thesis_Oct_20

advertisement
Master Thesis
Introduction to Phenology
Phenology is a term derived from the Greek word phaino meaning to show or to appear. It
studies repeatable phenomena and their reasons. The phenomenon could be the life stages of
plants and animals, spectral reflectance of vegetation on satellite imageries, or the range of
glaciers. Plant phenology studies the botanical cycle events, and addresses the developmental
phases of plant organisms, recurring biological phases of species, biotic and abiotic causes, and
the interrelation of phases within or among species (Badeck et al., 2004; Leith, 1970; Rathcke
and Lacey, 1985). The timing of biological phases and its reaction to the variations of
environment is the major focus of this subject matter because of their significant responses to
fluctuation of climate (Menzel et al., 2006; Parmesan, 2006; Walther et al., 2002). Once the
response has genetically fixed, it became a phenological traits among populations. For the spring
phenology, these traits refer to flowering, budburst, or leave unfolding. These phenological traits
are usually used to monitor adaptive traits of the plants and evaluate their relationships to the
ecosystem where they grow (Arora and Boer, 2005; Howe et al., 2003). It has been applied in
agriculture to determine timing of planting and harvesting in order to achieve the maximum crop
yield (Sakamoto et al., 2005), and also used to decide the timings of applying pesticide and
herbicide (Moola and Mallik, 1998). In forest management, tree phenology has been used to
estimate forest productivity (Goetz and Prince, 1996) and variation of biochemical mass in
leaves (Kause et al., 1999), and to improve seed zoon selections (Hamann and Wang, 2006). In
addition, plant phenology is considered the key element that affects the carbon balance of
terrestrial ecosystems (Gill et al., 1998) and characterizes plant competition capabilities (Rathcke
and Lacey, 1985).
The basic knowledge of phenology has a long history of application in agriculture and forest.
Europe has the longest scientific phenological observation and research history that could be
tracked back as far as the 18th century (Leinonen and Hanninen, 2002; Luterbacher et al., 2007).
In North America, however, Thomas Mikesell started the earliest systematic phenology
observation between 1883 and 1921, about a century later, and recorded about 25 species during
that period of time (Lechowicz, 1995). The modern phenological recording was started by a
Swedish biologist Carolus Linnaeus and a British landowner Robert Marsham in the 18th
century (Lechowicz, 2001). The historical recorder still contributes for today’s research(Sparks
and Carey, 1995). Since then, records of explicit phenological observations were started, and
extensive plant phenological observation networks were established across the world, such as
Encyclopedia of Life (EOL) and Project BudBurst. Nowadays, one of the most influential
phenology watch networks, the Plantwatch, is based in Western Canada and is still very active1.
This network has documented phenological observations for decades and covered hundreds of
species' flowering and leave flushing. Beside the field observation, digital camera, aerial photo
(Carreiras et al., 2006), and satellite imagery(Fisher and Mustard, 2007; Reed et al., 1994) have
been introduced into the studying of phenology.
1
http://plantwatch.sunsite.ualberta.ca/misc/tracking.php
Environmental control of budburst
Temperature, thermal time,
Walther
Chilling requirement, photoperiod , CO2 ,
Temperature supply kinetic energy for biochemical reactions and enzyme’s catalyst. It controls
plant’s respiration, growth, and carbon uptake (Saxe et al., 2001). Spring phenology is also
driven by temperature (Menzel, 2003; Morin et al., 2009; Penuelas and Filella, 2001). However,
instead of using temperature itself, heatsum is convenient to describe the rate of development of
plants and their organs. Heatsum is the accumulation of degree-days for a phenological event in
its active period. Degree-days also known as heat units, thermal units, or day degrees and defined
as the accumulation of effective temperature. Generally, it sums the algebraic average of daily
temperature within a range of thematic threshold (Ghelardini et al., 2006; Hunter and Lechowicz,
1992). For example, budburst, the active period from the cessation of dormant to the beginning
of leaf flush, will not happen until temperature surpasses its thermal threshold and certain
amount of heatsum is achieved. Using heatsum as predictor, Reaumur successfully predicted the
occurrence of a phenological stage more than 200 years ago, suggesting heatsum is a constant
and could be used to project future or past phenological event of the same kind. Following this
idea, investigating the relationships between heatsum and phenological events of various species
has become a major topic in phenological researches. Some studies show the rate of botanic
development is linearly related to heatsum (Johnson and Thornley, 1985; Sharpe and Demichele,
1977). Considering the variation of heatsum requirement for different genotypes (Lappalainen,
1994), the distribution of heatsum may therefore reflect the spring phenology and genotypic
ranges of a specific species (Howe et al., 2003).
The accuracy of thematic threshold and starting date for a phenological event impacts on the
precision of heatsum computations, and different approaches have been adopted to estimate these
values. Yang et al. (1995) summarized four most common used approaches including the
smallest stantard deivation method, linear regression model, iteration method, and the triangle
method. The main idea of these approaches was to approximate a thematic threshold with
regression or iteration method based on the field-measured phenological data. Snyder et al.
(1999) found that the results from iteration method usually provid the smallest root mean square
error (RMSE) in most cases, indicating this is a better approach in effective temperature
estimates. Although these thresholds are theoretical rather than realistic value which are based on
biological tests or field observations, these thresholds are considered close enough in practice
(Snyder et al., 1999). Normally, temperature between 0°C and 10°C, optimum at about 5°C, is
considered sufficient for dormancy release for most species based on experiments (Perry, 1971),
and many studies suggested to set 0°C or 5°C for budburst effective temperature in predict
models (Snyder et al., 1999). For aspen, Heide (1993)set the threshold as 1°C and Beaubien
(2000) set as 0°C. Generally, the threshold simply using 0 is nearly as good for most modal
(Ring et al., 1983; Snyder et al., 1999). After setting the threshold, the heatsum could be
calculated by adding up all the effective temperatures from the starting to the end of a
phenological event, this computation is also called thermal time model (Delahaut, 2003; Reader,
1983) 2.
Threshold), 0 (Elisabeth), -3 to 15 effective as chilling (Santini 2004)
Temperature below threshold cannot attribute to the organism’s development anymore, but it still
influences phenological timing, especially budburst (e.g. Campbell and Sugano 1975, Murray et
al. 1989). Tree has a period to dormancy during the winter for avoid the cold damage. This
mechanism is prevalent under maritime climates, requiring a certain amount of chilling degreedays (accumulated temperatures below a certain threshold), before accumulating degree-days
toward the heatsum requirement starts. A chilling requirement prevents premature heatsum
accumulation and budbreak if fall and winter temperatures are unusually mild.
After growth cessation and before budburst, bud is in two states: the state of rest and the state of
quiescence.
Chilling hour
Photoperiod and others.
This mechanic protects plants to expose such risk
Spring phenology as an adaptive trait
Geographic patterns of budburst
(latitudinal, altitudinal, and coastal Davis 2001 )
Roughly, the phenotypic variations can be considers as the genotypic variation.
Title: Running to stand still: adaptation and the response of plants to rapid climate change
Author(s): Jump, AS; Penuelas, J
(Brissetter Barnes 1984)
Earlier researches noticed the geographic variation of spring phenology and attributed this
difference to their adaptations to local climates (Lechowicz, 1984). For example, the latitudinal
phenology variations were observed for budburst of Scots pine (Pinus sylvestris) and Norway
spruce (Picea abies) that plants from the south usually have earlier budburst while they are
planted in the same environment (Beuker, 1994; Leinonen and Hanninen, 2002), and similar
trends were found in some boreal tree species such as silver birch (Betula pendula) (Leinonen,
1996) and European elms (Ulmus minor, Ulmus glabra, and Ulmus laevis) (Santini et al., 2003).
The interpretation to all these phenomena is that these phenological variations are the results of
2
For a particular species, their geographical variation of heatsum requirements could be summarized as
the Linsser’s law (Ref. Reader 1983): the fractions of the heatsum required for a particular phenological
stage divided by the total annual heatsum for a plant at its site of origin are the same for plants from all
locations.
species’ adaptation to their local climates by minimizing their exposition to frost damage
(survival adaptation), while maximizing the duration of growing season (capacity adaptation)
(Leinonen and Hanninen, 2002). Mismatches between the spring weather and plant phenological
responses could potentially cause the plants from failing to produce seeds or fruits to increasing
the chance of mortality (Billington and Pelham, 1991). And plants that cannot respond to interannual climate variability to sufficiently use the growing season will be at a competition
disadvantage.
Recent researchers also found that adaptive traits of a species appear to be different through their
life stages, and plants of different ages usually choose different adaptive strategies to the same
environment (???). For example, because the frost damage is negatively correlated with the
timing of budburst, a seedling or sapling usually choose a 'threshold' strategy that it scarifies
taking the advantage of the full growing season, and takes a late budburst before its stem grows
longer and bigger; however, an adult tree may have more endurance to frost damage and may
take more risk to have an earlier budburst (Leinonen and Hanninen, 2002). [The same species
adapting to different environments or same plant at different age adapting the same environment
are overall the embodiment of phenotypic plasticity. Phenotypic plasticity is defined as the
adaptive traits of a plant to the altered environment.] With this concept, the species adaptation
could be better explained…………….
Usually, phenological events, such as budburst or flowering, could be statistically described by
timings of their occurrence, duration, and synchrony (Rathcke and Lacey, 1985) . Studies of
these phenological events usually based on onsite observations and recordings (Beaubien and
Freeland, 2000); however, some events, such as budburst, could be observed with the help of
remote sensing technologies (Sellers et al., 1995). Plant phenological phases are usually
categorized as budburst, bud set, flowering, and fruiting (???), although some studies classified
these events somewhat differently (Schwartz, 2003). Biologische Bundesanstalt and Chemical
industry (BBCH) unified the description and coding system by separating growth stages into 10
general phenological phases based on the development stages of bud, leave, stem, flower, and
fruit (Schwartz 2003) (detailed definitions see Appendix table 1), which has been widely adapted
by phenological studies. Following these criteria and standard procedures, plant phenology of
many species, especially those from the temperate zone, have been recorded and studied (e.g. ???
review), these information have provided great help to forest management (??) and agricultural
development (??).
Interpretation
Plant phenology is considered the result of plants adapting to their environment, especially for
those that grow in extreme environments such as the ecosystems u cold temperature or less
moisture (Perry, 1970). Abiotic factors usually have direct impacts on plant phenology. These
factors could be seasonal temperature variations (Beaubien and Freeland, 2000), frost damage in
spring and fall (Leinonen and Hanninen, 2002; Vitasse et al., 2009), or chilling effect in early
spring (Jonsson et al., 2004). However, differences in precipitation and soil moisture (Beaulieu et
al., 2002; Kramer et al., 2000; Reich, 1995) or variations in photoperiod (Partanen et al., 1998)
also have major influences on plant phenology. Biotic factors, on the other hand, are more
influential to plant regenerations. For example, the population dynamics of pollinators could
determine the timing of flowering, and the population variation of seed predators would also
affect whether the fruiting is successful (Elzinga et al., 2007; Kolb et al., 2007). Among all the
abiotic and biotic factors, temperature is the most important driver to plant phenology, especially
for the deciduous plants in the temperate zone (Badeck et al., 2004; Kramer et al., 2000). For
instance, in the moist temperate zone, most dormant trees require winter chilling to end the
dormancy in the beginning of spring and certain amount of heat to start bud bursting (Hunter and
Lechowicz, 1992). In addition, research found that the higher temperature speeds up the
phenology development (Saxe et al., 2001).
Genetic differentiation with respect to quantitative aspects of the phenotypes are not reflected in
patterning of enzyme variation, indicating that populations diverged in relation to local climate
despite gene flow (Davis and Shaw, 2001).
Selection and rapidly differentiate populations along an environmental gradient as a species is
expanding its rang
Phenology and forest management
Forest management is to guide forests toward a society's goals: preserving the environment,
meeting the current and future forest products needs of human society, or the combinations of
these former goals. Besides growing trees, forest management deals with other benefits provided
by forested land, the non-wood forest products, such as habitats for wildlife, food resources,
biodiversity, agroforestry, or recreation (Zeide, 2008) . Forest management is a long-range
viewpoint of a planner (Davis et al., 2005), it considers the predictable changes (such as human
population and climates) in the future and finds the resolutions, for example, to answer the
question about how we can share benefits from forests (e.g. forest service) with our descendants.
A sustainable human-forest ecosystem is desirable under this context and the core of modern
forest management (Davis et al., 2005).
Among all the factors affecting forest management, climate change is one of the biggest threats
for the forest industry. Climate change impacts forest reproductions by failing tree flowering and
fruiting. Kudo (2004), for example, found that bee-pollinated species had less seed-set in 2000 in
Japan because of a shortage of pollinators for the earlier flowering in the warm spring.
Considerable climate change has been observed around the world (Parmesan and Yohe, 2003),
and the trend is predicted to be continuous for the next century (Mbogga et al., 2009)(IPCC,
2007). Studies showed that the temperature has been increased dramatically since the 1980's
(Karl et al 2005); however, climate change also increased the frequencies of storms, fire,
precipitation, flood, snow, and other extreme events (Groisman et al., 2005; Saxe et al., 2001).
These changes have major impacts on species abundance, biological process, organic matter
decomposition, species range shift, as well as species adaptations (e.g. plant phenology) (Badeck
et al., 2004; Parmesan and Yohe, 2003; Saxe et al., 2001; Walther et al., 2002). Species change
their phenology to cope with the changing environments have been widely observed, such as the
changes of budburst and flowering timings (Beaubien and Freeland, 2000), leaf coloring (Estrella
and Menzel, 2006), and length of growing season (???). However, the climate change is too fast
for some species to keep up with, the physical migrations and gene flow from warm-adapted
population will be more important than species' evolution for maintaining the level of forest
ecosystem services (Billington 2008). Therefore, assistant plant migrations are necessary and the
major tasks for the future forest management.
Species with large ranges growing under a variety of environmental conditions will likely show
phenotypes differences in their adaptive traits (Howe et al., 2003), which are usually important
factors to be considered in the movement of planting stock for reforestation and in genetic tree
improvement programs. For example, if genotypes are selected for growth traits in short-term
experiments, adaptive traits may be sub-optimal. Consequently, the better growth may be the
result of risking late spring and early fall frost damage for an extended growing season (Brissette
and Barnes, 1984). Therefore, phenotypes with high mortality risk due to susceptibility to frost
damage or drought may not be a suitable choice for reforestation. Ideally, we would like to
choose phenotypes that show lower adaptive risks while maintaining superior growth.
Provenance
Genetic variation is an evolutionary result of plant adaption to the environmental heterogeneity
(Jelinski, 1997), and can be maintained through reproduction if the diversity was acquired
through recombination, introgression, or somatic mutation (Rasmussen and Kollmann, 2007).
The genetic variations are regulated by forces such as mutation, genetic drift, gene flow, and
natural selection (Ohsawa and Ide, 2008). Studying of genetic variation can help us to identify
species, assess their spatial distribution, examine the genetic structure, or probe their phenotypes.
Moreover, it has been used to select the high-quality timber resources or seed zones for forest
industry.
Genetic variations can be detected by many methods such as field observation, molecular genetic
marker, or quantitative traits locus (QTL) mapping (Gonzalez-Martinez et al., 2006). Molecular
genetic markers can be used to directly detect the genetic variations of a species because they
could be found at a known location on a chromosome and associated with a particular gene or
phenological trait. However, researches also found that genetic variations do not alway march to
the phenological or growth variations (Hall et al., 2007), and genetic variations detected by these
markers do not always associate with the suitable growth or phenological traits that are desirable
for the forest industry. For example, research found that the genetic variations in the conifer
species are much less than their adaptive traits (Gonzalez-Martinez et al., 2004). QTL mapping is
relatively straightforward than the other methods (Damerval et al., 1994). With this method,
many adaptive traits of tree species have been successfully identified, such as poplar (Ferris et
al., 2002) and Douglas fir (Wheeler et al., 2005). However, this method requires a large sample
size, and it is very time-consuming and expensive to construct.
All phenological traits show significant genetic differentiation among population and the results
were similar at the common garden sites (Hall et al., 2007).
(Ohsawa and Ide)
Field observation of phenological traits is the most traditional way to reveal the genetic
variations of a species; however, these observed differences might be confounded by
environmental variations. To eliminate the environmental factors and protrude the genetic
variations, the provenance trial, which is also called common garden, progeny test, or clonal test,
has been routinely performed (e.g. Hamann et al., 2000; Kleinschmit et al., 2004; Savva et al.,
2007). The essence of the provenance trials is to compare phenological and growth traits of
different genotypes within or among species from different sites in a same experimental site,
where they can be exposed to the same environmental conditions—soils, climate, water, and
photoperiod—with a systematic experimental design that accounts for random site variation
(Bower and Aitken, 2008). Because different genotypes may respond to the same environment
conditions differently in phenological traits, the observed differences can reflect within-species
genetic variations. This information can be used to create guidelines of seed transfers and to
delineate seed zones. The objective of limiting seed movement in reforestation is to ensure that
planting stock is not mal-adapted to environmental conditions of the planting site. For example,
northern provenances of Norway spruce (Picea abies), have earlier budburst and should
therefore not be used in southern planting environments to avoid late spring frost damage
(Leinonen and Hanninen, 2002).
- The advantage and disadvantage of provenance trials
Studying genetic variation through provenance trials has advantages and disadvantages. Major
advantage of provenance trials is with environmental variables controlled, a variety of growth
and adaptive traits can be evaluated for genetic variation, e.g.: growth traits (Lesser et al., 2004),
wood properties (Beaulieu et al., 2002), and adaptive traits (i.e. phenological characters)
(Backman, 1991; Li et al., 1997; Lobo et al., 2003). It should be kept in mind, however, that the
failure to detect genetic differences among populations in a common garden trial does not mean
that genotypes are identical. Genetic differences may be revealed under one set of environmental
conditions, but not under another. Therefore, provenance trials are typically replicated over
several environments. Testing multiple genotypes over multiple environments makes provenance
trial series expensive research efforts. To evaluate growth traits in trees at rotation age, they are
also very time-consuming. Studying budburst, however, is simpler. The trait can be observed
early on in seedlings or saplings (assuming that there is no change in phenology between
juvenile and adult trees), and environmental factors such as soil conditions and soil moisture are
thought to play a minor role (Backman, 1991). Therefore, results from a single provenance trial
observed in a single year should provide sufficient information. This offers the opportunity to
abandon the provenance trial approach entirely and attempt to study genetic variation in situ: this
study proposed a new approach of using remote sensed data to differentiate genotypes. This
would allow for the first time to generate seamless maps of genetic variation in populations
rather than obtaining information for a very limited set of samples.
Remote sensing
What it’s been used for
Remotely sensed observation phenology is land surface phenology
Cloud and other noise, tempera resolution, imagery calibration, and mixture pixels.
The sharp increases in NDVI that can be related to the onset of significant photosynthetic
activity (Reed et al., 1994)
Onset and offset of ‘green period’
New application
Objectives
Spatial patterns of genetic variations;
Remote sensing approach;
Reason for spatial patterns: understand the mechanic of adaptation to climate, application for
forest management; and predict the shift in the future responded to the climate change.
References:
Arora V.K., Boer G.J. (2005) A parameterization of leaf phenology for the terrestrial ecosystem
component of climate models. Global Change Biology 11:39-59. DOI: DOI
10.1111/j.1365-2486.2004.00890.x.
Backman T.W.H. (1991) Genotypic and Phenotypic Variability of Zostera-Marina on the WestCoast of North-America. Canadian Journal of Botany-Revue Canadienne De Botanique
69:1361-1371.
Badeck F.W., Bondeau A., Bottcher K., Doktor D., Lucht W., Schaber J., Sitch S. (2004)
Responses of spring phenology to climate change. New Phytologist 162:295-309. DOI:
DOI 10.1111/j.1469-8137.2004.01059.x.
Beaubien E.G., Freeland H.J. (2000) Spring phenology trends in Alberta, Canada: links to ocean
temperature. International Journal of Biometeorology 44:53-59.
Beaulieu J., Girard B., Fortin Y. (2002) Effect of drying treatments on warping of 36-year-old
white spruce seed sources tested in a provenance trial. Annals of Forest Science 59:503509. DOI: DOI 10.1051/forest:2002034.
Beuker E. (1994) Adaptation to Climatic Changes of the Timing of Bud Burst in Populations of
Pinus-Sylvestris L and Picea-Abies (L) Karst. Tree Physiology 14:961-970.
Billington H.L., Pelham J. (1991) Genetic-Variation in the Date of Budburst in Scottish Birch
Populations - Implications for Climate Change. Functional Ecology 5:403-409.
Brissette J.C., Barnes B.V. (1984) Comparisons of Phenology and Growth of Michigan and
Western North-American Sources of Populus-Tremuloides. Canadian Journal of Forest
Research-Revue Canadienne De Recherche Forestiere 14:789-793.
Carreiras J.M.B., Pereira J.M.C., Pereira J.S. (2006) Estimation of tree canopy cover in
evergreen oak woodlands using remote sensing. Forest Ecology and Management
223:45-53.
Damerval C., Maurice A., Josse J.M., Devienne D. (1994) QUANTITATIVE TRAIT LOCI
UNDERLYING GENE-PRODUCT VARIATION - A NOVEL PERSPECTIVE FOR
ANALYZING REGULATION OF GENOME EXPRESSION. Genetics 137:289-301.
Davis L.S., Johnson K.N., Bettinger P., Howard T.E. (2005) Forest management : to sustain
ecological, economic, and social values Waveland Press, Inc., Dubuque.
Davis M.B., Shaw R.G. (2001) Range shifts and adaptive responses to Quaternary climate
change. Science 292:673-679.
Delahaut K. (2003) Phenology : an integrative environmental science Schwartz, Mark Donald
ed. Kluwer Academic Publishers, Boston.
Elzinga J.A., Atlan A., Biere A., Gigord L., Weis A.E., Bernasconi G. (2007) Time after time:
flowering phenology and biotic interactions. Trends in Ecology & Evolution 22:432-439.
DOI: DOI 10.1016/j.tree.2007.05.006.
Estrella N., Menzel A. (2006) Responses of leaf colouring in four deciduous tree species to
climate and weather in Germany. Climate Research 32:253-267.
Ferris R., Long L., Bunn S.M., Robinson K.M., Bradshaw H.D., Rae A.M., Taylor G. (2002)
Leaf stomatal and epidermal cell development: identification of putative quantitative trait
loci in relation to elevated carbon dioxide concentration in poplar. Tree Physiology
22:633-640.
Fisher J.I., Mustard J.F. (2007) Cross-scalar satellite phenology from ground, Landsat, and
MODIS data. Remote Sensing of Environment 109:261-273. DOI: DOI
10.1016/j.rse.2007.01.004.
Ghelardini L., Falusi M., Santini A. (2006) Variation in timing of bud-burst of Ulmus minor
clones from different geographical origins. Canadian Journal of Forest Research-Revue
Canadienne De Recherche Forestiere 36:1982-1991. DOI: Doi 10.1139/X06-092.
Gill D.S., Amthor J.S., Bormann F.H. (1998) Leaf phenology, photosynthesis, and the persistence
of saplings and shrubs in a mature northern hardwood forest. Tree Physiology 18:281289.
Goetz S.J., Prince S.D. (1996) Remote sensing of net primary production in boreal forest stands.
Agricultural and Forest Meteorology 78:149-179.
Gonzalez-Martinez S.C., Krutovsky K.V., Neale D.B. (2006) Forest-tree population genomics
and adaptive evolution. New Phytologist 170:227-238. DOI: DOI 10.1111/j.14698137.2006.01686.x.
Gonzalez-Martinez S.C., Mariette S., Ribeiro M.M., Burban C., Raffin A., Chambel M.R.,
Ribeiro C.A.M., Aguiar A., Plomion C., Alia R., Gil L., Vendramin G.G., Kremer A.
(2004) Genetic resources in maritime pine (Pinus pinaster Aiton): molecular and
quantitative measures of genetic variation and differentiation among maternal lineages.
Forest Ecology and Management 197:103-115. DOI: DOI 10.1016/j.foreco.2004.05.008.
Groisman P.Y., Knight R.W., Easterling D.R., Karl T.R., Hegerl G.C., Razuvaev V.A.N. (2005)
Trends in intense precipitation in the climate record. Journal of Climate 18:1326-1350.
Hall D., Luquez V., Garcia V.M., St Onge K.R., Jansson S., Ingvarsson P.K. (2007) Adaptive
population differentiation in phenology across a latitudinal gradient in European Aspen
(Populus tremula, L.): A comparison of neutral markers, candidate genes and phenotypic
traits. Evolution 61:2849-2860. DOI: DOI 10.1111/j.1558-5646.2007.00230.x.
Hamann A., Wang T.L. (2006) Potential effects of climate change on ecosystem and tree species
distribution in British Columbia. Ecology 87:2773-2786.
Heide O.M. (1993) Daylength and Thermal Time Responses of Budburst during Dormancy
Release in Some Northern Deciduous Trees. Physiologia Plantarum 88:531-540.
Howe G.T., Aitken S.N., Neale D.B., Jermstad K.D., Wheeler N.C., Chen T.H.H. (2003) From
genotype to phenotype: unraveling the complexities of cold adaptation in forest trees.
Canadian Journal of Botany-Revue Canadienne De Botanique 81:1247-1266. DOI: Doi
10.1139/B03-141.
Hunter A.F., Lechowicz M.J. (1992) Predicting the Timing of Budburst in Temperate Trees.
Journal of Applied Ecology 29:597-604.
Jelinski D.E. (1997) On genes and geography: a landscape perspective on genetic variation in
natural plant populations. Landscape and Urban Planning 39:11-23.
Johnson I.R., Thornley J.H.M. (1985) Temperature-Dependence of Plant and Crop Processes.
Annals of Botany 55:1-24.
Jonsson A.M., Linderson M.L., Stjernquist I., Schlyter P., Barring L. (2004) Climate change and
the effect of temperature backlashes causing frost damage in Picea abies. Global and
Planetary Change 44:195-207. DOI: DOI 10.1016/j.gloplacha.2004.06.012.
Kause A., Ossipov V., Haukioja E., Lempa K., Hanhimaki S., Ossipova S. (1999) Multiplicity of
biochemical factors determining quality of growing birch leaves. Oecologia 120:102-112.
Kolb A., Ehrlen J., Eriksson O. (2007) Ecological and evolutionary consequences of spatial and
temporal variation in pre-dispersal seed predation. Perspectives in Plant Ecology
Evolution and Systematics 9:79-100. DOI: DOI 10.1016/j.ppees.2007.09.001.
Kramer K., Leinonen I., Loustau D. (2000) The importance of phenology for the evaluation of
impact of climate change on growth of boreal, temperate and Mediterranean forests
ecosystems: an overview. International Journal of Biometeorology 44:67-75.
Kudo G., Nishikawa Y., Kasagi T., Kosuge S. (2004) Does seed production of spring ephemerals
decrease when spring comes early? Ecological Research 19:255-259.
Lappalainen H. (1994) Examples of plant phenological events in Finland and their relation to
temperature. Memoranda Societatis pro Fauna et Flora Fennica 70:105-121.
Lechowicz J. (2001) Phenology, Encyclopedia of Global Environmental Change, Wiley, Landon.
Lechowicz M.J. (1984) Why Do Temperate Deciduous Trees Leaf out at Different Times Adaptation and Ecology of Forest Communities. American Naturalist 124:821-842.
Lechowicz M.J. (1995) Seasonality of Flowering and Fruiting in Temperate Forest Trees.
Canadian Journal of Botany-Revue Canadienne De Botanique 73:175-182.
Leinonen I. (1996) Dependence of dormancy release on temperature in different origins of Pinus
sylvestris and Betula pendula seedlings. Scandinavian Journal of Forest Research 11:122128.
Leinonen I., Hanninen H. (2002) Adaptation of the timing of bud burst of Norway spruce to
temperate and boreal climates. Silva Fennica 36:695-701.
Leith. (1970) Phenology in productivity studies Springer-Verlag, New York.
Lesser M.R., Cherry M., Parker W.H. (2004) Investigation of limestone ecotypes of white spruce
based on a provenance test series. Canadian Journal of Forest Research-Revue
Canadienne De Recherche Forestiere 34:1119-1127. DOI: Doi 10.1139/X03-286.
Li P., Beaulieu J., Bousquet J. (1997) Genetic structure and patterns of genetic variation among
populations in eastern white spruce (Picea glauca). Canadian Journal of Forest ResearchRevue Canadienne De Recherche Forestiere 27:189-198.
Lobo J.A., Quesada M., Stoner K.E., Fuchs E.J., Herrerias-Diego Y., Rojas J., Saborio G. (2003)
Factors affecting phenological patterns of bombacaceous trees in seasonal forests in
Costa Rica and Mexico. American Journal of Botany 90:1054-1063.
Luterbacher J., Liniger M.A., Menzel A., Estrella N., Della-Marta P.M., Pfister C., Rutishauser
T., Xoplaki E. (2007) Exceptional European warmth of autumn 2006 and winter 2007:
Historical context, the underlying dynamics, and its phenological impacts. Geophysical
Research Letters 34:-. DOI: Artn L12704
Doi 10.1029/2007gl029951.
Mbogga M.S., Hamann A., Wang T.L. (2009) Historical and projected climate data for natural
resource management in western Canada. Agricultural and Forest Meteorology 149:881890. DOI: DOI 10.1016/j.agrformet.2008.11.009.
Menzel A. (2003) Plant phenological anomalies in Germany and their relation to air temperature
and NAO. Climatic Change 57:243-263.
Menzel A., Sparks T.H., Estrella N., Koch E., Aasa A., Ahas R., Alm-Kubler K., Bissolli P.,
Braslavska O., Briede A., Chmielewski F.M., Crepinsek Z., Curnel Y., Dahl A., Defila C.,
Donnelly A., Filella Y., Jatcza K., Mage F., Mestre A., Nordli O., Penuelas J., Pirinen P.,
Remisova V., Scheifinger H., Striz M., Susnik A., Van Vliet A.J.H., Wielgolaski F.E.,
Zach S., Zust A. (2006) European phenological response to climate change matches the
warming pattern. Global Change Biology 12:1969-1976.
Moola F.M., Mallik A.U. (1998) Phenology of Vaccinium spp. in a black spruce (Picea mariana)
plantation in northwestern Ontario: possible implications for the timing of forest
herbicide treatments. Canadian Journal of Forest Research-Revue Canadienne De
Recherche Forestiere 28:1579-1585.
Morin X., Lechowicz M.J., Augspurger C., O' Keefe J., Viner D., Chuine I. (2009) Leaf
phenology in 22 North American tree species during the 21st century. Global Change
Biology 15:961-975. DOI: DOI 10.1111/j.1365-2486.2008.01735.x.
Ohsawa T., Ide Y. (2008) Global patterns of genetic variation in plant species along vertical and
horizontal gradients on mountains. Global Ecology and Biogeography 17:152-163. DOI:
DOI 10.1111/j.1466-8238.2007.00357.x.
Parmesan C. (2006) Ecological and evolutionary responses to recent climate change. Annual
Review of Ecology Evolution and Systematics 37:637-669. DOI: DOI
10.1146/annurev.ecolsys.37.091305.110100.
Parmesan C., Yohe G. (2003) A globally coherent fingerprint of climate change impacts across
natural systems. Nature 421:37-42. DOI: Doi 10.1038/Nature01286.
Partanen J., Koski V., Hanninen H. (1998) Effects of photoperiod and temperature on the timing
of bud burst in Norway spruce (Picea abies). Tree Physiology 18:811-816.
Penuelas J., Filella I. (2001) Phenology - Responses to a warming world. Science 294:793-+.
Perry T.O. (1970) Immobility of Dormancy Regulators in Interracial Grafts of Acer-Rubrum L.
Plant Physiology 46:33-&.
Perry T.O. (1971) Dormancy of Trees in Winter. Science 171:29-&.
Rasmussen K.K., Kollmann J. (2007) Genetic diversity, spatial patterns, and growth of root
sprouts in a temperate tree at the northern distribution limit. Ecoscience 14:250-258.
Rathcke B., Lacey E.P. (1985) Phenological Patterns of Terrestrial Plants. Annual Review of
Ecology and Systematics 16:179-214.
Reader R.J. (1983) Using Heatsum Models to Account for Geographic-Variation in the Floral
Phenology of 2 Ericaceous Shrubs. Journal of Biogeography 10:47-64.
Reed B.C., Brown J.F., Vanderzee D., Loveland T.R., Merchant J.W., Ohlen D.O. (1994)
Measuring Phenological Variability from Satellite Imagery. Journal of Vegetation Science
5:703-714.
Reich P.B. (1995) Phenology of Tropical Forests - Patterns, Causes, and Consequences.
Canadian Journal of Botany-Revue Canadienne De Botanique 73:164-174.
Ring D.R., Calcote V.R., Cooper J.N., Olszak R., Begnaud J.E., Fuchs T.W., Neeb C.W., Parker
R.D., Henson J.L., Jackman J.A., Flynn M.S., Harris M.K. (1983) Generalization and
Application of a Degree-Day Model Predicting Pecan Nut Casebearer (Lepidoptera,
Pyralidae) Activity. Journal of Economic Entomology 76:831-835.
Sakamoto T., Yokozawa M., Toritani H., Shibayama M., Ishitsuka N., Ohno H. (2005) A crop
phenology detection method using time-series MODIS data. Remote Sensing of
Environment 96:366-374. DOI: DOI 10.1016/j.rse.2005.03.008.
Santini A., Ghelardini L., Falusi M., Bohnens J., Buron M., Collin E., Solla A., Vanden B.A.
(2003) Vegetative bud-burst variability of European elms, New approaches to elm
conservation. Second International Elm Conference, Valsain, Spain, 20-23 May 2003.,
Instituto Nacional de Investigacion y Tecnologia Agraria y Alimentaria (INIA). pp. 3745.
Saxe H., Cannell M.G.R., Johnsen B., Ryan M.G., Vourlitis G. (2001) Tree and forest
functioning in response to global warming. New Phytologist 149:369-399.
Schwartz M.D. (2003) Phenology : an integrative environmental science Kluwer Academic
Publishers, Boston.
Sellers P.J., Meeson B.W., Hall F.G., Asrar G., Murphy R.E., Schiffer R.A., Bretherton F.P.,
Dickinson R.E., Ellingson R.G., Field C.B., Huemmrich K.F., Justice C.O., Melack J.M.,
Roulet N.T., Schimel D.S., Try P.D. (1995) Remote-Sensing of the Land-Surface for
Studies of Global Change - Models, Algorithms, Experiments. Remote Sensing of
Environment 51:3-26.
Sharpe P.J.H., Demichele D.W. (1977) Reaction-Kinetics of Poikilotherm Development. Journal
of Theoretical Biology 64:649-670.
Snyder R.L., Spano D., Cesaraccio C., Duce P. (1999) Determining degree-day thresholds from
field observations. International Journal of Biometeorology 42:177-182.
Sparks T.H., Carey P.D. (1995) The Responses of Species to Climate over 2 Centuries - an
Analysis of the Marsham Phenological Record, 1736-1947. Journal of Ecology 83:321329.
Vitasse Y., Porte A.J., Kremer A., Michalet R., Delzon S. (2009) Responses of canopy duration to
temperature changes in four temperate tree species: relative contributions of spring and
autumn leaf phenology. Oecologia 161:187-198. DOI: DOI 10.1007/s00442-009-1363-4.
Walther G.R., Post E., Convey P., Menzel A., Parmesan C., Beebee T.J.C., Fromentin J.M.,
Hoegh-Guldberg O., Bairlein F. (2002) Ecological responses to recent climate change.
Nature 416:389-395.
Wheeler N.C., Jermstad K.D., Krutovsky K., Aitken S.N., Howe G.T., Krakowski J., Neale D.B.
(2005) Mapping of quantitative trait loci controlling adaptive traits in coastal Douglas-fir.
IV. Cold-hardiness QTL verification and candidate gene mapping. Molecular Breeding
15:145-156. DOI: DOI 10.1007/s11032-004-3978-9.
Yang S.S., Logan J., Coffey D.L. (1995) MATHEMATICAL FORMULAS FOR
CALCULATING THE BASE TEMPERATURE FOR GROWING DEGREE-DAYS.
Agricultural and Forest Meteorology 74:61-74.
Zeide B. (2008) The science of forestry. Journal of Sustainable Forestry 27:345-473. DOI:
10.1080/10549810802339225.
Download