nitrogen 20.0 and 14.2; and cation exchange capacity 17.1 and 25.8 m-eq/100 g (Whittaker 1968; Whittaker and Niering 1968). Minimum soil moisture is uniformly low (<5% at 15 cm) below 2000 m in elevation but increases to 28% on north slopes and 9% on south slopes near the peak (Shreve 1915). Haase (1970) likewise found the south to southwest aspects to be significantly more arid during all times of the year, while soil with southern exposure and at lower elevations experience a greater diurnal temperature range (Shreve 1924). Mature drainages with steep sides dissect the slopes and carry water and eroded crustal material downhill to the extensive bajadas of the Tucson and San Pedro Basins. Because the topographically complex and arid Southwest is geographically close to the Pacific Ocean, Gulf of California, and Gulf of Mexico, as well as located between mid-latitude and subtropical atmospheric circulation regimes, the region experiences a unique bimodal precipitation pattern (see review by Sheppard et al. 2002). A subtropical high-pressure ridge (“Bermuda 14 High") advances northward, while the predominant westerlies retreat, and high solar insolation over large upland areas bordered by lowlands generates convection for large amounts of oceanic moisture to move into the Southwest, especially Arizona. Over 60% of the total annual precipitation for the Santa Catalina Mountains falls during this North American monsoon from July through September. Much of this rain falls as intense convective storms with frequent lightning over the mountains in the late afternoon, rushes downslope as flashfloods, and evaporates from the soil surface (personal observation). Most of the remaining moisture falls during the winter months as large cyclonic storms come off the Pacific (Sheppard et al. 2002). These rain events are typically widespread with soaking rains at low elevations and snowfall at high elevations for consecutive days. Climate variability in the Southwest is strongly influenced by the interactions and synergism of the El Niflo - Southern Oscillation (ENSO) and Pacific Decadal Oscillation (PDO), which acts to strengthen ENSO effects in positive years and dampen ENSO effects in negative years (Sheppard et al. 2002). For example, years of combined El Nino and positive PDO result in cool and wet winters, while negative PDO enhances the warm and dry winter effects of La Nifla years (Sheppard et al. 2002). Over the last century, the Southwest has experienced low-frequency climate variability with periods of abundant moisture (1905-1930), extensive drought (1942-1964), and warm, wet winters . with irregular summers since 1976 (Sheppard et al. 2002). The influence of changing environment with elevation on vegetation patterns in the Santa Catalina Mountains has been reported in classic studies by 15 several well-known plant ecologists (Shreve 1915, 1922, 1924; Fuller 1916; Whittaker and Niering 1965, 1968, 1975). In the broadest sense, Shreve (1915) described four vegetation types from low to high elevation: desert, encinal (from encina, Spanish for evergreen oak, by J.W. Harshberger; Shreve 1915), and forest, which he further divided into pine and fir forests. In the archetypal gradient analysis, Whittaker and Niering (1965) described 9 vegetation types by elevation, aspect, aridity, and topography (i.e., canyons versus slopes) summarized by their diagram in Figure 1-1. Further work by Whittaker and Niering (1968) showed a similar transition from Desert Scrub through pine forest on limestone and granite derived soils with an upward elevational shift in species distributions on the limestone soils probably due to increased bedrock fissuring that drains available moisture, as well as a Cercrocarpus Scrub zone below the Pine-Oak Woodland. Whittaker and Niering’s (1965) Pine Forest was dominated by ponderosa pine with interspersed western white pine (Pinus strobiformis). At high elevation, dense ponderosa pine stands develop and become successional to Douglas-fir (Pseudotsuga menziesir) (Peet 1988). Other forest components above dominant ponderosa pine include white fir (Abies concolor), subalpine fir (Abies Iasiocarpa), corkbark fir (Abies Iasiocarpa var. arizonica), and quaking aspen (Populus tremuloides) (Whittaker and Niering 1965, Peet 1988). Within mid- elevations, ponderosa pine naturally forms a woodland mosaic of interspersed grassy patches (Peet 1988). Lower in elevation, ponderosa pine is replaced by a more open woodland of encinal (Quercus spp.) and pygmy conifer forest 16 containing Chihuahuan pine (Pinus chihuahuana), pinon pine (P. cembroides), junipers (Juniperus spp.), and Arizona cypress (Cupressus arizonica) (Whittaker and Niering 1965, Peet 1988). Stand density and structure of all these forest types are strongly influenced by fire and interannual climate variability (Peet 1988). 17 Veqetanon of the Santa Catalma mountain: (South slope. Data above 9000 feet from Pinaleno Mountains.) T l l l l l l Tl I l Picea engelmanii Subalpine Forest 1000( — 3000 L Abies Iasiocarpa " Mixed Conifer Forest Pseudotsuga menziesii Alnus tenuifolia p _____ 9000 —- —- -Montane Fir Forest —- —Pinus strobiformis — - Pinus ponderosa Abies ooncolor Acer Iabmm zsoor Acer gmndjdentum Pine Forest Quercus rugosa 8000 J Ace! grandidentum . - Pinus ponderosa Pine-Oak Forest 7000 Pinus ponderosa . _ Quercus gambe/ii Quercus hypo/eucordes {nus cembroides 2000 Alnus oblong/fo/ia Juniperus deDDeana - . Quercus arizonica PIne-Oak W_oodland 6000 — (Upper Encinal) _ I I Pinus chihuahuana I Pinus ponderosa Quercus hypoleucoides ‘— Juglans major Cupressus arizonica Canyon 1500» 5000 — WOOdIand Plateaus wn’ghtii Juniper-us deppeana Quercus arizonica Quercus emoryr Open Oak Woodland Vauquelinia califomica (Lower Encinal) Fraxinus . . velurina Quercus oblongrfolla - Desert Grassland 4000 — Fouquien‘a splendens Pmsopis juliflora P l s ”22%;,” Sonoran Desert Scrub - 1000 ._ .- Cercidium microphyllum Elev. Elev. 99”" Camegiea giganrea m. rt. rel/cuffs . P l_~ Fouquien'a splendens A I I Ravines Draws NNE NE ENE E ESE SE SSE S SSW N NNW NW WNW W WSW SW Mesic Xeric Figure 1-1. Vegetation of the Santa Catalina Mountains as drawn by Whittaker and Niering (1965) from 400 vegetation samples in a gradient analysis. Note that Pinus arizonica is not recognized as a species separate from P. ponderosa but replaces the latter species between the Pine and Pine-Oak Forests (personal observation). Also note that Picea engelmannii does not occur in the Santa Catalina Mountains. 18 The 19‘"-20th century changes in fire regime and widespread introduction of grazing by domestic livestock have had a profound effect on the forest structure and composition in the Santa Catalina Mountains. Due in part to a lack of knowledge of pre-European lightning frequency as well as anti-Native American bias, pre-European humans were considered to be the primary driver, while climate and lightning were secondary drivers to the past fire regime in the mountainous regions of the Southwest (Swetnam and Baisan 2003). Based on fire scar records, large fires occurred synchronously about every 7.5 years from 1700-1900 in Southwestern ponderosa pine forests (Swetnam and Baisan 2003). While prehistoric fire frequencies were generally highest during times of greatest human conflict, the understory fuel amount, which takes a few wet years of good growth to accumulate between fires, is strongly associated with large fires in the xeric ponderosa pine forests (Swetnam and Baisan 2003). Domestic livestock (“hoofed locusts;” Muir 1911) introduced in the late 19th century reduced interannual fuel accumulation by grasses and forbs. As a consequence, fire frequency significantly decreased; the last widespread fire of the 20th century occurred in 1900 (Swetnam and Baisan 2003). Due to reduced fire frequency from livestock grazing and organized fire suppression, begun in the Santa Catalina Mountains by the Forest Service during the 1910 Alder Canyon fire (Alexander 1991), stand density of Southwestern ponderosa pine in Arizona increased by several orders of magnitude through the 20th century (Moore et al. 1999). Finally, after close to a century of fire suppression, the Bullock (2002) and 19 Aspen (2003) Fires burned close to 470 kmz, or 92% of the Santa Catalina Mountain area (Coronado National Forest 2002, 2003). Human settlement and use of the Santa Catalina Mountains has certainly had a profound effect on the system. Evidence of the Hohokam living throughout the Santa Catalina Mountains exists from 100 BC. to 1400 AD, after which they disappeared (Alexander 1991 ). Shortly afterward, Apache people occupied the northern foothills and lower slopes during the Summer (Alexander 1991). The earliest Spanish explorer was Francisco Vasquez de Coronado in the 1540s (Alexander 1991), and the 17'“ century arrival of Spanish settlers in the Santa Cruz Valley distressed the Apaches, whereby they raided settlers and clashed with government armies until the late 1880s (Alexander 1991). There are two stories about the European naming of the mountains: in 1697, Father Eusebio Kino called them the Santa Catarino Mountains; and, in the 16908, a Jesuit mission in the northern foothills was called Santa Catalina de Cuitabagu, or “well where people gather mesquite beans" (Alexander 1991). The Europeans, though, recognized the potential for mineral wealth in the mountains and began exploring for gold as early as 1871 (Alexander 1991). Homesteading associated with mine claims began in 1880, and businesses to support the mining trade soon sprung up in the mountains (Alexander 1991). The highest point, Mount Lemmon, was named for Sara Plummer Lemmon, who together with botanist John Gill Lemmon and Emerson Oliver Stratton, were the first Europeans to be recognized as reaching the peak (Alexander 1991). During this trip, J.G. 20 Lemmon first noted the presence of Arizona pine, which he recognized as a variety of ponderosa pine (Alexander 1991). In 1902, President Theodore Roosevelt established a 100,000-acre (404- kmz) forest reserve in the Santa Catalina Mountains and, in 1908, created the Coronado National Forest, which then included the Dragoon, Rincon, Santa Rita, and Whetstone Mountains of southern Arizona (Alexander 1991). Trails to the summit begun in 1897 were improved, fire towers at Lemmon Rock Lookout (1913) and Mount Bigelow (1916) as well as numerous look-out trees aided fire detection efforts, and eventually a road was constructed to the summit (Alexander 1991). The village of Summerhaven, Boy and Girl Scout camps, and Forest Service campgrounds and picnic areas were developed from early camping sites throughout the early 20th century (Alexander 1991). In 1953, several ski runs and a lodge were constructed near the summit, bringing even more tourists to the mountain. The US. Air Defense Command established and operated a radar station (Semi-Automatic Ground Environment, 684“1 Radar Squadron) at the summit from 1959 through 1969 (Radomes, Inc. 2005). In 1962 and 1970, the University of Arizona constructed astrophysical observatories on Mount Bigelow and at the newly abandoned radar station on Mount Lemmon, respectively (Alexander 1991). Following the recent reconstruction of the 28- mile-Iong paved Catalina Highway, originally completed in 1949 (Alexander 1991), 22,000 vehicles per year of recreationists, homeowners, land managers, contractors, and scientists travel into the Santa Catalina Mountains (PCDOT 2007). 21 The nature of the Santa Catalina Mountains has changed dramatically over time, none more so than since Europeans began exploring its economic and recreational potential. While the ebb and flow of glacial climates allowed slow intermixing of Rocky Mountain and Madrean species, we are now challenged with understanding the resultant genetic diversity in the face of altered fire regimes, human uses of the mountain resources, and relatively rapid climate change. I embarked on this quest to examine ecophysiological influences on the distribution for one closely related group of species in the Ponderosae with the aim of contributing to a better understanding of their distribution, ecology, and genetic diversity, and with important implications for their management. 22 CHAPTER 2 DISTRIBUTION AND ECOLOGICAL NICHE DIFFERENTIATION OF SYMPATRIC PONDEROSAE TAXA IN THE SANTA CATALINA MOUNTAINS OF ARIZONA Introduction Phylogeographers are challenged not only with the genetic vagaries of species delimitation, especially with closely related and sympatric taxa, but also the difficulties in delineating the distribution of species. The geographic range of a species is commonly considered to be a spatial manifestation of its ecological niche (Brown and Lomolino 1998). However, one must consider the various interpretations of the niche when discussing the potential distribution of species. Early uses of the term “niche” implied a place in the environment that could support a species, thus the environment held a certain number of niches (Elton 1927). Conversely, Hutchinson (1957) proposed the n—dimensional hypervolume in which a species can persist indefinitely within the boundaries of two measured independent environmental variables and cannot exist outside of those boundaries. This fundamental, or Grinnellian, niche will be considerably larger than its realized niche, or the region between those environmental variables where the species actually exists, due to biotic interactions (Hutchinson 1957). Dispersal, whether limited or extended beyond suitable environmental conditions, can obfuscate the relationship between the geographic range and realized niche (Pulliam 2000). However, beyond the local scale, the geographic distribution of 23 species is largely influenced by the environment, especially climate (Woodward 1987). The relationship between the geographic distribution and spatially structured environmental characteristics can be used to investigate the potential distribution and fundamental ecological requirements for a species. Because the influence of biotic interactions varies across a species’ geographic distribution, models representing the potential distribution, or suitable habitat (Guisan and Thullier 2005), for a species are arguably more analogous to its fundamental rather than realized niche (Peterson et al. 2002). The amount of research using predictive species distribution models (SDMs) to understand the ecology and evolution of species has exploded over the last 10 years due to advances in computing technology. For example, even with few records of occurrence, the distribution of rare Madagascaran geckos were successfully modeled with advanced forms of species distribution models (Pearson et al. 2007), which can then be used to focus search and conservation efforts. From comparisons of occurrences to mean climate values through simple linear and additive models, Vetaas (2002) inferred the relative role of cold versus hot temperatures on the distribution of Himalayan Rhododendron species from herbarium specimens and field data. The potential role of competition between closely related South American pocket mice was identified through overlapped predicted distributions as a function of relevant environmental variables (Anderson et al. 2002). Peterson et al. (1999) used reciprocal geographic distributions to demonstrate the role of geographic rather than ecologic isolation during speciation between 24 sister taxa of butterflies, birds, and mammals in Mexico. Environmental divergence during speciation of Andean frogs was demonstrated with principal components analysis (PCA) of environmental variables determined to be suitable habitat for different species through predictive models (Graham et al. 2004). Using an ecological niche modeling method, Costa et al. (2002) successfully identified four ecologically distinct populations of a major Chagas’ disease vector in Brazil resulting in a change in monitoring and control measures. Application of a variety of ecological niche and species distribution models has resulted in advancement in the understanding of species' biogeography, ecology, evolution, and conservation. An assortment of SDM methods have been developed that can produce successful models using only documented presence localities (“presence-only”). Sources of occurrence data, such as herbaria, museums, and field surveys, do not usually incorporate documentation of absences, especially spatially redundant for rare or cryptic species. Spatial interpolation using explanatory variables of geographic coordinates and neighborhood similarity as measured by abundance is capable of producing comparable models to those using environmental variables in a regression tree algorithm (Bahn and McGill 2007), which demonstrates the predictive value of spatially structured variables. However, most presence-only modeling methods use information about where a species is documented to occur and its associated environmental characteristics to generate a habitat suitability map (Guisan and Thuiller 2005). In these models, pseudo-absences are randomly drawn from the background to evaluate 25 the model’s ability to correctly predict the recorded occurrences for the species. Of course, the modeled species is assumed to be in pseudo-equilibrium with its environment, thus overfitted models may conservatively underestimate the spatial extent of the potential distribution for the species (Guisan and Thuiller 2005). Another major assumption of any SDM is that an appropriate scale, both resolution and extent, is used in the modeling exercise (Maurer 2002, Guisan and Thuiller 2005; but see Guisan et al. 2007). Although each environmental variable will interact with a population at a different scale (Thomas et al. 2002), modeling algorithms usually require constant scale across explanatory variables. Despite these and other assumptions and limitations to SDM (Guisan and Thuiller 2005), the continuing advances in statistical techniques have resulted in great strides in understanding the biogeography of organisms. The comparison of modeling algorithms, each with means to reduce autocorrelation effects and estimate goodness-of-fit, in a variety of species modeling challenges has been extensive in the literature (e.g, Segurado and AraL'Ijo 2004, Guisan and Thuiller 2005, Elith et al. 2006, Tsoar et al. 2007). Some of the more common presence-only algorithms include regression (generalized linear, additive, and multivariate models), bioclimatic envelope, boosted decision trees, artificial neural networks, classification trees, decisions made with genetic rule sets, and maximum entropy (Elith et al. 2006). The latter two methods are drawn from the machine-learning community and are considered to be the most advanced (Elith et al. 2006) at this time. As implemented in GARP (Genetic Algorithm for Rule-Set Prediction, 26 httpzllnhm.ku.edu/desktopgarp/index.html), the genetic algorithm uses species localities and environmental data to produce a niche-based model using conditional decision rules from a combination of approaches, including atomic, logistic regression, and bioclimatic envelope (Anderson et al. 2002). The maximum entropy algorithm, as employed by Maxent (Maximum Entropy Species Distribution Modeling, Phillips et al. 2006b), calculates a probability distribution of maximum entropy (i.e., closest to uniform) when constrained by incomplete information about the species distribution, which presents itself as species occurrences within pixels across the study area and the associated environmental variables (Phillips et al. 2006). Under the maximum entropy distribution, the expected value (probability) of each pixel is equal to its empirical average from the occurrence data (Phillips et al. 2006). Consequently, Maxent may be conservative by constraining probabilities in the environmental niche space where the species is not known to occur (Pearson et al. 2007). The Ponderosae of the mountain islands in the American Southwest presents an opportunity to employ SDMs to segregate the realized niche of taxa with overlapping distributions by their association with the heterogeneous environment. These wind-pollinated conifers are extensively distributed from central Mexico through British Columbia and throughout the major mountain ranges and foothills (Little and Critchfield 1969). Despite significant outcrossing (>90%; Mitton et al. 1977), the Rocky Mountain ponderosa pine (Pinus ponderosa var. scopulorum Engelmann) experiences significant intrapopulation allozyme variation associated with geographic clines, such as elevation (Hamrick 27 et al. 1989, Rehfeldt 1990) and slope aspect (Mitton et al. 1977, 1980). lntrapopulation genetic variation of morphological characteristics, such as needles per fascicle, needle length, shoot length, and heightzdiameter ratio (Rehfeldt 1999), are also associated with geographic clines. Factors varying with geographic clines such as elevation and slope aspect and that seem to influence genetic variation center on length of the frost—free season and patterns of precipitation (Rehfeldt 1990). For example, years with prolonged drought are linked to decreased number of needles per fascicle in the following year (Haller 1965). Ecological response surface models produced by Humphries and Bourgeron (2003) identified maximum January temperature, annual solar radiation, and maximum July temperature as important predictors for presence of mature ponderosa pine in granite-based soils. Limited pollen dispersal (Latta et al. 2001) and limited seed-bearing trees (Linhart and Mitton 1985), in combination with strong environmental selection gradients in mountainous areas (Rehfeldt 1986, 1990), generates considerable genetic and morphologic spatial structure at local scales in ponderosa pine (Epperson et al. 2001). Although less work has been done on the closely related Arizona pine (P. an'zonica Engelmann; formerly P. ponderosa var. arizonica Engelmann) (except see Barton et al. 2001), its mating system and interactions with the environment are expected to produce similar genetic structures. Due to extensive geographic variation, the systematics of the ponderosa pine and its closely related species in the Southwest’s mountain islands has not been resolved (Kral 1993). The Southwestern race (Conkle and Critchfield 1988) 28 of the Rocky Mountain ponderosa pine reaches its southernmost distribution in these disjunct habitats, while the Sierra Madrean Arizona pine extends northward into southern Arizona, New Mexico, and Texas (Farjon and Styles 1997). Putative hybrids with varying needle number are reported by Peloquin (1971, 1984) to occur where the species distributions come into contact; artificial hybridization between closely related Ponderosae has been demonstrated by Conkle and Critchfield (1988). Alternatively, based on allometric, needle, and phenologic properties expressed in a common garden, Rehfeldt (1996) suggests that a nonhybrid “Taxon X,” identifiable in the field by producing variable needle number (range 3-5, mean 3.3, standard deviation 0.4; Rehfeldt et al. 1996) in comparison to the mean 3- and 5-needled fascicles of Southwestern ponderosa and Arizona pine, respectively (Peloquin 1984), replaces the commonly identified ponderosa pine of the Southwest. In the Santa Catalina Mountains of southern Arizona, Epperson et al. (2001, 2003) demonstrated a very strong elevational cline in number of needles per fascicle (0.55-0.64% heritable; Rehfeldt 1993) with little differentiation in allozymes across a contact zone between Arizona and Southwestern ponderosa pine. Unpublished work on the same trees by Epperson et al. shows strong spatial structure in chloroplast DNA single- sequence repeats (chNA SSRs) with two unique populations with regard to chNA heritage: trees corresponding to Arizona pine and the Taxon X morphotype. The objectives for this paper are to document the geographic distribution and determine if there is any detectable ecological niche differentiation between 29 the closely related Ponderosae taxa in the Santa Catalina Mountains. These taxa form a relatively continuous cover interrupted only by uninhabitable rock from about 1760 m to the summit of Mount Lemmon at 2791 m, depending on aspect and associated changes in forest cover; on the north slope of the peak, ponderosa pine becomes a minor component of a mixed subalpine forest community. Populations producing a mean of 3 needles per fascicle at the summit and on the north slope of Mount Lemmon, 5 needles per fascicle below 2438 m, and 3-5 needles per fascicle with high intratree variance (“Taxon X” of Rehfeldt et al. (1996), or “mixed”) between these two elevations have been reported by Dodge (1963) and observed by others (F. Telewski and T. Harlan, personal communication). Following the Bullock (2002) and Aspen (2003) fires, the Coronado National Forest (W. Hart, personal communication) identified a need to recognize the genetic diversity of the Ponderosae in the Santa Catalina Mountains and to delineate their distributions for reforestation and conservation efforts. Given the close genetic relatedness within Ponderosae, I expect distributions for all of the taxa to be associated with similar climatic factors, including minimum winter and maximum arid foresummer temperatures, and monsoonal precipitation. However, I expect that the 3- and 5-needled trees should have little overlap in ecological niche space. Under the hybridization hypothesis, the mixed-needled trees should have an intermediate ecological distribution, or niche overlap, and be less sensitive or intermediate in response to relevant climatic factors than the other taxa (Anderson and Stebbins 1954). 30 Assuming Taxon X (Rehfeldt et al. 1996), the 3- and mixed-needled trees should have the same distribution, occupied niche space, and climatic sensitivity. Methods Study area The Santa Catalina Mountains are located within the Basin and Range Province of the American Southwest (Figure 2-1). These isolated mountains were formed from a metamorphic core complex uplifted as strong crustal extension forces sheared the overlying sedimentary rock (Force 1997, Baldridge 2004). Multidirectional extension, shearing, intrusions, and uplifts have produced a ruggedly complex landscape in this system (Force 1997). The resulting mountainous area is roughly triangular in shape, extends over 510 kmz, and rises from a basal elevation of less than 760 m in the Tucson Basin to 2,791 m (9,157 ft) on Mount Lemmon. Soils are shallow lithosols derived largely from acidic granites and local outcrops of sedimentary rock (Whittaker and Niering 1965, 1968; Force 1997). Mature drainages with steep sides dissect the slopes and carry water and eroded crustal material downhill. The semiarid climate is characterized by the Southwestern bimodal precipitation pattern with an arid foresummer (May-June) followed by summer monsoon (Sheppard et al. 2002). Annual precipitation is about 820 mm at Palisade Ranger Station (2426 m elevation) near the observed transition zone between the 3— and 5-needled trees (Western Regional Climate Center 2007), with approximately half falling during the summer (Whittaker and Niering 1965). 31 Figure 2-1. General distribution of ponderosa pine (top. Pinus ponderosa. Thompson et al. 1999), and location of study region in the Santa Catalina Mountains in southeastern Arizona (bottom. USGS 2004). 32 The minimum January, maximum June, and annual temperatures at this location are -4, 23, and 15° C, respectively (Western Regional Climate Center 2007). The vegetation is strongly influenced by elevation and associated changes in climate (Shreve 1915, Whittaker and Niering 1965). Based on Whittaker and Niering’s classic study using gradient analysis in the Santa Catalina Mountains, the first “ponderosa pine” trees at low elevation occur around 1500 m elevation mixed with pinon pine, (Pinus cembroides), Chihuahuan pine (P. chihuahuana), and oak scrub (Quercus spp.). From about 1900 to 2500 m elevation, depending on aspect and topography, ponderosa pine forests are dominant. Forests near the Mount Lemmon peak become dominated by Douglas-fir (Pseudotsuga menziesil), white fir (Abies concolor), and Arizona corkbark fir (Abies Iasiocarpa var. arizonica), with interspersed ponderosa pine Field data collection Georeferenced species occurrence (presence-only) and needle samples were derived from several sources. Most of the data (n=671 samples, “SCAT;” data in Appendix, Table A-1) were collected with the intent to examine the spatial distribution of the three morphotypes in the Santa Catalina Mountains. Sampling locations were determined in an arbitrary manner to efficiently capture the spatial variation of these morphotypes. Based on species cover maps (Terri Austin, CNF, personal communication), institutional knowledge (William Hart, CNF, personal communication), and past observations (Frank Telewski, and Thomas Harlan, University of Arizona, personal communication), sampling locations were 33 located prior to entering the field so as to confirm centers of and transitions in distribution between the three morphotypes. Accessibility was limited to proximity of road and trail due to the rugged terrain, thus some geographical bias was introduced by the sampling methods. At each predetermined location, over 300 geographic positioning system (GPS; GPSMAP 6008, Garmin International, Inc., Olathe, Kansas, USA) readings were averaged to increase georeferencing accuracy. The point-centered quarter method was used to select the four closest Ponderosae trees. This method was used to increase the efficiency of species data per effort in a less subjective manner (Cottam and Curtis 1956) and was employed by Dodge (1963) when examining biogeographic patterns of Ponderosae in Arizona. From each tree, diameter at breast height (1.37 m) and distance to the GPS unit were recorded, and 1-4 needle-bearing branches from the most southern accessible side were cut with 23-ft pole pruners and stored in labeled paper bags; the more xeric southern exposure was expected to force variable production of needle number per fascicle, but a subsequent examination by aspect from trees across a transition zone (Epperson et al. 2001) yielded no difference (two-sample t-test, p>0.12, n=62 trees (19713 fascicles». In some cases, the closest tree had been recently killed by fire (CNF 2002, 2003) or branches were completely inaccessible, so the next closest tree in the quadrant was sampled. Predominant cover type and other notes were also recorded. From August 2004 to October 2006, 671 georeferenced samples were collected from 176 locations (Figure 2-2). 34 35 E 9.28-82 E x: H 33 £83.59... «53> :O_um>w_m Eoméoz . How a Z V $8.85 $50 ucm A .L.u3w 20:39.56 9: 9:3 c.3028 mEBQ mEEEmm Co 83955 .N-N p.59". 35 Samples were also collected in association with other components of this research program (“Non-SCAT,” Figure 2-2). Needles from at least the south side of each tree were similarly collected, and georeferenced locations were recorded by GPS at Kimball Peak Saddle (n=10 in 2001; GPS 12, Garmin lntemational, Inc., Olathe, Kansas, USA), Bear Canyon (n=15 in 2002; GPS 12), Wilderness of Rocks (n=10 in 2002; GPS 12), and Mount Lemmon (n=183 in 2004; G820 Personal Data Mapper, Leica Geosystems AG, St. Gallen, Switzerland). I identified the coordinates for additional trees (n=23) sampled and intensively studied at Palisade Rock with the aid of field notes and an online mapping program (Google Earth, URL: http:l/earth.google.com, accessed May 2007) Needles from each sample were examined by year to determine the average needle number for the tree (Epperson et al. 2001). Needle number was calculated both by year and across sampled years. Species, or morphotype, determination was by Peloquin’s (1984) classification: trees with mean needle number less than 3.2 are Southwestern ponderosa pine, greater than 4.6 needles per fascicle are Arizona pine, and between 3.2 and 4.6 needles per fascicle are Mixed (hybrids (Peloquin 1984) or Taxon X (Rehfeldt et al. 1996)). In sum, 75,083 fascicles from 671 trees in the distribution study (“SCAT”) were individually examined. The species occurrence data files (.csv) used for SDM contained a combination of geographic coordinates (WGS 1984) with either mean needle number or morphotype, depending on the model. 36 Environmental data The explanatory variables used in the species distribution modeling were derived from online sources. Digital elevation data were downloaded from the National Elevation Dataset (“NED,” 1/3 arc-second, or ~10-m resolution; USGS 2004). Annual and monthly average minimum temperature, maximum temperature, and precipitation data (30 arc-second, or 800-m resolution) for 1971-2000 were downloaded from PRISM (2006). The data were projected to a geographic coordinate system using the WGS 1984 datum in ArcCatalog (ESRI 2006). I imported the raster coverages into ArcView (ESRI 2006) as grids, clipped the grids to a rectangular polygon surrounding the Santa Catalina Mountains (32.6181481469°N, -110.984166666°W to -110.564629628°N, 32.2984259247°W) using Hawth’s Analysis Tools (Beyer 2006), and then converted the clipped grids into ascii format in ArcCatalog (ESRI 2006). Both NED (USGS 2004) and PRISM (2006) data are globally available and commonly used in the SDM literature. Modeling algorithm The SDM software called Maxent (Version 2.3, Phillips et al. 2006b) was the only method selected to focus less on modeling techniques and more on the ecophysiological relevance of the output. Maxent is proven useful in conservation planning, easy to use, freely downloadable from the internet, and successful at predicting species distributions in comparative studies (eg. Elith et al. 2006), even with small (<25) sample sizes (Hernandez et al. 2006). In 37 addition, Maxent does not explicitly require absence data, incorporates interactions between variables, minimizes model overfitting through regularization, and produces spatially continuous output for comparative interpretation (Phillips et al. 2006a). Furthermore, model complexity can be controlled by selecting from a combination of linear, quadratic, product, binary, and hinge mathematical functions (“features”). Recommended user-specified parameters included: 30% of the occurrence data set aside for model validation (“testing”); maximum iterations = 500; convergence threshold = 105; regularization multiplier = 1.0; automatically select mathematical features depending on sample size; and remove duplicate occurrences within an environmental pixel, except for comparisons (see Results). The major drawbacks are its use of an exponential distribution, which is inherently unbounded above, and its recent development in terms of fully understanding the implications of regularization, sample size, and spatial extent as well as its performance when extrapolating to environmental features outside the model’s calibration (Phillips et al. 2006a). Model selection The number of occurrences for a given species affects the predictive potential for any statistical procedure, including SDM. The asymptote of maximum predictive power depends not only on the sample size, or number of occupied pixels across the study region, but also the quality of and heterogeneity within the explanatory variables, the strength and shape of the species response, 38 and the spatial resolution of the analysis (Hernandez et al. 2006). However, due to the sampling structure, multiple trees for a given species will share a geographic coordinate, thus spatially pseudoreplicating that species’ occurrence at that location. In addition, the useful sample size can be reduced due to the resolution of the explanatory variables. Any duplicate occurrences for a species within the 064-ka PRISM pixel would also constitute spatial pseudoreplication in the modeling algorithm. Consequently, in the SCAT dataset, the number of occurrences for the 3-needled morphotype decreased from 146 to 31, with similar reductions for the other morphotypes. This sample size should be adequate given machine-learning methods, such as Maxent, correctly predict occurrences around 90% of the asymptote with 10 sample points and near maximum with 50 data points (Stockwell and Peterson 2002). Most algorithms allow inclusion of multiple environmental variables to use as explanatory variables for species distribution modeling. With extensive data freely available on-line, the modeler is tempted to include as many variables as the algorithm can tolerate to increase the precision and accuracy of predicting the spatial extent and probability of species occurrence within the region of study. However, ecological theory must be part of the model building exercise or else the output will have little relevance or application (Austin 2002). For example, in most models and ecology textbooks, species response to environmental gradients is assumed to be unimodal and symmetric. Unlike earlier algorithms like GLM, Maxent incorporates a number of mathematical functions (e.g., linear, 39 quadratic, threshold, and hinge) that can be combined into a response to each variable. The selection of explanatory variables depends in part on the ultimate application or hypotheses to be tested by the model. Conservation biologists interested in parameterizing occupied habitat for projection to potentially suitable habitat elsewhere will likely utilize a larger suite of environmental variables to capture and better transfer the multidimensional niche space for the species with more concern on successful prediction and less concern with the direct physiological implications of the variables. When the objective is to understand more about the environmental constraints on the distribution of a species, fewer but ecophysiologically important variables should be used to reduce interaction effects. In addition, decreasing the number of spatially and temporally autocorrelated variables, such as precipitation in July and August (r=0.92 for the Santa Catalina Mountains; Appendix, Table A-2), to those representing seasonally important features of the habitat further reduces potentially complex interaction effects. In this research, physiologically relevant variables representing the amount and pattern of precipitation and temperature (PRISM 2006) were used to parameterize the SDMs because of their demonstrated relation to the distribution of ponderosa pine across the Rocky Mountains of the American West. Daubenmire (1943) showed experimentally that soil drought during the summer is the primary determinant of lower elevation in seedlings, and thus adults, of Rocky Mountain conifers, including ponderosa pine. Although lower elevationally 40 distributed conifers had a higher tolerance for high summer temperatures, the ability to access moist soil (i.e., rate of root growth) during summer drought was most important (Daubenmire 1943). In California, Haller (1959) showed that lower elevational distribution in ponderosa pine was determined by drought, while the upper limit was determined by cold temperatures. He also noted that occupied elevational bands decrease in elevational range with decreasing latitude due to more rapid altitudinal climb in isohyets than isotherms further south (Haller 1959), thus the effects of summer drought and cold winter temperatures should be less spatially distal in the Santa Catalina Mountains. Independently, Yeaton et al. (1980) showed that summer drought stress limits the ability of ponderosa pine seedlings to grow below their lower elevation distributional limit in the Sierra Nevada; he also demonstrated unimodal mortality along an elevational gradient. Tree-ring data from P. ponderosa var. scopulorum suggest that summer drought becomes more important with decreasing elevation (Adams and Kolb 2005). At larger spatial scales, Oliver and Ryker (1990) surmised that precipitation in May and June were most associated with ponderosa pine distributions, and Thompson et al. (1999) showed that a combination of annual precipitation, January and July precipitation, and January and July temperature could effectively contain the realized niche for ponderosa pine across the American West. Similarly, Norris et al. (2006) demonstrated the importance of temperature in January and July, as well as growing season precipitation, in characterizing the realized niche space for P. ponderosa var. scopulorum. Further support for the importance of maximum January and July 41 temperature, in addition to the overall solar radiation, in predicting presence of ponderosa pine growing on granite in the Southwest is provided by Humphries and Bourgeron (2003). Consequently, the variables selected for this research in the Santa Catalina Mountains characterize winter cold and summer drought (Table 2-1). Models using variables suggested by the literature (“Lit”) focus on the influence of cold temperatures in restricting upper elevational distributions via the mean minimum and maximum January temperature variables (PRISM 2006). The influence of growing season drought, combining warm temperatures and soil moisture, was encapsulated by mean maximum June temperature (“arid foresummer”) and mean July precipitation (“monsoon”) (PRISM 2006). Models using the full set and various combinations of climatic (PRISM 2006) or topographic (USGS 2004) variables were generated to compare predictive versus physiological tolerance types of output. Topographic variables were not combined with climatic variables due to expected correlation between the variables. As mentioned earlier, spatial considerations influence the model outcome. For example, the spatial extent of the species’ range with regard to the environment can influence the success of the predictive model. Wide-ranging species will be less associated with certain environmental characteristics and thus could result in reduced predictive potential (Stockwell and Peterson 2002). On the other hand, McPherson et al. (2004) conclude that effects of range size may be statistically artifactual. Although Pinus ponderosa has a large 42 geographic range (Little 1971), its horizontal range within the craggy Santa Catalina Mountains relative to the scale of the explanatory variables is considerably smaller. Table 2-1. Models and environmental variables used in the modeling of species distributions, or suitable habitat (Phillips et al. 2006a), using Maxent (Phillips et al. 2006b). Climate data are either mean annual or monthly normals for 1971-2000 at 800- m resolution (PRISM 2006). Topographic data are 10-m resolution (USGS 2004). Model Variables Full Annual temperature Minimum January-December temperature Maximum January-December temperature Annual precipitation January- December precipitation Win.Sum Minimum December, January, February temperature Maximum June temperature January, February, July, August precipitation Lit Minimum January temperature Maximum January, June temperature July precipitation Lit.Thresh Minimum January temperature Maximum January, June temperature July precipitation Lit. Red1 Maximum January, July temperature Lit.Red2 Maximum January, June temperature NED.Asp Elevation Aspect The extent of the explanatory variables used in training (calibrating) the model has received much less attention in the SDM literature. Restricting the spatial extent of the environmental data to a species’ likely habitat reduces the number of “pseudo-absent” background points and thus affects the tails of the response curves (Thuiller et al. 2004). However, higher probabilities are given for the restricted versus complete environmental data range, thus the complete range is more conservative. R. Pearson (personal communication) has stated that the influence of the spatial extent of the explanatory variables on model 43 outcome for different species is currently attracting much attention by SDM modelers. Model evaluation Successful models depend on their objective and two types of statistics - threshold-dependent and -independent - against which competing models can be compared. Similar to the Type II" error matrix from hypothesis testing, predictive models can be evaluated based on their ability to correctly predict occurrences and absences from real data, whether apportioned from the existing data set (“test data”) or validated against independent data. Sensitivity and specificity are successful positive and negative (absent, or “pseudo-absent”) predictions, respectively (Phillips et al. 2006a). The analogous Type II error is the omission rate which is the fraction of actual observations (by pixel) predicted absent, while the commission index measures the fraction of predicted occurrences where no observation was made (Anderson et al. 2003). A threshold-dependent evaluation includes comparison of the omission rate and proportional predicted suitable area at a particular threshold across the models with a binomial probability calculated to determine if each model is better than random (Phillips et al. 2006a). At an arbitrarily determined cumulative threshold, t, t% of resampled pixels will have cumulative probability of tor less, or t% omission of test localities, and some minimum (correctly) predicted background area (Engler et al. 2004, Phillips et al. 2006a). Put another way, to measure one model output’s ability to discriminate between occupied and unoccupied sites against another model’s output, a threshold of predicted probability of occurrence must be selected at which sensitivity and specificity will be calculated (Pearce and Ferrier 2000). A higher decision threshold will have lower sensitivity (proportion of correct positive predictions) but higher specificity (proportion of correct negative or absent predictions). A variety of thresholds can be selected depending on known rarity of the species and the model application (Pearce and Ferrier 2000, Anderson et al. 2003, Liu et al. 2005). To focus on identifying differences in suitable habitat between the taxa in the Santa Catalina Mountains, I will use the “minimum training presence” (Phillips et al. 2006b), or “lowest presence threshold” (Pearson et al. 2007). This conservative approach will identify pixels where the environment (based on the explanatory variables) is at least as suitable as where the species was recorded with zero omission rate (Pearson et al. 2007), and thus will constrain the potentially suitable niche space for each species. The threshold-independent approach measures model performance by plotting the sensitivity against 1-specificity for all possible thresholds and then calculating the area under the curve (AUC). An AUC value of 1.0 indicates the model (or Maxent distribution) fits randomly selected recorded presences and absences perfectly, while an AUC equal to 0.5 indicates that a random model fits just as well (Phillips et al. 2006a, b). Due to using presence-only data, and thus Maxent generating pseudo-absences, the AUC must remain less than 1.0 (Phillips et al. 2004). The use of AUC on such plots, called receiver operating characteristic (ROC) curves, has a long history in medical research, and its 45 significance is approximated by the nonparametric Wilcoxon statistic (Hanley and McNeil 1982). Ecological niche space Ecological niche differentiation was measured by the ability of one species’ model to predict the occurrence of another species. Species occurrences were overlain onto the Maxent output (.asc) for each of the species in ArcView (ESRI 2006), and then Hawth’s Analysis Tools (Beyer 2006) was used to identify the output pixel’s value (probability) for each occurrence locality. Summary statistics for each combination were calculated in R (R Development Core Team 2007). Visualization of occupied ecological niche space was accomplished by plotting species occurrence points as a function of two- dimensional climate space (Thompson et al. 1999, Norris et al. 2006). The pixel value for the environmental variable on which an occurrence was located was determined as in the previous analysis. Results Spatial distribution of morphotypes The distribution of mean needle number per fascicle across all trees from the distribution study was bimodal (Figure 2-3). Of these 671 trees, 145 (22%) were classified as Southwestern ponderosa pine (PIN_PON), 136 (20%) were classified as mixed-needle (Mixed), 281 (42%) fall into Taxon X (Rehfeldt et al. 1996), and 390 (58%) were classified as Arizona pine (PIN_ARI). Within tree 46 variability in needle production peaked at a mean needle number around 4.0 (Figure 2-4), but 41 PIN_PON and 35 PIN_ARI trees had zero intratree variance in needle number. lntrannual and interannual variability in needle number was minimal for PIN_PON and PIN_ARI, while Mixed trees had 1.8-5.4 times the lntrannual variation as the other morphotypes (Figure 2-5). More needles per fascicle were produced by Mixed trees in 2003 than either in 1999-2002 or 2004- 2006 (two-sample t—test, p<0.01), and more needles per fascicle were produced in 2003-2006 than in 1999-2002 (two—sample t-test, p<0.01). 180 a 160 c 140 a 120 < 100 - 80* Frequency 603 20‘ | .. Jinn . I ws» #01 .0) 9’. 95109901599999? 4 o: oaxroocoo—smutsurm l- .“ PJ‘PS’ISPP'SHPI N «cacao-awe); 63 .w 0 Mean needle number per fascicle Figure 2-3. Frequency of trees (n=671) by mean needle number per fascicle. Data are from the distribution study (“SCAT”). 47 0 g o °° . :E: °° co co 9 8 ‘ 08« 8 eo°o Q. “0%; 2 ' “game. 0 6% g : 8 00 0 ; 2 35 ° ‘ s 8 1 C N '5 > 0.4— 0 0 r: E; APIN_PON oMixed oPlN_ARl 0 + . i i e i 2.5 3.0 3.5 4.0 4.5 5.0 5.5 Mean number of needles per fasclcle Figure 2-4. Variability in needle number within a tree as a function of its mean needle number. Hashed lines indicate breaks between one morphotype and another (Peloquin 1984). Forty-one PIN_PON and 35 PIN_ARl have no intratree variance. Data are from the distribution study (“SCAT") 6 - r 400 " 350 ‘t 300 ~ fig i o E E j I: , . a ‘ j . ‘ ‘ ‘ eStudy area 1 2 IP|N_PON I 6 4 .2 o 2 4 Mean monthly mlnlmum January temperature (’C) ‘33 a . 5 e 5 3 1 E g I z- n i i 2 g i I o ‘ I E b i 3 t i C . 5 4 I l I 4 eStudy area i 2, lMlxed a 4 -2 o 2 4 6 Mean monthly mlnlmum January temperature ('0) T 7T" * “rib ‘W7# _—"‘1 l l . I ‘ . 1 . l E: 3 1 .§ 2 ' '3 x l “g : i E. : i _ g, i I E; I I 0 i E 2‘ = 3 g 5 . " l . i eStudy area I ; 2 ePlN_AR| ‘ 6 4 2 o 2 4 a Mean monthly mlnlmum January temperature ('C) Figure 2-8. Distribution of the three morphotypes in climatic space with regard to mean monthly minimum and maximum January temperatures. Data are from the distribution study (“SCAT”). 52 and maximum January temperatures where Mixed and PIN_ARI were found were —0.6 and 10.4 °C, and -0.6 and 10.3 °C, respectively. If the anomalous 3- needled trees are removed, then the warmest minimum and maximum January temperature where PIN_PON were found were —2.0 and 9.5 °C, respectively; otherwise, these temperatures are -0.8 and 12.3 °C, respectively. For summer climate space, the three morphotypes were found in the coolest and wetters portions of the study region (Figure 2-9). Excluding the anomalous trees, PIN_PON was located in areas with no warmer maximum June temperature than 26.6 °C, while Mixed and PIN_ARl were found in areas as warm as 28.6 °C. Although overlapping in the wettest areas of the study region, Mixed and PIN_ARI were found in drier areas (down to 89.2 mm of July precipitation) than PIN_PON (down to 96.0 mm of July precipitation). Including the anomalous PIN_PON trees in these bivariate plots suggests that this taxon would occupy a larger climatic niche space than the other taxa. 53 Mean monthly July preclpltatlon (mm) a PIN_PON 40 20 25 :n :5 Mean monthly maximum June temperature ('0) $10 . 2 V E c : o 5 ; 3 S n _>~ . 3. so I _,, 1 5 C i ° i 5.. i g 0 Study area 3 a Mixed 40 20 25 :0 35 43 Mean monthly maxlmum June temperature ('C) 1‘0 A ‘ — “ ”i ’ — “i i 120 g 0 C ‘ . o , 3 i e m i g , 2* 5'. so _>« 5 C e E i 9 Study area 5 - PIN_ARI 40 20 30 40 25 35 Mean monthly maxlmum June temperature ('C) Figure 2-9. Distribution of the three morphotypes in climatic space with regard to mean monthly maximum June temperature and monthly July precipitation. Data are from the distribution study (“SCAT") 54 Model comparison The Maxent modeling algorithm does not appear to be sensitive to the spatial extent of calibration (“training”) data (Figure 2-10). The rasters for environmental explanatory variables of maximum June temperature and annual precipitation were clipped in either rectangular or irregular polygon shapes around the Santa Catalina Mountains, in both cases excluding much of the lower elevation desert and foothills from the modeled area. Models were generated for each polygon shape using the PIN_PON data, excluding duplicate occurrences in the same pixel, from the distribution study (“SCAT”). Based on both threshold— dependent and -independent statistics, models generated from the rectangular- shaped environmental data performed better than the polygon-shaped data (Table 2-3). Consequently, all further models were based on environmental data conforming to the former shape. Table 2-3. Comparison of influence of spatial extent by shape of calibration area on model output. Species data are PIN_PON from SCAT; environmental data are maximum June temperature and annual precipitation. Threshold-dependent statistics are based on the lowest presence threshold as described in the Methods: cumulative threshold (“cum.thr"), fractional predicted area (“fr.prd.ar"), and binomial probability (“P- value"). The threshold-independent statistics are based on the area under the receiver operating (ROC) curve (AUC) (see Methods). Lowest presence threshold AUC from test data Shape Cum.thr Fr.prd.ar P-value N AUC Std dev Polygon 3.689 0.198 4.589E-07 9 0.974 0.009 Rectangle 3.656 0.140 3.036E-08 9 0.982 0.006 55 N [:10-10 _10-25 _25-50 _50-100 Figure 2-10. Distribution of cumulative probability of suitable habitat for PIN_PON relative to maximum June temperature and annual precipitation in two spatial extents (“polygon" and “rectangle") for Maxent model calibration. Data are from the distribution study (“SCAT") 56 Fourteen models using the Maxent algorithm were generated for comparison for each morphotype, with all but one model being highly statistically significant (Table 2-4). Incorporating the occurrence data from finer spatial scaled studies with the distribution data (“SCAT”) for the modeling often increased the significance of the model but generally decreased the AUC, or ability of the model to maximize both sensitivity and specificity, probably due to increasing the overlap of taxa in the same environmental pixel across steep transition zones (e.g., south slope of Mount Lemmon, Epperson et al. (2001)). Removing duplicate occurrences within the same environmental pixel, or spatial pseudoreplication, generally decreased the model’s ability to fit the data but also decreases the likelihood of model overfitting. Understandably, the models with the highest AUC were generated with the topographic data at 10-m resolution, in comparison to the 800-m resolution for the environmental data. 57 tnwmwd owz mood oood or word Rod am .Eow owz god Sod ov oo+mood nod oowd am :4. ...:22 oood oood o mom—odd mid mood momma: > ._. h ._. h __< ...ocz22 wood wood do oo+mood Ford mom; 5 z __< 2.58 + ...9522 oood oood o oomood mid mood E:w.:_>> > beam 82% + 232% + #232 «Se 83 S 8.93 v3.0 83 same? > __< 222% + :92: + ...9522 oFod onod o ooflmwd «to wood __P. > ._. :4. ...2322 + ....922 + ._.>m_2_2 +5522 mood mood do oo+mood word rovd __P. z __< zOdIzE mozmtg E3395 >oo Em 02< : magi and: .583 Emu Sand Son 36on Emu «moo Eco 02< 202%...“ 8535 “$26.. .c2_>cw o>oEom $6on dos—5:36 $3305 .2: o5 8 mEo_oEooo “modem. 9.: 9:35:00 805 2m moi—mtg Eaton—E .o-~ 03m... c_ 89:32. 9m 83¢.sz “conceded? new Eon:oaou.v.ocwoz._. .TN 258. E contowoo 2m Emu _ScoEcescw do>oE2 .o 3255 556 m_ .35 _EcoEcESco cm 552, «85.580 “6 538.39 .22me .A..._.25 05 £531.22.» $82.38 3289 __m 55%.. $205 Emu 863w ..x2 9: c_ nontomou Em $89. $603 .569. 52m Edocfioo. 05 .2 £982 3 doomeocoo 280E cooofioa ..8 woomoflm ocm woomtofimcmco oocmEoton. .YN can... 58 omz oood Eod on Nomooo word onto am #35 owz vood onod om oo+wood oNFd rode am =4. #522 oood oood «N -.w#w.v oood Food momma: > # #(0w 5522 + #522 oood oood rm om.moo.o mood one 52:52 > # # ..< 5522 + #522 oood oood o3 oo+mood ood «3+ 5 z __< 8%.: + #8:: + #9522 wood wood a 2-9:.» 85 35 55.52, > Zoo 525 + #52: + 5522 #ood oood Nu mwmood oood vod .5053 > =< #nouE + 5.225 + #892 + #522 oood oood Pu --w§.~ vood nod =3... > # __< #52: + #592 +#.>.oz_2 + #522 Wood ooowo oi :ob+moo.o «to Emd __P. z __< _m # .2 #522 mood onod or #FmoNN oood oofim momma: > # # .525 5522 + #522 mood mnod 2. 2m id oood oood 2.. > #<0w 5522 + #53 vood Koo :V oo+mood mord orme 5 2 #(0m #52: + #522 + 5522 oood wood 2 3-5: oood Ed 5 > __< 5522 +#522 oood mood vo oo+wood to Node 5 z =< #nouE + 55% + 525 + 5522 oood o#o.o or oTwoNd 9 rd wood E:w.:_>> > # __< #52:+ #502 +5223 + #522 oood wood or otmooé mood ow: =3“. > # .2 #522 + +522 + 5.225 + #522 oood oood vo oo+wood Edd 5o... :3... z __< 55.2 8359 2559:. 5o Em 02< : 029.1 5.23: .5250 Emu 396 San $6on Son «mod .5: 02< . 205.25 8535 2526.. .555 96,51 36on «8:550» .5 22¢ 59 Excluding climate-based models with spatial pseudoreplication, the simplest models in terms of number of variables were the most significant and had highest AUC values (Table 2-4). When available in the suite of explanatory variables, maximum June and January temperatures had the highest coefficients and individually contributed most to the models. As the number of variables were reduced, maximum June temperature became the most important for PIN_PON, while maximum January temperature became the most important for Mixed and PIN_ARI. Monsoonal precipitation was consistently less important than maximum June and winter temperatures in predicting suitable habitat for all three morphotypes. Although statistically very close, models using only maximum January and June (or July) temperatures had highest success in correctly predicting the test data. Two models from each morphotype were more closely analyzed for differences in predicted suitable habitat, or niche differentiation, between the taxa. The Lit model, whose output in Figure 2-11 is shown with large class breaks to highlight coarse patterns, demonstrated marked differences between morphotypes. Suitable habitat (>50% cumulative probability) for PIN_PON was restricted to 1728 ha in the two high-elevation regions within the Santa Catalina Mountains: Mount Lemmon and Mount Bigelow/Kellogg complex. Suitable habitat for PIN_ARI included 3456 ha and was located on the west and south slopes of Mount Lemmon and Mount Bigelow/Kellogg complex, and southeast of Palisade Ranger Station. Suitable habitat for Mixed included 3008 ha, overlapped that for PIN_PON, and overlapped the southeastern portion of 60 . PIN_PON Mixed N [:1 0-10 10-25 J. - 25 - 50 - 50 - 100 PIN_ARI Figure 2-11. Distribution of cumulative probability of suitable habitat for the three morphotypes in the region for the Lit model by Maxent. Data are from the distribution study (“SCAT"). 61 PIN_ARl’s predicted distribution. Area estimates for finer class breaks are shown in Table 2-5. Table 2-5. Climatically suitable area estimates for the three morphotypes based on cumulative probabilities from the Lit and Lit.Thresh models. 50-75% 75-100% Morphotype Model Hectares Acres Hectares Acres PIN_PON Lit 1 152 2847 576 1423 Lit.Thresh 0 0 21 12 5219 Mixed Lit 1728 4270 1280 3163 Lit.Thresh 2304 5693 2368 5851 PIN_ARl Lit 2112 5219 1344 3321 Lit.Thresh 2816 6958 3008 7433 The second set of models used the same data but were limited to the threshold (“Lit.Thresh”) or binary mathematical function in the Maxent algorithm to identify climatic thresholds above or below which suitable habitat would not be predicted. These models had lower significance and/or similar AUC than the Lit models. The spatial extent of suitable habitat was approximately the same for PIN_PON but larger for Mixed and PIN_ARI as compared to using the full suite of mathematical functions in the Lit model, but the overlap of the other taxa’s niche space by Mixed is much more apparent in this set of models (Figure 212). Although there was little differentiation in threshold climatic values between PIN_ARI and Mixed, the latter and PIN_PON are clearly separated by minimum January temperature (A=8.1 °C) and possibly even July precipitation (A=1.6 mm) (Table 2-6). 62 . PIN_PON Mixed N [3040 _10-25 1‘ _25-50 1 - 50 - 100 PIN_ARI Figure 2—12. Distribution of cumulative probability of suitable habitat for the three morphotypes in the region for the Lit.Threshold model by Maxent. Data are from the distribution study ("SCAT"). 63 Table 2-6. Thresholds for the three morphotypes based on the Lit.Thresh model. No coefficients were generated for minimum January temperature for the PIN_ARI model. Data are from the distribution study (“SCAT”). Minimum January Maximum January Maximum June July precipitation Morphotype temperature (°C) temperature (°C) temperature (°C) (mm) PIN_PON -20.1 10.6 26.6 11.0 Mixed -12.0 10.2 28.3 9.4 PIN_ARI n/a 10.2 28.1 9.3 Model prediction Based on the ability of one species’ model to correctly predict occurrences of each of the taxa, there is clear ecological differentiation between PIN_PON and PIN_ARI, but Mixed appears to be an ecological intermediate (Table 2-7). In all cases, the Lit.Thresh models were more likely to predict occurrences than were the Lit models. Although the Mixed and PIN_ARl models predicted their own species’ occurrences more than the other’s species, the PIN_PON models were more likely to predict Mixed than PIN_PON occurrences. All of the models had similar ability to predict PIN_ARI occurrences. Table 2-7. Ecological similarity matrix based on the ability of one taxon’s model (columns) to predict the documented occurrences of each of the taxa (rows). Values are mean probabilities of pixels wherein an occurrence was documented for the given model; duplicate locations by species are excluded. Data are from the distribution study (“SCAT”). PIN_PON Mixed PIN_ARl Model Lit Lit.thresh Lit Lit.thresh Lit Lit.thresh PIN_PON 0.64 0.80 0.73 0.86 0.53 0.73 Mixed 0.50 0.64 0.62 0.75 0.53 0.71 PIN_ARl 0.33 0.51 0.49 0.63 0.59 0.70 Discussion The data and analyses presented in this paper give biogeographic and ecological evidence for possible hybridization between the 3-needled Southwestern Rocky Mountain ponderosa pine (PIN_PON) and the 5-needled Arizona pine (PIN_ARI). Close to 90% of the trees sampled in the distribution study produce fascicles intermediate to pure (sz=0.0) PIN_PON and PIN_ARl. While PIN_PON generally occupied a cooler and moister climatic niche, the same bivariate plots showed mixed-needle (Mixed) trees overlapping PIN_ARI. The Maxent models demonstrated the limiting influence of winter temperature on Mixed and PIN_ARI, while PIN_PON was more limited by temperature during the arid foresummer. Furthermore, PIN_PON appears to tolerate areas with an 8° C colder minimum January temperature than the Mixed trees. However, because of the broader distribution for the Mixed trees, suitable habitat predicted for this taxon overlapped the other species. The interpretation for these results must be tempered by a number of factors, including quality of data entering the models, model assumptions and selection, and ecophysiological considerations. Data quality The original intent for collecting species occurrence data was to characterize the distribution for the three taxa, and thus a particular subjective sampling strategy was employed. It was understood that their distributions varied across elevation (Dodge 1963, Epperson et al. 2001), but that elevation was not the sole factor determining location for each taxon (T elewski 65 unpublished data). For example, in March 2000, Telewski (unpublished data) located a tree producing mean 3.02 needles per fascicle at low elevation (~1800 m) near Bear Canyon, while the rest of the trees were mixed and 5-needled. Within my distribution study, I located 8 (“anomalous”) trees that also unexpectedly produced less than mean 3.2 needles per fascicle: Romero Pass, Catalina Camp, and near Lizard Rock. Interestingly, the climate at these lower elevation sites could be moderated by cold-air drainage patterns (Adams 2007): Mule Ears is adjacent to a northerly exposed pass through Samaniego Ridge; Romero Pass is located at the origin to a large canyon draining northwest to Romero Canyon and southeast to Upper Sabino Canyon; Catalina Camp is situated in the valley between two prominent ridges extending north from the main Santa Catalina ridgeline; and the Lizard Rock site is situated in a low-lying pocket with a substantial watercourse. These disjunct PIN_PON trees could simply be remnants of a once larger population at these elevations, surviving but lacking a suitable regeneration niche. The Maxent models included these and potentially other disjunct low—elevation trees, thus interpretations of the observed differences between distributions of the taxa are strengthened. In retrospect, the sampling strategy did not prove to be efficient for characterizing and modeling the distribution of the taxa. All of the trees to be sampled were selected by various, usually subjective, means. For example, at Palisade Rock, PIN_PON and PIN_ARl were selected as pairs in close proximity for ecophysiological studies. Along the south and north face of Mount Lemmon, as well as for Kimball Peak Saddle, Bear Canyon, and Wilderness of Rocks, 66 trees were selected either for accessibility, reproductive status, or systematically for spatial analyses (Epperson et al. 2001, 2003). Each of these trees had a unique set of geographic coordinates. In an effort to maximize efficiency in species data collection, up to four trees per point-location were sampled for the distribution study. These points were located within centers and expected edges (Telewski unpublished data) of each taxon’s distribution. At each predetermined point-location, four trees from a point-centered quarter could be represented by one to three taxa, which results in spatial pseudoreplication for this point-location. This error is compounded by the fact that multiple sampling points were located within a given pixel of environmental data, thus removing such duplication due to relatively low resolution of the explanatory variables resulted in drastic reduction of the sample sizes for each taxon in the modeling. For example, including duplicate samples in the modeling consistently increased the predictive performance of the models and likely produced overfitted models. Furthermore, I believe that performance statistics were consistently lower for models using all of the occurrence data (SCAT and non-SCAT), even with larger sample sizes, as compared to just the distribution study data (Table 2-4) because the former includes data from the very steep transition zones of PIN_PON to PIN_ARI on Mount Lemmon (Epperson et al. 2001) and Palisade Rock (Kilgore unpublished data), yet each case occurs within a single environmental pixel. The resultant smaller sample sizes, even with 30% of the occurrences allocated to model verification, are not expected to have had a significant effect on the model output (Stockwell and 67 Peterson 2002, Engler et al. 2004, and Pearson et al. 2007). Overall, the distribution study data were more dispersed, thus allowing better discrimination between their distributions across the field of environmental data. Additional bias in the sampling strategies was introduced clue to geography and two major disturbance events. The Santa Catalina Mountains are treacherously rugged but have a major paved road to the summit as well as extensive hiking trails. All of the sampled trees were close to primary and secondary roads or hiking trails. With the exception of the Front Range and the north slope of the Mount Bigelow-Kellogg ridge, most of the potential habitat for the three taxa was sampled, relative to the dimensions of the environmental pixels. However, the Bullock Fire (2002) extensively burned the forests on the north slope, while the Aspen Fire (2003) burned through most of the remainder of the forests in the Santa Catalina Mountains, except for the extreme southwest side of the Front Range (CNF 2002, 2003). Consequently, in most areas (especially Samaniego Ridge, north slope of Mount Lemmon, Wilderness of Rocks, Box Camp Ridge, and north of Mount Bigelow), live or dead trees with accessible needles were sometimes difficult or impossible to locate. Distribution data for some of these areas are functionally lost but can be inferred from the habitat suitability models. The parametric assumption of randomly sampling the entire study region for distribution modeling is unrealistic and unfeasible. Hirzel and Guisan (2002) found that random sampling was the least accurate and robust method as compared to regular and environmentally-informed systematic sampling in 68 correctly predicting presences of a virtual species. Instead, Araujo and Guisan (2006) argue that a model-based environmental stratification approach be used to direct additional sampling, assuming that the abiotic range of the taxon has been preliminarily sampled. Increasing sampling-effort at the extremes of a range, or tails of the unimodal species response curve, increases the accuracy in estimating species distributions (Mohler 1983). Fortuitously, this was the approach taken for the distribution study, given the limitations of accessibility and living forest. In the case of the Santa Catalina Mountains, the model outputs suggest that sampling effort be focused on the Front Range where suitable habitat for all three morphotypes is predicted (Figures 2-11 and 2-12). Although the Maxent model output provided additional support to the literature for the limiting influence of summer drought and winter cold on the distribution of ponderosa pine, the relatively large resolution of the environmental data restricted the full use of species data, as explained earlier. As expected, models using the 10-m resolution topographic data produced much finer scaled predictions (Figure 2-13), but a species’ association to elevation in the Santa Catalina Mountains is not transferable outside this system (Phillips et al. 2006). Instead, climate constrains the macroscale distribution of species (Woodward 1987), reducing historical contingencies, and associations with climate can be projected to another system. Furthermore, the objective here is to determine if any difference exists between species in their response to limiting climatic factors. 69 PIN_PON Mixed N C] 0 - 10 - 10-25 - 25 - 50 h’ - 50 - 100 PIN_ARI Figure 2-13. Distribution of cumulative probability of suitable habitat for the three morphotypes in the region for the NED.Asp model by Maxent. Data are from the distribution study (“SCAT”). 70 Unfortunately, the highest resolution climatic data available for the study region is 30—arcsec, or 800-m, resolution (PRISM 2006). These continuous data were generated by interpolating climatic data from monitoring stations using an algorithm developed for mountainous terrain: Parameter-elevation Regressions on Independent Slopes Model (PRISM 2006). While taking into account spatial scale and pattern of orographic processes using a digital elevation model (PRISM 2006), which should detect topographically depressed drainageways, the climatic data are scaled up to BOO-m resolution and thereby lose much of that information. Also, the accuracy of the resultant gridded climatic data is limited to the actual data collected. In the Santa Catalina Mountains, an on-Iine clearinghouse (Western Regional Climate Center, URL: www.wrcc.dri.edu/summary/Climsmaz.html) lists climate data for only two locations: Palisade Ranger Station (2426 m elevation) for 1965—1981; and Mount Lemmon (2344-2374 m elevation) for 1950-1956 and 1958-1996. Additional “unofficial” data from Mount Lemmon Ski Valley (A. Corona, National Oceanic and Atmospheric Administration, Tucson, personal communication) and Green Mountain (W. Hart, CNF, personal communication) exist but were not likely used by PRISM (2006). Consequently, the accuracy of the explanatory variables used in the modeling is determined by a combination of interpolation of limited climatic data and orographic extrapolation in the PRISM (2006) algorithm. 71 Model assumptions and selection A fundamental assumption of species distribution models (SDMs) is that they accurately approximate the species’ ecological niche from known occurrences and their association with particular values of climatic variables. In general, ecological niches are highly conserved, at least since the last glacial maximum (Peterson et al. 1999, Martinez-Meyer and Peterson 2006), so projecting the models to other locations should be informative. The output to such correlative models is a map of potential suitable habitat and, in fact, not potential geographic distribution because the models do not take into account spatially explicit features such as dispersal (Araujo and Guisan 2006). Other factors such as competition, historical contingency, and human fragmentation of populations and habitat can also alter the potential geographic distribution for a species. For example, Brown et al. (1996) suggest that species are limited along a key environmental gradient by abiotic stresses at one extreme and biotic stresses at the other extreme. Vetaas (2002) supported this theory with cold temperatures as an absolute boundary, while a mixture of competition and warm temperature forms the other boundary in Himalayan Rhododendron species. However, the Maxent models suggested that primarily warm temperatures during a particularly dry period (arid foresummer) for PIN_PON and cold temperatures during winter for PIN_ARI explain a sufficient amount of the variability in each taxon’s distribution to geographically segregate their distribution based on potential suitable habitat. Consequently, competition between these taxa may occur but does not need to be included in the models to differentiate their 72 ecological niches. Like PIN_ARI, the upper elevation distribution of the Mixed trees appears to be limited by winter temperatures, but its distribution overlaps both taxa. If the Mixed trees represent hybridization between PIN_PON and PIN_ARI (Peloquin 1984) and if hybrid trees possess “new adaptive systems” as suggested by Anderson and Stebbins (1954), then the Mixed trees in this system may have absorbed adaptive features such as greater cold tolerance from PIN_PON and higher heat tolerance from PIN_ARI. Competition between for limited resources during establishment of these sympatric taxa and their offspring is therefore expected. Input data can be processed to remove areas where a species cannot occur for one reason or another despite climatic suitability, but, in general, Maxent models approximate Hutchinson’s (1957) realized niche on a coarse climatic scale (Phillips et al. 2006). As a consequence, Maxent and other SDMs will predict suitable habitat larger than a species’ actual distribution (Phillips et al. 2006). Species are often absent from suitable habitat, and vice versa (Pulliam 2000). However, predicting larger potential ranges proves useful in targeting future sampling efforts and identifying potential restoration locations. A second fundamental assumption of SDMs is that modeled populations or species are in equilibrium with their environment (Guisan and Thuiller 2005). Alternatively, modelers are interested in how distant from equilibrium are current distributions (Araujo and Pearson 2005). Natural dispersal ability, landscape barriers, human fragmentation of habitat, and species interactions are factors that can restrict the ability of a species to track climate change (Davis et al. 1998, 73 Pearson and Dawson 2003). For example, Leathwick (1998) found that the distribution of disjunct Nothofagus species in New Zealand corresponds more to suboptimal habitat due to slow dispersal ability. Due to producing wind-borne pollen and winged seeds, ponderosa pine can disperse relatively long distances (Conkle and Critchfield 1988, Epperson et al. 2001). However, trees greater than 1.5 m in height were selected for the distribution study, which may bias the climatic interpretation of the model output toward conditions present at the time of their establishment rather than current conditions. Ponderosa pine are long- lived trees; for example, most of the trees cored on Green Mountain were around 300 years old, but one dated back to 1460 (Graybill 1986). On the other hand, dendrochronologists usually seek out the oldest trees for extending chronologies, while the nearest tree to the point-center was selected for this study. The trees sampled for the distribution study (n=671, mean diameter at breast height = 34.4 cm, range = 0.9-399.0 cm, standard deviation = 22.4 cm) more likely represent climatic conditions from the last 20-250 years. Given the 2° C rise in mean temperature over the last century in this region (USEPA 1998), which translates to an elevational shift of 267 m (Shreve 1915), focusing on presence and taxonomy of young trees would bring the model predictions closer to equilibrium with current climatic conditions. The Santa Catalina Mountains present numerous physical challenges to surmount for non-vagile organisms to migrate in response to climate change. The steep canyon walls, xeric southern exposures, and shallow lithosols reduce the ability for seeds to disperse and establish higher in elevation. Conversely, 74 the deeply etched valleys that bear floodwaters from higher elevations provide mesic refugia, sometimes from fire as well, for trees left behind in the march upward. Since larger surviving individuals produce a disproportionately large number of pistillate cones, these disjunct trees receive pollen from the invading lower elevation pines which produce pollen in all mature age classes (Conkle and Critchfield 1988). Assuming vigorous offspring, the stage for hybridization is prepared in part due to the landscape structure, whose complex topography allows for disequilibrium in distribution at the landscape scale. Nevertheless, at the available spatial resolution, certain climatic variables consistently gained larger coefficients than other variables across the variety of models for each species. Surprisingly, precipitation variables were always at least secondary in importance to either summer (all taxa) or winter (PIN_ARI and Mixed) temperature. Use and examination of the most parsimonious models (Lit and Lit.Thresh) supported by physiological tolerances indicated in the literature appears to be appropriate for these taxa. Winter temperature was likely not limiting to PIN_PON because little of the study region existed above the upper elevation distribution for this taxon; of the Southwestern mountain islands, only the Pinalenos and Chiricahua Mountains have substantial habitat above the ponderosa pine zone. Because of the bimodal distribution of precipitation in the Southwest, early summer (June) is the most xeric period of the growing season. Spring-germinating seeds are thus quickly exposed to drought conditions, an interaction of high temperature and low soil moisture. Consequently, cold winter temperatures constrain upward migration of Mixed and PIN_ARI and warm 75 temperatures prevent establishment of PIN_PON at its lower elevation distribution. These hypotheses could easily be tested by establishing Climatically monitored common gardens at different elevations (sensu Clausen et al. 1940). Exclusive use of the binary, or threshold, mathematical function (Lit.Thresh model) for detecting differences in cutoffs in species distribution appears to be unique in this study. Physiological tolerance for variation in a climatic variable is not expected to be binary, in fact the response is typically close to unimodal, but forcing the coefficients in the Maxent algorithm into binary mode seems to have allowed quantification of ecological niche differentiation between taxa with overlapping distributions. In this study, suitable (or at least occupied) habitat differed between PIN_PON and the other taxa by -8.1° C in January, -1.5° C in June, and 1.6 mm precipitation in July (Table 2-6), implying that cold temperatures in winter restrict PIN_ARI and Mixed from PIN_PON habitat, and that the warm arid foresummer restricts PIN_PON from growing in PIN_ARI and Mixed habitat. These conclusions are supported by the other models, but now a quantified difference can be related to other studies examining ecophysiological differences between the taxa and incorporated into models examining the effects of climate change on species distribution. Distribution and ecological differentiation The spatial distribution of occurrences (Figure 2-6), climatic niche space (Figures 2-8 and 2-9), and predicted suitable habitat maps (Figures 2-11 through 2-13) are all based on a single morphologic character (Peloquin 1984) with high 76 heritability (~0.6, Rehfeldt et al. 1996): mean needles per fascicle. In addition, elevation (Dodge 1963) and previous year’s July-August precipitation (Haller 1965) are correlated to the number of needles per fascicle in trees producing variable numbers of needles. Despite these confounding factors, the separation of distribution and ecological niche space is apparent for those trees classified as PIN_PON and PIN_ARI, even when the intratree variation in mean needle number (Figure 2-4) is relatively high. Ecophysiologic results support the ecological differentiation between PIN_PON and PIN_ARI. Dodge (1963) monitored radial expansion and reported tree-ring characteristics from sympatric PIN_PON and PIN_ARI trees near Mount Bigelow. Although both taxa ceased radial expansion simultaneously, PIN_ARI lagged initiation by 6 weeks yet produced significantly larger ring-widths over time. Furthermore, PIN_PON chronologies had higher autocorrelation (0.65 versus 0.24), and PIN_PON produced twice as many intra-annual rings, which was interpreted to be a genetic expression of adaptation to more consistent precipitation patterns during the growing season typical of the Rocky Mountains (Dodge 1963) or greater sensitivity to drought at the lower edge of its distribution (Fritts 1974). Besides needle number, Peloquin (1971) also noted significant differences between PIN_PON and PIN_ARl in number of resin canals in the needles and monoterpene ratios from sap. Morphologic, phenologic, and growth responses are clearly differentiated between PIN_PON and PIN_ARI. The biogeographic and phylogenetic quandary are the trees classified as Mixed, or those producing 3.2 to 4.6 needles per fascicle (Peloquin 1984). 77 Dodge (1963) found a needle cline, implying mixed-needle trees, in the Santa Catalina, Rincon, Chiricahua, Huachuca, and Santa Rita Mountains, but not in the nearby Galiuro or Pinaleno Mountains, which had only 3—needled trees. Based on community associations, Lowe (1961) describes the more southern mountains (Chiricahua, Huachuca, and Santa Rita Mountains) as outlying Sierra Madrean systems, while the other more northern mountains are extensions of the Rocky Mountains. Based on the distribution of needle and cone morphologies, Dodge (1963) concluded that the Santa Catalina and Rincon Mountains express the junction of the Rocky Mountain — Sierra Madrean flora more than the other mountains. Consequently, these mountains will have the highest likelihood for populations of Southwestern ponderosa and Arizona pines to interact and produce the needle number cllne (Dodge 1963). From monoterpenoid analysis, as well as needle and cone morphologies, Peloquin (1971) concluded that trees producing mixed-needled fascicles, even those sampled below Mount Bigelow, were F1 and advanced hybrids between Southwestern ponderosa and Arizona pines. Dodge (1963) reported that dendrochronological characteristics for Mixed trees at the Mount Bigelow site were intermediate but more similar to PIN_PON than to PIN_ARI. F urtherrnore, Conkle and Critchfield (1988) point out the moderate crossability of Ponderosae, in general, and especially across species that shared glacial refugia. Supporting Peloquin’s (1971) results with a common garden approach, Rehfeldt (1993) concluded that introgression between the Ponderosae in southern Arizona is common. However, Rehfeldt partially recanted with the development of a novel Taxon X (Rehfeldt et al. 1996) and 78 ascribed the geographic cline in needle number to within population genetic variability (Rehfeldt 1999). In the Santa Catalina Mountains, spatial structuring of needle number could indicate hybridization (Epperson et al. 2001), while the lack of allozyme structure and differentiation support either advanced introgression or insufficient time since isolation for differentiation to have occurred (Epperson et al. 2003). Unpublished chNA data (Epperson et al.) indicate either unidirectional cytoplasmic hybridization (Watano et al. 1995) between Arizona and Southwestern ponderosa pines or genetic barriers to unlimited pollen flow between Arizona pine and Taxon X. The hybridization hypothesis is supported by the results of this study. Despite arbitrary divisions between continuous needle numbers, the spatial extent of predicted suitable habitat for Mixed trees was similar at high elevation but visibly extended lower in elevation than suitable habitat for PIN_PON (Figures 2-11 through 213). Although Rehfeldt (1999) aligned the Mixed trees with PIN_PON, they in fact appear to be more climatically associated with PIN_ARI than PIN_PON (Table 2-6). Therefore, PIN_PON and Mixed do not share climatic tolerances, despite their sympatry. Given the results of prior work with these taxa (Dodge 1963, Peloquin 1971, Conkle and Critchfield 1988, Rehfeldt 1993, and Epperson et al. 2001), as well as similar patterns of hybridization in other Pinus (e.g., Watano et al. 1996), and the biogeographic evidence provided in this study, I deduce that the Mixed trees represent hybridization between PIN_PON and PIN_ARI (sensu Anderson and Stebbins 1954). Modeling the spatial variability of characteristics distinguishing PIN_PON 79 and PIN_ARI, such as monoterpenoid ratios and genetic markers (e.g., cp/mtlnuclear DNA), would lead to greater understanding of the historical progression of migration and putative introgression of the Ponderosae in the Santa Catalina Mountains. Projection A natural extension of SDMs is to spatially or temporally project the habitat, or climate, suitability model derived for each species (Guisan and Thuiller 2005, Hijmans and Graham 2006). Spatial projection to other mountain islands in the Southwest would identify areas to sample for not only validating the model but also conservation purposes. Although the Galiuro and Pinaleno Mountains are close to the Santa Catalina Mountains, Dodge (1963) did not find any PIN_ARI despite seemingly suitable habitat. Projection of the models derived in the Santa Catalina Mountains to these other systems can identify climatically limiting conditions and contribute to the understanding of biogeographic patterns. Conversely, such projection would highlight the importance of other factors, such as historical contingency in the lack of migration during the last glacial maximum to these systems from the Sierra Madrean PIN_ARI. Projecting to a geographically proximal and climatically similar but independent location could improve the model through validation and subsequent fine-tuning of variable selection and parameterization. This process could also be used to assess the ability of the model to project to different climate scenarios (Guisan and Thuiller 2005). In both cases, biotic interactions may differ and thus the projected 80 “realized niche” may vary depending on the parameterization of the original model. Since these models do not currently incorporate spatially explicit processes like dispersal, their output is potential suitable habitat, not potential geographic distribution. However, in comparison to a mechanistic model and other niche modeling algorithms, Maxent consistently performs well in predicting current, past, and future ranges in response to climate change (Hijmans and Graham 2006). Such an approach would be useful in predicting the potential response by taxa isolated on mountain islands to climate change, as well as identifying potential habitat for climatically induced plant migration. 81 CHAPTER 3 INFLUENCE OF SEED GERMINATION ECOLOGY ON THE DISTRIBUTION OF PONDEROSAE IN THE SANTA CATALINA MOUNTAINS OF ARIZONA Introduction The realized distribution for any species is determined by the establishment of seedlings, thus germination and survival through establishment are critical life history stages (Bazzaz 1979). Steep environmental gradients, such as those found in montane systems, present particular challenges for seeds dispersed beyond a relatively narrow bioclimatic window. At higher elevation, seeds are faced with more mesic conditions, such as colder air and soil temperatures, higher litter depth, higher soil moisture, and lower light due to shading from the more productive canopy (Whittaker and Niering 1975, Barton 1993). Seeds dispersed below the current range are faced with more xeric conditions than those of its seed-producing parent. In addition, biotic interactions, such as competition from shallow-rooted forbs. and disturbance regime (e.g., fire frequency) different from a seed-tree’s locally adapted habitat could affect eventual establishment. On the contrary, dispersal and establishment outside of a species’ current distribution could suggest that the species is not in equilibrium with its environment. However, due to restricted species niches in mountains, potential species migration here is sensitive to variations and change in climate (Tranquillini 1979), and thus the germination ecology for a species will influence the outcome of potential species migration. 82 Most montane tree seeds exhibit some sort of physiological dormancy, likely selected to avoid germination at a time with unfavorable environmental conditions for seedling establishment (Baskin and Baskin 1998). This dormancy is normally manifested as a chilling requirement, or cold stratification, before seed germination occurs. In some cases, such as the Carices (Schutze and Rave 1999), germination occurs regardless of stratification but is higher with colder stratification. Given that most seeds mature and disperse by the end of a growing season (Young and Young 1992, Baskin and Baskin 1998), a chilling requirement in cold, or higher elevation, conditions should increase the likelihood of seedling establishment in the subsequent growing season(s). Even after a year’s growth, though, disruption of the delicate root system of the seedling through frost-heaving can lead to high mortality (Heidmann 1987). The timing of germination during the growing season relative to local climate conditions, as well as irregular disturbances such as fire or logging, can also influence germination and establishment. For example, Larson (1961, as referenced by Howard 2003b) noted that Southwestern ponderosa pine seedling mortality is highest in regions with a bimodal precipitation pattern, which is characterized by winter precipitation, an arid foresummer drought (May-June), summer monsoon precipitation (July-August), and a moderate fall drought (Sheppard et al. 2002). Barton (1993) showed that the early summer drought limits pine seedling establishment lower in elevation, as higher elevation species germinated but did not survive the first year. In contrast, lower elevation pines germinated and survived for at least two years above their distributional limit 83 (Barton 1993). The fall drought also affects ponderosa pine seedling mortality (Jones 1967, Larson and Schubert 1970, cited in Howard 2003b). In Arizona, Schmid and Mitchell (1986, cited in Howard 2003b) measured highest first-year survivorship (13-19%) by seedlings that germinated during the early part of the monsoon, with lower (2-7%) survivorship by those germinating later in August. Like the climatic niche, the temporal window for germination and successful establishment confines opportunities for species migration. The closely related Ponderosae are roughly segregated, but sympatric in places, by elevation and factors contributing to more or less xeric conditions in the Santa Catalina Mountains of southern Arizona. Southwestern ponderosa pine (Pinus ponderosa var. scopulorum Engelmannii) co-occurs with Arizona corkbark fir (Abies Iasiocarpa var. arizonica) at high elevation (>2650 m) and with Douglas-fir (Pseudotsuga menziesii Franco) and Southwestern white pine (Pinus strobifonnis) throughout most of its elevational distribution (Whittaker and Niering 1965). At its steep (Epperson et al. 2001) xeric boundary (~2430-2530 m elevation), Southwestern ponderosa pine (“PIN_PON”) transitions to Arizona pine (Pinus arizonica Engelmannii), which extends down to ~1760 m elevation. This narrow boundary suggests that abiotic conditions may be limiting the upward migration of Arizona pine (“PIN_ARI") and downward migration of PIN_PON. Since prechilling is required to germinate stored (i.e., over-wintered) seed from PIN_PON, but notedly not PIN_ARI (Krugman and Jenkinson 1974), the lack of cold conditions at low elevation could also prevent germination of PIN_PON seeds. Further, putative hybrids (“Mixed”) between PIN_PON and PIN_ARI have 84 been reported throughout the range of both species as producing an intermediate number of needles per fascicle (Dodge 1963, Peloquin 1984, Rehfeldt et al. 1996, Epperson et al. 2001). If a product of hybridization, then the ability of this morphotype to overcome the lower and upper elevational barriers experienced by the other taxa (Kilgore, unpublished data) suggests that differences in germination ecophysiology that may exist between the parental taxa may have been reduced through hybridization. The principal aim for this study is to determine if there is a genetic difference in seed gerrninability and response to stratification by PIN_PON and PIN_ARI, and to relate these expected differences to potential species migration beyond current distributions. Given the close genetic relatedness of these taxa, little difference in total germination is expected. Although prechilling is not necessary for PIN_ARI seed (Young and Young 1992), stratification is expected to increase germination rate of seeds from both taxa (Baskin and Baskin 1998). Further, if lower elevation conifers can establish at high elevation, but not conversely (Barton 1993), then PIN_ARI germinants should have high survivorship at high elevation, and PIN_PON germinants should have relatively low survivorship at low elevation. Methods Seeds used in these experiments were collected from mature trees in the Santa Catalina Mountains in September-October 2005. In general, PIN_PON cones mature and seeds disperse in August-September, while PIN_ARI cone 85 ripening and seed dispersal are delayed by a month (Krugman and Jenkinson 1974). The PIN_PON trees near the summit of Mount Lemmon have been studied for needle number production for at least 10 years and are part of a larger research program (Epperson et al. 2001, 2003). Cones from PIN_ARI were collected from Rose Canyon, an area previously considered to be composed of “pure” 5-needled trees (F. Telewski, personal communication). The site effect will be reduced by collecting seed from all three morphotypes at a transition zone near Palisade Rock; at least maternal genetic contribution will be determined by known morphotype. Annual variability in seed lots will be controlled by using seed from one cone production year. With the aid of a 23-ft pole-pruner, branches containing abundant needles and multiple ripe cones were harvested from the mid- to upper-crown of each tree. Even for trees of known species, needles were counted to determine species designation (Peloquin 1984, Epperson et al. 2001). Cones were air- dried at room temperature in paper bags, and seeds were removed from the cones by shaking and removing cone bracts in early January 2006. Unopened cones were heated 14—27 h at 40° C to encourage opening (Young and Young 1992). Empty, parasitized, or otherwise unhealthy appearing seeds were discarded, while whole seeds were dewinged by hand, inventoried (Table 3-1), and stored in labeled small manila envelopes at room temperature. 86 Cold stratification experiment The controlled seed germination test as a function of cold stratification treatments was conducted in the laboratory at Michigan State University. The experimental design was a factorial with three sites, three taxa, and three stratification periods (0, 15, and 30 days). For each stratification period, an equal number (n=115) of seeds from 3-8 trees from each site and needle type were selected based on number of trees and seeds available (Table 3-1). 87 Table 3-1. Seeds inventoried from cones collected from the Santa Catalina Mountains in September-October 2005. Site refers to south slope of Mount Lemmon (MTL), south slope between Palisade Rock and Mount Bigelow (PAL), northeast of Lizard Rock (LIZ), and entrance drive to Rose Canyon Lake (RC). Morphotype refers to producing mean 3- (PIN_PON), mixed, or 5-needled (PIN_ARI) fascicles (Peloquin 1984). Tree refers to specific tree sampled as part of other research or specifically for this experiment. The first number in the last column refers to the number of seeds used from each tree per cold stratification treatment, while a second number refers to seeds used in the reciprocal transplant experiment (see Methods). Number of Number seeds per Site Morphotype Tree of seeds treatment MTL PIN_PON MTL 9 296 15 / 20 MTL 5 388 15 I 20 MTL 24 800 15 / 20 MTL 23 722 15 / 20 MTL 15 31 10 MTL 132 458 15 l 20 MTL 11 2388 15 / 20 MTL 104 282 15 MTL 1.5m behind junction 911 0 MTL junction of MTL+MeadowTrail 481 0 MTL 9.2m@280° from MTL 104 554 0 MTL 20m@210° from 60"dome 360 0 MTL 23m@330° from RadioTowerFence 299 7 0 PAL PIN_PON PAL 1 133 41 PAL 41 53 17 PAL 3 62 20 PAL 12 74 24 PAL 40 ...- 40 13 PAL Mixed PAL 28 270 87 PAL 32 48 16 PAL 30 36 12 PAL PIN_ARI PAL 14 332 30 / 20 PAL 2 104 30 / 10 PAL 10 366 32 l 20 PAL 16 71 23 PAL 49 11 0 / 10 RC PIN_ARI RC PINARI GroupPicnic 29.5cm dbh 214 37/20 RC Tree#15 28 0 l 20 .- RC Tree#16 _. n/a 0 I 20 g LIZ PIN_ARI LIZ 1 43 14 LIZ 3 17 5 LIZ PINARI 10cm dbh 3 0 LIZ PINARI 353° from LIZ 2 71 23 LIZ PINARI 61.50m dbh 155 36 LIZ PINARI 5 0 Seed handling and stratification methods for ponderosa pine vary widely in the literature (e.g., Stefferud 1961, Smith 1986, Weber and Sorensen 1990, 88 Young and Young 1992, Evans et al. 2001, Wenny and Dumroese 2001, Dreesen 2003), while little is published on germination characteristics for Arizona pine (except see Young and Young 1992). In general, seeds are rinsed with a fungicide, soaked for some period of time to induce imbibition and reduce germination inhibitors (Evans et al. 2001), placed into media or plastic bag, and stored (2-8 wk) at a constant cool temperature (0-5° C). Smith (1986) reported no effect of pregermination fungicides on germination success of ponderosa pine seeds. A contract grower for the Coronado National Forest and Trees for Mount Lemmon stores (“cold stratifies”) ponderosa pine seeds from southern Arizona at 1-2° C for variable lengths of time, rinses the seeds with 10% bleach solution for 5 minutes before rinsing and soaking the seeds for 12-24 h, and then immediately plants seeds into soil with ~85% germination success (B. Blake, Northern Arizona University, personal communication). The seed handling methods used in this paper combine elements from the literature, especially Weber and Sorensen (1990), with practices in use by CNF growers. In mid-January 2006, seeds from each tree by stratification treatment (see Table 3-1) were placed into a perforated and labeled 4-mL Whirl-Pak Write- On Bag (Nasco, Fort Atkinson, Wisconsin, USA); perforation of the plastic bag walls was accomplished with a wallpaper removing tool to allow water and air infiltration to the seeds. Approximately 25 h before each stratification treatment, or sowing for the no stratification treatment, the seeds were surface-sterilized in 10% bleach solution for 30 min, rinsed with distilled water, and soaked in a running distilled water bath at room temperature for 24 h. The seeds were air- 89 dried on individually labeled paper towels, while the bags were oven—dried at 40° C. The dry seeds were returned to their appropriate bags, which were pooled into a 1-gal Zip-Loo (SC Johnson, Racine, WI, USA) bag containing 1 L perlite moistened by 200 mL distilled water. The Zip-Loc bag containing the Whirl-Pak bags of seeds was stored in a standard refrigerator kept at 2-4° C. Each stratification treatment was temporally staggered so that all seeds would begin incubation at the same time; the seeds for the 0- and 15-day treatments were stored in Whirl-Pak bags at room temperature before surface sterilization. In mid-February 2006, stratified seeds were removed from refrigeration, and the no-stratification seeds were surface sterilized in preparation for incubation and germination; no seeds germinated prior to sowing. Following Weber and Sorensen (1990), seeds from each Whirl-Pak bag were sown into polystyrene Petri plates (100mm Not TC-Treated Culture Dish, Product #430591, Corning Life Sciences, Corning, NY, USA) containing 15 mL fine (#4) vermiculite saturated with 13 mL deionized (DI) distilled water (NANOpure ultrapure water system, Model D4741, Barnstead lntemational, Dubuque, IA, USA) and covered by a 90-mm round filter paper (Whatman Grade 1; Whatman, lnc., Florham Park, New Jersey, USA). Appropriately labeled lids allow gas exchange by design. The plates containing seeds were randomly arranged within an aerated growth chamber (Model E-15, Environmental Growth Chambers, Chagrin Falls, OH, USA) and kept in the dark at 20° C; Li et al. (1994) report no effect of light on germination of stratified ponderosa pine seed. DI water was added to plates as 90 necessary to maintain moist (but not saturated) filter papers, and plates were regularly shifted in the growth chamber to avoid chamber effects. Germinated seeds were assessed more frequently (e.g., twice daily) early in the experiment and then at least daily as the number of new germinants curtailed. Germination was counted when the radicle reached ~1 mm in length (Weber and Sorensen 1990). Seeds covered by fungi were not removed because many of them germinated even when infected. As temporally feasible, germinated seeds were removed with forceps sterilized with ethanol and planted into individual pots with soil for another component of this dissertation research. Ungerminated seeds were counted at the end of the experiment and considered functionally nonviable in the analyses. The time to initial germination, total percent, and time to complete actual germination were determined from each tree within each stratification treatment. Effects of site on the response variables within 3-needle (MTL and PAL sites) and 5-needle (PAL and RC/LIZ sites) were estimated by two-factor analysis of variance (ANOVA), then effects of species and stratification were similarly estimated for the PAL site. All response data were transformed by natural logarithm to normalize residuals. Finally, the pooled effects of species and stratification were estimated with two-factor ANOVA. Given the unbalanced design, Type II sums of squares (SS; Langsrud 2003) were used to test hypotheses using the ‘car’ (Companion to Applied Regression; Fox 2006) package in R (R Development Core Team 2007). Means were compared at each factor level using Tukey’s Honest Significant Difference (HSD) in R, which 91 incorporates an adjustment for unbalanced designs (R Development Core Team 2007). Reciprocal germination experiment Two sites were selected for the reciprocal germination experiment: the top of a ridge extending off the north slope of Mount Lemmon (32.445°N 110.789°W, ~2720 m elevation), and the top of a small west-facing ridge near Willow Canyon (32.387°N 110.694°W, ~2185 m elevation). These sites represent the top of the mountain beyond the known distribution of PIN_ARI and below the known distribution of PIN_PON, respectively. The soil at the Mount Lemmon site is derived from micaceous sandstone and sandy phyllite, while the soil around Willow Canyon is derived from muscovite—garnet leucogranite (Force 1997). The vegetation at the Mount Lemmon site is dominated by Douglas-fir (Pseudotsuga menziesir) and PIN_PON; the 2003 Aspen Fire eliminated regeneration and killed most of the overstory at this site. Willow Canyon vegetation is dominated by PIN_ARI, Arizona madrone (Arbutus arizonica), and oaks (Quercus spp.); the Aspen Fire eliminated ground vegetation, soil organic matter, and small-diameter trees at this site but little of the overstory was killed. In January 2006, canopy openness measured by spherical densiometer (Lemmon 1957) was not significantly different between the two sites (Mount Lemmon: 1? =16.6%, s=2.6%; Willow Canyon: 56 =22.9%, s=12.2%; Wilcoxon Rank Sum Test, W=38.5, p-value = 0.4055; R Development Core Team 2007). Seeds from six PIN_PON (mean 2752 m elevation) and seven PIN_ARI (Palisade Rock — mean 2526 m elevation; Rose Canyon - mean 2159 m 92 elevation) trees were selected for the reciprocal germination experiment (Table 3-1). In late January 2006, a total of twelve seeds, one from each contributing tree, were sown on a prepared seedbed free from litter within each of ten predator-exclusion cages. Six aluminum nails centered longitudinally and spaced by 10 cm were pushed into the soil to serve as sowing and monitoring guides (Figure 3-1). Each seed was placed 5 cm above or below the nail in a predetermined randomized pattern. Seeds were covered with 1 cm of light duff before cages were placed over the seeds (Figure 3-2). The cages were constructed from 0.5-in (1 .3-cm) hardware cloth to provide a 20-cm wide x 61-cm long seedbed with 15 cm of space for vertical growth and 3 cm of cloth folded back under the soil. Landscape pins and tags listing cage number and contact information were installed at two opposing comers of the cage. 93 Figure 3-1. Prepared seedbed within the predator—exclusion cage at the Willow Canyon site. Note the six aluminum nails either side of which seeds were sown. Figure 3-2. Predator-exclusion cage after placement over sown seeds at the Willow Canyon site (January 2006). Seed germination was assessed at both sites in June 2006 and mid- October 2006. No germinants were observed in any cage at either site in June 2006 (G. Friedlander, personal communication), thus no documentation was required. In October 2006, the germinants in each cage were digitally photographed from above, and live and dead pine germinants within each cage were mapped relative to the aluminum nails and sides of the cage. Patterns of erosion evidenced by lack of duff were also mapped. The roofs of two cages at the Mount Lemmon site were bent and repaired in the field; substantial force from a round object (e.g., bear paw) was required to create the pattern along the crease of the roofs and corners. No germinants appeared to have been damaged. Without referring to the identity of seeds arranged within each cage, probable planted (versus naturally sown) germinants were identified on the maps and then color-coded by species of seed. Each cage was considered to be a replicate in a balanced, two-factor (site and species) design, and thus percent germination was calculated for each species in each cage. Differences in germination resulting from site, species, and their interaction effects were summarized and analyzed by two-factor ANOVA, and means were compared at each factor level by Tukey’s HSD in R (R Development Core Team 2007). 95 Results Cold stratification experiment Seed coats began cracking two days after incubation began, and the first germinants appeared by the third day. Most of the germination occurred within the first 8 days with a number of seeds germinating for another 12 days; the experiment was terminated after day 22. No trend in total percent germination across site, species, or stratification was observed (Figure 3-3); range in values was 75% (PIN_ARI from PAL site in 0-day stratification) to 93% (PIN_ARI from RC/LIZ site in 15-day stratification). Stratification increased the mean time to first and last germination events for all morphotypes; values ranged from 77.6 h (3.2 d; PIN_PON from MTL site in O-day stratification) to 111.4 h (4.6 d; PIN_ARI from RC/LIZ site in 30-day stratification). There was a large amount of variability in time to last seed germination for PIN_ARI, especially in the shorter stratification treatments. Mean time to last germination ranged from 99.7 (4.2 d; PIN_PON from MTL site in 0-day stratification) to 271.1 h (11.3 d; Mixed from PAL site in 30-day stratification). 96 100 - 23 so - S E 60 1 Clo—day 5 B15—day 8. ISO-day - 40 g I 201 0L PIN_PON PIN_ARI Q: = I V 2 ¢/ E 7 Clo-day g / EMS—day a / I30—day E / a s / '- / - é Mmd PIN ARI 500 1 5 450+ r 5 4001 E — 350~ E 8: 300 . ‘6 Clo-day 5 250- D15~day 5 200 « ”may E 3 1504 8 E 100 . '— 50 1 o I PIN_PON Med PIN_ARI Figure 3-3. Mean seed germination (top), time to first germination (middle), and elapsed time to the last seed germination (bottom) for the full data set. Species data are pooled across sites. Data are mean values; bars are one standard deviation. 97 Site effects within each of the PIN_PON and PIN_ARI groups were either not significant. In the PIN_PON group, responses to stratification treatments by seeds from MTL and PAL overlapped for all response variables (Figure 3-4). However, the time to first and completion of realized germination were significantly affected by stratification treatment (Table 3-2); all painivise comparisons between stratification treatments were significant (p<0.05) for time to first germination, while only the 0- and 15—day stratification treatments were not different for time to completion of germination. Removing the non-significant interaction term, as in all of the analyses presented in this section, had no effect on identifying significant model terms, but rather only slightly affected their level of significance, which will not be reported. 1000 O O 500 OWL-initiate O 450 APAL-inltiate 8 Q Q OMTL-corrplete ‘ 55‘ 80W 0 g 40° APAL-corrplete - O gs ‘ 2 6° 3 8 30° . g Q 0 .133 zso . g "' 0 § 40 .3 £200 : 2 t- O “ 150 g A I I 20 7‘ 100; Q g OMTL 50 APAL o . r . o . . . o 15 30 o 15 30 Stratification treatment (day) Stratification treatrmnt (day) Figure 3-4. Germination as a function of stratification treatment for seeds from PIN_PON trees at MTL and PAL. 98 Table 3-2. Type II tests of 2-factor ANOVA of percent germination (pergerm), time to first germination (initiate), and time to last realized germination (complete) for PIN_PON seed as a function of site (MTL and PAL), stratification (strat; 0-, 15-, and 30—day), and their interaction. Determination of site effect was made by Tukey’s HSD. Adjusted R2 for the three models are -0.1029, 0.4235, and 0.4122, respectively. Significant (a<0.05) terms are bolded. Site effect Response Adj-p Term Df SS F-value P-value Ln(pergerm) 0.6454 Site 1 0.01405 0.2156 0.6454 Strat 2 0.02647 0.2031 0.8172 Site'strat 2 0.05426 0.4163 0.6629 Residuals 33 2.15052 Ln(initiate) 0.7028 Site 1 0.00088 0.0797 0.7794 Strat 2 0.35205 1 5.91 25 1.451E-05 Site'strat 2 0.01 1 19 0.5060 0.6075 Residuals 33 0.36505 Ln(complete) 0.3221 Site 1 0.08894 1.0103 0.3221 Strat 2 2.4821 8 14.0971 3.756E-05 Site'strat 2 0.21532 1.2229 0.3074 Residuals 33 2.90527 In the PIN_ARI group, responses were similarly distributed, but stratification appeared less important across the dependent variables (Figure 3- 5). Site and stratification were not significant (p>0.05) effects on percent germination or time to completion of germination, but both were significant (p=0.043 and 0.010, respectively) for time before initiation of germination (Table 3-3). The response of the seeds from RC/LIZ in the 30-day cold stratification treatment influences these results more than any of the other relevant pairwise comparisons; the 0- and 30-day stratification treatments within RC/LIZ was slightly less than significant (p=0.084) and across both sites was highly significant (p=0.008). Interestingly, the Type-Ill test reports nonsignificant p- values for site (p=0.450) and stratification (p=0.198) in the model for initiation of germination. Furthermore, a one-way ANOVA test of site on time to initial 99 germination was minimally significant (ln(initiate)~site, p=0.071). Due to the sole influence of the 30-day stratified seed from RC/LIZ, site effects between PAL and RC/LIZ are considered to be inconsequential for the analysis of the full data set. A PAL - initiate 10° 3 B 8 600 D LIZ/RC - initiate a E ‘ 4 PAL - corrplete .. I LIZ/RC - corrplete ‘ ‘ a! 80 3 U i 500 a I :1 . I g . A 5 5 400 g 60 A 8 8 . .. E ‘ 1" .2 300 I § A g 40 E i I I l- E 200 I . I 20 F 5 I . m 10° E i E D LIZ/RC 0 V 1 ‘ 0 r I 1 0 15 30 0 15 30 Stratification treatment (day) Stratification treatment (day) Figure 3-5. Germination as a function of stratification treatment for seeds from PIN_ARI trees at PAL and RC/LIZ. Table 3-3. Type II tests of 2-factor ANOVA of percent germination (pergerm), time to first germination (initiate), and time to last realized germination (complete) for PIN_ARI seed as a function of site (PAL and RC/LIZ), stratification (strat; 0-, 15-, and 30-day), and their interaction. Determination of site effect was made by Tukey's HSD. Adjusted R2 for the three models are 0.0396, 0.3054, and -0.1518, respectively. Significant (a<0.05) terms are bolded. Site effect Resmnse Adj-p Term Df SS F-value P-value Ln(pergerm) 0.0871 Site 1 0.09728 3.2220 0.08706 Strat 2 0.07894 1 .3073 0.29169 Site'strat 2 0.00712 0.1 178 0.88943 Residuals 21 0.63405 Ln(initiate) 0.0427 Site 1 0.07790 4.6543 0.04271 Strat 2 0.1 9128 5.7141 0.01044 Site'strat 2 0.00587 0.1754 0.84031 Residuals 21 0.35149 Ln(complete) 0.9280 Site 1 0.0024 0.0084 0.9280 Strat 2 0.2285 0.3962 0.6778 Site*strat 2 0.2228 0.3863 0.6843 Residuals 21 6.0550 100 Within the PAL site containing all three morphotypes, there were no effects of species on any of the response variables (Table 3-4). Time to initiation was influenced by stratification, especially the 30-day treatment; the mean difference between the 0- and 30-day and the 15- and 30-day treatments were significant (p=0.001 and 0.033, respectively; Figure 3-6). 0 O-day - initiate A 15-day - initiate 600 a 30-day - initiate 100 I C 0 I O-day - complete A ‘ 500 A 15—day - complete I 80 I I 30-day - complete ‘ 25 a ' ' c I ‘5: 400 g 60 g 3 S A ‘ E o g E 300 ‘ a A .5 E ‘ 7 - 4o 8 g g 200 7 ' 2 o O-day F - ' 0 A 15-day 100 a a I 30-day o r r 1 0 fi PIN_PON Mixed PIN-ARI PIN_PON Mixed PIN ARI Ilo ho Morphotype '13 MN Figure 3-6. Germination as a function of stratification treatment for combined morphotypes at the PAL site. 101 Table 3-4. Type II tests of 2-factor ANOVA of percent germination (pergerm), time to first germination (initiate), and time to last realized germination (complete) for the PAL site as a function of morphotype (PIN_PON, Mixed, and PIN_ARI), stratification (strat; 0-, 15-, and 30-day), and their interaction. Painivise comparison of PIN_PON and PIN_ARI (3-5 effect) was made by Tukey’s HSD. Adjusted R2 for the three models are -0.1774, 0.2311, and 0.0015, respectively. Significant (a<0.05) terms are bolded. 3-5 effect Response Adj-p Term Df SS F-value P-value Ln(pergerm) 0.9909 Morphotype 2 0.02273 0.2831 0.7556 Strat 2 0.01380 0.1719 0.8430 Site*strat 4 0.07290 0.4540 0.7686 Residuals 27 1 .08392 Ln(initiate) 0.9028 Morphotype 2 0.01074 0.3453 0.71 1 1 Strat 2 0.25405 8.1721 0.0017 Site*strat 4 0.02307 0.371 1 0.8271 Residuals 27 0.41969 Ln(complete) 0.2327 Morphotype 2 0.8094 1.5417 0.2323 Strat 2 0.9196 1.7516 0.1926 Site'strat 4 0.3844 0.3662 0.8305 Residuals 27 7.0872 Because site effect was not significant, the data were pooled across sites (Figure 3-3) and analyzed by morphotype and stratification treatments. None of the terms, including interaction, explained a significant (p>0.05) portion of the variance in percent germinated seed (Table 3-5). For time to initiation of germination, stratification was highly significant (p<0.001), but the effect of morphotype was not significant (p=0.28; PIN_ARI versus PIN_PON pairwise comparison p=0.26). Time to complete germination was significantly affected by both species and stratification (both p=0.001). Painivise comparisons in this model between PIN_ARI and PIN_PON (p=0.001), 15- and 0-day stratification (p=0.010), and 30- and 0-day stratification (p=0.002) were highly significant; the Mixed-needled trees were only marginally different (p=0.106) from PIN_PON. However, PIN_PON and PIN_ARI were only significantly different in the 0-day stratification treatment (Table 3-6). Although there was no significant interaction 102 (p=0.28), morphotype and stratification treatment affected the time to completion of germination. Table 3-5. Type II tests of 2-factor ANOVA of natural-log-transformed percent realized germination (pergerm), time to first germination (initiate), and time to last germination (complete) for pooled sites as a function of morphotype (PIN_PON, Mixed, and PIN_ARI), stratification (strat; 0-, 15- and 30-day), and their interaction. Pairwise comparison of PIN_PON and PIN_ARl (3-5 effect) was made by Tukey’s HSD. Adjusted R2 for the three models are -0.0666, 0.388, and 0.2657, respectively. Significant (a<0.05) terms are bolded. 3-5 effect Response Adj-p Term Df 88 F-value P-value Ln(pergerm) 0.5939 Morphotype 2 0.04697 0.4984 0.6098 Strat 2 0.08355 0.8867 0.4169 Morphotype'strat 4 0.02873 0.1525 0.9613 Residuals 66 3.10959 Ln(initiate) 0.2598 Morphotype 2 0.03384 1.3004 0.2793 Strat 2 0.66130 25.4099 6.56E-09 Morphotype*strat 4 0.01932 0.3712 0.8283 Residuals 27 0.85883 Ln(complete) 0.0011 Morphotype 2 2.5232 7.5459 0.0011 Strat 2 2.4268 7.2575 00014 Morphotype‘strat 4 0.8651 1.2935 0.2816 Residuals 27 Table 3-6. Pairwise comparisons (Tukey's HSD) of PIN_PON and PIN_ARI seeds for natural-log-transformed percent realized germination (pergerm), time to first germination (initiate), and time to last germination (complete) as a function of stratification treatment for pooled sites. Data are reported as adjusted p-values; significant values are bolded. Stratification pergerm initiate complete O-day 1.0000 0.9168 0.021 2 15-day 0.9860 0.9993 0.2557 30-day 0.9993 0.9974 0.9999 Reciprocal germination experiment Approximately one-half of the PIN_PON seeds germinated the first growing season at both sites, while only around one-fourth of the PIN_ARI seeds germinated between the two sites (Figure 3-7). Consequently, site was not 103 significant but species was significant in the ANOVA (Table 3-6). The overall interaction term was not significant (Table 3-7), but a significant difference exists between germination of PIN_PON and PIN_ARI seeds at the high elevation site on Mount Lemmon (Table 3-8). There was no difference in germination of PIN_ARl seeds from Palisade Rock and Rose Canyon (x2=1.6897, df=1, p- value=0.1936; R Development Core Team 2007). 100 - El PIN_PON I PIN_ARI V 80 ' 55 v 60 C .2 76 .5 E 5 40 0 l- 20 "l J o l— Mount Lemmon Willow Canyon Figure 3-7. Germination of PIN_PON and PlNl__ARl seeds planted at high (Mount Lemmon) and low (Willow Canyon) elevation. Values are mean percent germination; bars are one standard deviation. Table 3-7. ANOVA for germination success of PIN_PON and PIN_ARI seeds at the Mount Lemmon and Willow Canyon sites (adjusted R2=0.1625). Significant (a<0.05) terms are bolded. Term Df SS MS F-value P—value Site 1 837.2 837.2 1.0287 0.3172 Morphotype 1 7049.0 1049.0 8.6613 0.0057 Site'morphotype 1 714.0 714 0.8773 0.3552 Residuals 3 29298.7 813.9 Table 3-8. Multiple comparisons between site (Mount Lemmon - MTL and Willow Canyon - WC), morphotype, and interactions by Tukey’s HSD. Significant (a<0.05) comparisons are bolded. Group 1 Group 2 Mean difference P-value WC MTL 9.15 0.3172 PIN_ARI PIN_PON -26.55 0.0057 WC PIN_PON MTL PIN_PON 0.7 0.9999 WC PIN_ARl MTL PIN_ARI 17.6 0.5200 WC PIN_ARI WC PIN_PON -18.1 0.4962 MTL PIN_ARI MTL PIN_PON -35.0 0.0445 Discussion Effects of stratification on germination success and rate In the controlled cold stratification experiment, a similar percentage (75- 93°/o) of PIN_PON and PIN_ARI seeds germinated across stratification treatments. Although molds were present in some Petri dishes, seeds continued to germinate and thus were not considered to be affected by the molds. Seeds from PIN_ARI appeared (Figure 3-3) to have higher germination rates than PIN_PON with longer stratification periods, but this difference was not significant due to a large amount of intraspecies variation. In contrast to the stratification experiment, more PIN_PON than PIN_ARI seeds from the same seed lots used in the laboratory experiment (Table 3-1) germinated at both high (Mount Lemmon site; 52 vs. 13%, p=0.0445) and low (Willow Canyon site; 52 vs. 34%, p=0.4962) sites, although the difference at the latter site was not significant (Figure 3-7). Since the seeds in the present study were sown into the field in January, approximately the coldest time of year, “natural” cold stratification is assumed. In October, a large number of germinants were growing around the predator-exclosure cages at the Willow Canyon site 105 (personal observation), likely due to release of seeds by overlying trees during the winter. Consequently, selection of germinants from seeds sown for this experiment was highly conservative; i.e., I included only those germinants falling exactly in the expected location, which ostensibly led to a lower perceived germination rate for both taxa. Also, runoff appeared to have removed the light duff over the seeds in portions of two cages at the Willow Canyon site; expected seeds from these portions were removed from the calculations. No ponderosa pine germinants were observed outside of the cages at the Mount Lemmon site; instead, forbs had sprouted within and around the cages, possibly shading or displacing seeds away from the expected location. Like the stratification experiment, there was a large amount of intraspecies variability in total germination (Figure 3-7). In the controlled laboratory experiment, cold stratification significantly increased the amount of elapsed time after sowing before seeds began to germinate as well as the length of time before the last seeds germinated (Figure 3-3). There seemed to be no species difference in time to first germination, while PIN_ARI generally took longer to complete germination than PIN_PON in all stratification treatments and noticeably longer than Mixed in the 0-day control treatment. Distinct from all other patterns of stratification effects, PIN_ARI seeds in the longest stratification treatment actually completed germination in less time than those in the 15-day treatment and about the same amount of time as those in the 0—day control treatment (Figure 3-3). This idiosyncrasy is due to seeds from 2 of the 5 PIN_ARI trees from RC/LIZ completing their (100% actual) 106 germination relatively early (162 h; 6.7 d) in the incubation period and thus biasing the mean and trend observed in the other morphotypes. The effects of stratification on germination success and rate in ponderosa pine seeds varies, as was demonstrated in the present experiments. In laboratory tests, Barton (1930) also found higher initial but similar final germination by stratified and unstratified ponderosa pine seed. Smith (1986) also reported higher initial rate but found lower overall germination of stratified (f = 25% over 30-90 days) versus unstratified (f =38%) ponderosa pine seed from New Mexico and Colorado. The effects of stratification were supported by a second test of the same seed batch conducted at a production greenhouse (f = 7% over 31-55 days versus 3? =12% for unstratified seed; Smith 1986); the significantly lower germination at the greenhouse as compared to the laboratory could not be explained. Weber and Sorensen (1990) found that stratification increased germination rate but also uniformity and total percentage, especially at lower incubation temperatures, in ponderosa pine (Pinus ponderosa var. ponderosa) from central Oregon. All whole seeds germinated after 30 days of stratification followed by incubation at 20° C, the same temperature used in this study, with fewer days before the majority of seeds germinated in longer stratification periods (Weber and Sorensen 1990). Following sowing of ponderosa pine seed in Fort Valley Experimental Forest of northern Arizona, Schmid and Mitchell (1986, cited by Howard 2003b) found 61-90% germination beginning 12-37 days after sowing. 107 From the controlled stratification experiment, I conclude that there is no difference in germinability between any of the morphotypes tested in this study. In addition, stratification by itself did not appear to affect germination success as demonstrated in the cold stratification experiment. The difference in observed germination success between PIN_PON and PIN_ARI in the field must be due to some other factor, such as timing of germination (see below). However, emergence is only the first critical threshold in seedling establishment; germinant survival through the first arid foresummer (May-June 2007) has not been determined. In Barton’s (1993) study of elevational distribution of desertic pines in nearby Chiricahua Mountains, the species with the lowest elevational distribution (Pinus discolor) had higher emergence and first-year survival at high than low elevation, which is an opposite trend to PIN_ARI in the present study. However, all of the germinants of the higher elevation species (P. engelmannii and P. leiophylla var. chihuahuana) had died at lower elevation by the end of the first arid foresummer (Barton 1993). In the present study, the PIN_PON germinants, then, are expected to have lower survivorship at the Willow Canyon site as compared to the Mount Lemmon site. The long-term implication of survival by young seedlings through multiple seasons of drought is the determination of the community composition. Implications of variable germination period Seed germination is affected by a multitude of abiotic factors influencing the seed physiology, including temperature, water potential, light, smoke, and soil 108 chemistry (Bewley and Black 1985, Baskin and Baskin 1998). The survival of the resultant seedling is highly dependent on the narrow temporal congruence of these factors, especially temperature and water potential, or moisture content. Given that most montane trees produce and disperse mature seeds in the fall, it is no surprise that physiological dormancy is induced at maturity to endure the winter (Kozlowski et al. 1991, Baskin and Baskin 1998). Germinants would have precious few weeks to establish root systems and store photosynthates before the onset of freezing conditions. Where snow cover is absent, freezing injury severely reduces seedling survivorship (Germino et al. 2002). In pines, populations that are adapted to colder temperatures (i.e., higher in elevation) tend to have deeper dormancy and longer chilling requirements (Weber and Sorensen 1990, Skordilis and Thanos 1995, Allen and Meyer 1998). However, dormancy will vary within a given seed lot (Kozlowski et al. 1991) and even within mother plants (Andersson and Milbert 1998). Ponderosa pine is known for producing mature seeds in the fall that can germinate under light but that generally require prechilling for maximum (dark) germination (Li et al. 1994). However, more slowly drying seeds could become more dormant than those that dry fast (Baskin and Baskin 1998); in other words, seeds from more xeric habitats at low elevation could be less dormant than those from higher elevation. Although there was no species difference in time to initial germination in this study, PIN_PON did complete germination faster (i.e., more dormant and thus more influenced by the stratification effect) than PIN_ARI, the lower elevation species. 109 Physiological dormancy is usually broken in seeds by imbibing water during the stratification process (Baskin and Baskin 1998). Once the minimum cardinal temperature for germination is reached (Bewley and Black 1985), ponderosa pine germinants begin to appear in the spring through epigeal germination (Oliver and Ryker 1990). Since there is no innate soil seed bank in pines (Keeley and Zedler 1998; except for Pinus albicaulis in Tomback et al. 2001), all viable seeds should germinate when minimum moisture content and thermal loads coincide (Alvarado and Bradford 2002). However, pine seeds that germinate in the spring in the Santa Catalina Mountains and similar mountain islands of the semiarid Southwest will face the arid foresummer with little growing time to establish root systems that can access deeper water. Schubert (1974, cited in Oliver and Ryker 1990) noted that water potentials more negative than 0.7 MPa significantly reduced seed germination, as well as root penetration, root dry weight, and cotyledon length in southwestern ponderosa pine germinants. If seeds were to germinate during this cooler, moist period followed by the arid foresummer, drought stress would ultimately kill them. In fact, by 27 June 2006, there were no pine germinants observed around or in any of the predator- exclusion cages at either Mount Lemmon or Willow Canyon site (G. Friedlander, personal communication); this observation indicates that no pine seeds had yet germinated. Consequently, despite more northern Rocky Mountain ponderosa pine populations producing seed that germinate in the spring (Young and Young 1992), the Southwestern race (Conkle and Critchfield 1987) as well as Arizona 110 pine trees produce seed that undergo a secondary conditional dormancy. Baskin and Baskin (1998) describe primary dormancy as that which a seed possessed when freshly mature, while secondary conditional dormancy occurs when primary dormant seeds become partially or fully nondorrnant but then are induced into a second conditional dormancy under certain environmental conditions. The moist stratified seeds of PIN_PON and PIN_ARI are exposed to sufficiently warm but inconsistent (i.e., frequent night frosts) spring temperatures but do not germinate. Instead, as their water potential increases (i.e., the seeds lose moisture) with the coming arid foresummer, they undergo an induced secondary conditional dormancy. This secondary dormancy is then broken by the summer monsoon precipitation when the soil temperatures are consistently warm and soil is moist. Figure 3-8 illustrates the hypothesized path through secondary conditional dormancy and subsequent successful germination relative to the hydrologic seasons. We should then not be surprised that the species found at higher elevation, and thus presumably more sensitive to drought (Barton 1993), completes the entire germination process more rapidly and uniformly than lower elevation species. As the summer monsoon subsides, air and soil temperatures remain high, and soon the fall drought will impose stress on the young seedlings. Therefore, there is strong selection for rapid initiation and completion of germination in the narrow window of adequate soil moisture and consistently warm temperatures before both the fall drought ensues and winter delivers freezing conditions. Conversely, total loss of a year’s regeneration efforts could 111 occur as a result of precipitation events before the arid foresummer concludes, while species whose seeds take longer to germinate may endure the remaining drought prior to monsoon. Ponderosa pine seedlings rapidly grow tap roots, which enables them to survive in xeric soils (Howard 2003b), yet seedling mortality is highest in the arid Southwest (Howard 2003b). Ponderosa pine seedlings in Arizona that germinated earlier in the monsoon were more likely to survive the first year than those that germinated in August (Schmid and Mitchell 1986, cited by Howard 2003b). Given that there is greater selection for extreme conditions than for the mean (Ledig 1998) and that PIN_PON are faced with more limiting conditions as the growing season is shorter at high elevation, it is sensible that the germination and competition for soil-water resources occurs more rapidly for PIN_PON than for PIN_ARI, which is shown in part by the more rapid completion of germination after stratification by PIN_PON in this study. 112 Summer Monsoon ‘ Seeds mature in cones Seeds on ground ' ‘ imbibe moisture Foregsgimer Md erminate Seeds dry Germinants and enter die from sepergiaester Fa|| Drought finndr‘taio‘rlal drought stress dormancy A y / and disperse Early I Seeds on s rin “ ground P 9 \ germinate \ \ \ I/I‘ \ or / \ “I Seeds imbibe . 1 morsture \ Winter Rains/Snow Figure 3-8. A proposed path of selection for conditional (secondary) dormancy of seeds produced by Arizona and Southwestern ponderosa pine in the mountain islands of the Southwest experiencing bimodal precipitation patterns. Relevant hydrologic seasons outline the path. Biogeographic implications in a changing climate Based on projections from both the Intergovernmental Panel on Climate Change and United Kingdom Hadley Centre’s general circulation models, spring and fall temperatures in Arizona will increase by 1.7-2.2 °C, and winter and summer temperatures will increase by 2.8 °C by 2100 (USEPA 1998). Precipitation is projected to increase in spring (20%), decrease slightly in summer (-15-0%), increase in fall (30%), and increase in winter (60%) (USEPA 1998). The actual effects of these projected climatic changes on biological functions, commonly indicated by some measure of actual versus potential 113 evapotranspiration, are difficult to ascertain, especially when complex topography and elevation are considered. However, the projections do suggest that the Ponderosae in Arizona may experience a warming climate, especially during the winter and summer, and the bimodal precipitation pattern characteristic of southeastern Arizona may be somewhat more smoothed across the seasons, especially by reducing the intense monsoonal precipitation and severity of the fall drought. However, the projections are based on the entire state of Arizona, of which the southeastern portion that contains the mountains islands of Pondemsae is most affected by the rising subtropical high-pressure ridge (“Bermuda High”) and its associated intense convective storms (Sheppard et al. 2002). Consequently, I can only suggest with some confidence that all seasonal temperatures are expected to increase (“global warming”) and that precipitation patterns may be shifted toward a wetter fall and winter. In other parts of the world, global warming has already resulted in altitudinal shifts of over 100 m in treeline and over 300 m in advanced regeneration of tree species (e.g., Kullman 2002). In colder climates, the milder winters reduce freezing-related injuries and extend the growing season, especially important to germination (Hobbie and Chapin 1998) and resultant seedlings (Kullman 2002). In the Santa Catalina Mountains and other mountain islands of the Southwest, the overall warmer climate but wetter fall may extend the temporal window for germination and successful establishment for both PIN_PON and PIN_ARI. Dodge’s (1963) observation of dead trees and stumps of “ponderosa pine” below the current elevation of both PIN_PON and PIN_ARI 114 in the Santa Catalina Mountains indicates that at least one of the two species is migrating upward in elevation, likely in response to climate change. In fact, the annual average temperature has increased by 2 °C and the annual precipitation has increased by at least 15% in the Tucson area over the last century (USEPA 1998). Given the presumed strong influence of drought in limiting the more xeric lower elevational distribution for both species and that more humid fall conditions are projected for the next century, both taxa may be able to expand their overall range in the mountain islands. However, if the additional moisture is not physiologically useful (i.e., increased storm intensities), then drought will become even more severe in limiting their lower elevations. Given the low soil-water holding capacity of the shallow lithosols overlying steep bedrock gradients in the Santa Catalina Mountains, as well as the other mountain islands of the Southwest, though, distribution expansion is likely only to occur toward more mesic conditions at high elevation. 115 CHAPTER 4 DIFFERENTIAL COLD TOLERANCE SEGREGATES DISTRIBUTION OF PONDEROSAE IN THE SANTA CATALINA MOUNTAINS OF ARIZONA Introduction Cold conditions restrict the upper boundary, whether high latitude or elevation, of plant distribution through negative carbon balance (or inability to process acquired carbon; Hoch and Korner 2003), interrupted phenological cycle, and poor resistance to secondary factors, such as wind force, snow abrasion, parasites, and high heat (Tranquillini 1979). Since closely related species (Sutinen et al. 1992, Lindgren and Hallgren 1993, Hawkins et al. 2003, Nippert et al. 2004), populations within species (DeHayes 1977, Kolb et al. 1985, Schaberg et al. 1995), and even hybrid species (Lu et al. 2007) vary in their cold tolerance, their upper elevational limit should vary accordingly. Seedlings, in particular, are more sensitive to the severe abiotic conditions present at upper elevation limits (Cui and Smith 1991, Johnson et al. 2004) and should thus present stronger signals of differential interspecific cold tolerance. The development of cold tolerance, or hardening, is related to genetic endogenous rhythms and enhanced by exposure to environmental conditions that limit metabolism and carbon acquisition (Havranek and Tranquillini 1995). The endogenous control is observed when dormant plants recover little under favorable conditions until later in winter (Pisek and Schiessl 1947, cited by Havranek and Tranquillini 1995). The initiation of predormancy is triggered by shorter day lengths, which is associated with lower daily temperatures. 116 Consequently, chemical reaction rates (i.e., metabolism), biosynthesis, gas exchange, water and nutrient acquisition, and carbon assimilation decrease, with primary growth finally terminating at the end of the season (Larcher 2003). Even before freezing temperatures are experienced by plants, lower temperatures induce changes in cellular structures in preparation for hardening. The changes occur at different rates in different organs of the plant; e.g., cold hardiness develops slower and to higher temperatures in roots than in other organs (Kozlowski et al. 1991). Protoplastic osmotic potential increases, and thus intracellular freezing point decreases, from accumulation of sugars, amino acids, and ions, the vacuole fissions, and intracellular water moves to extracellular wall spaces, which also increases the osmotic potential and translocates initial (“extracellular”) ice crystal formation outside the cellular wall barrier (Larcher 2003). Exposure to near-zero freezing temperatures induces production of cold- stable phospholipids and stress proteins that stabilize biomembranes (Larcher 2003). Cellular survival is dependent on the ability to prevent ice crystal formation within the cell; not only do the crystals destroy membrane integrity, but they increase cellular dehydration due to the higher vapor pressure above the ice than the supercooled cytoplasm. As cytoplasmic desiccation increases, as in drought conditions, the concentration of salt ions and organic acids reach a toxic level which results in enzyme inactivation, biomembrane phospholipid deterioration, protein disassociation, and ATPase inactivity (Larcher 2003). During dormancy, conifer leaves continue to absorb light, even though low temperatures inhibit the enzyme-mediated rate of carbon assimilation (Berry and 117 Bjorkman 1980). Consequently, the light energy in excess of that used for photochemistry must be dissipated lest oxidized radicals damage the photosynthetic machinery. In fact, conditions where the amount of absorbed light energy is greater than biochemically useful can occur during other stressful times of the growing season (e.g., drought) or day (e.g., midday solar peak), as well. In photosynthetic leaves, light energy is absorbed by accessory pigments (e.g., carotenoids, chlorophyll b and c) and chlorophyll (chl) a in the thylakoid membrane of the chloroplast. When the pigment molecule absorbs a photon of light, an electron is elevated to an excited state; this energy is transferred by resonance to chl a and then to a specialized chl a (“trap”) molecule in one of two thylakoid membrane-bound photosystems (Nobel 2005). Production of a variety of pigments increases the ability to absorb photons from a broader range of the electromagnetic spectrum, but this can be harmful under limitation of electron sinks, such as during dormancy. Light energy absorbed by Chi a can be dissipated via excitation transfer through photochemistry or through nonphotochemical quenching (NPQ), such as ' fluorescence, phosphorescence, or radiationless transfer (i.e., heat; Nobel 2005). During noncyclic electron flow in photosynthesis, energy from an excited electron is transferred by resonance to photosystem I (PSI) or II (PSII) to drive electron transport and eventually reduce NADP+ (Larcher 2003). Under high light conditions, some of the energy is lost by deexcitation of the electron to its ground state with concomitant production of radiation slightly lower in energy (A=1 KJ moi-1), or longer wavelength (A=4 nm), than the 662-nm absorption peak (Nobel 118 2005). This radiation, fluorescence, originates largely from the antennae pigments associated with PSII (Larcher 2003). Phosphorescent radiation is emitted when a pigment molecule deexcites from a triplet state; the probability of an excited electron reversing its orbital spin (i.e., “triplet state) is low, thus phosphorescence is rare (Nobel 2005). Last, excess absorbed light energy can be dissipated by heat in a variety of ways. Higher-energy photons, such as those in the 430-nm absorption peak of chi a, excite electrons to a higher but unstable energy level; the electron quickly transitions to a lower-energy excited state with the energy difference resulting in production of heat (Nobel 2005). Another energy-dissipating mechanism through heat is by downregulating photosynthesis (Oquist and Huner 2003, Adams et al. 2004). When the thylakoid lumen acidifies from a buildup of protons due to reduced rate of proton pumping by electron transport, the carotenoid violaxanthin (V) is de-epoxidised to antheraxanthin (A) and then zeaxanthin (Z) (Bjorkman and Demmig-Adams 1994). This occurs when the electron transport chain becomes sink limited; i.e., environmental conditions reduce ATP consumption and photosynthate production. Normally, violaxanthin transfers absorbed light energy to pigments associated with chl a, while zeaxanthin thermally dissipates the energy (Adams et al. 2004). In a review by Oquist and Huner (2003), decreased total chlorophyll content, increased VAZ pool, and increased ratio of V to A+Z are concluded as the primary means of NPQ in lodgepole pine (Pinus contorta). As the plastoquinone pool becomes more reduced from photochemically excessive light, transcription of genes coding for the light-harvesting polypeptides and reaction- 119 center D1 protein in PSII is repressed, while the chlorophyll-binding PsbS protein from PSII becomes involved in the de-epoxidation of V to A+Z (Oquist and Huner 2003). The reverse reaction, Z+A to V, can occur within minutes in nondorrnant leaves or hours in fully hardened leaves (e.g., 100 h for Pinus ponderosa; Verhoeven et al. 1999). Damage by cold, or lack of cold tolerance, can be measured by a number of destructive and non-invasive methods following natural (e.g., Nippert et al. 2004) or controlled (e.g., Burr et al. 1990) freezing events, and different methods are often used complementarily (Burr et al. 2001). The non—destructive measurement of chlorophyll fluorescence yields an indicator of quantum efficiency, or efficiency of the photosynthetic apparatus. In dark-acclimated leaves, all reaction-centers in the ETC are oxidized, thus the steady-state fluorescence (F0) can be measured with a fluorometer. A rapid pulse, or series of pulses, of saturating light then reduces all reaction-centers (i.e., QA in PSII), and photochemically excess light energy is reradiated as fluorescence, which is recorded as Fm. The ratio of (Fm-Fo)/Fm, or Fv/Fm, is a measure of the potential maximum PSII quantum yield, or efficiency of PSII (Schreiber et al. 1994). During the daytime, the quantum yield of open (oxidized) PSII centers, also phrased as the excitation capture efficiency of PSII, can be determined by first reducing GA with a weak far red light source that exclusively drives PSI and thus drains electrons from PSII, measuring steady-state fluorescence (Fo’), flashing a saturation pulse to measure Fm’, and calculating Fv’lFm’ as for dark-acclimated leaves (Schreiber et al. 1994). Both measures of variable fluorescence give an 120 indication of injury or reduction in function at the photosystem-level in the chloroplast as a function of cold or other stress. Another method commonly used to assess cold damage and hardiness is by visual inspection of tissues after an incubation period (7-10 days) following the freezing event (Burr et al. 2001). While survival into future years following freezing at certain temperatures is a better test of cold hardiness, industry and researchers demand more rapid results. Once-frozen tissues, such as needles, buds, and cambia, are examined, and injury is quantified as percent ‘browning’ or mortality. This method requires expertise to accurately gauge existence and extent of tissue damage, but it can produce results consistent with other methods (e.g., Lindgren and Hallgren 1993). Furthermore, survival after freezing is an integrated response to cold damage and recovery. However, this method requires destruction of stems and/or roots to examine cambial damage. Measurement of electrolyte leakage following cellular membrane disruption by ice formation is another destructive method for assessing damage by freezing temperatures. Prepared tissue samples normally release some amount of cellular electrolytes into a solution; after membrane destruction by freezing, the electrical conductivity (ECO) of the prepared tissue samples in solution increases. Cells are Iysed by heating to release the remaining electrolytes, and the solution is remeasured (ECf). Relative conductivity is calculated variably as ECO/ECf (Ritchie 1991) or (ECO-B1)/(ECf'BZ) where B1 and 82 are EC values for blanks before and after heating (Burr et al. 2001). This method is relative fast, inexpensive, but does not produce a measure of mortality 121 (Burr et al. 2001). Also, the results can vary significantly by variation in plant health status (Burr et al. 2001). Other methods for assessing cold injury include differential thermal analysis (DTA) and electrical impedance spectroscopy (EIS). DTA is useful for identifying the extent to which supercooling has been employed by various tissues by measuring the exothermic spike in response to spontaneous nucleation of extra- and intracellular supercooled water (Burr et al. 2001). A carefully controlled block or chamber contains both the tissue sample and an inert reference material against which tissue temperature is compared as block temperature decreases. DTA is less useful for some plant organs, especially during early and late dormancy, and must be standardized to another method (Burr et al. 2001). EIS works on the principle that different structures within the plant, or even cell, pose impedance to an applied electrical current. Currently, the biological meaning of EIS values is not well understood, but technological advances in interpreting results from this method have eliminated an actual freezing test (Burr et al. 2001). Populations of the closely related Southwestern ponderosa pine (Pinus ponderosa var. scopulorum Engelmannii) and Arizona pine (Pinus arizonica Engelmannii) are known to segregate along an elevational gradient that varies in range and absolute value depending on topographic features and local climate conditions in the Santa Catalina Mountains of southern Arizona (Dodge 1963, Whittaker and Niering 1965, Barton 1993). At higher elevation (2430-2650 m), Southwestern ponderosa pine is codominant to Douglas-fir (Pseudotsuga 122 menziesii Franco) and Southwestern white pine (Pinus strobiformis) (Whittaker and Niering 1965). Around 2430-2550 m elevation (Epperson et al. 2001), the dominant Ponderosae transitions to Arizona pine, which extends down to ~1760 m elevation (personal observation). This differential distribution suggests that cold winter conditions could limit the upward migration of Arizona pine. While cold tolerance tests have been conducted on ‘ponderosa pine’ (e.g., Burr et al. 1989, 1990), no such tests have been reported for Arizona pine. The goal for this study was to determine if there is a genetic difference in cold tolerance between Southwestern ponderosa and Arizona pine in the Santa Catalina Mountains that could lead to differences in elevational distribution. lntraspeciflc variation in cold tolerance has been noted in Pinus (Hawkins et al. 2003), so local source of plant material was used in this study; however, extrapolation to other mountains with this or other pairs of Ponderosae taxa should be relevant. Given that Southwestern ponderosa pine has a higher elevational distribution than Arizona pine, l hypothesized that the former species will have higher resistance, or lower critical temperature values (LTso, temperature at which 50% mortality or failure is detected), to freezing conditions. This study incorporates three separate experiments involving 1-year-old potted seedlings in highly controlled laboratory conditions, 4-year-old potted seedlings exposed to Michigan’s climate, and small trees in the Santa Catalina Mountains. 123 Methods Controlled whole-seedling freezing test In September-October 2005, multiple ripe cones were harvested from trees considered to be located within “pure” Southwestern ponderosa pine (“PIN_PON”) or Arizona pine (“PIN_ARI") populations in the Santa Catalina Mountains. PIN_PON trees were located near the summit of Mount Lemmon (“MTL,” 32.4396°N, 110.7871°W (NADB3/WGS84), 2770 m elevation), while PIN_ARI trees were located either near Rose Canyon (“RC,” 32.3964°N, 110.6932°W (NAD83NVGSS4), 2175 m elevation) or northeast of Lizard Rock (“LIZ,” 32.3844°N, 110.6930°W (NAD83NVGS84), 2135 m elevation). Cones were air-dried at room temperature in paper bags, and seeds were removed from the cones by shaking and removing cone bracts in early January 2006. Unopened cones were heated 14-27 h at 40° C to encourage opening (Young and Young 1992). Empty, parasitized, or otherwise unhealthy appearing seeds were discarded, while whole seeds were dewinged by hand, inventoried, and stored in labeled small manila envelopes at room temperature. Seeds were germinated in February-March 2006 as part of another component of this dissertation. Germinants were carefully transferred to ShortOne TreePots (9.5 cm wide, 24 cm tall, 1.64 L volume; Stuewe & Sons, Inc., Corvallis, OR) containing a 1:1 :3 (v/v) mix of perlite, steam-sterilized loamy sand, and Fafard52 (33% bark, 18% peat, 6% perlite, and 3% vermiculite; Conrad Fafard, lnc., Agawam, MA). After planting, 50 mL (~5 mm) of medium-grit crushed granite was spread onto the soil. Seedlings were grown for 10 months in greenhouses 124 (17-30° C daily) at Michigan State University (East Lansing, MI; 42.7225°N, 84.4763°W (NADB3NVG884), 259 m elevation) with natural light (~9 h in December), regular watering, and two applications (19 May and 8 October 2007). of slow-release granular fertilization (Osmocote Classic, 14-14-14, 3-4 mo release at 70° C; The Scotts Company, Marysville, OH). From the available stock, 68-70 healthy seedlings of uniform size (i.e., similar number of fully expanded fascicles) from each species were selected for this experiment, which will use 60 seedlings per species. There were 8 mother trees for PIN_PON and 5 mother trees for PIN_ARI, and 8-9 PIN_PON and 14 PIN_ARI seedlings were selected from each seed source (Table 4-1). The proportion of trees coming from each germination treatment (see Chapter 3) did not differ (PIN_PON, 12:1 .4412, df=2, p-value=0.4865; PIN_ARI, x2=1-9143. df=2, p-value=0.384). Table 4-1. Seedlings and source trees from the Santa Catalina Mountains used in the controlled whole-seedling experiment. No. trees No. trees Site Morphotype Source tree available used Mount Lemmon PIN_PON MTL 9 35 9 MTL 5 38 9 MTL 24 42 9 MTL 23 38 9 MTL 15 24 8 MTL 132 16 6 MTL 11 35 9 MTL 104 34 9 Rose Canyon PIN_ARI RC PINARI GroupPicnic 90 8 Lizard Rock PIN_ARI LIZ 1 24 8 LIZ 3 20 14 LIZ PINARI 353 from LIZ 2 40 19 LIZ PINARI 61.5cm dbh 53 19 Measurement of cold-induced injury throughout the acclimation and deacclimation process would yield a complete picture of differential seasonal 125 susceptibility to cold, or cold tolerance. However, in ponderosa pine, Burr et al. (1990) observed little difference in tissue hardiness prior to acclimation and large differences during deacclimation. Due to limited numbers of seedlings available for this experiment, investigation of cold-induced injury was restricted to three periods corresponding to the hardening, full acclimation, and deacclimation stages. The seedlings underwent simulated field conditions for hardening, winter, and dehardening (Table 4-2) following the methods of Burr et al. (1989). Seedlings were hardened for 7 weeks on short-days (SD, 10-h light, 300 umol m' 2 s") at 10°l3° C in a standard walk-in growth chamber (TCZ controller, Environmental Growth Chambers, Chagrin, OH). Winter conditions were imposed at 5°/-3° C SD (280-340 umol rn'2 s") for 4 weeks in a controlled growth chamber (Model GC-20 BDAF, ENCONAIR, Ecological Chambers, Inc., Winnipeg, Manitoba). Dehardening occurred under long-days (16-h natural light extended by sodium vapor bulbs) at 16°l25° C in a greenhouse. When needed, seedlings were equally watered at all stages with distilled water. Table 4-2. Conditions for hardening, acclimating, and deacclimating PIN_PON and PIN_ARI seedlings for the whole-seedling freezing test. Relative timing and date for the freezing tests in each period are noted in the last two columns. Day Night Freezing test Length Length Temp temp Stage (wk) Date (h) (° C) (° C) Week Date Hardening 7 22Dec-9Feb 10 10 3 5.4 29Jan Winter 4 9Feb-9Mar 10 5 -3 10 23Feb Deacclimation 2 9-23Mar 16 25 16 14 23Mar 126 Cold injury was inflicted by whole-plant freezing at 5.4, 10, and 14 weeks (Tables 4-2), and response to freezing was measured by before-and-after dark- adapted chlorophyll fluorescence (DACF), freezing-induced electrolyte leakage (FIEL), and visual scoring of needles, buds, and cambia. Twelve hours prior to starting each test, 20 randomly selected seedlings from each species were watered to field capacity to reduce effects of different plant—water status (J. Flore, personal communication), and attached needles from 3-4 fascicles from every plant were carefully marked with small pieces of tape at the fascicle sheath. PSII quantum efficiency, or DACF, was measured from the taped needles with a pulse-amplitude modulated fluorometer (6400-40 Leaf Chamber F Iuorometer, LI- COR Biosciences, Inc., Lincoln, Nebraska, USA) attached to the base Ll-COR 6400 Portable Photosynthesis System. The fluorometer was calibrated before the first set of measurements, zero-ed before each set, and adjusted to recommended settings: Saturating pulse (flash) - 0.8 s saturating multiple flash of ~8800-9000 umol m'2 s", 20KHz modulation, and 50 Hz averaging filter; no Dark pulse since dark-acclimated needles were measured; and Measurement light - intensity 2, 0.25 KHz modulation, 1Hz averaging filter, and 10 gain factor (Ll-COR 2005a). Taped needles were then scored under fluorescent lighting for color (hue, value, and chroma; Munsell Tissue Color Chart, 1977 ed.) as in Boorse et al. (1998). Aftenrvard, the seedlings were moved from the growth chamber (or greenhouse for the deacclimation treatment) to a completely dark room (20 °C) for at least 30 minutes before DACF was remeasured from the same taped needles to determine if DACF is a function of temperature or 127 photosynthetic down-regulation associated with physiological acclimation (B. Pratt, personal communication). In both the Hardening and Acclimation stages (Table 4-2), DACF was significantly (paired t-test, df=19, p<0.05) higher when measured at room temperature, thus all analyses use values recorded at room temperature. The 20 selected seedlings from each species were then moved directly to a programmable (Model CN-2042 TC, Omega Engineering, Inc., Stamford, Connecticut, USA) large-capacity freezer. The freezing rate, duration at test temperature, and rate of thaw were meant to simulate conditions experienced by rooted seedlings in the field during extreme cold events. Sixteen of the seedlings per species were randomly arranged within the freezer, a fan at each end of the freezer circulated air, and at least 10 thermocouples were distributed in the seedling crowns throughout the freezer to monitor temperature distribution and level; four seedlings per species served as controls and were moved to a cold room (1° C). The ambient freezer temperature was lowered rapidly to 0° C and then at a rate of 5° C h'1 (Cannell and Sheppard 1982, Burr et al. 1989) to each test temperature. The same amount of damage is incurred between 1 and 4 h at a given freezing temperature (Cannell and Sheppard 1982); consequently, after the average crown temperature stabilized for 2 h at each of four test temperatures (Table 4-3), four seedlings of each species were quickly removed from the freezer, placed in a cooler with ice packs, and immediately transported to the cold room (1° C) to thaw (Burr et al. 1989). 128 Table 4-3. Temperature profiles for each of the freezing tests outlined in Table 4-2. The bolded steps are those times after which four seedlings from each species were removed from the freezer. Hardening Stage Winter Stage Deacclimation Stage Step Duration (h) Temp (°C) Duration (hL Temp (°C) Duration (h) Temp (°C) 00 ~1 :30 Down to 0 ~1:30 Down to 0 ~1:30 Down to 0 01 1:00 0 1:00 0 1:00 0 02 3:00 -15 2:00 -10 0:36 -3 03 2:10 -15 2:10 -10 2:10 -3 04 1:00 -20 1:00 -15 0:36 -6 05 2:10 -20 2:10 -15 2:10 -6 06 1 :00 -25 1 :00 -20 0:36 -9 07 2:10 -25 2:10 -20 2:10 -9 08 1 :00 -30 1 :00 -25 0:36 -12 09 2:10 -30 2:10 -25 2:10 -12 After thawing for 12-21 h, depending on relative time of removal from the freezer, the seedlings were moved to the laboratory (~20° C) for the FIEL measurements. Two-four healthy-appearing and non-taped fascicles from each tree were excised at the base of the fascicle sheath with a razor blade and then cut into 1-cm segments on a dry glass plate. The seedlings were moved to the greenhouse (described above). The cut segments were quickly rinsed in deionized (Dl) distilled water (NANOpure Ultrapure Water System, Model D4741, Barnstead International, Dubuque, IA, USA) in a Petri dish, and then 40 segments were carefully removed with forceps and placed into a 55-mL culture tube (25x150 mm, Pyrex No. 9826, Coming Incorporated Life Sciences, Lowell, MA) containing 10 mL of DI water. A pilot test of seedlings frozen down to -15° C indicated that 20 segments produced an electrical conductivity (EC) reading 3.7 times lower than the EC meter’s calibration setting (447 118), thus doubling the number of segments would safely increase the absolute range and precision of EC measurements. The tubes were capped, vortexed, incubated for 12 h in 20° 129 C distilled water, vortexed again, and measured for initial EC (CON 410, Oakton Instruments, Vernon Hills, IL). The EC meter was recalibrated using standard calibration (447 p8) solution before every set of measurements. With caps loosened 45° to prevent explosion, the tube were incubated at 90° C in an oven for 2 h to Iyse the cells and then for 12 h in 20° C water before the maximum EC (Ritchie 1991) was measured. Marked tubes did not lose any noticeable solution to evaporation during the incubation. The relative conductivity (RC) of the solution was calculated as the initial conductivity value divided by the maximum conductivity value (Ritchie 1991). The seedlings in the greenhouse were evaluated at both 3 and 7 days following the freezing test by DACF and visual scoring to assess any recovery of photosynthetic function. Bud and cambia were evaluated for visible tissue injury 7 days following the freezing test (Burr et al. 1989). In natural light conditions, apical buds were removed by razor and cut longitudinally to determine if the bud were live (bright green) or dead (dull green or brown); percent mortality was calculated to compare against freezer temperature. Seedlings were measured for length from root-collar to base of apical bud, and most of the stem was shaved with a razor to expose cambium. Live cambium, using the control trees as a standard, was very bright green, while functionally dead cambium was dull green or brown. The length of stem with dead cambium was measured and converted to percent injury relative to the total length of stem. Determination of live or dead bud and cambium requires close inspection, expertise, and conference with an experienced person; L. Sage (Department of Horticulture, 130 MSU) provided the initial calibration, while F. Telewski assisted with every bud and cambial scoring test. PSII quantum efficiency (DACF), freezing-induced electrolyte leakage (RC, relative conductivity), cambial and bud mortality, and change in Munsell value after freezing were modeled using analysis of variance (ANOVA) in R (R Development Core Team 2007). The full models included the main factors of species and freezing treatment, as well as their interaction. For each phase, the critical temperature (LTso) was estimated for each response variable, except change in Munsell value, by visual interpolation from plots of the fitted values from the full model (‘predict’ function; R Development Core Team 2007). The y- axis value for estimating LT5o was determined as one-half of the mean species response in the control trees at 3° C for DACF and RC, while 50% was used for cambial and bud mortality. Observational whole-seedling freezing test In fall 2004, the nongovernmental organization named Trees for Mount Lemmon (contact: Barbara Eisele) donated 391 two-year-old “ponderosa pine” seedlings for our research purposes. The seeds were originally collected within the Santa Catalina Mountains (exact locations not recorded) by a private contractor for Coronado National Forest to replant areas affected by the Catalina Highway reconstruction (B. Eisele, personal communication) and grown in greenhouses near Flagstaff (W. Hart, personal communication). In early November 2004, I retrieved the potted (n=3-7 seedlings per small pot) seedlings 131 from outdoor storage at high elevation (Summerhaven School and Radio Ridge), counted the number of needles in every fascicle (2003-2004), labeled each seedling with morphotype (sensu Peloquin 1984) and a unique identifying number, and barerooted all seedlings with roots covered by moist paper toweling and packed into plastic bags in cardboard boxes. The 6 wardrobe boxes were shipped ground freight to Michigan State University. All seedlings were then potted within 2 days of arrival in a standard conifer soil mix (Fafard52, Conrad Fafard, lnc., Agawam, MA) in 2-gal plastic pots. The seedlings were kept outdoors for about one week and then moved to a greenhouse heated sufficiently through the winter to prevent ice formation in the water pipes. Daytime temperatures inside the greenhouse during the winter were usually 4-10° C, but high solar radiation raised the temperature to ~20° C on occasion. Through spring, daytime temperatures inside the greenhouse reached 32° C; trees were generally watered as needed, but high seedling mortality in the greenhouse suggested that watering had been insufficient and/or daytime temperatures had been too high. On 21 May 2005, all seedlings were moved outside the greenhouse, with supplemental watering as needed; survival was high during the summer. Temperatures remained above freezing until mild frost on 27-29 October and 14 November (NCDC 2007). Beginning 22 November, nighttime (except 29 November) and daytime (except 28-29 November) temperatures stayed below 0° C, even reaching -11° C on 25 November (NCDC 2007). Consequently, soil water in the pots had frozen 132 diurnally for about a week and then likely remained frozen for at least 2 days before measurements of PSII excitation capture efficiency began. Light-acclimated chlorophyll fluorescence (LACF; Fv’lFm’) was measured from every living seedling as they were moved into the heated (f =13° C) greenhouse from 2-5 December 2005. Seedling selection was determined by a disinterested assistant, thus the order of moving trees into the greenhouse for measurement was not advertently biased by morphotype. LACF was measured by the Ll-COR fluorometer, as described in the first experiment, but a 6-s dark pulse was added to measure Fo' with the following additional settings: far red intensity of 8 turned on 1 s before and remain on for 1 5 beyond actinic light off, 0.25 KHz modulation, and 1 Hz averaging filter (LI-COR 2005a). As before, the fluorometer was calibrated with a standard before the first set of measurements and zero-ed before each set. After securely clamping a sufficient number of non- overlapping healthy needles to completely cover the aperture in the fluorometer, measurements were initiated after base fluorescence had stabilized (dF/dT<5; Ll- COR 2005a). Temperature inside the cuvette was simultaneously measured by a thermocouple. LACF was modeled with morphotype as a fixed effect and temperature as a covariate by analysis of covariance (ANCOVA) in R (R Development Core Team 2007). Afterward, temperature was removed from the model and LACF was fitted to morphotype in a one-way analysis of variance (ANOVA). Given the unbalanced design, Type II sums of squares (SS; Langsrud 2003) were used to 133 test hypotheses using the ‘car’ (Companion to Applied Regression; Fox 2006) package in R (R Development Core Team 2007). Small trees on the mountain Response to natural variations in cold temperatures was measured for small PIN_PON and PIN_ARI trees in the Santa Catalina Mountains in January 2006. Since the small-diameter trees would have been severely damaged by coring, their ages were not determined; size does not correlate well with age for semiarid mountainous trees (T. Harlan, personal communication). However, standardizing for size across sites (elevation) should reduce effects related to ability to acquire water and to tolerate above-ground climate conditions. Trees selected for winter measurements at the MTL (n=3 PIN_PON), Palisade Rock (“PAL,” 32.4138°N, 110.7149°W (NAD83/WGSB4), 2475 m elevation; n=4 per species), and LIZ (n=3 PIN_ARI) sites were part of ongoing research discussed elsewhere in this dissertation. Trees at each site were measured within one day over 24-27 January. For each tree at both predawn and midday, the ratio of variable to maximum chlorophyll fluorescence (Fv/Fm and Fv’lFm’, respectively) was measured from the same attached 2005 needles (recently taped at the fascicle sheath) on the lowest living branch with a LI-COR fluorometer calibrated with a standard before the first set of measurements, adjusted to settings used in the previous experiment, and zero-ed before each set of measurements. Net photosynthesis (A) was simultaneously measured during the midday fluorescence measurements (LACF). At predawn, DACF was 134 measured once per tree, while LACF and A were measured 3-4 times from the same needles and averaged. Predawn and midday xylem water pressure potentials (l-Pp) were measured from detached 2005 needles from the same branch as the fluorescence measurements. WP give an accurate estimate of plant water status (Kaufmann 1968, Ritchie 1975), especially diurnal changes in status, and has been used to asses cell rupture by freezing (Brown et al. 1972). Five-ten fascicles per tree were cut at the base of the fascicle sheath with a razor blade, immediately placed in a small resealable plastic bag containing moist paper toweling, and stored in a cooler with ice packs prior to measurements. Within 2-6.5 h, each fascicle was cleanly cut with a razor ~1 cm from the basipetal end of the fascicle sheath, inserted into a rubber grommet, and pressurized with N2 gas in a pressure chamber (Model 1000, PMS Instruments, Corvallis, Oregon, USA) to determine the xylem pressure potential. Measurements were taken from 3-6 fascicles from each tree and averaged by observer and replicates to obtain a mean Wp per tree. DACF and LACF were modeled with site and species as fixed factors and temperature as a covariate by ANCOVA in R (R DeVelopment Core Team 2007). Aftenrvard, temperature was removed from the models, and DACF, LACF, lllp, and A were modeled using two-factor ANOVA in R (R Development Core Team 2007). The full models included the main fixed factors of site and species, but interaction was not included due to the small sample size at MTL and LIZ. Given the unbalanced design, Type II SS (Langsrud 2003) were used to test the ANOVA hypotheses using the ‘car’ (Companion to Applied Regression; Fox 135 2006) package in R (R Development Core Team 2007), while the default Type I SS were used for the ANCOVA tests. Results Controlled whole-seedling freezing test Freezing caused significant damage to the seedlings, especially beyond threshold temperatures (Figures 4-1 and 4-2). The threshold was most obvious during the Hardening stage when the difference between the controls (3° C) and the first freezing treatment (~15° C) was greater than for the other stages; this difference is likely due to a lack of temperature treatments closer to zero during the Hardening freezing test. In the ANOVAs, the temperature treatments were significant (p<0.0004, but p<0.002 for change in Munsell value) for all response variables, but the interaction of temperature treatment and species was not significant (p>0.523, but p>0.3306 for change in Munsell value). Given the lack of pattern and precision in the change in Munsell value, these data will not be reported. Little difference in response by species was observed within the Hardening and Deacclimation phases, except for a larger degree of cambial mortality in PIN_ARI seedlings from the -15° C treatment during the Hardening stage (Figure 4-2). The largest apparent species difference occurred during the Winter stage. In this stage, PIN_PON had higher DACF in every temperature treatment group (Figure 4-1) and lower cambial and bud mortality in the -15° and -20° C freezing treatments (Figure 4-2). With the exception of the -3° C 136 treatment, PIN_ARI also had higher cambial and bud mortality in the -15° and - 20° C freezing treatments during the Deacclimation stage (Figure 4-2). However, the ANOVAs indicated that the only significant (p<0.05) difference between species is the amount of cambial mortality during the Winter stage (Table 4-4). Table 4-4. ANOVA results for dark-acclimated chlorophyll fluorescence (DACF), relative conductivity (RC), and cambial (Camb) and bud (Bud) mortality for PIN_PON and PIN_ARI seedlings exposed to freezing treatments (Tx) during the acclimation phase. Significant (a<0.05) terms are bolded. Response Adj Rz F(p-value) Term Df SS F-value P-value DACF 0.7111 1.96E-10 Species 1 0.0904 2.3771 0.1319 Tx 1 3.6766 96.6332 9.80542 Species'Tx 1 0.0001 0.0037 0.9517 Residuals 36 1.3697 RC 0.708 4.36E-10 Species 1 0.0113 0.9321 0.3409 Tx 1 1.1360 94.0674 1.88E-11 Species*Tx 1 0.0015 0.1202 0.7309 Residuals 36 0.4227 Camb 0.6135 3.48E-08 Species 1 0.4655 5.4571 0.0252 Tx 1 5.0310 58.9809 4.3E-09 Species*Tx 1 0.0401 0.4707 0.4971 Residuals 36 3.0708 Bud 0.3182 0.0007 Species 1 0.2250 1.2903 0.2635 Tx 1 3.4490 19.7786 8.025-05 Species*Tx 1 0.0234 0.1343 0.7162 Residuals 36 6.2776 Seasonal effects, or differences between acclimation stages, on response to freezing temperatures can be demonstrated by comparing the regressions of the response variables against treatment temperature for all stages (Figure 4-3). Although the expected response is sigmoidal, the R2 for the linear regressions were generally high (3? =0.81, s=0.12). The interpretation of these plots is simpler when summarized by plots of slope and intercept against time during the acclimation process (Figure 4-4); no standard errors are associated with these 137 plots, thus significance testing was not performed. In general, slope values are related to the degree of hardening, or sensitivity to freezing; larger absolute values of slopes indicates a stronger response to freezing. In other words, slopes should converge to zero when the plant tissue has thoroughly hardened. Intercepts indicate the degree of injury expected at 0° C; i.e., the difference between intercepts suggests the (genetic) difference in cold tolerance between species. Slope values were similar for both species across the Hardening and Winter stages but were steeper during Deacclimation (Figure 4-4). This suggests that both species were close to the degree of hardening during the first stage as compared to the second stage, but both species were functionally deacclimating during the third stage and thus becoming more sensitive to freezing temperatures. There was no species difference in sensitivity to freezing as measured by DACF and RC, but cambium in PIN_PON appeared to be more sensitive across stages than in PIN_ARI. However, PIN_ARI buds were more sensitive than PIN_PON buds during the Winter and Deacclimation stages. As for the slope values, the intercepts for the two species in the DACF and RC plots were similar, with PIN_PON having slightly higher PSII quantum efficiency and vaguely lower electrolyte leakage than PIN_ARI during the Hardening and Winter stages (Figure 4-4). The hump-shaped DACF curve for PIN_PON is considered anomalous because 2 seedlings had relatively high DACF following the -10° C during the Winter stage and thus biased the regression curve (Figure 4-3). With significance (p=0.025) calculated by ANOVA 138 (Table 4-4), PIN_ARI seedlings had much higher cambial mortality than PIN_PON seedlings; furthermore, mortality increased through the acclimation process, peaking during the Deacclimation stage, like PIN_PON. Bud mortality also increased for both species through the acclimation process, with PIN_ARI seedlings experiencing higher bud mortality during the Winter stage than PIN_PON seedlings. Table 4-5 shows the results of the second approach to examining species differences in cold tolerance, with LTso’s interpolated from the fitted values from the ANOVAs. The mean LTso’s for whole PIN_PON seedlings across the three stages were -12.6, -12.9, and 96° C, while the mean LT5o’s for PIN_ARI seedlings were -11.1, -8.4, and -7.6° C. PIN_PON was consistently more cold tolerant across measured variables than PIN_ARI, especially during the Winter (A=-4.4° C) when both species should have been fully hardened. The rank-order of variables according to their insensitivity to freezing (i.e., lower LT50) changed from buds, cambia, and needles (DACF/RC) during Hardening and Winter to buds, needles (DACF/RC), and cambia during Deacclimation. The species differences are generally supported by the previous method. PIN_PON had lower LT5o's for loss of PSII quantum efficiency during Hardening and Winter, for electrolyte leakage and cambial mortality during all stages, and for bud mortality during the Winter stage. 139 —o— PIN_PON —e— PN_PON \ --B—-PIN_ARI 0.1 * ——a—-PIN_ARI 0" 1 Y {31‘} ' ir fir fir f 0% 1 -40 -30 -2o -10 o 10 40 so -20 -10 o 10 o 9 -—e——- PIN_PON — —a - — PIN_ARI I —e— PIN_PON r 0.1 l 01* --B--PIN_ARI l I I . as A fl . . f a . 9e . .30 -25 .20 .15 .10 .5 o 5 -30 .25 -2o -15 .10 -5 o 5 o 9 o a I I a 0.7-1' ’ ' ”#0 —o—PIN_PON ———o——-—PIN_PON 0,1 . --B—-PIN_ARI -—B—-PIN_ARI ‘ - as fl nn '15 ‘10 '5 ° 5 -15 -1o —5 o 5 Terrperature (° C) Terrperature (°C) Figure 4-1. Dark-acclimated chlorophyll fluorescence (left) and relative conductivity (right) for PIN_PON and PIN_ARI after different freezing treatments across three stages of cold acclimation: Hardening (top), Acclimation (middle), and Deacclimation (bottom). Note the varying temperature scale between phases. 140 \ so \\ w \ 70 \ so it 5. \ \ 4o \ \ so \ —c— PIN_PON - Carib \ 2° —l:i— PIN_ARI - Cunt: \ +PN_PON - Bud \ 1o — I- - PN_AR| - Bud \3 = o 35 so 25 -2o 15 10 s o 5 ~c— PN_PON - Cunt) —D— PN_AR| - Cami: —O—PIN_PON - Bud — I- - PN ARI- Bud -30 -25 -20 -15 -10 -5 0 5 100 so so 70 so so 40 .r w + PIN_PON - Camb - - —CI—-— PIN_ARI - Camb " 3° —e— PIN_PON - Bud 10 — I- - PIN_ARI - Bud o -15 -10 -5 Ten-persure (’ C) Figure 4-2. Cambial and bud mortality for PIN_PON and PIN_ARI after different freezing treatments across three stages of cold acclimation: Hardening (lap). Winter (middle), and Deacclimation (bottom). Note the varying temperature scale between phases. 141 <> PIN_PON - Hardening - y = 0.0261x + 0.6365. R2 = 0.8543 PIN_PON - Winter - y = 0.0308x + 0.7797. R2 = 0.8981 ‘* 0-9 e PIN_PON - Deaccimaiion - y = 0.0516x + 0.5725. R2 = 0.8384 _//<> 0.8 g D PIN_ARI - Hardening - y = 0.0237x+ 0.5798. R2 = 0.8617 0 7 g ' O PIN_ARI - Winter - y = 0.0297x + 0.6322, R2 = 0.8928 g I PIN_ARI - Deaccimalion - y = 0.0575x + 0.6373, R2 = 0.8639 0.6 E A E 0.5 if, g ‘8 “c. 0.4 B 0.3 E 3? 0.2 E a o 0.1 a 1 0.0 -35 -30 -25 ~20 -15 -1o -5 o 5 Temperature (°C) ~ 0.8 ‘\\ x '1' 0.7 9 we a~ \ “- \ I ‘~ 9 . - \\ I r 0.5 \ \ \ e \ \o\ g. x ‘ ~ \ .2 \ x ‘ \D \ ‘ 0.5 g \ \ O x ‘ ‘2 ‘.\ ‘ ~ ~. \ 3 \ \ ;‘ \ ‘ ‘\ \ .0- o 4 g o PIN_PON - Hardening -y= -0.017x+ 0.2186, R2 = 0.9786 \ \ “~\ g \ ‘\ PIN_PON - Winter - y = -0.017x + 0.1646, R2 = 0.8658 " \ \ x ‘ -\ t 03 a: \ . O PIN_PON - Deacclimation - y = 00344;: + 0.2285, R2 = 0.8242 \ \ . t ‘ 0 2 . \ . x _ i3 PIN_ARI - Hardening - y = -0.0172x + 0.2311, R2 = 0.9773 5 \ PIN_ARI - Winter - y = 00183:: + 0.1764, R2 = 0.8734 ’ 0.1 I PIN_ARI - Deeccimation - y = 00395x + 0.2168, R’ = 0.8246 I T T I l T r I 1 0.0 -35 -30 -25 ~20 -15 -10 -5 0 5 Temperature (°C) Figure 4-3. Linear regression of dark-acclimated chlorophyll fluorescence, relative conductivity, and cambial and bud mortality against treatment temperature for PIN_PON (solid line) and PIN_ARI (dashed line) across three stages of cold acclimation: Hardening (thin line), Winter (normal line), and Deacclimation (thick line). Note the varying temperature scale between phases. 142 PIN_PON - I-hrdening - y = -3.52x- 0.5041, R2 = 0.6912 PIN_PON - Winter- y = -3.9142x - 4.1281, R2 = 0.7454 PIN_PON - Deacclimation - y = 4.7165x + 41.922, R2 = 0.8336 PIN_ARI - Hardening -y= -3.1349x+16.912, R2 = 0.8971 PIN_ARI - Winter - y = -3.2172x + 30.787, R2 = 0.8609 PIN_ARI - Deacclimation - y = 4.3392x + 51.929. R2; 0.6226 <> PIN_PON - Hardenino- y = -3.433x - 4.735, it2 = 0.7242 - PIN_PON - Winter - y = -2.6995x + 3.827, R2 = 0.4074 7 -30 T -25 Y -20 V -15 Temperature (°C) -10 -35 Q p|N_poN - Deacclimation - y=-3.433x-4.735, R2 = 0.7242 \ . CI PIN_ARI - Hardening - y = 4.556x + 34.826, R2 = 0.7999 ’ \ \ a pm AR. - Winter - y = -3.2242x + 15.796, R2 = 0.7834 _ \ \ I pm AR. _ Deaccmnafion - y = -5.9685x + 32.77, R2 = 0.8252 ‘ - m \ n y 1 T V l..] I v -30 -25 -20 -15 -10 -5 0 Temperature (°C) Figure 4-3. Continued. 143 100 100 90 80 70 60 50 40 Cambial mortality (%) Bud mortality (%) 0.07 1 «r 0.9 «r 0 8 A 0.08 - ‘83 1. 07 < e 0.05 < i 0.04 i . o 5 .9 a 5 0.03 J " 0" 8 ----- 8’ PIN PON-Slope 1 03 ‘5 0.02 1 " I _§- 0 PIN_ARl-Slope - 0.2 V) °-°1 ‘ o PIN_PON- ntercept, 01 0 . 5 PIN Ant-intercept (o 0 5 1o 15 Weeks since Dec 22 0 5 10 15 0 . 1 60 0 PN_PON - Slope . A 1:1 PN_ARI- Slope / ’ 5° ‘5’ -1 : / 8 O PN_PON - meroept / I .. .2 -2 1 3° g » 20 8 '3 i 10 ‘5 § .. o (7) 41 -10 5 ~ -20 Intercept of reoression lines (DACF) Intercept of rearession lines (Cambl Slope of regression lines (RC) Slope of regression lines (Bud) Weeks since Dec 22 5 10 15 -0.02 * 0.25 .0 N 1- 0.1 A r O _. o PIN_PON-Slope 1:1 PIN ARI-S -003 1 ' b” e PIN_PON- intercept I PIN_ARI- htercept \ -004 — Cl 0 o o -11 D -21 O I _3 -1 _4 4 _5 -4 -6 .. 5 o 8 Intercom of regression lines (RC) Intercept of regression lines (Bud) Figure 4-4. Slope (thin line) and intercept (thick line) for regressions of dark-acclimated chlorophyll fluorescence, relative conductivity, and cambial and bud mortality against treatment temperature for PIN_PON (solid line) and PIN_ARI (dashed line) across weeks of cold acclimation. Note the varying axis scales across graphs. Table 4-5. Cold tolerance (LTso) of PIN_PON and PIN_ARI estimated as a function of needle dark-acclimated chlorophyll fluorescence (ndl - DACF), needle relative conductivity (ndl - RC), and cambium and bud mortality across three stages of cold acclimation: Hardening, Winter, and Deacclimation. Hardening Winter Deacclimation Response PIN_PON PIN_ARI PIN_PON PIN_ARI PIN_PON PIN_ARI Ndl - DACF -11.25° C -10.0° C -12.25° C Ndl - RC -9.0 -8.0 -8.5 Cambium -14.25 -10.5 -13.75 Bud -16.0 -16.0 -17.0 -8.5° C -7.25 -6.75 -11.25 -12.5° C -12.25° C -9.25 -6.0 -6.25 -2.5 -10.5 -9.75 144 Observational whole-seedling freezing test The quantum yield of open PSII reaction centers, measured by LACF (Fv’lFm’), varied from 0.07 to 0.399 with a mean of 0.270 (s=0.053) across all seedlings (n=163) (Figure 4-5). Most of the data are slightly lower than the theoretical overall quantum yield of unstressed plants under incident light (Fv’lFm’=0.3-0.35; Schreiber et al. 1994). Mean (:80) quantum yields for PIN_PON, Mixed, and PIN_ARI were 0.282 (1:0.051), 0.255 ($0.051), and 0.241 (1:0.057), respectively. Air temperature when the seedlings were measured (p=0.105) and its interaction with morphotype (p=0.524) did not significantly affect LACF in the ANCOVA (Table 4-6). However, there was a highly significant (p=0.0009) effect of seedling morphotype on the quantum yield. Pain/vise comparison across morphotypes using Tukey’s HSD from a one-way ANOVA of morphotype indicated that LACF measured from PIN_PON seedlings was significantly higher than from both Mixed (p=0.023) and PIN_ARI (p=0.001), but LACF measurements from Mixed and PIN_ARI were not significantly different (p=0.569). 145 0.5 1 <> PIN_PON- y = 0.0012x + 0.2644 0 Mixed - y = 0.0002x + 0.2376 0.4 1 o 0 CI - = PIN_ARI y 0.0038x+ 0.1461 O o 00 o E 0.3 4 it: '63 5 0.2 + 0.1 1 o 0 I r T I 0 5 10 15 20 Temperature (° C) Figure 4-5. PSII quantum yield (LACF) as a function of air temperature when measured from 3-year-old potted seedlings with frozen soil water: PIN_PON (solid line), Mixed (dotted line), and PIN_ARI (dashed line). Coefficients are from full ANCOVA model. Table 4-6. Type II tests of ANCOVA for PSII quantum yield (LACF) fitted to morphotype and the covariate air temperature (Temp) for 3-year-old potted seedlings with frozen soil water. Significant (a<0.05) terms are bolded. Adj R2 F(p-value) Term Df SS F-value P-value 0.0876 0.0016 Morphotype 2 0.0372 7.3279 0.0009 Temp 1 0.0067 2.6570 0.1051 Morphotype *Temp 2 0.0033 0.6481 0.5244 Residuals 157 0.3980 Small trees on the mountain The potential maximum PSII quantum yield, measured as DACF (Fv/Fm), varied from 0555-0746 across all sites, with the highest means recorded at MTL and LIZ, while the realized PSII quantum yield under incipient light, measured as LACF (Fv’lFm’), ranged from 0300-0509, with the trees at the intermediate PAL site exhibiting the highest Fv’lFm’ readings (Figure 4-7). The values of DACF 146 were larger than of LACF because the PS reaction-centers are not fully oxidized thus increasing the relative minimal yield (F0) and heat dissipation, which reduces maximum fluorescence (Schreiber in Larcher 2005). Consequently, the ratio of the difference between F0 and Fm over Fm decreases accordingly. 0.8 1 Q . I. 0.7 — ' o 0.6 - ' .I i 8 o 8 0 5 <> 0 t5 ' a o O 3 0.44 <5 “7743““ E ~~~~~~~~ 1:11. a. O E] g 0.3 ~ 0 5 o PIN_PON - predawn - y = -0.0039x + 0.6779, R2 = 0.0164 0.2 - I PIN_ARl - predawn - y = -0.0043x + 0.6764, R2 = 0.0201 0.1 T o PIN_PON-midday-y=0.0048x+0.3957,R2=0.0102 n [:1 PIN_ARI-midday-y=-0.0111x+0.5054, R2=0.7343 -2 0 2 4 6 8 10 12 14 16 Temperature (° C) Figure 4-7. Variable chlorophyll fluorescence (FvlFm, solid; Fv’lFm’, empty) as a function of leaf temperature for PIN_PON and PIN_ARI trees combined across sites in January 2006. As a function of temperature, little difference in the ratio of variable to maximum chlorophyll fluorescence is apparent between species within time periods (Figure 4-7). However, the three midday measurements taken from PIN_ARI trees at LIZ at lower elevation (and thus higher temperature) bias the linear regression toward a negative slope (-0.0111) and larger intercept (A=0.1097). Consequently, temperature was a significant (p=0.038) covariate for the midday variable fluorescence measurements (Table 4-7). The interaction of 147 temperature and site were also significant terms for the midday xylem pressure potential (W,,) and net photosynthesis (A) models. However, temperature varied consistently by site, which generated 5-56 times higher SS than temperature across all of the ANCOVA models, and was thus confounded site; consequently, temperature was dropped from all further analyses. Table 4-7. ANCOVA results for predawn dark-acclimated chlorophyll fluorescence (pDACF), midday light-acclimated chlorophyll fluorescence (mLACF), predawn (ppsi) and midday (mpsi) xylem pressure potential, and midday net photosynthesis (mphoto) for PIN_PON and PIN_ARI trees at three sites (MTL, PAL, and LIZ) in the Santa Catalina Mountains in January 2006. Significant (a<0.05) terms are bolded. Response Adj R2 F(p-value) Term Df SS F-value P-value pDACF -0.3344 0.9644 Species 1 0.0001 0.0332 0.8580 Site 2 0.0010 0.1353 0.8746 Temp(covariate) 1 0.0000 0.0103 0.9206 Species'Temp 1 0.0000 0.0010 0.9747 Site'Temp 2 0.0053 0.7112 0.5079 Residuals 14 0.0518 mLACF 0.5936 0.0037 Species 1 0.0045 3.6257 0.0777 Site 2 0.0331 1 3.3687 0.0006 Tem p(covariate) 1 0.0065 5.241 7 0.0381 Species'Temp 1 0.0009 0.7450 0.4026 Site'Tem p 2 0.0016 0.6595 0.5324 Residuals 14 0.0173 ppsi 0.5209 0.0106 Species 1 0.4541 9.3342 0.0086 Site 2 0.9304 9.5616 0.0024 Temp(covariate) 1 0.0412 0.8477 0.3728 Species'Temp 1 0.0097 0.1999 0.6616 Site'Temp 2 0.0158 0.1627 0.8514 Residuals 14 0.6811 mpsi 0.7777 0.0003 Species 1 0.4700 20.9632 0.0006 Site 2 0.9589 21.3854 0.0001 Tem p(covariate) 1 0.0170 0.7583 0.4001 Species'Temp 1 0.0004 0.0167 0.8992 Site‘Temp 2 0.2008 4.4782 0.0352 Residuals 14 0.2690 mphoto 0.6589 0.0012 Species 1 12.2521 13.8766 0.0023 Site 2 16.9186 9.581 0 0.0024 Temp(covariate) 1 0.3441 0.3897 0.5425 Species'Temp 1 3.3358 3.7781 0.0723 Site'Temp 2 9.1430 5.1 777 0.0207 Residuals 14 12.3610 148 Since the interaction of species and site was not calculated by R due to the small number of trees (n=3) measured at each site, the resultant ANOVA models contained only the fixed factors of species and site. Using all of the data, the effect of species was not significant (p<0.05) in any model (Table 4-8). The significant effect of site on 4", and A are apparent in Figures 4-8 and 4-9, with trees at lower elevation having lower W9 and A. Predawn and midday Wp varied significantly (paired t—test, t=5.1532, df=19, p=5.65E-05) across all sites, with the largest difference observed within both species at PAL (Figure 4-8). However, analysis of the midday LACF and A models indicated a few major outliers. An MTL tree had 36% higher Fv’lFm’ than the other two trees, while a different MTL tree had 53% lower A than the other two, and one LlZ tree had 4% of the A as the other two trees. Despite consistent Fv’lFm’ readings, this LIZ tree had highly variable A ()7 =0.253, s=0.759). When these data points were removed, the species effect shifted slightly toward significance for both models (Table 4-8). 149 Table 4-8. Type II ANOVA results for predawn dark-acclimated chlorophyll fluorescence (pDACF), midday light-acclimated chlorophyll fluorescence (mLACF), predawn (ppsi) and midday (mpsi) xylem pressure potential, and midday net photosynthesis (mphoto) for PIN_PON and PIN_ARl trees at three sites (MTL, PAL, and LIZ) in the Santa Catalina Mountains in January 2006. The numeral “2" following a response name indicates removal of outlier data points. Significant (a<0.05) terms are bolded. Response Adj Rz F(p-value) Term Df SS F-value P-value pDACF -0.1441 0.9483 Species 1 0.0001 0.0215 0.8851 Site 2 0.0010 0.1578 0.8852 Residuals 18 0.0571 mLACF 0.5189 0.0009 Species 1 0.0062 4.2386 0.0543 Site 2 0.0331 11.2944 0.0007 Residuals 18 0.0264 mLACF2 0.6782 4.892e-05 Species 1 0.0062 6.0578 0.02484 Site 2 0.0421 20.5039 2.95E-05 Residuals 17 0.0174 ppsi 0.5908 0.0002 Species 1 0.0084 0.2027 0.6579 Site 2 0.9304 11.1955 0.0007 Residuals 18 0.7479 mpsi 0.6981 5.13E-05 Species 1 0.0051 0.1668 0.6884 Site 2 0.9589 15.7453 0.0002 Residuals 18 0.4872 mphoto 0.4595 0.0026 Species 1 3.9908 2.8524 0.1085 Site 2 16.9186 6.0462 0.0098 Residuals 18 25.1839 mphotoZ 0.4673 0.0042 Specles 1 3.9908 5.9438 0.0268 Site 2 3.9617 2.9502 0.0812 Residuals 18 10.7429 150 MTL - PIN_PON PAL - PIN_PON PAL - PIN_ARI LIZ - PIN_ARI I Predawn El Midday Xylem pressure potential (MPa) Figure 4-8. Xylem pressure potential (l-Pp) from excised needles of PIN_PON and PIN_ARI trees at three sites (MTL, PAL, and LIZ) in the Santa Catalina Mountains in January 2006. Values are means; bars are :l:1 standard deviation. Net photosynthesis (pmol m'2 s") o r l 1 I MTL - PIN_PON PAL - PIN_PON PAL - PIN_ARI LIZ- PIN_ARI Figure 4-9. Midday net photosynthesis (A) results for PIN_PON and PIN_ARI trees at three sites (MTL, PAL, and LIZ) in the Santa Catalina Mountains in January 2006. Values are means; bars are :1 standard deviation. 151 Following removal of the outlying data points, the species difference in midday LACF and A become more apparent (Table 4-9). PIN_PON trees had higher potential and realized PSII quantum yield and net carbon acquisition, while leaf water status was more favorable (i.e., less negative) than in PIN_ARI trees. All responses indicate higher light and water efficiency by PIN_PON. Table 4-9. Mean responses by PIN_PON and PIN_ARI trees by site (MTL, PAL, and LIZ) and across sites (Combined) following removal of three outlying data values. Responses listed in Table 4-8. Standard deviations in parentheses. Significant (p<0.05) responses (Table 4-8) are bolded. PIN_PON PIN_ARI Response MTL PAL Combined PAL LIZ Combined pDACF 0.68 (0.49) 0.66 (0.07) 0.68 (0.06) 0.67 (0.05) 0.66 (0.05) 0.66 (0.05) mLACF2 0.32 (0.02) 0.46 (0.04) 0.43 (0.07) 0.35 (0.03) 0.42 (0.02) 0.40 (0.04) ppsi -1.3 (0.3) -1.6 (0.1) -1.5 (0.2) -2.2 (0.4) -1.6 (0.1) -1.8 (0.3) mpsi -1.3 (0.2) -1.9 (0.1) -1.8 (0.3) -2.3 (0.3) -2.0 (0.1) -2.1 (0.2) mphoto2 6.2 (0.6) 6.0 (0.9) 6.0 (0.8) 3.4 (1.2) 5.0 (0.7) 4.7 (1.0) Discussion Control/ed whole-seedling freezing test Differential cold tolerance of organs across the three “seasons” of cold hardiness was expected (Sakai and Larcher 1987). In this study, needle tissue was less cold tolerant than cambial and bud tissue during the Hardening and Winter stages, while cambium became the least cold tolerant during Deacclimation (Table 4-10). Bud tissue was the most hardy until Deacclimation when its LTso was similar to that of needle tissue. In fact, little difference in needle cold hardiness across seasons was observed, with species differences only during the Winter stage (Table 4-5). In contrast, Burr et al. (1990) calculated much lower LT50 values as well as a 3° C increase in needle cold tolerance from 152 hardening through simulated winter, and a subsequent 8° C decrease in cold tolerance during deacclimation (Figure 4-10). Their LT50 was estimated by interpolating 50% index of injury from an inverted, modified Gauss sigmoid model of electrolyte leakage against test temperature (Burr et al. 1990), while this study estimated LT50 for needles by interpolating the temperature from a linear regression at 50% of the mean control (3° C) response. Low response values at higher test temperatures increase the (-) slope (Figure 4-3), thus increasing the interpolated LTso value, which could contribute to the higher LT50’s observed in this study in comparison to Burr et al. (1990). Consequently, tissue and species differences in LT5o’s are more informative than their absolute values. Table 4-10. Cold tolerance (LTso) of combined PIN_PON and PIN_ARI tissues from this study and from Burr et al. (1990) across three stages of cold acclimation: Hardening, Winter, and Deacclimation. The Winter value for Southwestern ponderosa pine needles for Burr et al. (1990) was interpolated between two values. Response Hardening Winter Deacclimation Needle -9.6° C 91° C -10.0° C Needle (Burr et al. 1990) -18.7 [-22] -13.6 Cambium -12.4 -10.2 -4.4 Bud -16.0 -14.1 -10.1 The lower cold tolerance of needle tissue by both PIN_PON and PIN_ARI during the Hardening and Winter stages suggests that hardening, and thus photosynthetic downregulation, had not fully occurred. The PSII quantum efficiency for the combined species (0.78) in the control treatments was close to the theoretical optimum of 0.832 for healthy plants (Bjérkman and Demmig 1987); the mean value dropped to 0.71 during deacclimation. However, PIN_PON had 4-6% higher PSII quantum efficiency in the first two stages and 153 then 10% lower efficiency during deacclimation than PIN_ARI in the control groups (Figure 4-1). Photosynthetic downregulation in conifers is inducted by a combination of environmental influences, including day length, temperature, and radiation, in concert with respiratory demand for processed carbon. If the plant remains metabolically active, then reduction in photosynthesis becomes less sink regulated. While the temperature and photoperiod in this study closely followed Burr et al. (1990), l was unable to maintain the same light level. Consequently, leaf-level light was ~300 pmol m'2 s'1 for the Acclimation and Winter stages, with natural lighting exceeding this level during Deacclimation; in contrast, Burr et al. (1989, 1990) sustained 518 pmol m'2 s", which is closer to the light saturation value (600 umol m'2 s") reported by Helms (1972) for mature ponderosa pine trees. Although pigment analysis was not conducted, I would expect little conversion of V to A+Z. Instead, the seedlings in this study were photosynthetically active to maintain respiratory carbon demand, thus the rate of hardening relative to other tissues was delayed with slow recovery during deacclimation (Table 4-5). Resistance to freezing temperatures by the bud tissue was consistently higher than by the cambium across all stages of cold tolerance (Tables 4-5 and 4-9). The method for examining but not measuring injury was the same across these two tissues; response was continuous for cambium and binary for bud. Consequently, the error distribution should have varied, yet the data were parallel within species (Figure 4-2), suggesting only a difference in tissue cold tolerance 154 and not a problem with data analysis. Sakai and Larcher (1987) note that bud resistance develops slightly later than needle resistance; this study found a 3-5° C increase in resistance by needles than buds, again suggesting that photosynthesis was not fully downregulated. Cambial tissue of PIN_ARI had successively more injury across acclimation stages even in the control groups, suggesting that sustained cold growing conditions can impose substantial mortality in PIN_ARI seedlings. Observational whole-seedling freezing test Despite having severely reduced water supply due to the frozen soil, and likely frozen stern storage, the 3-year-old seedlings had remarkably high PSll quantum efficiency (5 =0.270, s=0.053). On a colder day (-8 to -10° C versus mean maximum -2° C), Verhoeven et al. (1999) measured Fv’lFm’ for detached ponderosa pine needles between 015-020. PSII quantum efficiency should decrease with temperature, and ponderosa pine has been noted to retain some photosynthetic capacity through winter (Marshall et al. 2001). However, given the duration (week) of subfreezing maximum temperatures preceding this test, extracellular and xylem freezing was expected to have reduced PSII quantum efficiency beyond that observed in this study. Incipient light levels were not measured, but the seedlings were exposed to at least diffuse bright to sunny days. The “leaf temperature” plotted in Figure 4-5 represents that measured by the thermocouple inside the fluorometer’s cuvette, which was representative of air temperature inside the greenhouse. Seedlings were measured immediately 155 (i.e., within 1.5 min) as they were transferred to the lower-light greenhouse (13° C); the time between outdoor and greenhouse temperature did not likely raise inner leaf temperatures sufficiently to affect PSll efficiency. This supposition is borne out by the lack of significance of the temperature term (Table 4-6). Verhoeven et al. (1999) recorded an increase in the ratio F.,/Fm of 0.23 15 min after moving samples from -23° C to room temperature, a much larger change in temperature of approximately 43° C. Consequently, temperature likely had little effect on PSII quantum efficiency observed in this study. The significant finding from this test is that PIN_PON seedlings exhibited higher PSII quantum efficiency than the Mixed and PIN_ARI seedlings (Table 4- 6). In addition, the Mixed trees, which may be hybrids between PIN_PON and PIN_ARI (Peloquin 1984, Rehfeldt et al. 1996, Epperson et al. 2001), had a mean PSII quantum efficiency intermediate to the two purported parent species, although the difference to PIN_ARl was not significant (p=0.569). lnterrnediate cold tolerance has also been observed in white x Himalayan blue pine hybrids (Lu et al. 2007). Small trees on the mountain Removal of three significant outlier data points improved the demonstration of differential realized PSII quantum yield and net carbon acquisition by ‘small’ PIN_PON and PIN_ARI trees growing in the Santa Catalina Mountains. Although the sample sizes were small (n=2-3) at the elevational 156 extremes, sufficient number of sympatric trees were present at PAL to detect a species effect. The pattern of net photosynthesis, in combination with the xylem pressure potential and chlorophyll fluorescence data, across sites suggests different limiting factors to survival during the winter in these mountains. The mean maximum PSII quantum yields for both species were lower than recorded by the hardening and acclimated seedlings in the controlled laboratory experiment, but similar to the deacclimating seedlings. This suggests that the Santa Catalina Mountain trees were more innately hardened but yet able to utilize light resources during the day. The PIN_ARI trees at LlZ had higher net photosynthesis but also larger change in needle water status than the PIN_ARI trees at PAL; the higher temperatures lower in elevation stimulate higher rates of respiration and thus demand for carbon gain, despite the more xeric conditions at LlZ. The PIN_PON trees at the high-elevation MTL site also exhibited higher net photosynthesis than those at PAL but had much lower realized PSII quantum efficiency with no change in needle water status; these trees are likely much more hardened than those at PAL, suggesting a strong site effect (Table 4-7). However, gas exchange is occurring, as evidenced by substantial net photosynthesis and diurnal increases in xylem pressure potential (Table 4-9). Most obvious when compared at PAL, PIN_PON trees have higher PSII excitation capture efficiency, higher net photosynthetic rates, and smaller changes in needle water status, a critical component to surviving winter desiccation (Sakai and Larcher 1987), than do PIN_ARI trees. 157 Implications of differential cold tolerance on distribution All of the experiments in this study have demonstrated that PIN_PON is more cold tolerant than PIN_ARI, while only the controlled laboratory experiment quantified differences in cold tolerance. Using the most sensitive (i.e., highest LTso) tissues to calculate species differences, one finds a 1°, 125°, and 075° C difference across the three acclimation stages for needle electrolyte leakage for the first two values and bud mortality for the third value. Given that needle and bud primordia in Pinus are similarly tolerant to cold (Sakai and Larcher 1987), a mean of 1° C difference can be assumed from this experiment. This seemingly minor temperature differential translates to 133 m in elevation, using Shreve’s (1915) data of an adiabatic lapse rate of 75° C per 1000 m elevation for the Santa Catalina Mountains. Given a current upper elevational boundary for PIN_ARI of 2548 m on the south face of Mount Lemmon (Epperson et al. 2001) and the 133-m differential predicted by differences in cold tolerance in this study, then the upper boundary for PIN_PON should be ~112 m below the very peak of the mountain. In fact, on this aspect, PIN_PON extends to the top and north side of the mountain, suggesting that the difference in cold tolerance estimated by this experiment is conservative. Instead, assuming that the top of the mountain is the upper elevational boundary for PIN_PON, then the differential cold tolerance would be 18° C, based on Shreve’s (1915) model. This difference is not very large nor far from the species differences estimated from the controlled laboratory experiment (Table 4-5). In fact, the difference in cold tolerance as related to elevational difference will become moot if winter temperatures in the 158 region increase by the projected 2.8° C by 2100 (USEPA 1998). Kupfer et al. (2005) predict an upward migration of communities with a 53% loss in montane forest on Madrean mountain islands such as the Santa Catalina Mountains given projected local climatic changes (USEPA 1998). Consequently, assuming minimal dispersal limitation and disturbance (or lack of disturbance), winter conditions at the top of the Santa Catalina Mountains will become suitable for PIN_ARI within this century. 159 CHAPTER 5 THE ROLE OF DROUGHT TOLERANCE IN STRUCTURING PONDEROSAE DISTRIBUTION IN THE SANTA CATALINA MOUNTAINS OF ARIZONA Introduction While tolerance to cold conditions in determining high latitude or elevation limits in pine distribution is well known, the cause of the lower elevation boundary is less well understood (Rundel and Yoder 1998). The tacit assumption is that intolerance to more xeric conditions, or drought, constrains regeneration at lower elevation (Barton 1993). Daubenmire (1943) showed that soil drought is the primary determinant for lower elevation distribution for Rocky Mountain conifer seedlings, while Haller (1959) and Yeaton et al. (1980) demonstrated the lethal effect of drought stress on ponderosa pine (Pinus ponderosa) seedlings. However, the extensive range of habitats in which ponderosa pine can be found implies that a suite of genetically controlled and phenotypically plastic mechanisms have evolved that allow survival in a variety of soil-water conditions (Rundel and Yoder 1998). Plants must trade-off the maintenance of whole-plant water status for the requirement of the carbon substrate for photosynthesis. The entry of carbon into the plant requires concomitant loss of water through transpiration at the stomata. As air temperature increases, the water vapor gradient increases at the stomata. thereby increasing stomatal conductance (9.), or loss of water through the stomata. This increases the tension (negative pressure, \Pp) of water within the root-stem-leaf pathway, which is exacerbated by low soil water availability. 160 When WP becomes sufficiently negative, cavitation, or air infusion into the xylem, occurs and an embolism is formed (Sperry and Tyree 1990). Cavitation results in a sustained increased resistance to water transport in those vessels or tracheids and, if sufficient amount of hydraulic conduits cavitate, wilting with eventual plant death (Tyree and Sperry 1988). The most direct means of preventing cavitation is by stomatal closure (Sperry et al. 1993) and recovery of plant water status through water uptake in the roots. It is thought that stomatal closure is stimulated by negative feedback from cells surrounding the stomatal complex in response to cell water potential (Hubbard et al. 2001) and/or root signaling via abscisic acid (Jackson et al. 1995). Plants have evolved a number of other mechanisms by which water loss relative to carbon acquisition is balanced. Rooting architecture varies genetically and depending on soil characteristics in ways to help access available soil water. For example, ponderosa pine can develop very deep roots (24 m; Stone and Kalisz 1991) to access deep water sources and interstitial bedrock water (Rundel and Yoder 1998). Roots also vary in their ability to extract available water; Fowells and Kirk (1945) reported ponderosa pine seedlings extracting water from soil below the wilting point of sunflowers. Pines also maintain a significant portion of functional xylem sapwood with storage of reserve water (Rundel and Yoder 1998). In comparison to populations growing in more mesic sites, genetically undifferentiated populations of ponderosa pine have reduced total leaf surface area (A) relative to sapwood area (As), even with higher whole-tree leaf specific hydraulic conductance (KL), thereby buffering total water loss (Maherali 161 et al. 2002). The anatomical properties of xylem also contribute to cavitation resistance. For example, larger conduits reduce overall hydraulic resistance, but larger diameter vessels are more susceptible to irreparable cavitation than tracheids. Stronger pit membranes are less porous and thus cause higher resistance to both water flow and air infusion (Piflol and Sala 2000). At the leaf- level, higher water-use efficiency (WUE), whether by increasing photosynthetic rate (A), decreasing 9., or effectively dissipating heat, results in better maintenance of whole-plant water status relative to carbon gain (Monson and Grant 1989, Rundel and Yoder 1998). In contrast, Zhang et al. (1997) found no relationship between drought tolerance and high WUE, measured by instantaneous gas exchange and carbon isotope ratio (6130), a measure of integrated WUE (Farquhar et al. 1989), in highly controlled ponderosa pine seedlings. Furthermore, in greenhouse studies, ponderosa pine seedlings from geographically variable sources differed in biomass allocation patterns but not in 9,, A, or needle WP (Cregg 1994). The combination of anatomical and ecophysiological attributes contributing to drought tolerance, especially the apparent trade-off between KL and resistance to cavitation, is an active area for research. Recent evidence indicates that ponderosa pine is highly plastic (Maherali 2002), reduces ALzAs in xeric habitats (Carey et al. 1998), maintains high WP in drought conditions (Hubbard et al. 2001), and has similar vulnerability to cavitation across habitats and climates (Maherali and DeLucia 2000) but increased stomatal control (Pinol and Sala 2000) and sapwood water storage (Stout and Sala 2003) relative to 162 codominant Douglas-fir (Pseudotsuga menziesir). These combinations of adaptations to low soil-water and high vapor pressure deficit contribute to differential survival across environmental gradients. For example, within a group of Southwestern montane conifers distributed along an elevational gradient, seedlings from lower elevation species suffered less photosynthetic depression yet maintained higher plant water status under imposed drought (Barton and Teeri 1993). Mechanistically linking distribution limits to ecophysiological responses should improve our understanding of species responses to climate. In the Santa Catalina Mountains of southern Arizona, populations of the Southwestern ponderosa pine (Pinus ponderosa var. scopulorum Engelmannii) exhibit a discontinuous lower elevational boundary suggesting an abiotic influence. Although located in the Sonoran Desert, the forests in the Santa Catalina Mountains experience reduced frostfree seasons (17 weeks at 2438 m; Shreve 1915) due to high elevation (2793 in peak). In addition, the climate is characterized by a bimodal distribution of precipitation as a result of an infusion of subtropical high pressure from the south during the summer months (Sheppard et al. 2002); consequently, a pronounced drought (“arid foresummer”) occurs through late spring until the mid-summer rains (“monsoon”) arrive. At the lower elevational boundary for Southwestern ponderosa pine, a closely related Ponderosae, Arizona pine (Pinus arizonica Engelmannii), overlaps in distribution forming a transition zone at approximately 2430-2550 m in elevation on a south aspect (Epperson et al. 2001). This transition zone has been noted in Santa Catalina Mountains (Whittaker and Niering 1965) and other mountain islands of 163 the Southwest (Dodge 1963, Barton 1993), but only Barton and Teeri (1993) have sought an ecophysiological explanation for the lower elevational limits. Arizona pine was included in Barton’s dissertation work but was not represented by sufficient replicates for analysis (A. Barton, personal communication). The objective for this study was to determine if critical measures of drought tolerance indicated a genetic difference between Southwestern ponderosa pine and Arizona pine. If evidence suggests that the higher elevation Southwestern ponderosa pine (“PIN_PON”) were less drought tolerant than Arizona pine (“PIN_ARI”), then the correlation of their differential distribution by elevation would be supported by physiological (mechanistic) differences. In one series of experiments, I measured diurnal A, transpiration (Tr), and W9 from coincident ‘small’ PIN_PON and PIN_ARI trees across three hydrologically important seasons: winter (January-February), arid foresummer (June), and monsoon (August); W,, was compared to trees from other elevations, as well. A second experiment tested for differences in integrated WUE (61°C) and nutrient dynamics in needle (leaf) tissue of coincident PIN_PON and PIN_ARI. The third experiment investigated photosynthetic function, measured as Photosystem ll (PSII) excitation capture efficiency (Fv’lFm’; Schreiber et al. 1994), relative to dehydrating needles of PIN_PON and PIN_ARI. In the final experiment, xylem vulnerability to cavitation was measured from PIN_PON and PIN_ARI stems across an elevational gradient; the laboratory work for this portion of the study was conducted by Anna Jacobsen (MSU, Michigan State University). 164 Methods Tree and site selection At three locations across an elevational gradient, I selected “small” PIN_PON and PIN_ARI trees for the ecophysiological experiments in this study (Table 5-1). Small trees were chosen because they are more sensitive to changes in soil water availability and temperature than larger trees with more developed root systems and larger stem storage capability (Rundel and Yonder 1998). The Aspen fire (CNF 2003) eliminated most of the small-diameter trees and virtually all seedlings from most of the Ponderosae forest in the Santa Catalina Mountains (personal observation). Consequently, I attempted to standardize the size by diameter across sites. A second criterion was to locate similarly sized PIN_PON and PIN_ARI trees as close as possible to each other (“paired”) without other nearby trees influencing potential soil water availability. At each site, species designation was determined by calculating the mean number of needles per fascicle across 5 years of needles in one terminal shoot; a mean less than 3.2 was considered to be PIN_PON, and a mean greater than 4.6 needles per fascicle was a PIN_ARI tree (Peloquin 1984). Trees with mean needle number between 3.2 and 4.6 were not used in this study because they may represent hybrids between PIN_PON and PIN_ARI (Peloquin 1984, Rehfeldt et al. 1996, Epperson et al. 2001). 165 Table 5-1. Characteristics of ‘small’ trees used for gas exchange and xylem pressure potential measurements in 2005-06 at the Mount Lemmon (MTL), Palisade Rock (PAL), and Lizard Rock (LIZ) sites. DBH Height Site Location (NAD83) Elev (m) Tree Species (cm) (m) Mount Lemmon 32.4396°N, 110.7871°W 2770 MTL 221 PIN_PON 19.3 6.2 MTL 222 PIN_PON 5.1 2.3 MTL 223 PIN_PON 6.8 3.0 Palisade Rock 32.4138°N, 110.7149°W 2475 PAL 1 PIN_PON 12.9 4.4 PAL 2 PIN_ARI 13.0 3.9 PAL 3 PIN_PON 19.3 5.3 PAL 4 PIN_ARI 12.4 4.1 PAL 5 PIN_PON 17.0 2.4 PAL 6 PIN_ARI 8.1 3.4 Lizard Rock 32.3844°N, 110.6930°W 2135 LIZ 1 PIN_ARI 17.7 3.6 LIZ 2 PIN_ARI 12.2 3.1 LIZ 3 PIN_ARI 7.0 2.7 I selected three major sites for this study: Mount Lemmon (MTL), Palisade Rock (PAL), and Lizard Rock (LIZ). The MTL site is located on a gentle slope just south of the summit of Mount Lemmon. Three PIN_PON trees at the north edge of the forest cover were selected; PIN_ARI has not been documented at this elevation. The PAL site is located downslope of the Mount Bigelow radio towers and upslope from the Palisade Ranger Station (Coronado National Forest). Brown (1968) and Budelsky (1969) measured gas exchange from “ponderosa pine” at this site, but they did not differentiate between PIN_PON and PIN_ARl. This site was not selected due to this previous work but rather because it was identified as a possible steep transition zone between the two taxa; Epperson et al. (2001) documented the steep transition on the south face of Mount Lemmon at similar elevation. Three “pairs” of PIN_PON — PIN_ARI trees growing closely (<3 m) were located and chosen for this study. Another 44 trees were identified, with 5 from each species used to increase sample size for the WP 166 measurements. The LIZ site is located along a small ridge northeast of Lizard Rock and across the Catalina Highway. The xeric ridge is bare bedrock with pockets of disintegrated granite, while the valleys on both sides of the ridge receive floodwater and are thus much more mesic. Three PIN_ARI trees were selected from this site; PIN_PON has not been documented in this area. All of the trees were labeled with both aluminum write-on and nailed disc tags. Additional PIN_ARI trees for the xylem vulnerability study were selected from a ridge south of Rose Canyon Lake (RC, 32.3964°N, 110.6932°W (NAD83), 2175 m elevation) but were not labeled. Seasonal gas exchange and xylem pressure potential Diurnal net photosynthetic (A) and transpiration (Tr) rates were measured from PIN_PON and PIN_ARI needles to investigate relative changes in gas exchange through the course of the day and across seasons. Needles from three “pairs” of PIN_PON — PIN_ARI trees were selected for measurements from the terminal shoot on the lowest living branch with southern exposure. Within an hour prior to measurements, three PIN_PON fascicles (9 needles) and two PIN_ARI fascicles (10 needles) were taped at the fascicle sheath so as to prevent needle crossing within the cuvette, but the needles remained attached to the branches; this number of needles per species covered the entire cuvette with no overlap. One pair of trees was measured per day for the arid foresummer (14/16/18 June 2005) and monsoon (2/4/6 August 2005) seasons, while all three pairs were measured in the same day during the winter (31 January 2006). 167 Consequently, except for winter, species comparisons are limited to measurements from the paired trees. Frequency of measurements varied but were approximately hourly, with 10 points taken per tree, while measurements were taken from predawn through afternoon during the monsoon and through dusk during winter and the arid foresummer; regular afternoon thunderstorms (and even funnel clouds) during the monsoon prevented many opportunities for data collection. Gas exchange was measured using the standard 2x3-cm sensor head in a Ll-COR 6400 Portable Photosynthesis System (LI-COR Biosciences, Inc., Lincoln, Nebraska, USA). The actinic light, infrared gas analyzer (IRGA), and 002 mixer were calibrated prior to measurements each day, and the lRGAs were matched prior to each set of paired measurements within a time period. Carbon dioxide concentration was maintained at approximately 400 pmol C02 mol", which was usually the ambient concentration, with C02 gas cartridges (6400-01 C02 Injector, Ll-COR Biosciences, Inc.). Relative humidity was stabilized across measurements from the paired trees within a time period. A constant block temperature was used to stabilize leaf temperature across trees within a set of paired measurements. Incident light (6400-028, Ll-COR Biosciences, Inc.) for paired measurements was selected based on maximum light intensity (9901-013 External Quantum Sensor, Ll-COR Biosciences, Inc.) under full sky prior to each set of measurements. Leaf temperature was measured by contact thermocouple (6400-04 Leaf Temperature Thermocouple, Ll-COR Biosciences, Inc.) inside the 168 cuvette. Power was supplied by a number of rechargeable batteries (6400-03 Rechargeable Battery, Ll-COR Biosciences, Inc.). Diurnal patterns in gas exchange are difficult to statistically compare across trees, species, days, and seasons because changes in instantaneous gas exchange occur rapidly in response to highly variable conditions, such as solar radiation, temperature, and wind (Jarvis 1976). Consequently, analysis of this experiment is limited to intrapair comparisons of patterns in response to mean climatic fluctuations (e.g, incident light and temperature). The rate of change (i.e., slope) in A and Tr in response to leaf temperature was compared by species and season with paired t-test in R (R Development Core Team 2007), assuming species within a day’s measurements represents the before-and-after treatment. Coincident with the diurnal gas exchange measurements, xylem pressure potential (W,,) was measured from PIN_PON and PIN_ARI fascicles to assess whole-tree water relations. At predawn, midday, and dusk (arid foresummer only), fascicles were harvested from the trees used for the gas exchange experiment, as well as 5 additional trees of each species to increase sample size. On other days, diurnal WP was also measured from PIN_PON trees at MTL and PIN_ARI trees at LIZ (Table 5-1). Fascicles were selected from a different branch but same whorl number as that from which diurnal gas exchange was being measured. Each of 5-10 fascicles was excised at the base of the fascicle sheath with a razor blade, immediately placed into a prelabeled resealable plastic bag with moist paper toweling, and then placed in a small cooler containing ice 169 packs. Within 2.5 h, LPy, was measured from up to 7 fascicles per tree with a pressure chamber (PMS Instruments, Inc., Corvallis, Oregon, USA) and averaged for each tree. Seasonal differences in predawn and midday 91,, across sites were compared by the Welch two-sample t-test, while species differences by season and time of day were examined with univariate ANOVA; site effects were nested within species. All statistical analyses were conducted in R (R Development Core Team 2007). Integrated water-use efficiency and nitrogen dynamics Needle tissue harvested from sympatric PIN_PON and PIN_ARI trees in the Santa Catalina Mountains was analytically measured for species and annual differences in integrated water-use efficiency (WUE), as measured by carbon isotope composition (61°C; Farquhar et al. 1989). Given the physiological relationships between foliar nitrogen content (%N) and 51°C (Sparks and Ehleringer 1997) and nitrogen isotope composition (615N; Hobbie et al. 2000), these parameters were also measured in the same analysis to investigate potential differences in potential photosynthetic capacity (Reich et al. 1995) and nitrogen acquisition (%N, 515M). Amount of structural carbon (%C) in the samples was analyzed as a signal for differences in carbon-to-nitrogen (CzN) aflocafion. Needle tissue was collected in the late afternoon through early evening of 26 October 2005 for both this and the following experiment (Photosynthetic function during dehydration). I harvested one healthy lower branch ~0.8-m in 170 length from each of 11 PIN_PON trees near the summit of Mount Lemmon, 12 PIN_PON trees and 8 PIN_ARI trees at PAL, 6 PIN_ARI trees from ridges surrounding Rose Canyon Lake (RCL), and 4 PIN_ARI trees at LIZ. Sampled trees included those listed in Table 5-1. Immediately after removal from the tree, each branch was placed in a separate plastic bag containing moist paper toweling inside a large cooler with ice packs. To reduce the effect of differential soil water availability, the base of the branches were submersed in water in a 5- gal bucket, ~1 cm was cut off from each end to restore sapflow into the branch, and the base of the branches remained in water for ~ 7 h before being sealed with instant cyanoacrylate adhesive (All Purpose Instant Krazy Glue Gel, Krazy Glue, Columbus, Ohio, USA). Sealed branches were returned to the plastic bags with moist toweling, placed into the cooler with ice packs, and shipped by air to MSU. Approximately 20 h after the branch ends were sealed, the branch ends were recut under water in a 5-gal bucket and rehydrated in a humid walk-in cooler room (4° C) for ~58 h. After removing and processing fascicles for the next experiment (see Photosynthetic function during dehydration), the remaining fascicles on the terminal shoot were excised by razor blade at the base of the fascicle sheath, separated by year of production, and placed into prelabeled envelopes (No. 10 28-lb Heavy-Duty Brown Kraft Envelope, Columbian, Stamford, Connecticutt, USA) to dry at room temperature. In early June 2006, I selected samples from 10 trees from each species growing at the PAL site based on trees having the largest range of annual production (e.g., 2001-2005 fascicles) in sampled fascicles; needles from an 11th 171 PIN_PON tree were also included, but, at the time, there was some doubt as to its taxonomic nature because of finding more >3-needled fascicles in the - samples than when in the field. All of the samples within their envelopes were placed into the drying oven (60° C) for at least 24 h before processing (CPSIL 2006). At least 5 (up to 12) fascicles with the appr0priate number of complete, uninjured, and consistently colored needles (e.g., PIN_PON fascicles had to have 3 needles) were selected from each year for each tree, trimmed by razor 1 cm from the terminus of the fascicle and from the distal portion of the needle, and cut into ~1-cm pieces on a dry glass plate. The pieces were poured into a small stainless steel canister containing a small stainless steel rod, both of which were thoroughly cleaned, rinsed with deionized distilled water (NANOpure Ultrapure Water System, Model D4741, Barnstead International, Dubuque, IA, USA), and dried between samples. The tissue was pulverized within the covered canister by a ball-mill grinder (Model 6, WlG-L-BUG Amalgamator, Crescent Dental Manufacturing Company, Lyons, Illinois, USA) for 5-10 min per sample until becoming a fine powder (CPSIL 2006). [Milling with a standard hand-held coffee grinder and UDY Cyclone Sample Mill (0.5- and 1.0-mm screens; Seedburo Equipment Company, Chicago, Illinois, USA), even for 15+ min, produced visibly heterogeneous tissue particles unsuitable for isotopic analysis (CPSIL 2006).] The pulverized tissue was stored in new 1-dram glass vials (Gerresheimer Glass, lnc., Vineland, New Jersey, USA) that were color-taped by needle-year and labeled by tree, species, and year. The capped vials were stored at room temperature. 172 The powdered samples were prepared for laboratory analysis by removing and packaging ~4.000-mg aliquots into 5x9-mm pressed tin capsules (COSTECH Analytical Technologies, Inc., Valencia, California, USA). The capsules were weighed before and after adding a sample followed by crimping with a microanalytical balance (Model CP2P, Sartorius AG, Goettingen, Germany). Sterile technique was used to prevent sample contamination since the entire capsule is consumed in a mass spectrometer. Crimping was accomplished by placing the preweighed capsule in one of three depressions in an aluminum block (10 x 5 cm) and carefully pinching and then shaping the capsule to a tight ball with two pairs of sterile forceps (CPSIL 2006). Eighty-five capsules, in addition to a number of duplicates to check the analysis, containing pulverized needle tissue from 21 trees were placed into a 96-well cell plate (Greiner Bio- One North America, Inc., Monroe, North Carolina, USA) and shipped overnight on 20 June 2006 for analysis at the Colorado Plateau Stable Isotope Laboratory (CPSIL, Northern Arizona University, Flagstaff, Arizona, USA). The samples were processed on 5 August 2006 by a continuous-flow gas isotope-ratio mass spectrometer (Therrno-Finnigan Deltap'“° Advantage, Therrno Fisher Scientific, Inc., Waltham, Massachusetts, USA) interfaced with an elemental analyzer (ECS4010, COSTECH Analytical Technologies, Inc., Valencia, California, USA); specifications can be found at CPSIL (2006). CPSIL used a variety of standards to check raw isotope data (NIST peach leaves, pine needles, apple leaves, tomato leaves, bovine liver, and mussel tissue; and caffeine - Aldrich), normalize raw isotope data (NAEA CH6, CH7, N1, and N2), and correct raw %C and %N 173 data (acetanilide, BBOT, cystine, methionine, sulfanilamide, cyclohexanone, and nicotinamide). Their external precision from the standards is reported as :l:0.10%o or better for 613C and 10.20%» or better for 615N; 613C and 615N are expressed relative to VPDB and air, respectively (CPSIL 2006). The analysis of 61°C, 6‘5N, %C, %N, and C:N could be approached by assuming independence in the variables across years within a given tree or by assuming that they represent repeated measures from the same individuals across years. The physiological questions relate to the relationship of timing of current-year needle production to current-year nitrogen and carbon acquisition, and the lability of nitrogen and carbon from one year’s needles to another year. Budelsky (1969) showed that ponderosa pine needles at the PAL site begin slowly elongating in mid-May, have the highest rate of elongation during July and early August, and complete extension by the end of August. Consequently, I will assume that current-year needle structural carbon was acquired during the growing season. There is no a priori reason to assume that structural carbon is labile; Billow et al. (1994) report strong seasonal differences in foliar starch and sugar concentration, but no difference from August-October in New Mexican Douglas-fir. Total nitrogen shifts seasonally, at least in Douglas-fir (Billow et al. 1994), and purportedly translocates from older to younger foliage (Rundel and Yoder 1998). Consequently, both %N and (WW are expected to show some degree of temporal autocorrelation. Potvin and Lechowicz (1990) demonstrated complex approaches to modeling repeated measures from individuals in ecophysiological experiments. However, the simplest statistical approach to 174 recognize repeated measures is to compare mean responses by species within a given year by one-way ANOVA, with species as the fixed factor, which was achieved using R (R Development Core Team 2007). Another approach is to conduct an ANOVA on each response by nesting individual tree responses in year, which in turn is nested in species: response ~ species (year (individual tree»; this approach was also calculated in R. In addition, relationships between isotopic composition and nutrient content by species were assessed in two— factorial ANCOVAs with nutrient content as a continuous covariate to trees nested in species. All of the nested analyses indicated no significant (p>0.26) effects of individual trees on patterns observed across years and species (Table 5-5), thus the nested approach was considered satisfactory for investigating species x year responses. Photosynthetic function during dehydration Needles collected from PIN_PON and PIN_ARI trees across an elevational gradient in the Santa Catalina Mountains were slowly dehydrated with concomitant measurement of PSII excitation capture efficiency to determine the relationship of photosynthetic capacity during severe drought. PSII excitation capture efficiency was estimated by the ratio of variable to maximum chlorophyll fluorescence when acclimated to light. I used a pulse-amplitude modulated fluorometer (6400-40 Leaf Chamber Fluorometer, Ll-COR Biosciences, Inc., Lincoln, Nebraska, USA) attached to the base LI-COR 6400 Portable Photosynthesis System. The fluorometer was calibrated before the first set of 175 measurements, zero-ed before each set, and adjusted to recommended settings: Saturating pulse (flash) - 0.8 s saturating multiple flash of ~8800-9000 umol m'2 s", 20KHz modulation, and 50 Hz averaging filter; a 6-s pulse of far red (“dark”) intensity of 8 turned on 1 s before and remain on for 1 s beyond actinic light off, 0.25 KHz modulation, and 1 Hz averaging filter; and Measurement light — intensity 2, 0.25 KHz modulation, 1Hz averaging filter, and 10 gain factor (Ll- COR 2005a). Needles from branches harvested on 26 October 2005 were used in this experiment; see Methods in previous experiment for collection, processing, and transport details. After hydrating overnight in the walk-in cooler (4° C), the buckets of branches were moved to the laboratory. At room temperature, one 2005 fascicle per branch (i.e., tree) was excised from the terminal shoot with a razor blade at the base of the fascicle sheath, weighed using an analytical balance (Model AJ100, Mettler Instrument Corporation, Hightstown, New Jersey, USA), immediately sealed with instant cyanoacrylate adhesive, and lightly marked at the base with colored typographic correction fluid (Liquid Paper, Sanford, Oak Brook, Illinois, USA) by species and relative elevation. After all fascicles (n=41) were processed (~1.7 h), PSII excitation capture efficiency (Fv’lFm’) was measured twice with the fluorometer, with a weight measurement between readings, and averaged for the initial value. Over the next 16 days, fascicle mass and FV’IFm’ were remeasured 30 times. Laboratory conditions were monitored 2-3 times within each measurement period with a porometer (Ll- 1600 Steady State Porometer, LI-COR Biosciences, Inc., Lincoln, Nebraska, 176 USA). Needles were exposed to fluorescent lighting (3 =9.5:I:1.1SD pmol m-2 6") during normal laboratory hours; air temperature was relatively high (a? =26.718.7SD °C), and relative humidity varied with Michigan’s weather ()7 =29.2i14.BSD %). After the last mass and Fv’lFm’ measurements, fascicles were dried in individual paper envelopes (No. 10 28-lb Heavy-Duty Brown Kraft Envelope, Columbian, Stamford, Connecticutt, USA) for 6 days at 60° C in a ventilated drying oven (Model 637G lsotemp Oven, Fisher Scientific, Inc., Pittsburgh, Pennsylvania, USA) and then weighed to determine if all nonlabile water had been removed from the needles. Relative loss of fresh (wet) mass and PSII excitation capture efficiency were modeled using 3-factor ANCOVA in R (R Development Core Team 2007). The former model used site and species as fixed factors and time since excision from the branch as a (continuous) covariate. The latter model also used site and species as fixed factors but included percent loss of fresh mass as the covariate. Nonstructural water remaining after the last mass and Fv’lFm’ measurements was tested against final mass by paired t-test (R Development Core Team 2007). Xylem conductivity and vulnerability Branch stems were collected from mature PIN_PON and PIN_ARI trees in the Santa Catalina Mountains on 20 October 2005 for determination of stem xylem vulnerability to cavitation and xylem specific conductivity. Throughout the morning, stern sections approximately 0.5 m in length were cut from the lowest living branch from trees at the MTL (n=11 trees), PAL (n= 9 trees per species), 177 RCL (n=7 trees), and LIZ (n=3 trees) sites. Stems with few curves, diameter between 5 and 10 mm, and no branches for at least 15 cm were selected. The cut stems were immediately double-bagged (Zip-Loo) with a moist paper towel, placed on ice, and transported via overnight shipping to Anna Jacobsen at MSU. Stems were refrigerated until measured, and all stems were measured within 3 days of collection. A. Jacobsen performed all sample preparation, conductivity measurements, methods documentation, and conductivity calculations for this experiment. Each stem was prepared for conductivity measurements by trimming under water from each end until a straight, unbranched segment 6-9 mm in diameter and 14 cm in length was obtained. Stems were connected to a tubing system and flushed with low pH degassed water (pH 2 HCI) that had been passed through a 0.1 pm filter to remove gas emboli from stems. Stems were flushed for 1 h at 30 kPa. This relatively low pressure was used in order to avoid aspiration of tori in pit membranes. Hydraulic conductivity (K1,) of stems was then measured, and stems were flushed for additional 20 min intervals until a constant maximum hydraulic conductivity (Khmax) was obtained (usually less than 2 h). Hydraulic conductivity of stems was measured gravimetrically (Sperry et al. 1988) using an analytical balance (Model BP 121 S, Sartorius AG, Goettingen, Germany). Following determination of Khmax, stems were spun in a centrifuge (Sorvall RC-SB, DuPont Instruments, Wilmington, Delaware, USA) using a custom built rotor to generate known negative pressures (Alder et al. 1997). Stems were then reconnected to the tubing system and the new hydraulic 178 conductivity (Kn) determined. This process was repeated with successive spins generating more negative pressures until stems experienced >80% loss in hydraulic conductivity. Percent loss in hydraulic conductivity (PLC) was calculated as [Khmax- Khjl Khmax. Vulnerability-to-cavitation curves were constructed by plotting the PLC against the water potential generated using the centrifuge. For each stem, curves were fit with a second-order polynomial model (Jacobsen et al. 2007) to predict the water potential at 50% loss in hydraulic conductivity (W50). Xylem specific hydraulic conductivity (K3) of stems was determined using the Khmax and the cross-sectional xylem area. Cross-sectional xylem area was measured using a dissecting microscope and digital calipers. The xylem diameter (without bark) and pith diameter were measured in two perpendicular directions, and the area of the calculated pith area (an oval defined by the radii in two directions) was subtracted from the xylem area (also calculated as an oval defined by the radii in two directions). The Khmax was then divided by this area for each stem to yield the xylem specific conductivity (kg rn'1 MPa'1 s"). W50 and K, were modeled using two-factor ANOVA in R (R Development Core Team 2007). The full models included the main fixed factors of site and species, as well as their interaction. Given the unbalanced design, Type II sums of squares (SS; Langsrud 2003) were used to test hypotheses using the ‘car’ (Companion to Applied Regression; Fox 2006) package in R (R Development Core Team 2007). Means were compared at each factor level using Tukey’s Honest Significant Difference (HSD) in R, which incorporates an adjustment for 179 unbalanced designs (R Development Core Team 2007). Results Seasonal gas exchange and xylem pressure potential Species differences in diurnal gas exchange during the winter were similar during the morning and evening but highly variable during the midday (Figure 5- 1). Incident radiation (PAR) peaked at 3000 pmol rn'2 s'1 at midday, while leaf temperature (T leaf) varied from -1° C at dusk to 14° C at midday (Figure 5-1). Transpiration (Tr) values were negative around dawn and after sunset, suggesting either that needles were absorbing moisture from the air or that the instrument calibration was in error. At cold temperature, condensation of water vapor on the warmer-than-ambient IRGA sensors could have generated erroneous measurements of reference versus sample water vapor; this error is not expected for C02 given its much lower freezing point. Tr values were highly variable within and across species during the midday. Net photosynthesis (A) was negative at predawn and dusk, indicating respiration. A increased rapidly for all trees following daybreak and decreased rapidly in the late afternoon; like Tr, A was highly variable across all trees during the midday. The rate of change in A as a function of Tjeaf was the same for the two species (paired t-test, t=0.971, df=2, p-value=0.434). The paired trees demonstrated different responses in gas exchange through the course of the day during the arid foresummer (Figure 5-2). PAR across all three days was consistently ~2000 umol rn'2 s'1 by early morning and 180 Tleaf ('C) 1 I + PIN_PON (PAL 1) -o— PIN_PON (PAL 3) 1 , +PIN_P0N(PAL 5) , --— ~ PIN_ARI (PAL 2) — r - PIN_ARI (PAL 4) —-— PIN_ARI (PAL 6) Tr (mmol H20 m'2 s“) O ‘ +PIN_PON(PAL 1) . + PIN_PON (PAL 3) y‘\ —+— PIN.PON (PAL 5) ,0 . -—I-- PIN_ARI (PAL 2) - I» — PIN_ARI (PAL 4) a « —I— PIN_ARI (PAL 6) A (pmol m‘2 s") Time of day 4:00 6:00 8:00 10:00 12:00 14:00 16:00 18:00 20200 Figure 5-1. Diurnal variation in net photosynthetic rate (A) and transpiration (Tr) as a function of incident radiation (PAR) and temperature (Tleaf) for paired PIN_PON and PIN_ARI trees at PAL during winter (January 2006). 181 to late evening before sunset; once the sun broke the horizon at ~5:45, intensity increased rapidly. Tleaf was consistently 11-16° C in early morning, increased to ~22° C by midday, and then decreased to ~17° C by dusk. Both species experienced active Tr and A in the early morning, followed by a midday depression before renewal of limited gas exchange in the later afternoon. In particular, the PIN_PON tree labeled PAL 1 precipitously dropped Tr and A before 10:00, indicating almost complete closure of stomata (Figure 5-2, first page). Its paired PIN_ARI tree also decreased gas exchange by almost 75% of its early morning rate but did continue losing water through transpiration. Under similar climatic conditions, though, another PIN_PON tree (PAL 3) decreased Tr and A through the afternoon but did not demonstrate the severe midday depression. Likewise, its paired PIN_ARI tree steadily decreased gas exchange through the afternoon. The third pair of trees demonstrated a great deal of variability in gas exchange throughout the day but did peak at around the same time (~9:00) as the other pairs of trees. As during the winter, the two species did not differ in the rate of change in A as a function of T..." (t=0.9375, df=2, p- value=0.4475). Changes in gas exchange for the same three pairs of trees tracked climatic conditions independent of species during the monsoon season (Figure 5- 3). PAR was higher than during the arid foresummer but lower than during winter, but it was highly variable throughout each day and peaked near 2010, 2290, and 2620 umol rn'2 6'1 across the three days of measuring the paired trees, respectively. Treat was higher than during the arid foresummer, with temperatures 182 i _h PAR (pmol m'2 s“) Tleaf (°C) Tr (mmol i-to in2 s") A (pmol m2 s") i —+—-— PIN_PON (PAL 1) — o- - PIN_ARt (PAL 2) J —e— PIN_PON (PAL 1) — .- -PIN_ARI (PAL 2) 081 06* 04‘ 02‘ -0.2 J +PIN_PON (PAL 1) — '- -PIN_ARI (PAL 2) —O-PIN_P(N (PAL 1) {1‘ - o- - PIN_ARI (PAL 2) Time of day 6.00 8.00 10:00 12:00 14:00 16:00 18:00 20:00 Figure 5-2. Diurnal variation in net photosynthetic rate (A) and transpiration (Tr) as a function of incident radiation (PAR) and temperature (Tleaf) for paired PIN_PON and PIN_ARI trees at PAL during the arid foresummer (June 2005). Note that the ordinate scales for A and Tr vary across pages. 183 250° ‘ —O—-PN__PON (PAL 3) — .— - PN_ARI (PAL 4) 15* Tleaf (°C) + PIN_PON (PAL 3) 5 ~— .- -PN_ARI (PAL 4) o . . - , , , 1 + PN_PON (PAL 3) 2.0 4 — I- - PN_ARI (PAL 4) 1.5 ‘ Tr (mmol I-bO rn‘2 s") 9 .- U'I o 0.0 0.5 1 14 ‘ —-O—PIN_PCN (PAL 3) — n- - PIN_ARI (PAL 4) A (pmol m" s") Time of day 4:00 6:00 8200 10:00 12:00 14:00 16:00 18:00 20:00 Figure 5-2. Continued. 184 250° ‘ —o— PN_PON (PAL 5) —- .— - PN_AR| (PAL 6) + PN_PON (PAL 5) — I- - PIN_ARI (PAL 6) + PN_PON (PAL 5) — n- - PN_ARI (PAL 6) Tr (mmol H20 m‘2 s") O P n —0.4 -O.6 7 . —O— PIN_PW (PAL 5) 5 . — 3' 'PIN_ARI (PAL 6) 5 . A (umol m4 s") u ~I._ '1 / I’ Time of day 4:00 6:00 8:00 10:00 12:00 14:00 16:00 18:00 20:00 Figure 5-2. Continued. 185 +PN_PON (PAL 1) — u- - PN_ARI (PAL 2) ,2000 'm ‘3‘ E1500 _0 E 31000 n: < D. 500 o c—r L 30 25 A 20 9 u; 15 2 I‘ 10 + PIN_PON (PAL 1) 5 — e- -PIN_ARI (PAL 2) O r 2.5 - -—-O—- PIN_PON (PAL 1) — I- - PIN_ARI (PAL 2) Tr (mmol HzO m'2 s") 12 . +PN_PON (PAL 1) 10 _ — n. — PN_ARI (PAL 2) :~ s4 ‘ / “I‘m a 4 / \ E l a 6 5 . . < 2 I 0 /. . . . .2 Time of day 4:00 6:00 8:00 10:00 12:00 14:00 16:00 18:00 Figure 5—3. Diurnal variation in net photosynthetic rate (A) and transpiration (Tr) as a function of incident radiation (PAR) and temperature (Tleaf) for paired PIN_PON and PIN_ARI trees at PAL during the monsoon season (August 2005). Note that the ordinate scales for A and Tr vary across pages. 186 Im' A (ll 0 O 1 15* Tleaf (°C) 10* + PIN_PON (PAL 3) — n- - PIN_ARI (PAL 4) —o— PIN_PON (PAL 3) — I- - PIN_ARI (PAL 4) Tr (mmol I-bO rn'2 s") _. $5 on or J ... —A ..s _l O N § m J A ().imol rn‘2 s") a Y T l T r ' I —e— PIN_PON (PAL 3) — I- - PIN_ARI (PAL 4) —o— PN_PON (PAL 3) — u- — PN_ARI (PAL 4) T T I I Time of day 6:00 8:00 10:00 12:00 14:00 16:00 18:00 Figure 5-3. Continued. 187 3000 1 N 0| 0 O A O O O PAR (pnlgl m'is") O 0'! O O O O 500 1 251 20.. 151 Tleaf (°C) 10‘ —e— PIN_PON (PAL 5) — u- - PIN_ARI (PAL 6) —e— PN_PON (PAL 5) -- s- - PIN_ARI (PAL 6) 3.0 l 2.5 1 2.0 1 1.5 1 1.0 1 Tr (mmol HzO rn'2 s") 0.5 1 T T r T T l 1 + PIN_PON (PAL 5) - '- - PIN_ARI (PAL 6) 0.0 161 14« A (pmol rn‘2 s“) —o— PN_PON (PAL 5) — .- - PN_ARI (PAL 6) ‘l-v.‘ V1 \ 1 7 fl Time of day 6:00 0:00 10:00 12:00 14:00 16:00 18:00 Figure 5-3. Continued. 188 varying from 14-17° C in the morning to peak readings of 25-29° C later in the day. Tr and A for both species paralleled changes in PAR; Tr and A increased when PAR increased. Given this very strong apparent relationship of gas exchange with PAR, and no visible evidence of such a relationship with Tiger, the rate of change in A to Thar by species is not given. Across all seasons, needle xylem pressure potential (%) from PIN_PON and PIN_ARI trees became more negative from predawn to midday and then partially recovered by evening (PIN_PON, t=-1.9774, df=14.239, p-value=0.068; PIN_ARI, t=-2.3072, df=19.79, p-value=0.032), at least during the arid foresummer (Figure 5-4). ‘Pp was most negative for both species in winter and least negative during the monsoon season, with significant (p<0.04) differences at each time period across every season (Table 5-2). For both individual species, there was no significant (p<0.05) difference in predawn l-Pp between the arid foresummer and monsoon, indicating similar soil water availability during these two seasons; however, the difference for PIN_ARI was marginally significant (p=0.087), suggesting a lower ability to acquire soil water relative to PIN_PON. The differences in midday ‘Pp for PIN_PON trees across winter to the arid foresummer and across the arid foresummer to monsoon were marginally significant (p=0.076 and 0.063, respectivelY); relatively large variation in winter and monsoon WP prevented finding a significant difference, although the trend remains. Across the elevational gradient, these data suggest that the order of “drought" conditions was winter > arid foresummer > monsoon. 189 -10 1i w. (MPa) -15 4 -20 .. -25 J Figure 54. Diurnal by seasonal needle xylem pressure potential (Wp) for PIN_PON and Predawn 1 I Midday 1 Evening —<>— PIN_PON — Winter + PIN_PON - Arid foresummer —-O— PIN_PON - Monsoon - -B- - PIN_ARI - Winter - I- - PIN_ARI - Arid foresummer —'l— PIN_ARI - Monsoon PIN_ARI trees growing in the Santa Catalina Mountains. Table 5-2. Comparison of seasonal xylem water potential (W,,) of combined (both), PIN_PON, and PIN_ARI fascicles harvested at predawn, midday, and evening across an elevational gradient. Significant (p<0.05) terms are bolded. Season 1 Season 2 Time Species t-value df p-value Winter Arid Foresummer Predawn Both -6.5491 40.973 7.2E-08 Predawn PIN_PON 4.1688 17.22 0.0006 Predawn PIN_ARI -5.6791 14.742 4.7E-05 Winter Arid Foresummer Midday Both -3.5004 31.469 0.0014 Midday PIN_PON -1.9242 13.234 0.0761 Midday PIN_ARI 4.0673 17.900 0.0007 Winter Monsoon Pred awn Both -8.6737 40.420 9.1 B1 1 Predawn PIN_PON -7.9012 18.309 2.6E-07 Predawn PIN_ARI -5.9794 19.991 7.6E-06 Winter Monsoon Midday Both —8.6737 40.420 9.1E-11 Midday PIN_PON -3.0359 18.207 0.0070 Midday PIN_ARI 4.4044 13.659 0.0006 Arid Foresummer Monsoon Predawn Both -2.2027 41.931 0.0332 Predawn PIN_PON -1.2752 14.332 0.2225 Predawn PIN_ARI -1.8300 14.919 0.0873 Arid Foresummer Monsoon Midday Both -3.1052 28.169 0.0043 Midday PIN_PON -2.0384 12.813 0.0627 Midday PIN_ARI -2.3528 12.552 0.0357 190 The diurnal pattern of intraspecies variation in ll", within each season gives an indication of individual tree variation in water loss (Figure 5-5). In winter, more variation is observed from predawn to midday in PIN_PON but not PIN_ARI, which in fact demonstrated less variability by midday; this result could be due to cold conditions limiting photosynthesis for PIN_PON trees at the highest elevation while milder conditions at PAL allowed photosynthesis, and thus transpiratory water loss. During the arid foresummer, variation in W9 decreased for PIN_PON midday but increased by evening, while PIN_ARI trees were quite consistent in their response throughout the day. In the monsoon season, midday variation increased substantially for both species relative to predawn variation. During winter and arid foresummer, less variation is expected by midday as trees respond to the drought conditions; conversely, when soil (liquid) water availability is high such as during the monsoon, variation is expected to be greater during the midday as both species are actively photosynthesizing. lntraspecies variation in recovery of water status during the evening of the arid foresummer was larger for PIN_PON than PIN_ARI trees. 191 6 - —<>—— PIN_PON - Winter —0— PIN_PON - Arid foresumner -0— PIN_PON - Monsoon 5 a. / --e—- PIN_ARI - Winter , ’ -— I - PIN_ARI - Arid foresurrmer 4 ~ ’ —I- PIN_ARI - Monsoon so of w. (MPa) 0) Predawn Midday Evening Figure 5-5. Variation (standard deviation, SD) in diurnal needle xylem pressure potential (\Pp) by season for PIN_PON and PIN_ARI trees growing in the Santa Catalina Mountains. Separation of WI, measurements by site (i.e., nesting) allows consideration of species x site interactions (Figure 5-6). During winter, needles from trees growing at higher elevation site had consistently lower (i.e., less negative) Wp than those from lower elevation; the diurnal change in WP was largest at PAL (A=-3.6 MPa), considered to be the intermediate site in terms of cold at high and aridity at low elevation. Both species and site effects were significant (p<0.004) in explaining variation in 9),, (Table 5-3). The same pattern of diurnal ll", was observed during the arid foresummer, with higher elevation correlated with lower WP, but the needles at the highest elevation site had the largest amount of diurnal variation (A=-8.2 MPa). While site effects were significant (p<0.005) across predawn, midday, and evening periods, species differences were only significant at midday and evening (p<0.005); large variation in predawn 91,, for PIN_PON trees at PAL reduced the species effect (Figure 5-5). During the monsoon 192 season, predawn Wp was unexpectedly highest at the lowest elevation site (LIZ, Wp=-6.5 MPa), followed by trees from MTL and PAL (mean LlJp=-8.8 MPa), and then the negative elevation-to-‘I’p pattern developed by midday as all WP increased substantially (mean A=4.8 MPa). However, \Pp did not significantly vary by species (p>0.50) nor site (p>0.09) at predawn or midday. Predawn Midday Predawn Midday 0 1 1 0 1 —<>—— PIN_PON - MTL —e— PIN_PON - PAL -5 1 -5 1 — I- -PIN_ARI - PAL '\ -I— PIN_ARI - LIZ ‘13.10 I. “““ $-10 « é ‘\ §~ E o— —<> 0. \ Q. 9 -15 4 ' 3 -15 ~ —<>— PIN_PON - MTL ~ - ‘ +PIN_PON-PAL ““~- '20 ’ — I - PIN_ARI - PAL '20 ‘ —I-- PIN_ARI - LIZ "" —————— I -25 -25 1 Predawn Midday Evening 0 L . -5 «l 0.5-10 \ g \ £45 1 \ \ x \ __' \ \ -' ‘— ...-l ——<>— PIN_PON - MTL \ ____ .— —’ -20 . +PIN_PON- PAL " ‘— — I- - PIN_ARI - PAL -I- PIN_ARI . UZ -25 1 Figure 5-6. Needle xylem pressure potential (I-Pp) by winter (top, left), arid foresummer (bottom), and monsoon (top,n'ght) for PIN_PON and PIN_ARI trees growing along an elevational gradient in the Santa Catalina Mountains. 193 Table 5-3. Nested ANOVA results xylem pressure potential (Wp) in needles from PIN_PON and PIN_ARI trees across an elevation gradient, 3 seasons (“Arid” is arid foresummer), and 2-3 daytime periods. Significant (a<0.05) models and terms are bolded. Season Period Adj R2 F(p-value) Term df SS F-value P-value Winter Predawn 0.591 0.0002 Species 1 45.413 10.929 0.0039 Species(site) 2 93.039 11.1 95 0.0007 Residuals 18 74.794 4.155 Winter Midday 0.698 5.1E-05 Species 1 46.997 15.434 0.0012 Species(site) 2 95.887 15.745 0.0002 Residuals 16 48.719 Arid Predawn 0.506 0.0012 Species 1 10.128 2.7815 0.1127 Species(site) 2 79.278 10.8861 0.0008 Residuals 18 65.542 Arid Midday 0.511 0.0011 Species 1 20.119 10.2943 0.0049 Species(site) 2 28.687 7.3389 0.0047 Residuals 18 35.180 Arid Evening 0.767 1 .6E-06 Species 1 35.112 14.587 0.0013 Species(site) 2 138.178 28.703 2.5E-06 Residuals 18 43.326 Monsoon Predawn 0.126 0.1491 Species 1 2.756 0.4635 0.5047 Species(site) 2 33.049 2.7786 0.0888 Residuals 18 107.047 Monsoon Midday -0.132 0.9071 Species 1 8.30 0.3226 0.5770 Species(site) 2 5.76 0.1 1 19 0.8948 Residuals 18 463.18 Integrated water-use efficiency and nitrogen dynamics Carbon isotopic composition (613C) varied strongly across years, while nitrogen isotopic composition (615N) was relatively more stable (Figure 5-7). Mean 61°C was -25.7%o (s=0.8%o) across years for both species, with a mean increase of 0.891» in 2004-05 relative to the previous two years. There was no significant species effect on 613C within each year (p>0.52, Table 54) nor across years (p=0.94, Table 55). Mean 0‘5N was 4.4%. (s=0.8%o) across years for both species, but PIN_PON needles discriminated against 15N more than PIN_ARI 194 needles (A=18%) for the years 2003-05, with significant (p>0.05) differences only for 2003 (p=0.013) and 2005 (p=0.047) (Table 54). 2001 2002 2003 2004 2005 -25.0 i i t i -3.0 —<>— PIN_PON - 0130 ’25'2 + "0- PIN_ARl-613C 1" '3'2 -254 4» —O—PIN_PON - 615N T -3.4 _25_6 fl +PIN_ARI-615N -- _3_6 -25.8 «~ -. -30 «)0 .... 1. ‘02 .0 -26.0 ( .40 70 -26.2 .- ,. 4.2 -26.4 4 «r .44 266 ~— 1 4.6 -26.8 4» .- 4.8 -27.0 i ._ -5.0 Figure 5-7. Annual variation in mean carbon (6‘3C, solid lines) and nitrogen (6‘5N, dashed lines) isotope composition for leaf tissue of PIN_PON and PIN_ARI trees at PAL. Note the different ordinate scales. Table 54. Univariate ANOVA results for carbon isotope (513C), nitrogen isotope (6‘5N), carbon (%C), nitrogen (%N), and carbon-to-nitrogen (C:N) content in needles for each year from PIN_PON and PIN_ARI trees at PAL. Responses with significant (a<0.05) differences between species for at least one year and significant terms are bolded. Response Year Adj Rz Term Df SS F-value P-value 013C 2001 -0.1096 Species 1 0.1280 0.5061 0.5161 Residuals 4 1.0114 2002 -0.1161 Species 1 0.2022 0.2719 0.6207 Residuals 6 0.7439 1 1 1 1 2003 -0.0503 Species 0.0113 0.0418 0.8401 Residuals 9 5.1448 2004 -0.0525 Species 0.0011 0.0027 0.9588 Residuals 9 7.5773 2005 -0.0435 Species 1 0.0847 0.1668 0.6875 Residuals 19 9.6468 195 Table 5-4. Continued. Response Year Adj Rz Term Df SS F-value P-value 5"N 2001 -0.0199 Species 1 0.1368 0.9023 0.3960 Residuals 4 0.6066 2002 -0. 1402 Species 1 0.0901 0.1391 0.7220 Residuals 6 3.8848 2003 0.2428 Species 1 3.3051 7.4131 0.0135 Residuals 19 8.4710 2004 0.0815 Species 1 1.5587 2.7744 0.1122 Residuals 19 10.6746 2005 0.1501 Species 1 3.4746 4.532 0.0466 Residuals 14.5671 %C 2001 -0.1146 Species 1 0.0977 0.4861 0.5241 Residuals 4 0.8043 2002 -0. 1627 Species 1 0.0044 0.0203 0.8913 Residuals 6 1 .2857 2003 0.1251 Species 1 1.4059 3.8610 0.0642 Residuals 19 6.9183 2004 0.1217 Species 1 1.3576 3.7714 0.0671 Residuals 19 6.8397 2005 -0.0375 Species 1 0.0305 0.2778 0.6043 Residuals 19 2.0843 %N 2001 0.7836 Species 1 0.0109 19.1010 0.0120 Residuals 4 0.0022 2002 0.0788 Species 1 0.0102 1.5988 0.2530 Residuals 6 0.0382 2003 -0.0294 Species 1 0.0052 0.4294 0.5201 Residuals 19 0.2318 2004 0.1217 Species 1 1.3576 3.7714 0.0671 Residuals 19 6.8397 2005 0.0626 Species 1 0.0586 2.3361 0.1429 Residuals 19 0.4764 C:N 2001 0.8096 Species 1 27.8548 22.2640 0.0092 Residuals 4 5.0045 2002 0.0654 Species 1 27.9230 1.4902 0.2680 Residuals 6 1 12.2470 2003 -0.0082 Species 1 14.8900 0.8367 0.3718 Residuals 19 338.1900 2004 -0.0422 Species 1 3.6900 0.1906 0.6673 Residuals 19 367.7400 2005 0.0599 Species 1 27.4190 2.2751 0.1479 Residuals 19 228.9850 Like 613C, needle tissue carbon content (%C) was relatively stable for both species in the years 2001-03 but then dropped ~2% in the years 2004-05, which correspond to years with higher precipitation than the drought years of 2000-02 196 in the Santa Catalina Mountains (personal observation) (Figure 5-8). In 4 of 5 years, PIN_ARI needles had higher %C than PIN_PON needles. Consequently, both year and species were significant (Table 5-5). Conversely, nitrogen content (%N) was ~36% lower in the older (34 year-old) foliage relative to the 2005 needles; nonetheless, the difference was not significant (p=0.571). However, PIN_PON needles had consistently (Figure 5-8) higher %N than PIN_ARI needles across years. 53.0 .. ~— 1.8 -_ 1.6 52.5 ~~ -_ 1.4 A $5 g 52.0 .. 1.2 ;-; a C c .9 0 -- E 1.0 § 8 515 C c +118 0) 5 8 <0 L a. .1: 0 51'0 —<>—PIN_PON-%C 0'6 2 --Cl-— PIN_ARI- %c -. 0.4 50.5 ... —O—PIN_PON-%N 4 +PIN_ARI- %N ’ 0'2 50.0 i 4 4 i 0.0 2001 2002 2003 2004 2005 Figure 5-8. Annual variation in mean carbon (solid lines) and nitrogen (dashed lines) content for leaf tissue of PIN_PON and PIN_ARI trees at PAL. Note the different ordinate scales. ‘ 197 Table 5-5. Nested ANOVA results for carbon isotope (61°C), nitrogen isotope (6‘5N), carbon (%C), nitrogen (%N), and carbon-to-nitrogen (C:N) content in needles from PIN_PON and PIN_ARI trees at PAL. Individual tree (tree) is nested in year, which is nested in species. Significant (a<0.05) models and terms are bolded. Response Adj R2 F (p-valueLTerm Df SS F -value P-value 6"C 0.1989 0.0236 Species 1 0.0023 0.0049 0.9445 Species(year) 8 17.0604 4.5055 0.0003 Species(year(tree)) 10 0.8642 0.1826 0.9969 Residuals 57 26.9792 6‘5 N 0.1071 0.1286 Species 1 6.9880 12.1216 0.0010 Species(year) 8 3.8800 0.485 0.5706 Species(year(tree)) 10 5.3420 0.9266 0.5160 Residuals 57 32.8620 %C 0.5774 1 .91 E-08 Species 1 1.3775 5.0514 0.0285 Species(year) 8 29.7354 1 3.6303 8.24E-11 Species(year(tree)) 10 2.3886 0.8759 0.5604 Residuals 57 15.5436 %N 0.5964 6.08E-09 Species 1 0.0482 2.4496 0.1231 Species(year) 8 2.3547 14.9629 1 .53E-11 Species(year(tree)) 10 0.1798 0.9142 0.5267 Residuals 57 1.1212 C : N 0.6964 4.15E-12 Species 1 54.0200 3.5350 0.0652 Species(year) 8 2718.7100 22.2391 6.28E-15 Species(year(tree)) 10 181.32 1.1866 0.3191 Residuals 57 871.02 Nutrient content in leaf tissue was related to 613C but not to 615N. 61°C was significantly negatively associated with %C (Figure 5-9; p=0.0001), positively associated with %N (Figure 5-10; p=6.79E-05), and negatively associated with C:N (p=1.51E-05); nutrient dynamics of PIN_PON and PIN_ARI relative to 513C did not differ (p>0.45). Conversely, 615N was not significantly (p>0.17) associated with nutrient content, but the species effect was significant for all models (%C, p=0.0009; %N, Figure 5-11, p=0.0003; and C:N, Figure 5-12, p=0.0003). 198 Carbon content (%) 49.5 50 50.5 51 51.5 52 52.5 53 53.5 -23.5 + _L: —i- I I I I T I I I 4)- T —24.0 4 -245 « -25.0 « ..0 -25.5 4 7° -26.0 - -26.5 - -27.0 1 -27.5 1 o PIN_PON- y = -0.618x + 6.0526, R2 = 0.3007 6 28 0 . a PIN_ARI- y = -0.2495x - 12.803, R2 = 0.0935 Figure 5-9. Carbon isotopic composition (513C) relative to carbon content in needles from PIN_PON (solid line) and PIN_ARI (dashed line) trees at PAL. Nitrogen content (%) 0.7 0.9 1.1 1.3 1.5 1.7 1.9 -23.5 i i 4 4 i 4 24 0 - <> PIN_PON - y = 1.8718x - 28.22, R2 = 0.2708 ' 1 a PlN__ARl-y=1.1287x-27.154,R2=0.1121 <> 0 -245 4 ,3 0O (3 DD Cl U 6) o '25.0 " O 0 (>0 G [9 '° —26.0 . __ ,. .m— on Cl <> 0 -26.5 i Q) ‘3 D oo 0 D -270 ~ [II D O -27.5 1 0 o -28.0 — Figure 5-10. Carbon isotopic composition (5‘30) relative to nitrogen content in needles from PIN_PON (solid line) and PIN_ARI (dashed line) trees at PAL. 199 Nitrogen content (%) 0.7 0.9 1.1 1.3 1.5 1.7 1.9 0 I i I I i 4 o PIN_PON- y = 0.2597x - 5.0048. R2 = 0.0104 11 1 o PIN_ARI - y = 1.1971x- 5.6, R2 = 0.0772 -2 4 I3 Cl C] -3 a D IE) ‘3 2 91o r5J g q’ib ______ -4 7 O ..EJ-EID" "% fl -0?) C] D ”07.8.9413 9513993301: <90 0 £0 -5 4 '6 v ow O o 000 00 CI Cl C] -6 - C] o O o -7 A Figure 5-11. Nitrogen isotopic composition (6‘5N) relative to nitrogen content in needles from PIN_PON (solid line) and PIN_ARI (dashed line) trees at PAL. Carbon:nitrogen (C:N) 20 25 3o 35 40 45 50 55 60 0 I I I I I I I I o PIN_PON - y = -0.0077x - 4.3508, R2 = 0.0082 '1 I" o PIN_ARI - y = -0.0352x- 2.6065, R2 = 0.0821 -2 4» D C] _3 _,_ D C] DC] 2 D 0 31o __ _ d3 g c? 41— - ‘ 6 ~ 6 I O 800 E 004% A113)» AD o'Qe—Q~O.-_‘ -5~~ <2 00900 0585‘ ”o [3 CI -6 4» D o O E] o -7 «— Figure 5-12. Nitrogen isotopic composition (6‘5N) relative to carbon-to-nitrogen content (C:N) in needles from PIN_PON (solid line) and PIN_ARI (dashed line) trees at PAL. 200 Photosynthetic function during dehydration Needles from PIN_PON and PIN_ARI trees across the elevational gradient lost a similar amount of water mass across the first 8 d, but mass loss asymptoted earlier for PIN_ARI than PIN_PON needles consistently across sites (Figure 5-13). All measurements from one tree (LIZ 1) were removed due to exceptionally low initial Fv’lFm' (0.035); resultant sample sizes for PIN_PON and PIN_ARI were 23 and 18, respectively. There was a significant decrease in water loss following the final drying (paired t-test, t=5.7051, df=40, p- value=1.27E-06), indicating that some intracellular water remained after the final measurements. The species difference becomes more apparent when trees are grouped across sites by species (Figure 5-14). Two-way ANOVA (Adj-R2=0.507, F (p- value)=4.85E-06) of mass loss at the final (31") measurement indicates a significant species (p=1.81E-07) effect but no site effect (p=0.309). However, when analyzed as a function of time since excision from the hydrated branch, almost twice the amount of variation in mass lost is accounted for by site differences than species differences, although both terms were highly significant in the ANCOVA (Table 5-6). Attempts to reduce the right skewness of the otherwise normally distributed residuals through square, square-root, reciprocal, natural-log, natural-log of reciprocal, and arc-sine transformation of the response variable were not successful; i.e., skew remained and the adjusted-R2 decreased. Because all transformations identified the same significant terms, only the level of significance varied, the linear model was retained. The 201 significant interaction of site by time elapsed since branch excision is unfortunate and related to the order imposed to prevent measurement error; no difference in measurements as a function of elapsed time was expected due to the rapid rate in taking Fv’lFm’ measurements at each time period (1? =1.23 h, s=0.25 h). 100 90 j 80 ( SE 8 70 - N E 5 60 ~ E i o PIN_PON - MTL 50 « e PIN_PON- PAL - PIN_ARI - PAL 40 . x PIN_ARI - RCL - PIN_ARI - LIZ 30 T 7 T I I 1 0 100 200 300 400 500 600 Elapsed time (h) Figure 5-13. Dehydration of fascicles for PIN_PON (solid lines) and PIN_ARI (dashed lines) stems from four sites in the Santa Catalina Mountains: MTL, PAL, RCL, and Liz. Fitted curves are second-order polynomials. 202 90 49 80 . :s‘ m 70 « 8 E § 60 u‘: 50 ~ 0 PIN_PON ______ 40 1 o PIN_ARI 30 r I F 1 r 1 0 100 200 300 400 500 600 Elapsed time (h) Figure 5-14. Mean dehydration of fascicles for PIN_PON (solid lines) and PIN_ARI (dashed lines) stems across four sites in the Santa Catalina Mountains. Fitted curves are second-order polynomials. Table 5-6. ANCOVA results for fresh mass loss (pmass) and PSII excitation capture efficiency (FV’IFm’) of dehydrating needles for PIN_PON and PIN_ARI stems from four sites in the Santa Catalina Mountains: MTL, PAL, RCL, and LIZ. Covariates for the two models are elapsed time (Etime) and pmass, respectively. Significant (a<0.05) terms are bolded. Response Adj R2 F(p-value) Term Df SS F-value P-value pmass 0.8059 <2.2E-16 Site 1 1145 25.7966 4.366-07 Species 3 640 4.8117 0.0025 Etime 1 230338 5191.3640 <2.2E-16 Slte'Etime 1 2097 47.2516 9.79E-12 Species*Etime 3 210 1.5805 0.1923 Residuals 1261 55950 Fv’lFm’ 0.7939 <2.25-16 Site 1 0.1145 19.9976 8.45E-06 Species 3 0.1520 8.8483 8.33E—06 pmass 1 27.7698 4848.7223 <2.28-16 Site‘pmass 1 0.0002 0.0272 0.8689 Species*pmass 3 0.0352 2.0491 0.1052 Residuals 1261 7.2221 203 PSII excitation capture efficiency relative to progression of dehydration was consistently higher for needles from PIN_PON trees than from PIN_ARI trees, while site differences were only apparent for PIN_ARI (Figure 5-15). While very close, PIN_PON needles from PAL had higher Fv’lFm’ for a given relative loss in water mass, while higher elevation PIN_ARI clearly had higher retention of photosynthetic function through dehydration. Grouping by site increased the apparent species difference (Figure 5-16) that is supported by the ANCOVA results (Table 5-6). In contrast to rate and extent of dehydration, both site and species are equally significant (p=8E-06) in contributing to variation in PSII excitation capture efficiency; no interactions with the covariate in this model were significant (Table 5-6). Although statistical painivise comparison by site within species is not possible due to the significance of the covariate relative mass, the overall site effect was significant (Table 5-6). In Figure 5-16, the difference in photosynthetic function for a given relative amount of water loss appears minimal across sites for PIN_PON but clear for PIN_ARI, hence the overall significant site effect. 204 0.6 1 o PIN_PON - MTL 0,5 . e PIN_PON - PAL 0 - PIN_ARI - PAL x PIN_ARI - RCL °~4 ‘ - PIN_ARI - LIZ 0.3 1 0.2 1 Chlorophyll fluorescence (Fl/Fm') 0.1 1 30 40 50 60 70 80 90 100 Fresh mass (%) Figure 5-15. Photosynthetic efficiency as a function of needle dehydration for PIN_PON (solid lines) and PIN_ARI (dashed lines) stems from four sites in the Santa Catalina Mountains: MTL, PAL, RCL, and LIZ. Fitted curves are second-order polynomials. 0.5 1 0-45 1 o PIN_PON o PIN_ARI .0 a _o g _o on 0.25 1 .0 N Chlorophyll fluorescence (Rf/Fm') .0 5‘. .0 _I 0.05 1 30 4'0 50 60 70 8‘0 90 100 Fresh mass (%) Figure 5-16. Photosynthetic efficiency as a function of needle dehydration for PIN_PON (solid lines) and PIN_ARI (dashed lines) stems across four sites in the Santa Catalina Mountains. Fitted curves are second-order polynomials. 205 Xylem conductivity and vulnerability Stems from trees lower in elevation (i.e., more xeric) had consistently higher loss of xylem conductivity for a given imposed xylem tension (Figure 5- 17). PIN_ARI stems were also more vulnerable to cavitation than PIN_PON stems, but neither site nor species was significant in the ANOVA (Table 5-7). The interaction term was not calculated due to the extremely small sample size at the LIZ (n=1) and RCL (n=2) sites; stern disintegration (i.e., explosion) inside the centrifuge at moderate tensions (W=-2.5 MPa) prevented completion of individual vulnerability curves for a number of trees. One-way ANOVA of species effect at PAL only was also not significant (p=0.3527). Xylem specific hydraulic conductivity was likewise not different between species, but there was a clear (Figure 5-18) and significant (p=0.005) site effect (Table 5-7). Sample sizes for non-PAL sites were substantially larger, but the non-significant interaction term was removed from the final model. Based on Tukey’s HSD comparison of means, K8 of MTL stems was significantly higher than those from PAL (p=0.041) and RC/LIZ (p=0.005), while there was no difference in K; between stems at PAL and RC/LIZ (p=0.348). Increased sample size could have reduced the substantial variation in Ks found within the PIN_ARI trees and thereby separate the lower elevation sites. 206 r 100 r 100 90 '\ . 90 OPIN PON- MTL \ IPIN_ARI - PAL 7 30 ' ‘ ‘\ IPIN ARI - LIZ/RCL ‘ 8° OPIN PON- PAL 63 \ _ A _ v \\ 33 70 £1 \ ~ 70 e 1% \\ jS . \ L 60 a ‘ \ 50 g 8 50 o \ \ i 50 E E \ .9 ‘ - 40 g 40 e \ i\ '6 \ \ ‘5 3° 8 \ ‘ i” is? 20 3 I \ \bs ~ 20 —’ \\I \K ' 10 \\ q \ 1 10 r r r I o I T 1 Vi‘ O -5 -4 -3 -2 1 o -5 4 -3 -2 -1 0 Xylem pressure potential (MPa) Xylem pressure poterlial (MPa) r 100 <>\ " 9° \\ OPIN_PON 80 \ EIPIN_ARI £5 E ‘5 .3. Xylem pressue potertid (MPa) Figure 5-17. Xylem vulnerability curves for PIN_PON (solid lines) and PIN_ARI (dashed lines) stems collected from three sites in the Santa Catalina Mountains: MTL, PAL, and LIZ/RCL. The bottom curve combines sites by species. Table 5-7. ANOVA results for xylem pressure potential at 50% loss of hydraulic conductivity (Woo) and xylem specific hydraulic conductivity (Ks) for PIN_PON and PIN_ARI stems collected from three sites in the Santa Catalina Mountains: MTL, PAL, and LIZ/RCL. Significant (a<0.05) terms are bolded. Response Adj R2 F(p-value) Term Df SS F-value P-value Wp 0.0053 0.3924 Site 2 0.4493 1.0180 0.3777 Species 1 0.2424 1 .0984 0.3060 Residuals 22 4.8543 K. 0.2383 0.0121 Site 2 2.27E-05 6.4777 0.0047 Species 1 9.70E-08 0.0553 0.8157 Residuals 29 5.08E-05 207 0.007 1 0.006 ( 0.005 « i 0.003 - J 0.002 1 O O O A *— ——l Xylem specific hydraulic conductivity (kg rn’ MPa1 s‘) 0.001 1 0 T I 1r I PIN_PON - MTL PIN_PON - PAL PIN_ARI - PAL PIN_ARI - RC/LIZ Figure 5-18. Xylem specific hydraulic conductivity (Ks) for PIN_PON and PIN_ARI stems collected from three sites in the Santa Catalina Mountains: MTL, PAL, and LIZ/RCL. Values are means; bars are :1 sd. Discussion Seasonal gas exchange and xylem pressure potential The results from this experiment support the hypothesis (Brown 1968, Barton and Teeri 1993) that ponderosa pine, in this case Southwestern ponderosa pine, avoids drought effects by maintaining positive water balance during periods of low moisture availability and actively photosynthesizing during favorable conditions. The relative species differences, especially from the LIJp results, suggest that PIN_PON is more drought tolerant than PIN_ARI. These differences are most evident during the winter and less obvious during the arid foresummer. The 3-needled PIN_PON had consistently less negative Wp than 5- needled PIN_ARI across all seasons, sites, and periods during the day (Figure 5- 208 4). These results support Haller’s (1965) findings that ponderosa pine with fewer needles are associated with drought conditions, but they contrast with Malusa (1992) who found that single-needled fascicles in hybrid pinyon pine (Pinus califomiarum x P. edulis) had more negative WP than the double-needled fascicles and concluded that needle number did not affect water relations. This experiment found a significant difference in LPp, with the fewer-needled species having lower Wp, thus these data support the conclusion that PIN_PON is more drought tolerant than PIN_ARl based on differences in Wp. Given the hypothesis by Monson and Grant (1989) that ponderosa pine has adapted to drier habitats by lowering gs at the expense of reduced maximum A, the more drought tOIerant of the two species in this study should have lower A and 93, or Tr as measured here. Conclusions about absolute differences in gas exchange should not be made from the limited number of paired diurnal measurements in this experiment, but patterns and variation in these values can be discussed. The WP data suggest that winter leaf desiccation is more severe than during the arid foresummer (Figure 5-4); however, both species showed positive A and Tr at the intermediate-elevation site (PAL) during the high-PAR winter (Figure 5-1). On the other hand, 41,, did not change through the day for high-elevation PIN_PON and low-elevation PIN_ARI, suggesting no change in water status (i.e., no Tr) and possibly no A. Osmotic adjustment from cold hardening can increase total leaf xylem water potential by around -2 MPa (Abrams 1988), which is insufficient to compensate for differences observed here between winter and the other seasons at any of the sites. Consequently, we can 209 assume that both species are photosynthetically active during the winter at the intermediate PAL but less so at MTL and LIZ. Ascertaining the degree of photoinhibition and photosynthetic down-regulation, measured partly by ratio and total pool sizes of leaf violaxanthin-antherxanthin-zeaxanthin (Bjdrkman and Demmig-Adams 1994), during winter across the three sites would contribute to understanding more about cold temperature influences on the distributions of these species. While gas exchange during the monsoon appeared to be directly linked to PAR, thus soil water was not considered to be limiting, gas exchange during the arid foresummer was not related to PAR or Tleaf, but instead was likely related to soil moisture (Bassman 1987). In ponderosa pine, Tr rates are typically higher than other western conifers until a threshold soil '4’ (-0.2 MPa) is reached, after which Tr for ponderosa pine decreases immediately, long before other conifers (Bassman 1987). Greater stomatal control, or sensitivity to leaf water status, leading to higher overall WUE, is considered the mechanism by which ponderosa pine survive in more xeric habitats than expected given its higher Tr. In this study, WP was similar for coincident PIN_PON and PIN_ARI at predawn and midday, but PIN_PON had an average 42% higher standard deviation in daily A measurements and average 62% higher standard deviation in daily Tr than did PIN_ARl at PAL during the arid foresummer. These data support the conclusion that PIN_PON is more drought tolerant than PIN_ARI based on more rapid stomatal control of transpiration. 210 Density and architecture of stomata also exert considerable control over gas exchange (Bassman 1987, Rundel and Yoder 1998). Increased stomatal density, as well as increased needle surface area, should lead to increased Tr. Stomatal density was compared for a limited number of needles from PIN_PON and PIN_ARl; no significant (p>0.05) difference in density on any side of the needle was found. Hadley (1986) found that epicuticular waxes had a profound influence on stomatal conductivity. In a limited inspection of scanning electron microscopy (SEM) images (courtesy of Anna Jacobsen, MSU) from PIN_PON and PIN_ARl needles from PAL, abundant epicuticular wax is produced by both species, with little observed difference (Figure 5-19). Removal of the waxes with chloroform revealed a deeply sunken stomatal aperture (Figure 5-20). Figure 5-19. Stomatal openings in PIN_PON (left) and PIN_ARI (right) by SEM. Note the abundant epicuticular waxes surrounding the stomatal pit, as well as the network of waxes above the aperture. Scale bar equal to 20 microns. Photos courtesy of Anna Jacobsen. 211 Figure 5-20. Chloroform-cleared stomatal openings in PIN_ARl of top and transverse views by SEM. Scale bar equal to 20 microns. Photos courtesy of Anna Jacobsen. Integrated water-use efficiency and nutrient dynamics Discrimination against heavier isotopes occurs in plants due to physical properties of the heavier isotope and ecophysiological responses through fractionation (Fry 2006). Less discrimination of 13C is associated with higher WUE because plants that produce photosynthate with relatively higher 13C close their stomata during stressful conditions, thereby increasing the intercellular-to- atmospheric C02 concentration (ct/ca) and increasing relative 13C concentration inside the leaf (Dawson et al. 2002). Under normal conditions, the higher diffusivity (Melander and Saunders 1979) and reactivity with ribulose-1,5- bisphospate carboxylase (Park and Epstein 1961) of 1"’0 results in discrimination of 13C. When stomata close, ci decreases, relative concentration of 12C decreases, more 13C is processed, and photosynthates are produced without losing water; hence, plants with greater stomatal control should have higher 212 WUE and less negative 613C. In this experiment, 613C values were low (-25.2 to - 26.6 960) relative to ponderosa pine in Idaho (18.6960; Marshall and Zhang 1994) but comparable to ponderosa pine in northern Arizona (-24 to -25.5%o at low and high elevation; Adams and Kolb 2004). There was no significant difference in 6130 between species across years at the coincident PAL site (Table 5-4), but inspection of the values from 2001 and 2002 lead to some speculation (Figure 5- 7). The years 2000 and 2002 were considered drought years; the manifestation of drought was observed in the substantially shorter needles from these years than from 2001 and 2003. In 2001, the PIN_PON needles had relatively less 130 than the PIN_ARI needles, while the converse was true in 2002, the drought year. These data suggest that, with a larger sample size, PIN_ARl may in fact have higher WUE than PIN_PON, but the data are sparse to make such an inference. This conclusion departs from the supposition made after analyzing the LPp and gas exchange results. Comparison of 6130 and %N for both species across the elevational gradient would lead to greater insight on differential response to drought (Adams and Kolb 2004). Carbon (%C) and nitrogen (%N) content varied by year, while species differences were observed only for %C (Figure 5-8). PIN_ARI had greater allocation of mass to structural C across most years; the higher leaf surface area to volume ratio for this 5-needled pine should increase the relative amount of C allocated to structural than photosynthetic cells. The increase in %N in younger needles likely represents N reallocation within the plant rather than less N produced and stored in 2001-02 relative to later years. Foliar nitrogen is strongly 213 correlated to maximum A on a leaf mass (not area) basis for conifers (Reich et al. 1995). Although not significant, PIN_PON needles had higher %N than PIN_ARI needles, suggesting a slightly higher potential photosynthetic capacity by PIN_PON. The nitrogen isotope discrimination (615M) results suggest differential nitrogen resource dynamics between PIN_PON and PIN_ARI. Interestingly, 615N decreased during the drought year 2002 for both species, but more discrimination against 15N occurred in PIN_ARI trees in subsequent years (Figure 5-8). Discrimination of 15N in plants is much more complex than 13C (Dawson et al. 2002). Nitrogen is normally acquired by plants through the soil. While physical discrimination against the heavier isotope does not occur across living membranes, enzyme-mediated processes can lead to discrimination (Dawson et al. 2002). Low nitrogen availability, osmotic stress, or drought can lead to N isotopic fractionation, but the mechanism remains unclear (Dawson et al. 2002). Hobbie et al. (2000) suggest that fractionation occurs during transfer of nitrogen from mycorrhizal fungi to plants, especially at low-nitrogen sites. Given the significant effect of ectomycorrhizal fungi in the establishment, survival, and nutrient acquisition for Pinus (Read 1998), the results of this experiment suggest that these fungi may be influencing the nitrogen dynamics, and thus photosynthetic capacity, of PIN_PON and PIN_ARI trees at PAL. Although highly speculative, increased fractionation of 15N (i.e., more negative 515N) by PIN_PON trees may suggest lower accessibility to N and thus lower potential growth. 214 Photosynthetic function during dehydration Needles of trees from lower in elevation lost less mass from dehydration, or conversely had a lower relative water content, even when potentially fully hydrated, than needles from trees higher in elevation within a species (Figure 5- 13). Furthermore, needles from lower elevation trees, especially PIN_ARI, had lower PSll excitation capture efficiency than those originating from higher elevation (Figure 5-15). Given the identical pretreatment conditions, these results suggest a plastic response in total potential needle water storage and photosynthetic function during extreme dehydration (i.e., drought) across the elevational gradient. PIN_PON needles lost a larger portion of their initial mass through dehydration, suggesting that, despite full opportunity for hydration prior to the experiment, PIN_PON needles had a higher storage capacity for water than do PIN_ARl needles. Furthermore, for a given relative loss of water mass, PIN_PON trees had higher PSII excitation capture efficiency through the dehydration period (Figure 5-16). However, needle morphology was not characterized. Higher-needled fascicles for a given fascicle diameter and length have a geometrically higher surface area to volume ratio. The difference in total stored water mass per fascicle, and thus water available for photosynthesis, observed in this experiment could be accounted for by the difference in potential storage volume due to areazvolume relationships, hence these results should be cautiously interpreted. 215 Xylem conductivity and vulnerability Vulnerability of the xylem to cavitation has been suggested to be the primary driver for drought tolerance in plants (Tyree and Ewers 1991), but plants have evolved alternative avoidance strategies that are also important. For example, Pil‘lol and Sala (2000) showed that increased stomatal control of water loss compensated for lower resistance to cavitation by ponderosa pine than by codominant species. In this study, the lower elevation species, PIN_ARI, was more vulnerable to cavitation than PIN_PON as determined by 50% loss of conductivity as a function of imposed negative xylem pressure (Figure 5-17); mean W50 for PIN_PON was -3.44 MPa and for PIN_ARI was -3.16 MPa. The difference of -0.24 MPa is not substantial nor significant (Table 5—7). A mild site effect was also discernible but not significant. In a study of desert versus montane ponderosa pine populations, vulnerability to cavitation did not differ, but xylem specific hydraulic conductivity (K3) was higher for the more xeric trees (Maherali and DeLucia 2000). In this study, K8 was highest for PIN_PON stems from the highest elevation site (MTL), similar across species at PAL, and lowest for PIN_ARI at low elevation (Figure 5- 18). High Ks implies an increased capability to transport water due to lower conductive resistance, whether clue to larger tracheids, more porous pit membranes, or fewer cavitated tracheids. Standardizing for xylem area allows interspecies comparisons in potential hydraulic conductance. However, Ks did not differ for PIN_PON and PIN_ARI when coincident, while K8 was positively correlated with elevation. Maherali et al. (2002) also found Ks to be 216 phenotypically plastic in ponderosa pine across a climatic gradient. The results from this experiment indicate that K8 is similar for PIN_PON and PIN_ARI, that the ensemble of traits contributing to K8 is sufficiently plastic to acclimate across a steep climatic gradient within a species, and that drought stress is negatively associated with elevation. Implications of drought tolerance to distribution The balance of evidence from this study, composed of four experiments, suggests that PIN_PON is more drought tolerant than PIN_ARI. With only the exception of PIN_ARl trees at LIZ during the monsoon season, PIN_PON trees had less negative ‘41,, at all periods of the day and across major seasons (Figure 5-6). PIN_PON trees exhibited greater variation in A during the arid foresummer than did PIN_ARI trees (Figure 5-2); this result could be interpreted as PIN_PON stomata responding more closely to fluctuations in soil W or as PIN_PON trees being closer to the drought threshold at that time of year. However, the less negative “1,, for PIN_PON trees suggests the former interpretation. Neither species appears to be water-limited during the monsoon season; rather light is limiting to A (Figure 5-3). When exposed to severe drought (i.e., dehydration), PIN_PON needles retain a higher level photosynthetic function than do PIN_ARI needles (Figure 5-16). Although not significant, PIN_PON trees had higher %N, which is correlated with maximum A and therefore growth potential. Conversely, PIN_PON needles contain a relatively higher concentration of 15N, suggesting a larger amount of fractionation (or processing) occurs before N reaches the leaf; 217 this could indicate drought stress or, alternatively, different access to soil N pools than experienced by PIN_ARI trees. However, in sum, the results from this study suggest that PIN_PON trees are more drought tolerant than PIN_ARl trees. lf drought tolerance is not limiting the lower elevational distribution of PIN_PON, then what factors are limiting its range? This study did not examine potential competition nor the role of drought tolerance on seedlings, which are more susceptible to low soil water availability due to their developing root systems. In controlled greenhouse conditions as well as reciprocal in situ plantings, Barton and Teeri (1993) investigated drought tolerance of Southwestern conifers, excluding Arizona pine, and found that lower elevational distributions were determined by drought tolerance of seedlings. A similar approach with the taxa from this study would allow for standardized comparisons among these and other Southwestern species. Likewise, Daubenmire (1943) saw soil water availability as the limiting factor to Rocky Mountain conifers. Grasses and forbs are significant competitors against ponderosa pine seedlings for belowground resources, especially water and available nitrogen (White 1985, Elliott and White 1987, Riegel et al. 1992, and Kolb and Robberecht 1996). Seedling mortality due to moisture stress, in part from competition for water by all plants, can be over 90% within the first year (Elliott and White 1987). Relative differences in seedling root growth rates significantly affect survival (Kolb and Robberecht 1996). Although seedlings in the field were preferred subjects for this study, extensive recent ground fires, especially through the transition zone in the Santa Catalina Mountains, eliminated all small diameter seedlings and many 218 saplings, but the fires also created more habitat for grass, forb, and pine regeneration (personal observation). Stomatal control of water relations appear key to separating ponderosa pine from other conifers of the Rocky Mountains. Given the purported warm, humid origins of Arizona pine (Millar 1993) relative to the cooler, drier evolutionary origin for ponderosa pine (Axelrod and Raven 1985, Millar 1993), investigations into differences in stomatal control may shed insight on how the two species control water loss, especially during the predictable arid foresummer. Since roots directly acquire water and since ponderosa pine are known for extensive root systems, differential rooting architecture, water uptake, and nutrient absorption may be important in segregating species distribution across the shallow soils of the Santa Catalina Mountains. The intriguing nitrogen isotope data suggest that differential mycorrhizal relationships may influence nitrogen uptake (Perry et al. 1989) and consequent photosynthetic capacity and WUE. The value of the elevational gradient and transition zone in this region should not be overlooked when examining distributions of related species. As indicated by this study, a number of responses to drought can be integrated to contribute to an understanding of species distributions; for example, spatial analysis of indicators, including integrative measures like 613C and tree-rings, can yield models to predict expected changes in distribution from climate change. 219 CHAPTER 6 GENERAL CONCLUSIONS The objectives for this dissertation research were to characterize the spatial distribution and examine ecophysiological responses that could affect the respective distributions of the Ponderosae in the Santa Catalina Mountains of southern Arizona. A priori expectations were that Southwestern ponderosa pine were distributed higher in elevation than the evolutionarily close and spatially overlapping Arizona pine because the former species is less drought tolerant and the latter taxon is less cold tolerant. Differences in distribution and cold tolerance met these expectations, while results from the drought tolerance work demonstrated the opposite expectation. In general, Southwestern ponderosa pine was better able to maintain positive net photosynthesis and water balance during periods of low soil water availability, both during winter and the arid foresummer. Most Southwestern ponderosa pine were indeed located at higher elevation than Arizona pine trees, with a few low-elevation occurrences by the former species. As expected, spatial modeling of sampled trees by species, or morphotype to avoid taxonomic debate, suggested that Arizona pine and the Mixed morphotype are limited in upper and lower range by cold temperatures and climatic variables contributing to drought during the arid foresummer, respectively. Because the upper elevational range was not reached by Southwestern ponderosa pine in the Santa Catalina Mountains, its distribution 220 was understandably limited most by arid foresummer conditions. However, Southwestern ponderosa pine and the Mixed morphotype had higher probability of occurrence at higher elevation as compared to the distribution predicted for Arizona pine, while both Arizona pine and the Mixed morphotype had higher probability of occurrence at lower elevation as compared to the distribution predicted for Southwestern ponderosa pine. Since lower elevational limits were based on the same climatic factors, the higher placement of Southwestern ponderosa pine suggests a decreased ability to establish at its lower margin, where drought is more severe and Arizona pine is sympatric. Furthermore, the presence of the Mixed morphotype above the upper margin for Arizona pine and below the lower margin for Southwestern ponderosa pine suggests that this group possesses unique genetic adaptations to a broader climatic range than the other taxa. The ecophysiological experiments indicate that Southwestern ponderosa pine is undoubtedly more cold tolerant than Arizona pine. Arizona pine seed had much lower germination rate, especially at high elevation, while Southwestern ponderosa pine seed germinated equally well across elevations. Survival of germinants at both sites through the first full arid foresummer (2007) has yet to be determined. Sapling Southwestern ponderosa pine had higher photosynthetic function and rate and less negative leaf xylem pressure potential during the winter than did Arizona pine saplings, demonstrating less cold tolerance of Arizona pine trees; in one experiment, seedlings of the Mixed morphotype in frozen soil water exhibited photosynthetic function intermediate to Southwestern 221 ponderosa and Arizona pine seedlings. Additionally, fully hardened Arizona pine seedlings had higher threshold freezing temperatures than did similarly hardened Southwestern ponderosa pine seedlings, thereby further supporting the conclusion of greater genetic cold tolerance by Southwestern ponderosa pine. The studies examining indicators of drought tolerance offered no clear sign of Arizona pine being more drought tolerant than Southwestern ponderosa pine; in fact, most evidence suggested the opposite trend. Southwestern ponderosa pine had less negative leaf xylem pressure potentials across all times of clay and seasons, lower vulnerability to xylem cavitation in branches, higher immediate response to variations in climatic conditions during the arid foresummer, and higher photosynthetic function during dehydration. These indications suggest that Southwestern ponderosa pine is more drought tolerant than Arizona pine. Although not statistically significant, Southwestern ponderosa pine had higher foliar nitrogen across all drought and non-drought years, possibly indicating a higher maximum photosynthetic capacity. No difference in integrated water-use efficiency, as measured by carbon isotopic discrimination, was found between. sympatric species; however, sample sizes were small, and application of this technique across the elevational gradient may detect species x site differences. Differential leaf nitrogen isotope composition suggests that these species differ in their nitrogen uptake pathway, and, although nitrogen dynamics are highly complex, combining this type of analysis with studies of rooting architecture and mycorrhizal associations may contribute more to understanding 222 lower elevational distribution of these taxa in the nitrogen-poor granitic soils of the Santa Catalina Mountains. The rapidly developing field of species distribution modeling offers tools for inferring evolutionary processes, tracing population migration, and predicting responses to climate change. Heretofore unseen in the literature, a thorough spatial analysis of genetically linked morphological traits and combined cytoplasmic and nuclear DNA of the Ponderosae would contribute to an understanding of the systematics and evolution of this complex in the Southwest. In this dissertation, the unique spatial and modeled niche of trees producing intermediate needle numbers (“Mixed”) relative to Southwestern ponderosa and Arizona pine do not support the Taxon X hypothesis; instead, these results support the accepted taxonomy of the parent species (Conkle and Critchfield 1988, Price et al. 1998) and putative introgression (Peloquin 1984, Rehfeldt et al. 1996, Epperson et al. 2001). Further, the high intratree variance in the highly heritable morphologic feature of number of needles per fascicle (Rehfeldt et al. 1996) draws attention to intermediate, or putative hybrid, morphotypes. Population migration can be traced, for example, by back-calibrating models without the outlying low-elevation 3-needled trees to determine the likelihood of their being part of, or date of last contact with, the current Southwestern ponderosa pine population. Potential low-elevation areas containing “old” trees could be identified for dendrochronological research, or, in longer time periods, the progression of ponderosa pine through the region could be modeled and cross-calibrated with packrat midden data. Conversely, output from predictive 223 models can be projected to new scenarios, whether they be other mountain islands in the Southwest or altered climate, to generate probabilities of suitable habitat for the modeled species. Using ecophysiological methods to confirm the importance of climatic and other variables (e.g, mycorrhizal associations) can only improve variable selection and directions for continued research on dynamic species distributions. 224 APPENDIX 225 Table A-1. Geographic coordinates and mean needle number for trees from the distribution study ("SCAT"). Universal Transverse Mercator (UTM) coordinates are for Zone 128 projected to World Geodetic System 1984. Elevation was interpolated from Topozone.com. Morphotype is based on Peloquin (1984): Pinus ponderosa var. scopulorum (PIN_PON), P. arizonica (PIN_ARI), and putative hybrid (Mixed). 226 Mean Elevation needle WP UTM-X UTM—Y (m) number Morphotype 1 518198 3588043 2598.4 3.05 PIN_PON 1 518198 3588043 2598.4 3.63 Mixed 1 518198 3588043 2598.4 3.40 Mixed 1 518198 3588043 2598.4 3.61 Mixed 2 517666 3588386 2484.1 4.99 PIN_ARI 2 517666 3588386 2484.1 5.01 PIN_ARI 2 517666 3588386 2484.1 4.92 PIN_ARI 2 517666 3588386 2484.1 5.00 PIN_ARI 3 517260 3588735 2430.2 4.96 PIN_ARI 3 517260 3588735 2430.2 4.53 Mixed 3 517260 3588735 2430.2 4.88 PIN_ARI 3 517260 3588735 2430.2 5.05 PIN_ARI 4 516955 3589032 2346.0 4.86 PIN_ARI 4 516955 3589032 2346.0 5.00 PIN_ARI 4 516955 3589032 2346.0 4.89 PIN_ARI 4 516955 3589032 2346.0 4.91 PIN_ARI 5 516823 3589263 2368.8 5.00 PIN_ARI 5 516823 3589263 2368.8 4.98 PIN_ARI 5 516823 3589263 2368.8 4.65 PIN_ARI 5 516823 3589263 2368.8 5.00 Pl N_ARI 6 517210 3590125 2193.1 4.97 PIN_ARI 6 517210 3590125 2193.1 4.78 PIN_ARI 6 517210 3590125 2193.1 4.93 PIN_ARI 6 517210 3590125 2193.1 4.75 PIN_ARI 7 517328 3590761 2266.3 4.80 PIN_ARI 7 517328 3590761 2266.3 4.82 PIN_ARI 7 517328 3590761 2266.3 4.78 PIN_ARI 7 517328 3590761 2266.3 4.88 PIN_ARI 7 517328 3590761 2266.3 4.86 PIN_ARI 8 517298 3590996 2252.1 4.95 PIN_ARI 8 517298 3590996 2252.1 4.08 Mixed 8 517298 3590996 2252.1 4.03 Mixed 8 517298 3590996 2252.1 4.97 PIN_ARI 9 517636 3592413 2221.3 4.90 PIN_ARI 9 517636 3592413 2221.3 3.93 Mixed 9 517636 3592413 2221 .3 4.47 Mixed 9 517636 3592413 2221.3 4.59 Mixed 10 517417 3593325 2200.2 4.96 PIN_ARI 10 517417 3593325 2200.2 4.37 Mixed 10 517417 3593325 2200.2 4.74 PIN_ARI 10 517417 3593325 2200.2 4.54 Mixed 1 1 517429 3593667 2172.7 3.00 PIN_PON __11 51329 J 3593667 2172.7 3.00 PIN_:PQN Table A-1. Continued. 227 Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 11 517429 3593667 2172.7 3.02 PIN_PON 11 517429 3593667 2172.7 3.14 PIN_PON 17 525172 3587947 2251.3 4.99 PIN_ARI 17 525172 3587947 2251.3 4.71 PIN_ARI 18 525234 3588017 2246.9 4.53 Mixed 18 525234 3588017 2246.9 4.90 PIN_ARI 18 525234 3588017 2246.9 5.01 PIN_ARI 18 525234 3588017 2246.9 4.85 PIN_ARI 19 525617 3588455 2110.2 4.51 Mixed 19 525617 3588455 2110.2 4.80 PIN_ARI 19 525617 3588455 2110.2 4.37 Mixed 19 525617 3588455 2110.2 4.24 Mixed 20 525467 3588808 2060. 1 4. 88 PIN_ARI 20 525467 3588808 2060. 1 4.24 Mixed 20 525467 3588808 2060.1 4.98 PIN_ARI 20 525467 3588808 2060.1 4.60 Mixed 21 524432 3588773 2080.4 4. 81 PIN_ARI 21 524432 3588773 2080.4 4.77 PIN_ARI 21 524432 3588773 2080.4 4.55 Mixed 21 524432 3588773 2080.4 4.78 PIN_ARI 22 524420 3590284 2083.4 4.72 PIN_ARI 22 524420 3590284 2083.4 4.66 PIN_ARI 22 524420 3590284 2083.4 5. 00 PIN_ARl 22 524420 3590284 2083.4 4.91 PIN_ARI 23 524703 3590316 2083.6 4.80 PIN_ARI 23 524703 3590316 2083.6 4.70 PIN_ARI 23 524703 3590316 2083.6 4.98 PIN_ARI 23 524703 3590316 2083.6 4.94 PIN_ARI 24 524236 3590600 2134.2 4.72 PIN_ARI 24 524236 3590600 2134.2 4.65 PIN_ARI 24 524236 3590600 2134.2 4.69 PIN_ARI 24 524236 3590600 2134.2 4.99 PIN_ARI 25 523595 3590662 2256.7 4.60 Mixed 25 523595 3590662 2256.7 4.85 PIN_ARI 25 523595 3590662 2256.7 4.94 PIN_ARI 25 523595 3590662 2256.7 4. 79 PIN_ARI 26 523343 3590272 2328.0 5.00 PIN_ARI 26 523343 3590272 2328.0 4.99 PIN_ARI 26 523343 3590272 2328.0 4.47 Mixed 26 523343 3590272 2328.0 4.86 PIN_ARI 28 528669 3582880 2133.7 4.91 PIN_ARI 28 528669 3582880 2133.7 4. 70 PIN_ARI 28 528669 3582880 2133.7 4.54 Mixed 28 528669 3582880 2133.7 4.78 PIN_ARI 29 528529 3582872 2192.9 3.74 Mixed 29 528529 3582872 2192.9 4.97 PIN_ARI _29 528529 3582872 2192.9 4.84 . __ _ PINARI . Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (m) number Morghotype 29 528529 3582872 2192.9 4.77 PIN_ARI 30 528850 3582967 2135.3 4.93 PIN_ARI 30 528850 3582967 2135.3 4.77 PIN_ARI 30 528850 3582967 2135.3 4.32 Mixed 30 528850 3582967 2135.3 4.97 PIN_ARI 31 528339 3583418 2111.6 4.62 PIN_ARI 31 528339 3583418 2111.6 4.90 PIN_ARI 31 528339 3583418 2111.6 4.91 PIN_ARI 31 528339 3583418 2111.6 4.66 PIN_ARI 32 528636 3584094 2195.4 4.94 PIN_ARI 32 528636 3584094 2195.4 4.95 PIN_ARI 32 528636 3584094 2195.4 4.71 PIN_ARI 32 528636 3584094 2195.4 4.99 PIN_ARI 33 527366 3583739 2131.1 5.39 PIN_ARI 34 527280 3583339 2125.7 3.64 Mixed 34 527280 3583339 2125.7 5.04 PIN_ARI 34 527280 3583339 2125.7 4.72 PIN_ARI 34 527280 3583339 2125.7 4.95 PIN_ARI 35 525564 3586049 2504.2 3.56 Mixed 35 525564 3586049 2504.2 3.00 PIN_PON 35 525564 3586049 2504.2 3.02 PIN_PON 35 525564 3586049 2504.2 3.22 Mixed 36 525571 3585999 2492.3 3.07 PIN_PON 36 525571 3585999 2492.3 3.00 PIN_PON 37 526461 3585865 2400.7 5.15 PIN_ARI 37 526461 3585865 2400.7 5.04 PIN_ARI 37 526461 3585865 2400.7 4.61 PIN_ARI 37 526461 3585865 2400.7 4.99 PIN_ARI 37 526461 3585865 2400.7 5.01 PIN_ARI 38 527040 3585860 2428.2 4.97 PIN_ARI 38 527040 3585860 2428.2 4.79 PIN_ARI 38 527040 3585860 2428.2 4.96 PIN_ARI 38 527040 3585860 2428.2 4.99 PIN_ARI 39 526097 3586082 2487.4 3.46 Mixed 39 526097 3586082 2487.4 3.05 PIN_PON 39 526097 3586082 2487.4 4.91 PIN_ARI 39 526097 3586082 2487.4 3.06 PIN_PON 40 526177 3586133 2526.3 3.13 PIN_PON 40 526177 3586133 2526.3 3.38 Mixed 40 526177 3586133 2526.3 4.07 Mixed 40 526177 3586133 2526.3 3.01 PIN_PON 41 526231 3586172 2546.1 3.24 Mixed 41 526231 3586172 2546.1 3.02 PIN_PON 41 526231 3586172 2546.1 3.91 Mixed 41 526231 3586172 2546.1 3.37 Mixed 42 526206 3586225 2572.3 3.70 Mixed 42 526206 _3586225 2572.3 3.52_____ y Mixed .. 228 Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 42 526206 3586225 2572. 3 3. 00 PIN_PON 42 526206 3586225 2572.3 3.00 PIN_PON 43 524405 3588019 2335.8 3.21 Mixed 43 524405 3588019 2335.8 3.05 PIN_PON 43 524405 3588019 2335.8 3.29 Mixed 43 524405 3588019 2335.8 2.99 PIN_PON 44 524395 3587937 2342. 5 5.01 PIN_ARI 44 524395 3587937 2342. 5 3.24 Mixed 44 524395 3587937 2342.5 4. 99 PIN_ARI 44 524395 3587937 2342. 5 3.00 PIN_PON 47 525002 3587025 2475.0 3.00 PIN_PON 47 525002 3587025 2475.0 3.36 Mixed 47 525002 3587025 2475.0 3.02 PIN_PON 47 525002 3587025 2475.0 3.01 PIN_PON 48 524884 3586716 2499.4 3.60 Mixed 48 524884 3586716 2499.4 3.23 Mixed 48 524884 3586716 2499.4 3.03 PIN_PON 48 524884 3586716 2499.4 3.08 PIN_PON 49 525728 3586553 2547.8 3.26 Mixed 49 525728 3586553 2547.8 3.00 PIN_PON 49 525728 3586553 2547.8 3.01 PIN_PON 49 525728 3586553 2547.8 3.02 PIN_PON 50 526064 3586493 2541 .3 3.34 Mixed 50 526064 3586493 2541 .3 3.03 PIN_PON 50 526064 3586493 2541.3 4.13 Mixed 50 526064 3586493 2541 .3 4.08 Mixed 50 526064 3586493 2541 .3 4. 34 Mixed 51 526931 3586640 2581 .4 3.03 PIN_PON 51 526931 3586640 2581.4 3.04 PIN_PON 51 526931 3586640 2581.4 3.01 PIN_PON 51 526931 3586640 2581.4 3.01 PIN_PON 52 529524 3584657 2198.4 4.95 PIN_ARI 52 529524 3584657 2198.4 4.98 PIN_ARI 52 529524 3584657 2198.4 4.95 PIN_ARI 52 529524 3584657 2198.4 5.00 PIN_ARI 53 529829 3584400 2193.1 5.03 PIN_ARI 53 529829 3584400 2193.1 4.99 PIN_ARI 53 529829 3584400 2193.1 4.92 PIN_ARI 53 529829 3584400 2193.1 4.79 PIN_ARI 54 530378 358451 1 2010.5 4.64 PIN_ARI 54 530378 358451 1 2010.5 4.95 PIN_ARI 54 530378 3584511 2010.5 4.64 PIN_ARI 54 530378 358451 1 2010.5 4.94 PIN_ARI 55 530828 3584110 2110.0 5.08 PIN_ARI 55 530828 3584110 2110.0 4.82 PIN_ARI 55 530828 3584110 2110.0 4.97 PIN_ARI 55 530828 35841107, g1 10 0 _4_.81 _ _ PIN:ARI_ _ _ Table A-1. Continued. 230 Mean Elevation needle WP UTM-X UTM-Y (m) number Mclphotype 57 531064 3583808 2162.7 4.96 PIN_ARI 57 531064 3583808 2162.7 4.92 PIN_ARI 57 531064 3583808 2162.7 4.42 Mixed 57 531064 3583808 2162.7 4.99 PIN_ARI 58 531692 3583274 2212.7 4.03 Mixed 58 531692 3583274 2212.7 4.11 Mixed 58 531692 3583274 2212.7 4.38 Mixed 58 531692 3583274 2212.7 3.81 Mixed 59 532261 3583539 2171.8 4.20 Mixed 59 532261 3583539 2171.8 4.98 PIN_ARI 59 532261 3583539 2171.8 5.00 PIN_ARI 59 532261 3583539 2171.8 4.46 Mixed 60 530739 3583800 2073.1 4.98 PIN_ARI 60 530739 3583800 2073.1 4.99 PIN_ARI 60 530739 3583800 2073.1 4.80 PIN_ARI 60 530739 3583800 2073.1 4.82 PIN_ARI 61 530500 3583244 2031.5 4.52 Mixed 61 530500 3583244 2031.5 4.96 PIN_ARI 61 530500 3583244 2031.5 4.96 PIN_ARI 61 530500 3583244 2031.5 4.87 PIN_ARI 62 530500 3583244 2031 .5 4.82 PIN_ARI 62 530500 3583244 2031 .5 4.96 PIN_ARI 62 530500 3583244 2031 .5 4.95 PIN_ARI 62 530500 3583244 2031 .5 5.00 PIN_ARI 65 523114 3590540 2378.6 4.86 PIN_ARI 65 523114 3590540 2378.6 4.99 PIN_ARI 65 523114 3590540 2378.6 4.84 PIN_ARI 65 523114 3590540 2378.6 3.44 Mixed 66 523015 3590116 2413.9 3.74 Mixed 66 523015 3590116 2413.9 3.40 Mixed 66 523015 3590116 2413.9 3.01 PIN_PON 67 522749 3589887 2406.1 3.23 Mixed 67 522749 3589887 2406.1 3.05 PIN_PON 67 522749 3589887 2406.1 3.00 PIN_PON 67 522749 3589887 2406.1 3.08 PIN_PON 68 523072 3589232 2438.0 4.91 PIN_ARI 68 523072 3589232 2438.0 3.00 PIN_PON 68 523072 3589232 2438.0 3.33 Mixed 68 523072 3589232 2438.0 3.00 PIN_PON 68 523072 3589232 2438.0 4.99 PIN_ARI 69 523212 3589207 2436.1 6.50 PIN_ARI 69 523212 3589207 2436.1 3.00 PIN_PON 69 523212 3589207 2436.1 3.33 Mixed 69 523212 3589207 2436.1 3.03 PIN_PON 70 521773 3590551 2398.3 5.00 PIN_ARI 70 521773 3590551 2398.3 4.99 PIN_ARI ‘70 521773 3590551 2398.3 5.23 “__PlNgARLH Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 70 521773 3590551 2398.3 5.02 PIN_ARI 71 521997 3592420 2069.5 4.87 PIN_ARI 71 521997 3592420 2069.5 4.65 PIN_ARI 71 521997 3592420 2069.5 4.88 PIN_ARI 71 521997 3592420 2069.5 4.88 PIN_ARI 72 522002 3593674 1745.0 3.00 PIN_PON 72 522002 3593674 1745.0 3.00 PIN_PON 72 522002 3593674 1745.0 3.00 PIN_PON 72 522002 3593674 1745.0 3.00 PIN_PON 73 523628 3593676 2127.0 3.1 1 PIN_PON 73 523628 3593676 2127.0 3.07 PIN_PON 73 523628 3593676 2127.0 2.98 PIN_PON 73 523628 3593676 2127.0 3.02 PIN_PON 74 524046 3592746 2295.7 3.23 Mixed 74 524046 3592746 2295.7 3.04 PIN_PON 74 524046 3592746 2295.7 3.01 PIN_PON 74 524046 3592746 2295.7 3.00 PIN_PON 75 523059 3587849 2268.4 3.05 PIN_PON 75 523059 3587849 2268.4 2.99 PIN_PON 75 523059 3587849 2268.4 3.00 PIN_PON 75 523059 3587849 2268.4 3.00 PIN_PON 76 522747 3588693 2321 .7 3.49 Mixed 76 522747 3588693 2321 .7 3.29 Mixed 76 522747 3588693 2321 .7 3.24 Mixed 76 522747 3588693 2321 .7 3.05 PIN_PON 77 522552 3589178 2344.7 3.01 PIN_PON 77 522552 3589178 2344.7 3.00 PIN_PON 77 522552 3589178 2344.7 3.02 PIN_PON 77 522552 3589178 2344.7 3.00 PIN_PON 78 521964 3590069 2422.0 3.03 PIN_PON 78 521964 3590069 2422.0 3.00 PIN_PON 78 521964 3590069 2422.0 3.06 PIN_PON 78 521964 3590069 2422.0 3.35 Mixed 79 521731 3590186 2465.4 3.22 Mixed 79 521731 3590186 2465.4 3.03 PIN_PON 79 521731 3590186 2465.4 3.01 PIN_PON 79 521731 3590186 2465.4 2.99 PIN_PON 80 520297 3590584 2572.5 3.00 PIN_PON 80 520297 3590584 2572.5 4.31 Mixed 80 520297 3590584 2572.5 3.25 Mixed 80 520297 3590584 2572.5 3.06 PIN_PON 81 520952 3589507 2680.4 3.13 PIN_PON 81 520952 3589507 2680.4 3.1 1 PIN_PON 81 520952 3589507 2680.4 4.99 PIN_ARI 81 520952 3589507 2680.4 3.99 Mixed 82 520361 3589357 2756.9 4.02 Mixed 82 529361 3589357 2756.79 3.64 _Mixed __ 231 Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 82 520361 3589357 2756.9 3.06 PIN_PON 82 520361 3589357 2756.9 3.68 Mixed 83 521521 3589530 2625.6 3.00 PIN_PON 85 523068 3587618 2314.2 4.85 PIN_ARI 85 523068 3587618 2314.2 4.97 PIN_ARI 85 523068 3587618 2314.2 4.98 PIN_ARI 85 523068 3587618 2314.2 4.98 PIN_ARI 86 522678 3587265 2410.5 5.00 PIN_ARI 86 522678 3587265 2410.5 5.07 PIN_ARI 86 522678 3587265 2410.5 5.10 PIN_ARI 86 522678 3587265 2410.5 5.00 PIN_ARI 87 521895 3587265 2472.8 4.64 PIN_ARI 87 521895 3587265 2472.8 5.00 PIN_ARI 87 521895 3587265 2472.8 4.86 PIN_ARI 87 521895 3587265 2472.8 4.89 PIN_ARI 88 521510 3587614 2478.7 4.99 PIN_ARI 88 521510 3587614 2478.7 5.00 PIN_ARI 88 521510 3587614 2478.7 4.98 PIN_ARI 88 521510 3587614 2478.7 5.00 PIN_ARI 89 521313 3588163 2434.8 4.90 PIN_ARI 89 521313 3588163 2434.8 4.52 Mixed 89 521313 3588163 2434.8 4.97 PIN_ARI 89 521313 3588163 2434.8 4.64 PIN_ARI 90 521837 3588153 2363.9 3.00 PIN_PON 90 521837 3588153 2363.9 3.04 PIN_PON 90 521837 3588153 2363.9 3.06 PIN_PON 90 521837 3588153 2363.9 3.00 PIN_PON 91 521198 3589037 2613.4 3.01 PIN_PON 91 521198 3589037 2613.4 3.02 PIN_PON 91 521198 3589037 2613.4 3.20 PIN_PON 91 521198 3589037 2613.4 3.70 Mixed 92 521161 3588666 2481.9 4.95 PIN_ARI 92 521161 3588666 2481.9 5.40 PIN_ARI 92 521161 3588666 2481.9 3.00 PIN_PON 92 521161 3588666 2481.9 4.60 PIN_ARI 93 520867 3588430 2372.2 4.98 PIN_ARI 93 520867 3588430 2372.2 4.95 PIN_ARI 93 520867 3588430 2372.2 3.51 Mixed 93 520867 3588430 2372.2 4.97 PIN_ARI 94 522536 3588172 2417.8 4.97 PIN_ARI 94 522536 3588172 2417.8 5.01 PIN_ARI 94 522536 3588172 2417.8 4.92 PIN_ARI 94 522536 3588172 2417.8 4.81 PIN_ARI 95 522408 3587860 2331 .1 5.00 PIN_ARI 95 522408 3587860 2331.1 3.00 PIN_PON 95 522408 3587860 2331.1 3.05 PIN_PON 95 522408 3587860 2331.1 5.00 _PIN ARI _ __ _ 232 Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y Jim number Morphotype 97 519443 3589532 2714.6 3.11 PIN_PON 97 519443 3589532 2714.6 3.00 PIN_PON 97 519443 3589532 2714.6 3.00 PIN_PON 97 519443 3589532 2714.6 3.14 PIN_PON 98 525491 3584718 2305.2 4.92 PIN_ARI 98 525491 3584718 2305.2 4.94 PIN_ARI 98 525491 3584718 2305.2 4.86 PIN_ARI 98 525491 3584718 2305.2 4.95 PIN_ARI 99 525777 3584681 2317.9 4.96 PIN_ARI 99 525777 3584681 2317.9 4.89 PIN_ARI 99 525777 3584681 2317.9 5.01 PIN_ARl 99 525777 3584681 2317.9 5.00 PIN_ARI 100 525551 3584070 2291.8 4.93 PIN_ARI 100 525551 3584070 2291.8 4.98 PIN_ARI 100 525551 3584070 2291 .8 4.44 Mixed 100 525551 3584070 2291.8 4.94 PIN_ARI 101 526357 3585193 2365.6 4.98 PIN_ARI 101 526357 3585193 2365.6 4.41 Mixed 101 526357 3585193 2365.6 5.1 1 PIN_ARI 101 526357 3585193 2365.6 5.02 PIN_ARI 102 526350 3585532 2395.7 4.87 PIN_ARI 102 526350 3585532 2395.7 4.95 PIN_ARI 102 526350 3585532 2395. 7 4.97 PIN_ARI 102 526350 3585532 2395.7 5.00 PIN_ARI 103 525110 3585720 2478.0 4.60 PIN_ARI 103 5251 10 3585720 2478.0 4.55 Mixed 103 5251 10 3585720 2478.0 4.84 PIN_ARl 103 5251 10 3585720 2478.0 4.97 PIN_ARI 104 525208 3586246 2489.2 3.10 PIN_PON 104 525208 3586246 2489.2 3.97 Mixed 104 525208 3586246 2489.2 3.40 Mixed 104 525208 3586246 2489.2 3.59 Mixed 105 528133 3585823 2482.1 4.40 Mixed 105 528133 3585823 2482.1 4.78 PIN_ARI 105 528133 3585823 2482.1 4.98 PIN_ARI 106 528204 3586238 2313.3 4.90 PIN_ARI 106 528204 3586238 2313.3 4.96 PIN_ARI 106 528204 3586238 2313.3 4.95 PIN_ARI 107 527998 3586344 2340.9 4.93 PIN_ARI 107 527998 3586344 2340.9 4.98 PIN_ARI 107 527998 3586344 2340.9 4.38 Mixed 107 527998 3586344 2340.9 4.85 PIN_ARI 108 527884 3586038 2467.4 5.1 1 PIN_ARI 108 527884 3586038 2467.4 5.45 PIN_ARI 108 527884 3586038 2467.4 4.57 Mixed 108 527884 3586038 2467.4 4.92 PIN_ARI #109 524358 3585856 2355.0 4.99 WPIN_ARI 233 Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 109 524358 3585856 2355.0 4.95 PI N_ARI 109 524358 3585856 2355.0 4.97 Pl N_ARI 109 524358 3585856 2355.0 4.89 Pl N_ARI 1 10 524331 3586423 2379.9 4.99 PIN_ARI 1 10 524331 3586423 2379.9 4.79 PIN_ARI 1 10 524331 3586423 2379.9 5.00 PIN_ARI 1 10 524331 3586423 2379.9 4.83 PIN_ARI 1 11 524607 3586710 2419.0 4.90 PIN_ARI 1 11 524607 3586710 2419.0 4.88 PIN_ARI 1 1 1 524607 3586710 2419.0 3.72 Mixed 1 11 524607 3586710 2419.0 4.78 PIN_ARI 1 12 524527 3586971 2451.1 4.94 PIN_ARI 1 12 524527 3586971 2451 .1 3.00 PIN_PON 1 12 524527 3586971 2451.1 2.99 PIN_PON 1 12 524527 3586971 2451.1 4.49 Mixed 1 13 524083 3587031 2490.7 5.06 PIN_ARI 1 13 524083 3587031 2490.7 4.95 PIN_ARl 1 13 524083 3587031 2490.7 5.18 PIN_ARI 1 13 524083 3587031 2490.7 4.87 PIN_ARI 1 14 526916 3587500 2338.0 4.86 PIN_ARI 1 14 526916 3587500 2338.0 4.52 Mixed 1 14 526916 3587500 2338.0 4.98 PIN_ARI 1 15 527095 3587904 2309.0 4.65 PIN_ARI 1 15 527095 3587904 2309.0 4.58 Mixed 1 15 527095 3587904 2309.0 3.07 PIN_PON 1 15 527095 3587904 2309.0 4.85 PIN_ARl 1 16 527123 3588052 2232.9 4.90 PIN_ARI 1 16 527123 3588052 2232.9 4.95 PIN_ARI 1 16 527123 3588052 2232.9 4.74 PIN_ARI 1 17 526723 3588045 2172.7 4.97 PIN_ARI 1 17 526723 3588045 2172.7 4.76 PIN_ARI 117 526723 3588045 2172.7 5.13 PIN_ARI 1 18 526559 3587741 2108.5 5.00 PIN_ARI 1 18 526559 3587741 2108.5 4.66 PIN_ARI 1 18 526559 3587741 2108.5 4.88 PIN_ARI 1 19 525808 3587904 2142.4 4.99 PIN_ARI 1 19 525808 3587904 2142.4 5.02 PIN_ARI 1 19 525808 3587904 2142.4 4.80 PIN_ARI 120 527424 3586021 2419.8 4.81 PIN_ARI 120 527424 3586021 2419.8 4.99 PIN_ARI 120 527424 3586021 2419.8 4.97 PIN_ARI 120 527424 3586021 2419.8 4.98 PIN_ARI 121 527600 3586131 2451.6 4.96 PIN_ARI 121 527600 3586131 2451.6 4.95 PIN_ARI 121 527600 3586131 2451.6 5.00 PIN_ARI 121 527600 3586131 2451.6 4.75 PIN_ARI 122‘ _ 527450 J586192 __247_0_.8 4.94 PIN_ARI 234 Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (rm number Morphotype 122 527450 3586192 2470.8 4.94 PIN_ARI 122 527450 3586192 2470.8 4.97 PIN_ARI 122 527450 3586192 2470.8 4.70 PIN_ARI 123 527388 3586363 2554.3 3.16 PIN_PON 123 527388 3586363 2554.3 2.99 PIN_PON 124 527265 3586402 2551.7 3.02 PIN_PON 124 527265 3586402 2551.7 3.14 PIN_PON 124 527265 3586402 2551 .7 3.42 Mixed 124 527265 3586402 2551.7 3.00 PIN_PON 125 527129 3586406 2521.1 3.26 Mixed 125 527129 3586406 2521.1 3.03 PIN_PON 125 527129 3586406 2521.1 3.04 PIN_PON 125 527129 3586406 2521.1 3.01 PIN_PON 126 527059 3586353 2508.2 3.01 PIN_PON 126 527059 3586353 2508.2 4.94 PIN_ARI 126 527059 3586353 2508.2 4.98 PIN_ARI 126 527059 3586353 2508.2 4.45 Mixed 127 519668 3589289 2768.9 3.06 PIN_PON 127 519668 3589289 2768.9 3.52 Mixed 127 519668 3589289 2768.9 3.08 PIN_PON 128 519615 3589131 2767.8 3.70 Mixed 128 519615 3589131 2767.8 3.68 Mixed 128 519615 3589131 2767.8 3.05 PIN_PON 128 519615 3589131 2767.8 3.08 PIN_PON 129 519183 3588867 2700.6 3.38 Mixed 129 519183 3588867 2700.6 3.00 PIN_PON 129 519183 3588867 2700.6 3.03 PIN_PON 129 519183 3588867 2700.6 3.63 Mixed 130 519136 3588734 2682.0 3.00 PIN_PON 130 519136 3588734 2682.0 3.00 PIN_PON 130 519136 3588734 2682.0 3.37 Mixed 130 519136 3588734 2682.0 3.02 PIN_PON 131 518926 3588551 2677.5 3.07 PIN_PON 131 518926 3588551 2677.5 3.38 Mixed 131 518926 3588551 2677.5 3.03 PIN_PON 131 518926 3588551 2677.5 2.93 PIN_PON 132 519369 3588742 2689.8 3.52 Mixed 132 519369 3588742 2689.8 3.97 Mixed 132 519369 3588742 2689.8 3.56 Mixed 132 519369 3588742 2689.8 2.99 PIN_PON 132 519369 3588742 2689.8 3.28 Mixed 133 519716 3588880 2712.0 3.01 PIN_PON 133 519716 3588880 2712.0 3.01 PIN_PON 133 519716 3588880 2712.0 3.14 PIN_PON 133 519716 3588880 2712.0 3.00 PIN_PON 134 519878 3588961 2730.1 3.00 PIN_PON 134 519878 3588961 2739.1 3.38 Mixed 7 a.-- 235 Table A-1. Continued. 236 Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 134 519878 3588961 2730.1 3.09 PIN_PON 134 519878 3588961 2730.1 3.42 Mixed 135 529463 3584453 2334.9 4.84 PIN_ARI 135 529463 3584453 2334.9 4.75 PIN_ARI 135 529463 3584453 2334.9 4.42 Mixed 135 529463 3584453 2334.9 4.77 PIN_ARI 136 529496 3584309 2402.2 4.38 Mixed 1 36 529496 3584309 2402.2 5.22 PIN_ARI 136 529496 3584309 2402.2 4.68 PIN__ARI 136 529496 3584309 2402.2 4.32 Mixed 137 529406 3584161 2407.9 4.60 PIN_ARI 137 529406 3584161 2407.9 5.00 PIN_ARI 137 529406 3584161 2407.9 4.91 PIN_ARI 137 529406 3584161 2407.9 5.01 PIN_ARI 138 529469 3584041 2396.2 4.86 PIN_ARI 138 529469 3584041 2396.2 4.99 PIN_ARI 138 529469 3584041 2396.2 4.67 PIN_ARI 138 529469 3584041 2396.2 5.00 PIN_ARI 139 529396 3583820 2370.2 4.31 Mixed 139 529396 3583820 2370.2 4.99 PIN_ARI 139 529396 3583820 2370.2 5.00 PIN_ARI 139 529396 3583820 2370.2 5.00 PIN_ARI 140 529244 3583659 2319.5 5.02 PIN_ARI 140 529244 3583659 2319.5 4.95 PIN_ARI 140 529244 3583659 2319.5 5.00 PIN_ARI 140 529244 3583659 2319.5 5.25 PIN_ARI 141 528978 3583449 2228.5 5.00 PIN_ARI 141 528978 3583449 2228.5 5.10 PIN_ARI 141 528978 3583449 2228.5 4.80 PIN_ARI 141 528978 3583449 2228.5 4.95 PIN_ARI 142 528739 3583426 2173.4 4.94 PIN_ARI 142 528739 3583426 2173.4 4.89 PIN_ARI 142 528739 3583426 2173.4 4.89 PIN_ARI 142 528739 3583426 2173.4 4.88 PIN_ARI 143 528664 3584663 2190.9 4.96 PIN_ARI 143 528664 3584663 2190.9 4.51 Mixed 143 528664 3584663 2190.9 4.95 PIN_ARI 143 528664 3584663 2190.9 4.95 PIN_ARI 144 527887 3584178 2167.9 4.85 PIN_ARI 144 527887 3584178 2167.9 4.83 PIN_ARI 144 527887 3584178 2167.9 4.99 PIN_ARI 144 527887 3584178 2167.9 4.87 PIN_ARI 145 527156 3583655 2145.8 4.92 PIN_ARI 145 527156 3583655 2145.8 4.86 PIN_ARI 145 527156 3583655 2145.8 4.92 PIN_ARI 145 527156 3583655 2145.8 4.65 PIN_ARI _1:l§__.53844§ 358§ZZ§._24§_9.§_- -4-68 Plus: Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 146 528445 3585778 2439.8 4.98 PIN_ARI 146 528445 3585778 2439.8 5.03 PIN_ARI 146 528445 3585778 2439.8 4.92 PIN_ARI 147 528272 3585635 2397.0 4.70 PIN_ARI 147 528272 3585635 2397.0 4.85 PIN_ARI 147 528272 3585635 2397.0 5.04 PIN_ARI 147 528272 3585635 2397.0 4.81 PIN_ARI 148 528482 3585409 2376.9 4.97 PIN_ARI 148 528482 3585409 2376.9 4.97 PIN_ARI 148 528482 3585409 2376.9 4.71 PIN_ARI 148 528482 3585409 2376.9 4.97 PIN_ARI 149 528480 3585146 2409.1 5.00 PIN_ARI 149 528480 3585146 2409.1 4.93 PIN_ARI 149 528480 3585146 2409.1 5.22 PIN_ARI 149 528480 3585146 2409.1 5.02 PIN_ARI 150 528592 3585010 2341.7 4.97 PIN_ARI 150 528592 3585010 2341.7 5.13 PIN_ARI 150 528592 3585010 2341.7 4.76 PIN_ARI 150 528592 3585010 2341.7 4.98 PIN_ARI 151 528877 3585124 2293.6 4.89 PIN_ARI 151 528877 3585124 2293.6 5.37 PIN_ARI 151 528877 3585124 2293.6 4.99 PIN_ARI 151 528877 3585124 2293.6 4.78 PIN_ARI 152 529181 3584828 2244.9 5.12 PIN_ARI 152 529181 3584828 2244.9 4.73 PIN_ARI 152 529181 3584828 2244.9 4.89 PIN_ARI 152 529181 3584828 2244.9 4.94 PIN_ARI 153 527783 3584801 2268.5 4.97 PIN_ARI 153 527783 3584801 2268.5 5.20 PIN_ARI 153 527783 3584801 2268. 5 4.93 PIN_ARI 153 527783 3584801 2268.5 5.00 PIN_ARI 154 527563 3585623 2334.1 4.98 PIN_ARI 154 527563 3585623 2334.1 4.96 PIN_ARI 154 527563 3585623 2334.1 4.96 PIN_ARI 154 527563 3585623 2334.1 5.03 PIN_ARl 155 528120 3585438 2310.3 4.96 PIN_ARI 155 528120 3585438 2310.3 4.70 PIN_ARI 155 528120 3585438 2310.3 4.99 PIN_ARI 155 528120 3585438 2310.3 4.97 PIN_ARI 1 56 525955 3586763 2556.7 3.63 Mixed 156 525955 3586763 2556.7 3.07 PIN_PON 156 525955 3586763 2556.7 3.00 PIN_PON 1 56 525955 3586763 2556. 7 3.20 Mixed 157 525565 3586272 2595.4 3.06 PIN_PON 157 525565 3586272 2595.4 3.13 PIN_PON 1 57 525565 3586272 2595.4 3.94 Mixed _1_ 57 525565 3586272 25954 _3.07___7___ _ __PflflgPQN __ 237 Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 158 525782 3586502 2561.0 3.15 PIN_PON 158 525782 3586502 2561.0 3.44 Mixed 158 525782 3586502 2561.0 4.01 Mixed 158 525782 3586502 2561.0 3.17 PIN_PON 159 525333 3586998 2432.5 3.08 PIN_PON 159 525333 3586998 2432.5 3.03 PIN_PON 159 525333 3586998 2432.5 3.01 PIN_PON 159 525333 3586998 2432.5 3.00 PIN_PON 160 520065 3589172 2769.7 3.22 Mixed 160 520065 3589172 2769.7 3.01 PIN_PON 160 520065 3589172 2769.7 3.01 PIN_PON 160 520065 3589172 2769.7 3.40 Mixed 161 526520 3586370 2558.6 3.06 PIN_PON 161 526520 3586370 2558.6 3.01 PIN_PON 162 526504 3586244 2564.1 3.06 PIN_PON 162 526504 3586244 2564.1 3.35 Mixed 162 526504 3586244 2564.1 3.24 Mixed 162 526504 3586244 2564.1 4.51 Mixed 163 525218 3586605 2507.9 3.00 PIN_PON 163 525218 3586605 2507.9 3.00 PIN_PON 163 525218 3586605 2507.9 3.01 PIN_PON 163 525218 3586605 2507.9 3.02 PIN_PON 164 524790 3587570 2407.7 3.71 Mixed 164 524790 3587570 2407.7 4.86 PIN_ARI 164 524790 3587570 2407.7 3.00 PIN_PON 165 523505 3588328 2439.0 3.20 Mixed 165 523505 3588328 2439.0 5.10 PIN_ARI 165 523505 3588328 2439.0 4.98 PIN_ARI 165 523505 3588328 2439.0 4.02 Mixed 166 528831 3583112 2125.6 4.76 PIN_ARI 166 528831 3583112 2125.6 3.10 PIN_PON 166 528831 3583112 2125.6 4.99 PIN_ARI 166 528831 3583112 2125.6 3.39 Mixed 167 529014 3583212 2197.0 4.99 PIN_ARI 167 529014 3583212 2197.0 4.91 PIN_ARI 167 529014 3583212 2197.0 4.93 PIN_ARI 167 529014 3583212 2197.0 5.00 PIN_ARI 168 529195 3583607 2309.8 5.00 PIN_ARI 168 529195 3583607 2309.8 4.96 PIN_ARI 168 529195 3583607 2309.8 5.03 PIN_ARI 168 529195 3583607 2309.8 5.38 PIN_ARI 169 528976 3583398 2221.8 4.99 PIN_ARI 169 528976 3583398 2221.8 4.78 PIN_ARI 169 528976 3583398 2221.8 4.88 PIN_ARI 169 528976 3583398 2221.8 4.73 PIN_ARI 170 519676 3587162 2210.6 4.46 Mixed _170 519676 3587162 2210.6 4.89 _PlNJflM 238 Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 170 519676 3587162 2210.6 4.47 Mixed 170 519676 3587162 2210.6 5.00 PIN_ARI 171 518348 3586331 2130.8 5.00 PIN_ARI 171 518348 3586331 2130.8 4.86 PIN_ARI 171 518348 3586331 2130.8 5.00 PIN_ARI 171 518348 3586331 2130.8 4.99 PIN_ARI 172 517083 3585705 2233.9 4.98 PIN_ARI 172 517083 3585705 2233.9 5.1 1 PIN_ARI 172 517083 . 3585705 2233.9 4.88 PIN_ARI 172 517083 3585705 2233.9 4.96 PIN_ARI 173 516251 3585092 2014.8 3.00 PIN_PON 174 516431 3585241 2066.6 3.58 Mixed 174 516431 3585241 2066.6 3.13 PIN_PON 174 516431 3585241 2066.6 3.01 PIN_PON 174 516431 3585241 2066.6 4.26 Mixed 175 516888 3585400 2215.3 4.97 PIN_ARI 175 516888 3585400 2215.3 4.10 Mixed 175 516888 3585400 2215.3 4.99 PIN_ARI 175 516888 3585400 2215.3 4.99 PIN_ARI 177 517261 3586397 2301 .9 4.56 Mixed 177 517261 3586397 2301 .9 4.99 PIN_ARI 177 517261 3586397 2301.9 4.84 PIN_ARI 177 517261 3586397 2301.9 4.98 PIN_ARI 178 517653 3587136 2330.9 4.95 PIN_ARI 178 517653 3587136 2330.9 4.99 PIN_ARI 178 517653 3587136 2330.9 4.96 PIN_ARI 178 517653 3587136 2330.9 4.96 PIN_ARI 179 517969 3587762 2493.1 3.01 PIN_PON 179 517969 3587762 2493.1 4.83 PIN_ARI 179 517969 3587762 2493.1 4.87 PIN_ARI 179 517969 3587762 2493.1 5.00 PIN_ARI 180 524999 3586496 2449.3 3.01 PIN_PON 180 524999 3586496 2449.3 4.98 PIN_ARI 180 524999 3586496 2449.3 4.90 PIN_ARI 180 524999 3586496 2449.3 3.13 PIN_PON 181 524595 3587330 2458.1 4.42 Mixed 181 524595 3587330 2458.1 3.41 Mixed 181 524595 3587330 2458.1 4.96 PIN_ARI 181 524595 3587330 2458.1 4.78 PIN_ARI 182 523842 3587780 2354.8 4.50 Mixed 182 523842 3587780 2354.8 4.97 PIN_ARI 182 523842 3587780 2354.8 4.99 PIN_ARI 182 523842 3587780 2354.8 4.99 PIN_ARI 183 522592 3584399 2155.0 4.93 PIN_ARI 183 522592 3584399 2155.0 4.36 Mixed 183 522592 3584399 2155.0 3.99 Mixed .183 32239.2 9.594399 215.5..9__411_ PIN_ABI- 239 Table A-1. Continued. Mean Elevation needle WP UTM-X UTM-Y (m) number Morphotype 184 522830 3585258 2248.2 4.87 PIN_ARI 184 522830 3585258 2248.2 4.88 PIN_ARI 184 522830 3585258 2248.2 4.18 Mixed 184 522830 3585258 2248.2 4.86 PIN_ARI 185 523842 3585874 2413.9 4.97 PIN_ARI 185 523842 3585874 2413.9 4.98 PIN_ARI 185 523842 3585874 2413.9 4.64 PIN_ARI 185 523842 3585874 2413.9 5.01 PIN_ARI 186 529520 3582208 1814.2 4.94 PIN_ARI 186 529520 3582208 1814.2 4.83 PIN_ARI 186 529520 3582208 1814.2 4.81 PIN_ARI 186 529520 3582208 1814.2 4.99 PIN_ARI 187 528657 3581733 1782.6 4.50 Mixed 187 528657 3581733 1782.6 4.55 Mixed 187 528657 3581733 1782.6 3.98 Mixed 187 528657 3581733 1782.6 4.72 PIN_ARI 188 528111 3581568 1758.0 5.01 PIN_ARI 240 .. 00050.0 00050.0 00000.0 00000.0 00050.0 00000.0 00000.0 00000.0 0000.0 50000.0 00500.0 00550.0 000 00050.0 P 5 3000.0 50550.0 00000.0 50000.0 00500.0 00000.0 00500.0 00000.0 50000.0 00000.0 52 00050.0 5 5 3000.0 ..0550.0 00000.0 50000.0 00500.0 00000.0 00500.0 00000.0 50000.0 00000.0 80 00000.0 $000.0 $000.0 5 00000.0 00050.0 00500.0 05000.0 00000.0 00000.0 5000.0 .5030 00000.0 .000 00000.0 50550.0 50550.0 00000.0 5 00000.0 00000.0 00050.0 00000.0 00000.0 00050.0 00000.0 05000.0 0:0. 00050.0 00000.0 00000.0 00050.0 00000.0 5 05000.0 00000.0 00000.0 0000.0 0000.0 00000.0 00050.0 5.0.. 00000.0 50000.0 50000.0 00500.0 00000.0 0.0000 F 50000.0 00000.0 0000.0 00000.0 00000.0 05050.0 000.. 00000.0 00500.0 00500.0 05000.0 00050.0 00000.0 50000.0 5 00000.0 00500.0 00000.0 00000.0 00000.0 50.2 00000.0 00000.0 00000.0 00000.0 00000.0 00000.0 00000.0 00000.0 .. 00000.0 00000.0 300.0 5000.0 a< 0000.0 00500.0 00500.0 00000.0 00000.0 0000.0 0000.0 00500.0 00000.0 5 00000.0 00000.0 0.5000 00.2 50000.0 00000.0 00000.0 50000.0 00050.0 0000.0 00000.0 00000.0 00000.0 00000.0 F 05000.0 0000.0 00“. 005000 ..00000 50000.0 55050.0 00000.0 00000.0 00000.0 00000.0 $00.0 00000.0 05000.0 5 00000.0 00.. 00550.0 00000.0 00000.0 00000.0 05000.0 00050.0 05050.0 00000.0 5000.0 0500.0 0000.0 00000.0 5 .0:00< 000 52 50 .000 004. 5.3.. 000.. 50.2 a... 00.2 00“. 00.. .0000... QE050§ 5 05000.0 00000.0 00000.0 05000.0 50500.0 00050.0 00050.0 00050.0 00050.0 00050.0 55500.0 50050.0 000 05000.0 P 05000.0 05050.0 00000.0 50000.0 50050.0 00050.0 00000.0 $000.0 00000.0 50000.0 00050.0 52 00000.0 05000.0 5 50000.0 00500.0 00000.0 05000.0 00500.0 05000.0 0000.0 05050.0 00050.0 00000.0 80 00000.0 05050.0 50000.0 _. 50000.0 50000.0 0000.0 00005.0 00000.0 00500.0 50000.0 00050.0 00000.0 .000 05000.0 00000.0 00.00 50000.0 5 00050.0 00000.0 00550.0 55000.0 00500.0 50500.0 0005.0 0000.0 0:< 50500.0 50000.0 00000.0 50000.0 00050.0 5 0000.0 5055.0 05005.0 00005.0 000500 0.0050 50000.0 5.0.. 00050.0 50050.0 050000 00000 00000.0 0000.0 5 00050.0 00500.0 00000.0 0000.0 50000.0 0000.0 00.... 00050.0 00050.0 00500.0 0005.0 0030.0 5055.0 00050.0 5 05000.0 00000.0 00000.0 00050.0 00050.0 50.2 00050.0 00000.0 05000.0 00000.0 55000.0 05005.0 00500.0 05000.0 .. 50500.0 00000.0 0050.0 50000.0 5< 00050.0 3000.0 0000.0 00.000 00500.0 00005.0 00000.0 00000.0 50500.0 5 05500.0 0000.0 0000.0 .05. 00050.0 00000.0 0050.0 50000.0 50500.0 0000.0 0000.0 00000.0 00000.0 05500.0 5 50000.0 50050.0 00“. 55500.0 50000.0 00050.0 00050.0 0.0050 0.0050 50000.0 0000.0 0050.0 50000.0 50000.0 .. 00000.0 00... 50050.0 00050.0 00000.0 000000 0000.0 5000.0 0000.0 00050.0 50000.0 0000.0 50050.0 00000.0 5 .0005. 000 52 60 :00 W00. 5...... 000.. 50.2 .9. 00.2 00“. 00.. .00004. 0.0001 00.000 5030 0.0.0022 05.000 0.0000 05 .0. .003 0x00 .0E0...x0.2. 05.000900. E:E_x0E 000 .. Eouoc .0E0hc..20 05.000900. 05.0.0.0. .. no. .9005. 020000.00... 5.50oE 000 .00000 .0000 5.0.0.0 0000-550: 000E ..o. x500. 02.0.0000 .0-< 0.00... 241 5 50000.0 00000.0 05000.0 00000.0 00000.0 50000.0 500.0 0500.0 00000.0 00000.0 55500.0 00500.0 000 50000.0 5 00000.0 00000.0 50000.0 00500.0 00500.0 50000.0 00000.0 00500.0 50000.0 00500.0 50500.0 32 00000.0 00000.0 5 00000.0 00000.0 50500.0 05500.0 00000.0 00000.0 00000.0 00000.0 00500.0 00000.0 .00 05000.0 00000.0 00000.0 5 0000.0 50500.0 0000.0 00000.0 50000.0 05500.0 50000.0 00500.0 00000.0 .000 00000.0 50000.0 00000.0 0000.0 5 50000.0 00000.0 55000.0 50500.0 00000.0 00000.0 50000.0 00000.0 000. 00000.0 00500.0 50500.0 50500.0 50000.0 5 00000.0 00500.0 05000.0 00000.0 50000.0 05000.0 00000.0 5.0.. 50000.0 00500.0 05500.0 0000.0 00000.0 00000.0 5 50000.0 00000.0 05000.0 05000.0 00500.0 50500.0 000.. 500.0 50000.0 00000.0 00000.0 55000.0 00500.0 50000.0 5 050000 500.0 00000.0 00000.0 00000.0 50.2 0500.0 00000.0 00000.0 50000.0 50500.0 05000.0 00000.0 05000.0 5 50000.0 00500.0 5000.0 00000.0 .a< 00000.0 00500.0 00000.0 05500.0 00000.0 00000.0 05000.0 500.0 50000.0 5 05000.0 05000.0 500000 .05. 00000.0 50000.0 00000.0 50000.0 00000.0 50000.0 05000.0 00000.0 00500.0 05000.0 5 50000.0 00000.0 00“. 55500.0 00500.0 00500.0 00500.0 50000.0 05000.0 00500.0 00000.0 5000.0 05000.0 50000.0 5 50000.0 00.. 00500.0 50500.0 00000.0 00000.0 00000.0 00000.0 50500.0 00000.0 00000.0 50000.0 00000.0 50000.0 5 .0005. 000 52 .00 .000 004. 5.0.. 000.. 50.2 .04. .05. 00“. 00.. .0000< 0500.50.03. .b0::...=00 .N.< 0.00... 242 REFERENCES 243 REFERENCES Adams, H.D., and TE. Kolb. 2004. Drought responses of conifers in ecotone forests of northern Arizona: tree ring growth and leaf 613C. Oecologia 140:217—225. Adams, H.D., and TE. Kolb. 2005. Tree growth response to drought and temperature in a mountain landscape in northern Arizona, USA. Journal of Biogeography 32:1629-1640. Adams, J. 2007. Vegetation-Climate Interaction: How Vegetation Makes the Global Environment. Springer-Verlag, Berlin, Germany. 300p. Adams, W.W. lII, C.R. Zarter, V. Ebbert, and B. Demmig-Adams. 2004. Photoprotective strategies of overwintering evergreens. BioScience 54(1):41-49. Alder N.N., W.T. Pockman, J.S. Sperry, and S. Nuismer. 1997. Use of centrifugal force in the study of xylem cavitation. Journal of Experimental Botany 48: 665-674. Alexander, K. 1991. Paradise Found: The Settlement of the Santa Catalina Mountains. Skunkwork Productions, Mount Lemmon, Arizona, USA. 115p. Anderson, E., and G.L. Stebbins, Jr. 1954. Hybridization as an evolutionary stimulus. Evolution 81378-388. Anderson, R.P., A.T. Peterson, and M. Gemez-Laverde. 2002. Using niche- based GIS modeling to test geographic predictions of competitive exclusion and competitive release in South American pocket mice. Oikos 98:3-16. Anderson, R.P., D. Lew, and AT. Peterson. 2003. Evaluating predictive models of species’ distributions: criteria for selecting optimal models. Ecological Modelling 162:211-232. Andersson, L., and P. Milberg. 1998. Variation in seed dormancy among mother plants, populations and years of seed collection. Seed Science Research 8:29-38. Araujo, MB, and RC. Pearson. 2005. Equilibrium of species' distributions with climate. Ecography 28(5):693-695. 244 Aronsson, A., and L. Eliasson. 1970. Frost hardiness in Scots pine (Pinus sylvestris L.). l. Conditions for test on hardy plant tissues and for evaluation of injuries by conductivity measurements. Studia Forestalia Suecica Nr77. 30p. Austin, MP. 2002. Spatial prediction of species distribution: an interface between ecological theory and statistical modelling. Ecological Modelling 157:101- 1 18. Axelrod, DJ. 1986. Cenozoic history of some western American pines. Annals of the Missouri Botanical Garden 73:565-641. Axelrod, DI, and PH. Raven. 1985. Origins of the Cordilleran flora. Journal of Biogeography 12:21-47. Bahn, V., and B.J. McGill. 2007. Can niche-based distribution models outperform spatial interpolation? Global Ecology and Biogeography, OnLine Early Articles, 15 May 2007:1-10. Baldridge, W.S. 2004. Geology of the American Southwest. Cambridge University Press, Cambridge, UK. 280p. Barton, AM. 1993. Factors controlling plant distributions: drought, competition, and fire in montane pines in Arizona. Ecological Monographs 63(4):367- 397. Barton, A.M., and J.A. Teeri. 1993. The ecology of elevational positions in plants: drought resistance in five montane pine species in southeastern Arizona. American Journal of Botany 80(1):15-25. Bassman, J.H. 1987. Photosynthesis and water relations of ponderosa pine. Pages 45-58 in D.M. Baumgarnter and J.E. Lotan, editors. Ponderosa pine: the species and its management. Washington State University, Pullman, Washington, USA. 281p. Bazazz, FA. 1979. The physiological ecology of plant succession. Annual Review of Ecology and Systematics 10:351-371. Betancourt, J.L., T.R. van Devender, and PS. Martin, editors. 1990. Packrat Middens: The Last 40,000 Years of Biotic Change. The University of Arizona Press, Tucson, Arizona, USA. 467p. Bewley, JD, and M. Black. 1985. Seeds: Physiology of Development and Germination. Plenum Press, New York, New York, USA. 367p. 245 Beyer, H. L. 2006. Hawth's Analysis Tools for ArcGlS, Version 3.26. URL: http:llwww.spatialecology.com/htools (accessed 18 May 2007). Bjorkman, O., and B. Demmig. 1987. Photon yield of OZ-evolution and chlorophyll fluorescence characteristics at 77K among vascular plants of diverse origins. Planta 170:489-504. Bjorkman, O., and B. Demmig-Adams. 1994. Regulation of photosynthetic light energy capture, conversion, and dissipation in leaves. Pages 17-48 in E.- D. Schulze and MM. Caldwell, Editors. Ecophysiology of Photosynthesis. Springer-Vedag, Berlin, Germany. 576p. Boorse, G.C., T.L. Bosma, A.-C. Meyer, F .W. Ewers, and SD. Davis. 1998. Comparative methods of estimating freezing temperatures and freezing injury in leaves of Chaparral shrubs. International Journal of Plant Science 159(3):513-521. Brown, G.N., T.M. Hinckley, and J.A. Bixby. 1972. Xylem sap pressure relationship to seedling survival after freezing. Cryobiology 9:314. Brown, J.H., and M.V. Lomolino. 1998. Biogeography, Second Edition. Sinauer Associates, Inc., Sunderland, Massachusetts, USA. 691p. Burr, K.E., R.W. Tinus, S.J. Wallner, and RM. King. 1989. Relationships among cold hardiness, root growth potential and bud dormancy in three conifers. Tree Physiology 5:291-306. Burr, K.E., R.W. Tinus, S.J. Wallner, and RM. King. 1990. Comparison of three cold hardiness tests for conifer seedlings. Tree Physiology 6:351-369. Burr, K.E., S.J. Wallner, and R.W. Tinus. 1991. Ethylene and ethane evolution during cold acclimation and deacclimation of ponderosa pine. Canadian Journal of Forest Research 21:601-605. Burr, K.E., C.D.B. Hawkins, S.J. L'Hirondelle, W.D. Binder, M.F. George, and T. Repo. 2001. Methods for measuring cold hardiness of conifers. Pages 369-401 in F.J. Bigras and S.J. Colombo, Editors. Conifer Cold Hardiness. Kluwer Academic Publishers, Boston, Massachusetts, USA. 596p. Barton, L.V. 1930. Hastening the germination of some coniferous seeds. American Journal of Botany 17:88-115. Cannell, M.G.R., and L.J. Sheppard. 1982. Seasonal changes in the frost hardiness of provenances of Picea sitchensis in Scotland. Forestry 55:137-153. 246 Clausen, J., D.D. Keck, and WM. Hiesey. 1940. Experimental studies on the nature of species. I. Effect of varied environments on western North American plants. Carnegie Institution of Washington, Publication 520. Washington, DC, USA. 452p. Colorado Plateau Stable Isotope Laboratory (CPSIL). 2006. Sample submission — organics - 0‘30 and 0‘5N via EA-CFIRMS. URL: http:llwww4.nau.edu/cpsil/13c_15n_org.htm (accessed 19 May 2006). Conkle, MT, and WE. Critchfield. 1988. Genetic variation and hybridization of ponderosa pine. Pages 27-43 in D.M. Baumgarnter and J.E. Lotan, editors. Ponderosa pine: the species and its management. Washington State University, Pullman, Washington, USA. 281p. Coronado National Forest (CNF). 2002. Bullock Fire (Archival Site). URL: www.fs.fed.us/r3/coronado/bullock/ (accessed 14 May 2007). Coronado National Forest (CNF). 2003. Aspen Fire (Archival Site). URL: www.fs.fed.us/r3/coronado/aspen/ (accessed 14 May 2007). Cottam, G., and J.T. Curtis. 1956. The use of distance measures in phytosociological sampling. Ecology 37:451-460. Covington, W.W., R.L. Everett, R.W. Steele, L.|. II'WII‘I, T.A. Deer, and A.N.D. Auclair. 1994. Historical and anticipated changes in forest ecosystems of the Inland West of the United States. Journal of Sustainable Forestry 2:13- 63. Costa, J., A.T. Peterson, and CB. Beard. 2002. Ecologic niche modeling and differentiation of populations of Tn'atoma brasiliensis Neiva, 1911, the most important chagas’ disease vector in northeastern Brazil (Hemiptera, Reduviidae, Triatominae). American Journal of Tropical Medicine and Hygiene 67(5):516-520. Crawley, M.J. 2005. Statistics: An Introduction using R. John Wiley & Sons, Ltd., West Sussex, England. 327p. Cregg. BM. 1994. Carbon allocation, gas exchange, and needle morphology of Pinus ponderosa genotypes known to differ in growth and survival under imposed drought. Tree Physiology 14:883-898. Critchfield, W.B., and EL. Little, Jr. 1966. Geographic distribution of the pines of the world. USDA Misc. Publ. 991. 247 Cui, M., and WK. Smith. 1991. Photosynthesis, water relations and mortality in Abies Iasiocarpa seedlings during natural establishment. Tree Physiology 8:37-46. Daubenmire, RF. 1943. Soil temperature versus drought as a factor determining lower altitudinal limits of trees in the Rocky Mountains. Botanical Gazette 105(1):1-13. Davis, A.J., J.H. Lawton, B. Shorrocks, and LS. Jenkinson. 1998. lndividualistic species responses invalidate simple physiological models of community dynamics under global environmental change. Journal of Animal Ecology 67:600-612. Dawson, T.E., S. Mambelli, A.H. Plamboeck, P.H. Templer, and KP. Tu. 2002. Stable isotopes in plant ecology. Annual Review of Ecology and Systematics 33:507-559. DeHayes, DH. 1977. Genetic variation in cold hardiness and its effects on the performance of ponderosa pine (Pinus ponderosa) in Michigan. Unpublished dissertation, Michigan State University, East Lansing, Michigan, USA. 141p. Dodge, RA. 1963. Investigations into the Ecological Relationships of Ponderosa Pine in Southeast Arizona. Unpublished dissertation, The University of Arizona, Tucson, Arizona, USA. 118p. Dreesen, D. 2003. Propagation protocol for production of container Pinus ponderosa plants (One-Gallon Tree Pot, 4”x4”x14”); Los Luna Plant Materials Center, Los Lunas, New Mexico. In Native Plant Network. URL: www.nativeplantnetwork.org (accessed 14 October 2005). Elton, C. 1927. Animal Ecology. Macmillan Company, New York, New York, USA. 207p. Elliott, K.J., and AS. White. 1987. Competitive effects of various grasses and forbs on ponderosa pine seedlings. Forest Science 33(2):356-366. Engler, R., A. Guisan, and L. Rechsteiner. 2004. An improved approach for predicting the distribution of rare and endangered species from occurrence and pseudo-absence data. Journal of Applied Ecology 41:263- 274. Epperson, B.K. F.W. Telewski, A.E. PIovanich-Jones, and J.E. Grimes. 2001. Clinal differentiation and putative hybridization in a contact zone of Pinus ponderosa and P. arizonica (Pinaceae). American Journal of Botany 88(6):1052-1057. 248 Epperson, B.K., M.G. Chung, and F.W. Telewski. 2003. Spatial pattern of allozyme variation in a contact zone of Pinus pondemsa and P. arizonica (Pinaceae). American Journal of Botany 90(1):25-31. Environmental Systems Research Institute (ESRI), Inc. 2006. ArcGlS version 9.2. Redlands, California, USA. Evans, J., T. Luna, and D. Wick. 2001. Propagation protocol for production of container Pinus ponderosa Dougl. Plants (172 ml conetainers); Glacier National Park, West Glacier, Montana. ln Native Plant Network. URL: http:l/www.nativeplantnetwork.org (accessed 14 October 2005). Faraway, J.J. 2002. Practical regression and Anova using R. URL: http:llcran.r- project.org/doc/contrib/Faraway-PRA.pdf (accessed 15 June 2007). 213p. Farjon, A., and ET. Styles. 1997. Pinus (Pinaceae), Flora Neotropica, Monograph 75. The New York Botanical Garden, Bronx, New York, USA. 293p. Farquhar, G.D., J.R. Ehleringer, and KT. Hubick. 1989. Carbon isotope discrimination and photosynthesis. Annual Review of Plant Physiology 40:503-537. Farris, MA, and J.B. Mitton. 1984. Population density, outcrossing rate, and heterozygote superiority in ponderosa pine. Evolution 38(5):1151-1154. Force, ER. 1997. Geology and Mineral Resources of the Santa Catalina Mountains, Southeastern Arizona: A Cross-Sectional Approach. Monographs in Mineral Resource Science No. 1. The University of Arizona, Tucson, Arizona, USA. 135p. Fritts, H.C. 1963. Statistical evaluation of tree-ring series. Seventh Annual Meeting, Arizona Academy of Science, Biology Section, Tucson, Arizona, USA. Fritts, H.C. 1974. Relationships of ring widths in arid-site conifers to variations in monthly temperature and precipitation. Ecological Monographs 44:411- 440. Fry, B. 2006. Stable Isotope Ecology. Springer Science+Business Media, LLC, New York, New York, USA. 308p. Fuller, GD. 1916. Ecology of the Santa Catalina Mountains. Botanical Gazette 62(1):75-78. 249 Gemiino, M.J., W.K. Smith, and AC. Resor. 2002. Conifer seedling distribution and survival in an alpine-treeline ecotone. Plant Ecology 162:157-168. Gernandt, D.S., G.G. prez, S.O. Garcia, and A. Liston. 2005. Phylogeny and classification of Pinus. Taxon 54(1):29-42. Graybill, DA. 1986. Green Mountain, Arizona, ponderosa pine. International Tree-Ring Data Bank. URL: ftp:l/ftp.ncdc.noaa.gov/pub/data/paleo/treering/measurements/northameric a/usa/a2520.rwl (accessed 11 June 2007). Guisan, A., and W. Thuiller. 2005. Predicting species distribution: offering more than simple habitat models. Ecology Letters 8:993-1009. Guisan, A., C.H. Graham, J. Elith, F. Huettmann, and the NCEAS Species .Distribution Modelling Group. 2007. Sensitivity of predictive species distribution models to change in grain size. Diversity and Distributions 13:332-340. Haase, E.F. 1970. Environmental fluctuations on south-facing slopes in the Santa Catalina Mountains of Arizona. Ecology 51(6):959-974. Haller, JR. 1959. Factors affecting the distribution of ponderosa and Jeffrey pines in California. Madroflo 15(3):65-96. Haller, JR. 1965. The role of 2—needle fascicles in the adaptation and evolution of ponderosa pine. Brittonia 17:354-382. Hamrick, J.L., H.M. Blanton, and K.J. Hamrick. 1989. Genetic structure of geographically marginal populations of ponderosa pine. American Journal of Botany 76(11):1559-1568. Hanley, J.A., and B.J. McNeil. 1982. The meaning and use of the area under a receiver operating characteristic (ROC) curve. Radiology 143(1):29-36. Havranek, W.M., and W. Tranquillini. 1995. Physiological processes during winter dormancy and their ecological significance. Pages 95-124 in W.K. Smith and TM. Hinckley, Editors. Ecophysiology of Coniferous Forests. Academic Press, Inc., San Diego, California, USA. 338p. Hawkins, B.J., H.J. Guest, and K. Kolotelo. 2003. Freezing tolerance of conifer seeds and germinants. Tree Physiology 23:1237-1246. Heidmann, L.J. 1987. Regeneration strategies for ponderosa pine. Pages 227- 233 in D.M. Baumgarnter and J.E. Lotan, editors. Ponderosa pine: the 250 species and its management. Washington State University, Pullman, Washington, USA. Helms, J.A. 1972. Environmental control of net nhotosynthesis in naturally growing Pinus ponderosa Laws. Ecology 53(1): 92-101. Hernandez, P.A., C.H. Graham, L.L. Master, and BL Albert. 2006. The effect of sample size and species characteristics on performance of different species distribution modeling methods. Ecography 29:773-785. Hijmans, R.J., and CH. Graham. 2006. The ability of climate envelope models to predict the effect of climate change on species distributions. Global Change Biology 12:2272-2281. Hirzel, A., and A. Guisan. 2002. Which is the optimal sampling strategy for habitat suitability modelling? Ecological Modelling 157:331-341. Hobbie, E.A., S.A. Macko, and M. Williams. 2000. Correlations between foliar 615N and nitrogen concentrations may indicate plant-mycorrhizal interactions. Oecologia 122:273-283. Hoch, G., and C. Korner. 2003. The carbon charging of pines at the climatic treeline: a global comparison. Oecologia 135210-21. Howard, Janet L. 2003a. Pinus ponderosa var. arizonica. In Fire Effects Information System. US. Department of Agriculture, Forest Service, Rocky Mountain Research Station, Fire Sciences Laboratory. URL: http:llwww.fs.fed.us/database/feis (accessed 21 March 2005). Howard, Janet L. 2003b. Pinus ponderosa var. scopulorum. In Fire Effects Information System. US. Department of Agriculture, Forest Service, Rocky Mountain Research Station, Fire Sciences Laboratory. URL: http:llwww.fs.fed.us/database/feis (accessed 16 June 2007). Hubbard, R.M., M.G. Ryan, V. Stiller, and J.S. Sperry. 2001. Stomatal conductance and photosynthesis vary linearly with plant hydraulic conductance in ponderosa pine. Plant, Cell and Environment 24:113-121. Humphries, H.C., and PS. Bourgeron. 2003. Environmental responses of Pinus ponderosa and associated species in the south-western USA. Journal of Biogeography 30:257-276. Hutchinson, GE. 1957. Concluding remarks. Cold Spring Harbor Symposium on Quantitative Biology 22:415-427. 251 Jackson, G.E., I.J. Irvine, J. Grace, and A.A.M. Khalil. 1995. Abscisic acid concentrations and fluxes in droughted conifer saplings. Plant, Cell and Environment 18:13-22. Jacobsen A.L., R.B. Pratt, F.W. Ewers, and SD. Davis. 2007. Cavitation resistance among twenty-six Chaparral species of southern California. Ecological Monographs 77:99-115. Jarvis, PG. 1976. The interpretation of the variations in leaf water potential and stomatal conductance found in canopies in the field. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences 273(927):593-610. Johnson, D.M., M.J. Gerrnino, and WK. Smith. 2004. Abiotic factors limiting photosynthesis in Abies Iasiocarpa and Pica engelmannii seedlings below and above the alpine timberline. Tree Physiology 24:377-386. Jones, JR. 1967. Regeneration of mixed conifer clearcuttings on the Apache National Forest, Arizona. Res. Note RM-79. Fort Collins, CO: US. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station. 8 p. Kaufmann, MR. 1968. Evaluation of the pressure chamber technique for estimating plant water potential of forest tree species. Forest Science 14(4):369-374. Kipfer, J.A., J. Balmat, and J.L. Smith. 2005. Shifts in the potential distribution of sky island plant communities in response to climate change. Pages 485- 490 in G.J. Gottfried, B.S. Gebow, LG. Eskew, and CB. Edminster, Compilers. Connecting Mountain Islands and Desert Seas: Biodiversity and Management of the Madrean Archipelago lI. Proceedings RMRS-P- 36. US. Department of Agriculture, Forest Service, Rocky Mountain Research Station, Fort Collins, Colorado, USA. 631 p. Kolb, RF, and R. Robberecht. Pinus ponderosa seedling establishment and the influence of competition with the bunchgrass Agropyron spicatum. International Journal of Plant Science 157(4):509-515. Kolb, T.E., K.C. Steiner, and HF Barbour. 1985. Seasonal and genetic variations in loblolly pine cold tolerance. Forest Science 31(4):926-932. Kozlowski, T.T., P.J. Kramer, and 8.6. Pallardy. 1991. The Physiological Ecology of Plants. Academic Press, San Diego, California, USA. 657p. Kral, R. 1993. Pinus. Pages 373-398 in Flora of North America Editorial Committee, editors. Flora of North America North of Mexico. Volume 2. 252 Pteridophytes and Gymnosperrns. Oxford University Press, New York, New York, USA. 475pp. Krugman, S.L., and J.L. Jenkinson. 1974. Pinus L. Pine. Pages 598-638 in OS. Schopmeyer, Technical Coordinator. Seeds of Woody Plants in the United States. US. Department of Agriculture, Forest Service, Agriculture Handbook No. 450, Washington, DC, USA. Langsrud, fl. 2003. ANOVA for unbalanced data: Use Type II instead of Type III sums of squares. Statistics and Computing 13:163-167. Larcher, W. 2005. Physiological Plant Ecology: Ecophysiology and Stress Physiology of Functional Groups, 4th Edition. Springer-Verlag, Berlin, Germany. 513p. Larson, MM. 1961. Seed size, germination dates, and survival relationships of ponderosa pine in the Southwest. Research Note No. 66. Fort Collins, CO: US. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station. 4 p. Larson, MM, and G.H. Schubert. 1970. Cone crops of ponderosa pine in central Arizona, including the influence of Abert squirrels. Res. Pap. RM-58. Fort Collins, CO: US. Department of Agriculture, Forest Service, Rocky Mountain Forest and Range Experiment Station. 15 p. Latta, R.G., Y.B. Linhart, and J.B. Mitton. 2001. Cytonuclear disequilibrium and genetic drift in a natural population of ponderosa pine. Genetics 158:843- 850. Leathwick, JR. 1998. Are New Zealand’s Nothofagus species in equilibrium with their environment? Journal of Vegetation Science 92719-732. Lehdig, F.T. 1998. Genetic variation in pines. Pages 251-280 in D.M. Richardson, Editor. Ecology and Biogeography of Pinus. Cambridge University Press, Cambridge, UK. 527p. Lemmon, PE. 1957. A new instrument for measuring forest overstory density. Journal of Forestry 55(9):667-668. Li, X.J., P.J. Burton, and CL. Leadem. 1994. Interactive effects of light and stratification on the germination of some British Columbia conifers. Canadian Journal of Botany 72:1635-1646. Ll-COR. 2005a. Using the Ll-6400 Portable Photosynthesis System, Version 5: Book 5. Leaf Chamber Fluorometer. LI-COR Biosciences, Inc., Lincoln, Nebraska, USA. 100p. 253 Linhart, Y8, and J.B. Mitton. 1985. Relationships among reproduction, growth rates, and protein heterozygosity in ponderosa pine. American Journal of Botany 72(2):181-184. Little, E.L., Jr., and WE. Critchfield. 1960. Subdivisions of the genus Pinus (pines). United States Department of Agriculture Miscellaneous Publication 1144. Liu, 0., PM. Berry, T.P. Dawson, and R.G. Pearson. 2005. Selecting thresholds of occurrence in the prediction of species distributions. Ecography 28:385- 393. Lu, P., S.J. Colombo, and R.W. Sinclair. 2007. Cold hardiness of interspeciflc hybrids between Pinus strobus and P. wallichiana measured by post- freezing needle electrolyte leakage. Tree Physiology 27:243-250. Maherali, H., and EH. DeLucia. 2000. Xylem conductivity and vulnerability of ponderosa pine growing in contrasting climates. Tree Physiology 20:859- 867. Maherali, H., B.L. Williams, K.N. Paige, and EH. DeLucia. 2002. Hydraulic differentiation of Ponderosa pine populations along a climate gradient is not associated with ecotypic divergence. Functional Ecology 16:510-521. Marshall, JD, and J. Zhang. 1994. Carbon isotope discrimination and water use efficiency of native plants of the north-central Rockies. Ecology 7511887- 1895. Martin, P.S., and P.J. Mehringer, Jr. 1965. Pleistocene pollen analysis and biogeography of the Southwest. Pages 433-451 in HE. Wright, Jr. and DC. Frey, editors. The Quaternary of the United States: A Review Volume for the Vll Congress of the International Association for Quaternary Research. Princeton University Press, Princeton, New Jersey, USA. 922p. Martinez-Meyer, E., and AT. Peterson. 2006. Conservatism of ecological niche characteristics in North American plant species over the Pleistocene-to- Recent transition. Journal of Biogeography 33: 1 779-1789. Maurer, B. 2002. Predicting distribution and abundance: thinking within and between scales. Pages 125-133 in J.M. Scott, P.J. Heglund, M.L. Morrison, J.B. Haufler, M.G. Raphael, WA. Wall, and FE. Samson, Editors. Predicting Species Occurrences: Issues of Accuracy and Scale. Island Press, Washington, D.C., USA. 868p. 254 McPherson, J.M., W. Jetz, and DJ. Rogers. 2004. The effects of species’ range sizes on the accuracy of distribution models: ecological phenomenon or statistical artefact? Journal of Applied Ecology 41 :81 1-823. Melander, L., and W.H. Saunders. 1979. Reaction rates of isotopic molecules. Wiley, New York, New York, USA. Millar, CL 1993. Impact of the Eocene on the evolution of Pinus L. Annals of the Missouri Botanical Garden 80:471-498. Mitton, J.B., Y.B. Linhart, J.L. Hamrick, and J.S. Beckman. 1977. Observations on the genetic structure and mating system of ponderosa pine in the Colorado Front Range. Theoretical and Applied Genetics 51 :5-13. Mitton, J.B., K.B. Sturgeon, and ML. Davis. 1980. Genetic differentiation in ponderosa pine along a steep elevational gradient. Silvae Genetica 29(3- 4):100-103. Moore, M.M., W.W. Covington, and P2. Fulé. 1999. Reference conditions and ecological restoration: 3 Southwestern ponderosa pine perspective. Ecological Applications 9(4):1266-1277. Muir, J. 1911. My First Summer in the Sierra. Houghton Mifflin Company, New York, New York, USA. 353p. National Climatic Data Center (NCDC). 2007. NNDC Climate Data Online. Station: East LANSING 4 S, Ml, COOP-WBAN ID 202395-99999. National Climatic Data Center, Asheville, North Carolina, USA. URL: http:llwww.ncdc.noaa.gov/oa/ncdc.html (accessed 26 June 2007). Neilson, RP, and L.H. Wullstein. 1983. Biogeography of two southwest American oaks in relation to atmospheric dynamics. Journal of Biogeography 10:275-297. Niebling, CR, and MT. Conkle 1990. Diversity of Washoe pine and comparisons with allozymes of ponderosa pine races. Canadian Journal of Forest Research 20:298-308. Nobel, PS. 2005. Physicochemical and Environmental Plant Physiology, Third Edition. Elsevier Academic Press, Burlington, Massachusetts, USA. 567p. Norris, J.R., S.T. Jackson, and J.L. Betancourt. 2006. Classification tree and minimum-volume ellipsoid analyses of the distribution of ponderosa pine in the western USA. Journal of Biogeography 33:342-360. 255 Oliver, W.W., and RA. Ryker. 1990. Ponderosa Pine. In R.M. Burns and EH. Honkala, technical coordinators. Silvics of North America: Volume 1. Conifers. Agriculture Handbook 654. US. Department of Agriculture, Forest Service, Washington, DC. URL: www.na.fs.fed.us/spfo/pubs/silvics_manual (accessed 29 May 2007). Oquist, G., and N.P.A. Huner. 2003. Photosynthesis of ovenlvintering evergreen plants. Annual Review of Plant Biology 54:329-355. Park, R., and S. Epstein. 1961. Metabolic fractionation of C13 and C12 in plants. Plant Physiology 36: 1 33-1 38. Parker, J. 1955. Annual trends in cold hardiness of ponderosa pine and grand fir. Ecology 36(3):377-380. Pearson, R.G., C.J. Raxworthy, M. Nakamura, and AT. Peterson. 2007. Predicting species distributions from small numbers of occurrence records: a test case using cryptic geckos in Madagascar. Journal of Biogeography 342102-117. Peet, R.K. 1984. Forests of the Rocky Mountains. Pages 63-101 in MG. Barbour and W.D. Billings, editors. North American Terrestrial Vegetation. Cambridge University Press, Cambridge, UK. 434p. Peloquin, R.L. 1971. Variation and hybridization patterns in Pinus ponderosa and Pinus engelmannii. Ph.D. dissertation, University of California, Santa Barbara, California, USA. 196p. Peloquin, R.L. 1984. The identification of three-species hybrids in the ponderosa pine complex. Southwestern Naturalist 29(1):115-122. Perry, D.A., H. Margolis, C. Choquette, R. Molina, and J.M. Trappe. 1989. Ectomycorrhizal mediation of competition between coniferous tree species. New Phytologist 112:501-511. Perry, J.P., Jr. 1991. The Pines of Mexico and Central America. Timber Press, Inc., Portland, Oregon, USA. 231p. Peterson, A.T., J. Soberbn, and V. Sénchez-Cordero. 1999. Conservation of ecological niches in evolutionary time. Science 285:1265-1267. Peterson, A.T., D.R.B. Stockwell, and D.A.Kluza. 2002. Distributional prediction based on ecological niche modeling of primary occurrence data. Pages 617-623 in J.M. Scott, P.J. Heglund, M.L. Morrison, J.B. Haufler, M.G. Raphael, WA. Wall, and F .B. Samson, Editors. Predicting Species 256 Occurrences: Issues of Accuracy and Scale. Island Press, Washington, D.C., USA. 868p. Phillips, S.J., M. Dudik, and RE. Schapire. 2004. A maximum entropy approach to species distribution modeling. Pages 655-662 in Proceedings of the 21"t International Conference on Machine Learning. ACM Press, New York, New York, USA. Phillips, S.J., R.P. Anderson, and RE. Schapire. 2006a. Maximum entropy modeling of species geographic distributions. Ecological Modelling 190:231-259. Phillips, S.J., M. Dudik, and RE. Schapire. 2006b. Maximum Entropy Species Distribution Modeling, Version 2.3. URL: www.cs.princeton.edu/~schapirelmaxentl (accessed 20 February 2007). Pima County Department of Transportation (PCDOT). 2007. Permanent Traffic Counter Stations for 2005-2006. URL: http:llwww.dot.pima.gov/trafeng/trafcnt/ (accessed 29 June 2007). Piflol, J., and A. Sale. 2000. Ecological implications of xylem cavitation for several Pinaceae in the Pacific Northern USA. Functional Ecology 14:538- 545. Pisek, A., and R. Schiessl. 1947. Die Temperaturbeeinflussbarkeit der Frostharte von Nadelholzem und Zwergstrauchern an der alpinen Waldgrenze. Berlin Naturwiss.-Med. Ver. Innsbruck 47:33-52. Potvin, C., and M.J. Lechowicz. 1990. The statistical analysis of ecophysiological response curves obtained from experiments involving repeated measures. Ecology 71 (4): 1 389-1400. Price, R.A., A. Liston, and SH. Strauss. 1998. Phylogeny and systematics of Pinus. Pages 49-68 in D.M. Richardson, editor. Ecology and Biogeography of Pinus. Cambridge University Press, Cambridge, UK. 527p. PRISM. 2006. 30-arcsec (800m) Normals (1971-2000). PRISM Group, Oregon State University, Corvallis, Oregon, USA. URL: www.prismclimate.org (accessed 13 October 2006). Pulliam, HR. 2000. On the relationship between niche and distribution. Ecology Letters 3:349-361. R Development Core Team. 2007. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, 257 Austria. Version 2.5.0 (2007-4-23). URL: www.R-project.org (accessed 21 May 2007). Radomes, Inc. 2005. Online Air Defense Radar Museum. URL: www.radomes.org/museum/ (accessed 8 June 2007). Read, DJ 1998. The mycorrhizal status of Pinus. Pages. 324-340 in D.M. Richardson, Editor. Ecology and Biogeography of Pinus. Cambridge University Press, Cambridge, United Kingdom. 527p. Rehfeldt, GE. 1986. Adaptive variation in Pinus ponderosa from lntermountain regions. I. Snake and Salmon River basins. Forest Science 32:79-92. Rehfeldt, GE. 1990. Genetic differentiation among populations of Pinus ponderosa from the upper Colorado River Basin. Botanical Gazette 151(1):125-137. Rehfeldt, GE. 1993. Genetic variation in the Ponderosae of the Southwest. American Journal of Botany 80(3):330-343. Rehfeldt, GE. 1999. Systematics and genetic structure of Ponderosae taxa (Pinaceae) inhabiting the mountain islands of the Southwest. American Journal of Botany 86(5):?41-752. Rehfeldt, G.E., B.C. Wilson, S.P. Wells, and RM. Jeffers. 1996. Phytogeographic, taxonomic, and genetic implications of phenotypic variation in the Ponderosae of the Southwest. Southwestern Naturalist 41 :409-418. Reich, P.B., M.B. Walters, B.K. Kloeppel, and 0.8. Ellsworth. 1995. Different photosynthesis-nitrogen relations in deciduous hardwood and evergreen coniferous tree species. Oecologia 104:24-30. Riegel, G.M., R.F. Miller, and WC. Krueger. 1992. Competition for resources between understory vegetation and overstory Pinus ponderosa in northeastern Oregon. Ecological Applications 2(1):71-85. Ritchie, GA. 1975. The pressure chamber as an instrument for ecological research. Advances in Ecological Research 92165-254. Ritchie, GA. 1991. Measuring cold hardiness. Pages 557-582 in J.M. Lassoie and TM. Hinckley, Editors. Techniques and Approaches in Forest Tree Ecophysiology. CRC Press, Inc., Boca Raton, Louisiana, USA. 599p. 258 Rundel, P.W., and B.J. Yoder. 1998. Ecophysiology of Pinus. Pages 296-323 in D.M. Richardson, Editor. Ecology and Biogeography of Pinus. Cambridge University Press, Cambridge, United Kingdom. 527p. Schaberg, P.G., R.C. Wilkinson, J.B. Shane, J.R. Donnelly, and PF. Call. 1995. Winter photosynthesis of red spruce from three Vermont seed sources. Tree Physiology 15:345-350. Schmid, J.M., S.A. Mata, and JC. Mitchell. 1986. Number and condition of seeds in ponderosa pine cones in central Arizona. The Great Basin Naturalist. 46: 449-451. Schreiber, U., W. Bilger, and C. Neubauer. 1994. Chlorophyll fluorescence as a nonintrusive indicator for rapid assessment of in vivo photosynthesis. Pages 49-70 in ED. Schulze and MM. Caldwell, Editors. Ecophysiology of Photosynthesis. Springer-Verlag, Berlin, Germany. 576p. Schubert, G.H. 1974. Silviculture of southwestern ponderosa pine: the status of our knowledge. US. Department of Agriculture, Forest Service, Research Paper RM-123. Rocky Mountain Forest and Range Experiment Station, For Collins, Colorado, USA. 71p. Schlitz, W., and G. Rave. 1999. The effect of cold stratification and light on the seed germination of temperate sedges (Cerex) from various habitats and implications for regenerative strategies. Plant Ecology 144:215-230. Segurado, P., and MB. Aral’ljo. 2004. An evaluation of methods for modelling species distributions. Journal of Biogeography 31 :1555-1568. Sheppard, P.R., A.C. Comrie, G.D. Packin, K. Angersbach, and MK. Hughes. 2002. The climate of the US Southwest. Climate Research 21:219-238. Shreve, F. 1915. The Vegetation of a Desert Mountain Range as Conditioned by Climatic Factors. Carnegie Institution of Washington, Washington, D.C., USA. 112p. Shreve, F. 1922. Conditions indirectly affecting vertical distribution on desert mountains. Ecology 3(4):269-274. Shreve, F. 1924. Soil temperature as influenced by altitude and slope exposure. Ecology 5(2):128-136. Smith, RH. 1977. Monoterpenes of ponderosa pine xylem resin in western United States. USDA Forest Service, Technical Bulletin 1532. 48p. 259 Smith, TM. 1986. Stratification reduced germination of ponderosa pine seed collected in New Mexico and southern Colorado. Pages 17-18 in T.D. Landis. Proceedings: Combined Western Forest Nursery Council and Interrnountain Nursery Association Meeting. August 12-15, 1986. USDA Forest Service General Technical Report RM-137. Soberbn, J., and A.T. Peterson. 2005. Interpretation of models of fundamental ecological niches and species' distributional areas. Biodiversity Informatics 2:1-10. Sperry J.S., J.R. Donnelly, and MT Tyree MT. 1988. A method for measuring hydraulic conductivity and embolism in xylem. Plant, Cell and Environment 11:35-40. Stefferud, A., ed. 1961. The Yearbook of Agriculture 1961. US. Government Printing Office, Washington, DC. 591p. Stockwell, D.R.B., and A.T. Peterson. 2002. Effects of sample size on accuracy of species distribution models. Ecological Modelling 148:1-13. Stout, UL, and A. Sala. 2003. Xylem vulnerability to cavitation in Pseudotsuga menziesii and Pinus ponderosa from contrasting habitats. Tree Physiology 23:43-50. Sutinen, M.-L., J.P. Palta, and PB. Reich. 1992. Seasonal differences in freezing stress resistance of needles of Pinus nigra and Pinus resinosa: evaluation of the electrolyte leakage method. Tree Physiology 11:241-254. Swetnam, T.W., and OH. Baisan. 2003. Tree-ring reconstructions of fire and climate history in the Sierra Nevada and southwestern United States. Pages 154-191 in T.T. Veblen, W.L. Baker, G. Montenegro, and T.W. Swetnam, editors. Fire and Climate Change in Temperate Ecosystems of the Western Americas. Ecological Studies 160. Springer-Verlag, New York, New York, USA. 444p. Tanaka, Y. 1984. Assuring seed quality for seedling production: cone collection and seed processing, testing, storage, and stratification. Pages 27-39 in M.L. Duryea and TD. Landis, editors. Forest Nursery Manual: Production of Bareroot Seedlings. Martinus Nijhoff/Dr W. Junk Publishers. The Hague/Boston/Lancaster, for Forest Research Laboratory, Oregon State University, Corvallis, Oregon, USA. 386p. Theophrastus. 1916. Enquiry Into Plants, Volume I, Books 1-5, with an English Translation by Sir Arthur Hort. Harvard University Press, Cambridge, Massachusetts, USA. 504p. 260 Thompson, K., T. Keller-Wolf, and J. Franklin. 2002. A comparison of fine- and coarse-resolution environmental variables toward predicting vegetation distribution in the Mojave Desert. Pages 133-139 in J.M. Scott, P.J. Heglund, M.L. Morrison, J.B. Haufler, M.G. Raphael, WA. Wall, and PB. Samson, Editors. Predicting Species Occurrences: Issues of Accuracy and Scale. Island Press, Washington, D.C., USA. 868p. Thompson, R.S., K.H. Anderson, and P.J. Bartlein. 1999. Atlas of relations between climatic parameters and distributions of important trees and shrubs in North America. US. Geological Survey Professional Paper 1650 A83, Denver, Colorado, USA. URL: pubs.usgs.gov/pp/p1650-a/ (accessed 27 May 2007). Thompson, RS, and K.H. Anderson. 2000. Biomes of western North America at 18,000, 6000 and 0 1"C yr BP reconstructed from pollen and packrat midden data. Journal of Biogeography 27:555-584. Tomback, D.F., A.J. Anderies, K.S. Carsey, M.L. Powell, and S. Mellmann- Brown. 2001. delayed seed germination in whitebark pine and regeneration patterns following the Yellowstone fires. 82(9):2587-2600. Tranquillini, W. 1979. Physiological Ecology of the Alpine Timberline: Tree Existence at High Altitudes with Special Reference to the European Alps. Springer-Verlag, Berlin, Germany. 137p. Tsoar, A., O. Allouche, O. Steinitz, D. Rotem, and R. Kadmon. 2007. A comparative evaluation of presence-only methods for modeling species distribution. Diversity and Distributions, OnLine Early Articles, 27 March 2007, 1-9. Tyree, MT, and F.W. Ewers. 1991. The hydraulic architecture of trees and other woody plants. New Phytologist 119:345-360. US. Environmental Protection Agency (USEPA). 1998. Climate Change and Arizona. EPA 236-F-98-007c, Office of Policy, Washington, D.C., USA. 4p. US. Geological Survey (USGS). 2004. National Elevation Dataset (NED) 1/3 Arc Second. United States Geological Survey, Sioux Falls, SD. URL: http:l/seamless.usgs.gov/index.asp (accessed 18 May 2007). van Devender, TR. 1990. Late Quaternary vegetation and climate of the Sonoran Desert, United States and Mexico. Pages 134-163 in J.L. Betancourt, T.R. van Devender, and PS. Martin, editors. Packrat Middens: The Last 40,000 Years of Biotic Change. The University of Arizona Press, Tucson, Arizona, USA. 467p. 261 van Devender, TR, and W.G. Spaulding. 1979. Development of vegetation and climate in the southwestern United States. 204(4394):701-710. van Devender, T.R., J.L. Betancourt, and M. Wimberly. 1984. Biogeographic implications of a packrat midden sequence from the Sacramento Mountains, south-central New Mexico. Quaternary Research 22:344-360. Verhoeven, A.S., W.W. Adams Ill, and B. Demmig-Adams. 1999. The xanthophyll cycle and acclimation of Pinus ponderosa and Malva neglecta to winter stress. Oecologia 118:277-287. Vetaas, OR. 2002. Realized and potential climate niches: a comparison of four Rhododendron tree species. Journal of Biogeography 29:545-554. Wagner, DB. 1992. Nuclear, chloroplast, and mitochondrial DNA polymorphisms as biochemical markers in population genetic analyses of forest trees. New Forests 6(1-4):373-390 Warrington, I.J., and A.K.H. Jackson. 1981. Injury to radiata pine as influenced by freezing and thawing rate, and low temperature duration. New Zealand Journal of Forestry Science 11(1):37-44. Watano, Y., M. lmazu, and T. Shimizu. 1996. Spatial distribution of chNA and mtDNA haplotypes in a hybrid zone between Pinus pumila and P. parviflora var. pentaphylla (Pinaceae). Journal of Plant Research 1092403- 408. Wenny, D.L., and R.K. Dumroese. 2001. Propagation protocol for production of container Pinus ponderosa Laws. Var. ponderosa plants (90 ml (5 cu in) plugs); Forest Research Nursery, Moscow, Idaho. In Native Plant Network. URL: http:llwww.nativeplantnetwork.org (accessed 14 October 2005) Western Regional Climate Center. 2007. Western US. Historical Climate Summaries. URL: www.wrcc.dri.edulClimsum.html (accessed 1 June 2007) White, AS. 1985. Presettlement regeneration patterns in a Southwestern ponderosa pine stand. Ecology 66(2):589-594. Whittaker, RH. and WA. Niering. 1965. Vegetation of the Santa Catalina Mountains, Arizona: A gradient analysis of the south slope. Ecology 46(4):429-452. 262 Whittaker, RH. and WA. Niering. 1968. Vegetation of the Santa Catalina Mountains, Arizona: lV. Limestone and acid soils. Journal of Ecology 56(2):523-544. Whittaker, RH. and WA. Niering. 1975. Vegetation of the Santa Catalina Mountains, Arizona: V. Biomass, production, and diversity along an environmental gradient. Ecology 56(4):771-790. Woodward, F .l. 1987. Climate and Plant Distribution. Cambridge University Press, Cambridge, UK. 174p. Yeaton, R.l., R.W. Yeaton, and J.E. Horenstein. 1980. The altitudinal replacement of digger pine by ponderosa pine on the western slopes of the Sierra Nevada. Bulletin of the Torrey Botanical Club 107(4):487-495. 263