Eucalyptus - Forest Productivity Cooperative

Transcription

Eucalyptus - Forest Productivity Cooperative
2014 Research Summaries
Volume 2:
Eucalyptus
Forest Productivity Cooperative Members
Full Members
Agropical
American Forest Management
Arauco-Bioforest
Boise
Buckeye Technologies
Claritas
Copener
Cotopaxi
Deforsa
Dougherty & Dougherty Forestry
Evans Properties
F&W Forestry
FCB
Fibria
FOMEX
Forest Investment Associates
Forestal Mininco
Forestal Rio Biabo
Greenwood Resources
Hancock Forest Management
IFOM
International Paper
Jordan Lumber
Klabin
Lykes Ranch
Masisa
MWV
Milliken Forestry
Molpus Timberlands Management
Refocosta
RMK Timberland Group
Resource Management Service
Smurfit Cartón de Colombia-Venezuela
Superior Pine Products
Suzano
The Campbell Group
Timberland Investment Resources
Valor Florestal
Westervelt Company
Weyerhaeuser
Sustaining Members
Agrium
Aprilasia
North Carolina Forest Service
USDA Forest Service
Virginia Department of Forestry
Corresponding Members
AgXplore
ArborGen
Carolina Soil
COMPO Agro Chile
CONIF
C3 Partners SAS
Du Campo
DuPont Crop Protection
Florida Grown
Forestry & Land Resource Consultants
Gavilon
GreenTechnologies
Innovatech
International Forest Company
International Plant Nutrition Institute
Koch Agronomic Services
Mosaic Fertilizer
Payne’s Flying Service
Syngenta
Thrash Aviation
ii
2013 Research Summaries
Dear FPC Members,
The mission of the Forest Productivity Cooperative (FPC) is to create innovative solutions to enhance forest
productivity and value through sustainable management of site resources. Our primary research focus is
developing whole-rotation silvicultural regimes that are biologically sound, cost effective, and
environmentally sustainable.
Since its organization in 1969, the FPC has evolved into an international research and education program
working with the major pine and eucalypt plantation species in the southern United States, throughout
Latin America, including: Argentina, Brazil, Chile, Colombia, Ecuador, Mexico, Uruguay and Venezuela,
and Indonesia.
The success of the FPC has been driven by our conceptual approach to research, interaction with the
members, and undergraduate and graduate education. Collaborations with our members are crucial to
our success, and all FPC meetings (2 Contact Meetings and 1 Annual Meeting per year) provide an open
environment for exchange of ideas, experiences and sharing research results to improve operational,
tactical and strategic decisions. In addition to these regular meetings, frequent local visits to members
provide opportunities for technology transfer and practical use of research findings.
The FPC research initiatives are driven by the strategic plan approved by the members in 2009 which
established eleven major research topics: 1) nutrient availability, fertilizer uptake and efficiency; 2)
decision support tools and models; 3) remote sensing and leaf area index; 4) weed control and
fertilization; 5) fertilizer response; 6) eucalypt silviculture; 7) genetic x silviculture interactions; 8) response
modeling and growth and yield modeling; 9) ecophysiology, process models and potential productivity;
10) thinning; and 11) water and environmental impacts. In 2013 the FPC created working groups based
on species and geography and the research and technology transfer will be organized around four
working groups: i. SE US Pine, ii. SE US Eucalyptus, iii. Latin America Pine, iv. Latin America Eucalyptus.
To address these research interest areas, FPC has a large network of field research trials that includes 61
active experiments in the US and 45 in Latin America. This is one of the largest networks of forestry
research installations in the world. They provide the FPC a unique ability to produce practical solutions to
issues related to sustained productivity of intensively managed forests. We are also able to leverage the
investment made in these studies with outside grants and partners that address key global questions
related to ecophysiology of forest growth, remote sensing, biomass energy, carbon sequestration and
global change. Among our collaborators are the IPEF Forestry Institute in Brazil and the CAFS, Pinemap,
IBSS, SIFEB, Fondecyt, Conicyt projects.
These are the 2013 FPC Research Summaries, which we update annually every October, during our
Annual Meeting. The main purpose of this research summary is to provide an overview and summary of
the key results and management implications from all of the regionwide studies, special studies and
graduate student research projects currently underway within the FPC. We also include a list of the
papers published in peer reviewed journals over the last four years (average of 20 papers per year),
highlighting the productivity and quality of our research program. Members can access all of the
research results from the FPC website (www.forestproductivitycoop.org). Given the new working group
system, the FPC Research Summaries are published in two volumes: Vol 1 Pine and Vol 2 Eucalyptus.
Finally, we would like to THANK all FPC members for their continued support.
Sincerely yours,
The Forest Productivity Cooperative Team
iii
iv
Forest Productivity Cooperative
2013 Research Summaries
Eucalyptus
v
Table of Contents
Publications
Eucalyptus and Hardwood Publications: 2010 - 2013 ........................................................................................ 2
Regionwide Trials
Regionwide 13: Effects of N and P fertilization post-establishment on growth
of Eucalyptus nitens plantations in the south-central Chile .............................................................................. 8
Regionwide 18: Rate and optimal frequency of fertilization at mid-rotation Eucalyptus grandis stands in
Colombia ................................................................................................................................................................ 10
Regionwide 21: Eucalyptus Twin-Plot network: the Deforsa Venezuela Trial................................................ 12
Regionwide 23: Duration and intensity of weed control of Eucalyptus grandis
in Colombia ............................................................................................................................................................ 14
Regionwide 24: Eucalyptus Special Project - Identifying cold tolerant Eucalyptus
species and their productivity in the SEUS ........................................................................................................ 16
Regionwide 25: Eucalyptus Nutrient Omission Study........................................................................................ 18
Regionwide 25: Eucalyptus Nutrient Omission Study in Northern Brazil ......................................................... 20
Regionwide 25: Nutrient omission at the establishment and early development of Eucalyptus grandis in
Colombia. ............................................................................................................................................................... 22
Regionwide 25: Nutrient omission study on young Eucalyptus nitens plantations in the south-central
Chile ......................................................................................................................................................................... 24
Graduate Student Research
Effects of intensive silviculture on the productivity of Eucalyptus in Venezuelan
Llanos and a probabilistic analysis of its profitability ........................................................................................ 28
Eucalyptus forest plantation effects on soil organic carbon across Eastern Brazil...................................... 30
Soil carbon dynamics following reforestation of tropical pastures in Brazil .................................................. 32
Effects of Eucalyptus camaldulensis, E. globulus, E. nitens and Acacia melanoxylon short-rotation forest
crops (SRFC) for energy production in contrasting soils of the Bío-Bío Region, Chile ................................. 34
Using forest inventory as a tool of monitoring silvicultural quality in
Eucalyptus clonal plantations .............................................................................................................................. 36
Growth, water use and water use efficiency of Eucalyptus clones under different spacings: from
the tree to the watershed scales ...................................................................................................................... 38
Growth and yield model for Eucalyptus benthamii in the SE United States………………………………… 40
The radiation-temperature effect on cold tolerant Eucalyptus species ...................................................... 42
Effects of Intensive Silviculture in the Restoration Process of Northeastern
Atlantic Rainforest in Brazil .................................................................................................................................... 44
Fertilizer response to Eucalyptus plantations and its correlation with the soil, climate
and silviculture characteristics in different regions of São Paulo state ......................................................... 46
Zoning the productivity of Eucalyptus forest plantation in the northeastern
of São Paulo state in Brazil .................................................................................................................................... 48
Characterization of seasonal stem growth, canopy development and light use efficiency in
Eucalyptus clone………………………………………………………………………………………………………… 50
Leaf anatomy characterization of 16 Eucalyptus genotypes with different canopy ideotypes.............. 52
Physiological variables and γ-aminobutyric acid (GABA) accumulation in Eucalyptus clones under
water stress.............................................................................................................................................................. 54
vi
Effect of soil water availability on photosynthetically active radiation capture by
Eucalyptus globulus, Eucalyptus nitens and Eucalyptus camaldulesis hybrids ............................................ 56
Effect of water availability and temperature on leaf deployment in genotypes of Eucalyptus globulus,
Eucalyptus nitens and Eucalyptus camaldulensis hybrids ............................................................................... 58
Special Studies
Tolerance of Eucalyptus Clones to Hydric and Thermal Stresses - TECHS ..................................................... 62
Eucahydro: tools for early assessment of Eucalyptus genotypes for water use efficiency, water consumption, efficient use of water and drought resistance ......................................................................................... 64
Special study: Potential Productivity of Eucalyptus maidenii and E. globulus in Uruguay ........................ 66
Integrated Biomass Supply Systems —Comparative water and nutrient use in Eucalyptus benthamii and
Pinus taeda plantations ........................................................................................................................................ 68
Geodatabase System: maps for forestry and environmental applications ................................................. 70
Modeling monthly mean air temperature for Brazil.......................................................................................... 72
Stand-level patterns of carbon fluxes and partitioning in a Eucalyptus grandis
plantation across a gradient of productivity, in São Paulo State, Brazil ....................................................... 74
Stem production, light absorption and light use efficiency between dominant and
non-dominant trees of Eucalyptus grandis across a productivity gradient in Brazil.................................... 76
Nelder spacing trial to evaluate Acacia mangium development in
Northeast Llanos Area of Venezuela .................................................................................................................. 78
vii
Publications
2
Eucalyptus and Hardwoods Publications: 2010 - 2013
Since the 1970s, the faculty, staff and graduate students associated with the FPC have produced over
300 articles and reports related to forest productivity and soils. All of our articles, book chapters, theses
and reports are available for download from our website. Below is a list of our publications since 2010.
2014
Alvares, CA, JL Stape, PC Sentelhas, JLM Goncalves, G Sparove. 2014. Koppen’s climate classification
map for Brazil. Meteorologische Zeitschrift. DOI: 10.1127/0941-2948/2013/0507
Watt, M., R. Rubilar, M. Kimberley, D. Kriticos, V. Emhart, O. Mardones, M.Acevedo, M. Pincheira, J.
Stape, T. Fox. 2014. Using seasonal measurements to inform ecophysiology; Extracting cardinal
growth temperatures for process based growth models of five Eucalyptus species/crosses from
simple field trials.
New Zealand Journal of Forestry Science. 44:9. http://
www.nzjforestryscience.com/content/44/1/9.
Watt, M, Rafael Rubilar, Mark O Kimberley, Darren J Kriticos, Veronica Emhart, Oscar Mardones, Manuel
Acevedo, Matias Pincheira, Jose L. Stape, Thomas Fox. Using seasonal measurements to inform
ecophysiology: extracting cardinal growth temperatures for process-based growth models of
five Eucalyptus species/crosses from simple field trials. New Zealand Journal of Forestry Science
44(9): 1-11.
2013
Alvares, C.A., J.L. Stape, P.C. Sentelhas, J.L.D.M. Goncalves. 2013. Modeling monthly mean air
temperature for Brazil. Theor. Appl. Climatol. 113:407-427.
Campoe, O.C., J.L.Stape, Y.Nouvellon, J.Laclau, W.L.Bauerle, D.Binkley, G.le Maire. 2013. Stem
production, light absorption and light use efficiency between dominant and non-dominant trees of
Eucalyptus grandis across a productivity gradient in Brazil. Forest Ecology and Management 288: 1420
Cook, R., J.L.Stape, D.Binkley. In press. Soil carbon dynamics following reforestation of tropical pastures.
Soil Science Society of America Journal
Ferraz, S.F.B., M.Pereira, F.Paula, C.Vettorazzi, C.A. Alvares. 2013. Simulação de perdas de solo em
função de cenários de adequação ambiental em microbacias agrícolas. Scientia Forestalis 98: 271282
Gonçalves, J.L.M., C.A. Alvares, A.R. Higa, L.D. Silva, A.C. Alfenas, J. Stahl, S.F.B. Ferraz, W.P. Lima, P.H.S.
Brancalion, A. Hubner, J.P.D. Bouillet, J.P. Laclau, Y. Nouvellon, D. Epron. 2013. Integrating genetic
and silvicultural strategies to minimize abiotic and biotic constraints in Brazilian eucalypt plantations.
Forest Ecology and Management 301:6-27.
Laclau, J.P., E.Silva, G.R.Lambais, M.Bernoux, G. le Maire, J.L.Stape, J.P. Bouillet, J.L.M.Goncalves,
C.Jourdan, Y.Nouvellon. 2013. Dynamics of soil exploration by fine roots down to a depth of 10 m
throught the entire rotation in Eucalyptus grandis plantation. Frontiers in Plant Science, Functional
Plant Ecology / doi: 10.3389/fpls.2013.00243
3
Marsden, C., Y.Nouvellon, J.P.Laclau, M.Corbeels, R.E.McMurtrie, J.L.Stape, D.Epron, G.le Maire, 2013.
Modifying the G'DAY process-based model to simulate the spatial variability of Eucalyptus plantation
growth on deep tropical soils. Forest Ecology and Management 301: 112–128
Mora, F., R.A. Rubilar, V.I. Emhart, J. Saavedra. 2013. Prediccion Bayesiana de parametros geneticos en
clone de Eucalyptus globulus bajo condiciones de suplemento hidrico. Ciencia Florestal, Santa
Maria 23(2): 529-536.
Navarrete-Campos, D., L.A. Bravo, R.A. Rubilar, V.I. Emhart, R. Sanhueza. 2013. Drought effects on water
use efficiency, freezing tolerance and survival of Eucalyptus globulus and Eucalyptus globulus x
nitens cuttings. New Forests 44:119-134.
Otto, M.S.G., A.R.Vergani, A.N.Goncalves, A.Vrechi, S.R.Silva, J.L.Stape. 2013. Photosynthesis, stomatal
conductance and productivity of Eucalyptus clones under different soil and climatic conditions.
Arvore 36: 431-439
Sabatia, C.A., T.R. Fox, and H. Burkhart. 2013. Extending a model system to predict biomass in mixedspecies southern Appalachian hardwood forests. Southern Journal of Applied Forestry 37(2):122-126.
DOI 10.5849/sjaf.12-005.
Silva, P.H.M., A.C. Miranda, M.L.T. Moraes, E.L. Furtado, J.L. Stape, C.A. Alvares, P.C. Sentelhas, E.S. Mori,
A.M. Sebbenn. 2013. Selecting for rust (Puccinia psidii) resistance in Eucalyptus grandis in São Paulo
State, Brazil. Forest Ecology and Management 303:91-97.
Stanturf, J.A., E.D. Vance, T.R. Fox, M. Kirst. 2013. Eucalyptus beyond its native range: Environmental
issues in exotic bioenergy plantation. Forestry. DOI: 10.1155/2013/46030.
Zhou, J., C.Proisy, X.Descombes, G.le Maire, Y.Nouvellon, J.L.Stape, G.Viennois, J.Zerubia, P.Couteron.
2013. Mapping local density of young Eucalyptus plantations by individual tree detection in high
spatial resolution satellite images. Forest Ecology and Management 301: 129-141
2012
Alvares, C. A., Gonçalves, J. L. M., Vieira, S.R., Silva, C. R., Franciscatte, W. 2012. Spatial variability of
physical and chemical attributes of some forest soils in southeastern of Brazil. Scientia Agrícola (USP.
Impresso), v. 68, p. 697-705, 2011.
Brondani, G.E., F.J.B.Baccarin, H.W.W.Ondas, J.L.Stape, A.N.Goncalves, M.Almeida. 2012. Low
temperature, IBA concentrations and optimal time for adventitious rooting of Eucalyptus benthamii
mini-cuttings. Journal of Forestry Research 23: 583-592
Campoe, O.C., Stape, J.L., Laclau, J.-P., Marsden, C., Nouvellon, Y., 2012. Stand-level patterns of carbon
fluxes and partitioning in a Eucalyptus grandis plantation across a gradient of productivity, in São
Paulo state, Brazil. Tree Physiology 32, 696–706.
Cook, R.L. 2012. Long-term Effects of Forest Plantations on Soil Carbon in Brazil. Ph.D. Thesis. Dept. of
Forestry, NCSU, Raleigh, NC. 124 pp.
Fox, T.R. and J. Creighton. 2012. Silvics of Oak. In A Practical Guide to Managing Oak Forests in the
Eastern United States. University of Tennessee Press.
Gazell, A.C.F., C.A.Righi, J.L. Stape, O.C.Campoe. 2012. Tree species richness, does it play a key role on
a forest restoration plantation? Bosque 33: 3-4
Gonçalves, j. l. m., Alvares, C. A., Gonçalves, T.D., Moreira, R.M., Mendes, J.C.T., Gava, J.L. 2012.
Mapeamento de solos e da produtividade de plantações de Eucalyptus grandis em Itatinga, SP,
com uso de sistema de informação geográfica. Scientia Forestalis (IPEF), v. 94, p. 187-201, 2012.
4
Kiser, L.C. and T.R. Fox. In Press. Short Rotation Woody Crop Biomass Production for Energy. Chapter 6. In
B. Singh (Ed.). Biofuel Crop Sustainability. John Wiley and Sons.
Kiser, L.C. and T.R. Fox. 2012. Soil accumulation of nitrogen and phosphorus following annual fertilization
of loblolly pine and sweetgum on sandy sites. Soil Science Society of America Journal 76:2278-2288.
Navarrete-Campos, L. A. Bravo, R.A. Rubilar, V. Emhart, R. Sanhueza. 2012. Drought effects on water use
efficiency, freezing tolerance and survival of Eucalyptus globulus and Eucalyptus globulusXnitens
cuttings. New Forests DOI 10.1007/s11056-012-9305-0.
Peduzzi, A., R.H. Wynne, V.A. Thomas, R.F. Nelson, J.J. Reis and M. Sanford. 2012. Combined Use of
Airborne Lidar and DBInSAR Data to Esitmate LAI in Temperate Mixed Forests. Remote Sensing 4: 1758
-1780.
Sabatia, C.A., T.R. Fox and H. Burkhart. 2012. Extending a model system to predict biomass in mixedspecies southern Appalachian hardwood forests. Southern Journal of Applied Forestry.
Silva, A. M., Alvares, C. A., Watanabe, C. H. 2012. Natural Potential for Erosion for Brazilian Territory. In:
GODONE, D; STANCHI, S. (Org.). Soil Erosion Studies. Soil Erosion Studies. Rijeka: InTech - Open Access
Publisher, 2011, v. , p. 1-22.
Stape, J. L., Souza, V. C., Torrado, P. V., Rodriguez, L. C. E. 2012. Estimativas das taxas de sequestro de
carbono na Reserva Particular do Patrimônio Natural SESC Pantanal. Conhecendo o Pantanal 8. 1.
ed. Rio de Janeiro: SESC Serviço Social do Comércio, 2011. v. 1. 144p
Vickers, L.A. ,T.R. Fox, D. L. Loftis, and D. A. Boucugnani. 2012. Predicting Forest Regeneration in the
Central Appalachians Using the REGEN Expert System. Journal of Sustainable Forestry.30(8):790-822.
2011
Alvarez, J. H. Mitasova and H.L. Allen. 2011. Estimating Solar Radiation in South-Central Chile. Chilean J.
Agric. Res. Vol. 71(4): 601-609.
Atwood, C.J., T.R. Fox, and D. L. Loftis. 2011. Effects of Alternative Silvicultural Treatments on
Regeneration in the Southern Appalachians. Journal of Sustainable Forestry. 30:419-440.
Fox, T.R., B.W. Miller, R.R. Rubilar, J.L. Stape, T.J. Albaugh. 2011. Phosphorus nutrition of forest plantations:
the role of inorganic and organic P. Chapter 13 In: Buneman, E., A. Oberson, and E. Frossard (eds).
Phosphorus in action, Soil Biology 26. Springer-Verlag.
Harrison, R.H, D. Richter, and T. R. Fox. 2011. Deep soils. For. Sci. 57:1-2.
Le Maire G, Marsden C, Verhoef W, Ponzoni F, Seen D, Stape JL, Nouvellon Y. 2011. Leaf area index
estimation with MODIS reflectance time series and model Inversion during full rotations of Eucalyptus
plantations. Remote Sensing of Environment 115: 586-599.
Le Maire G, Marsden C, Nouvellon Y, Grinand C, Hakamada R, Stape JL, J.P.Laclau. 2011. MODIS NDVI
time-series allow the monitoring of Eucalyptus plantation biomass. Remote Sensing of Environment
115: 2613-2625
Miller, B.W. and T.R. Fox. 2011. Long-term fertilizer Effects on oxalate desorbable phosphorus pools in a
typic paleaquult. Soil Science Society of America Journal. 75 (2).
Silva E, Bouillet JP, Goncalves JL, Abreu C, Trivelin P, Nouvellon Y, Stape JL, Laclau JP. 2011. Functional
specialization of Eucalyptus fine roots: contrasted potential uptake rates for nitrogen, potassium and
calcium tracers at varying soil depth. Functional Ecology (doi: 10.1111/j.1365- 2435.2011.01867.x).
Sucre, E.B. J.W. Tuttle, and T.R. Fox. 2011. The use of ground-penetrating radar as a tool to accurately
estimate soil depth in rocky forest soils of the southern Appalachians. Forest Science. 57:59-66.
5
2010
Binkley, D., J.L. Stape, W.L. Bauerle, M.G. Ryan. 2010. Explaining growth of individual trees: Light
interception and efficiency of light use by Eucalyptus at four sites in Brazil. Forest Ecology and
Management 259(9): 1704-1713.
Binkley D, Laclau JP, Stape JL, Ryan MG. 2010. Applying ecological insights to increase productivity in
tropical plantations. Forest Ecology and Management. 259: 1682-1683.
Campoe, O.C., J.L. Stape, J.C.T. Mendes. 2010. Can intensive management accelerate the restoration
of Brazil’s Atlantic forest? Forest Ecology and Management 259: 1808-1814.
Cubbage, F., P. MacDonagh, G. Balmelli, R. Rubilar, R. de la Torre, V. Hoeflich, M. Murara, H. Kotze, R.
Gonzalez, O. Carrero, G. Frey, S. Koesbandana, V.M. Olmos, J. Turner, R. Lord, J. Huang, R. Abt. 2010.
Global forest plantation investment returns. XIII World Forestry Congress, Buenos Aires, Argentina, 1823 October 2009.
Doi, B., D. Binkley, J.L. Stape. 2010. Does reverse growth dominance develop in old plantations of
Eucalyptus saligna? Forest Ecology and Management 259(9): 1815-1818.
Fox, T.R., B.W. Miller, R.R. Rubilar, J.L. Stape, T.J.Albaugh. 2010. Phosphorus nutrition of forest plantations:
the role of inorganic and organic P. Chapter 13 In: Buneman, E., A. Oberson, and E. Frossard (eds).
Phosphorus in action – biological processes in soil phosphorus cycling. Springer
Homyack, J.A., E.B. Sucre, C.A. Haas, and T.R. Fox. 2010. Does Plenthodon cinereus affect leaf litter
decomposition and invertegrate abundances in mixed oak forests? Journal of Herpetology.44:447456.
Hubbard, R.M., J.L. Stape, M.G. Ryan, A.C. Almeida, J. Rojas. 2010. Effects of irrigation on water use and
water use efficiency in two fast growing Eucalyptus plantations. Forest Ecology and Management
259: 1714-1721.
Marsden C, G. Maire, J.L. Stape, D. Seen, O. Roupsard, Y. Nouvellon. 2010. Relating MODIS vegetation
index time-series with structure, light absorption and stem production of fast-growing Eucalyptus
plantations in Brazil. Forest Ecology and Management 259(9): 1741-1753.
Maurice, J., J.P. Laclau, D.S. Re, J.L. de Moraes Gonçalves, Y. Nouvellon, J.P. Bouillet, J.L. Stape, J.
Ranger, M. Behling, J.L. Chopart. 2010. Fine root isotrophy in Eucalytpus grandis plantations.Towards
the prediction of root length densities from root counts on trench walls. Plant Soil. DOI 10.1007/s11104
-010-0380-8.
Ryan, M.G., J.L. Stape, D. Binkley, S. Fonseca, R.A. Loos, E.N. Takahashi, C.R. Silva, S.R. Silva, R.E.
Hakamada, J.M. Ferreira, A.M.N. Lima, J.L. Gava, F.P. Leite, H.B. Andrade, J.M. Alves, G.G.C. Silva.
2010. Factors controlling Eucalyptus productivity: How water availability and stand structure alter
production and carbon allocation. Forest Ecology and Management 259: 1695-1703.
Stape, J.L., Binkley, D. 2010. Insights from full-rotation Nelder spacing trials with Eucalyptus in São Paulo,
Brazil. Southern Forests 2010, 72(2): 90–97.
Stape, J.L., D. Binkley, M.G. Ryan, S. Fonseca, R.A. Loos, E.N. Takahashi, C.R. Silva, S.R. Silva, R.E.
Hakamada, J.M.A. Ferreira, A.M.N. Lima, J.L. Gava, F.P. Leite, H.B. Andrade, J.M. Alves, G.G.C. Silva,
M.R. Azevedo. 2010. The Brazil Eucalyptus Potential Productivity Project: Influence of water, nutrients
and stand uniformity on wood production. Forest Ecology and Management 259: 1684-1694.
6
7
Regionwide Trials
8
Regionwide 13: Effects of N and P fertilization post-establishment on
growth of Eucalyptus nitens plantations in the south-central Chile
Objectives.

To determine the effect mid-rotation N and P
fertilization of Eucalyptus nitens.

To evaluate leaf area, growth efficiency and
foliar nutrient level responses to N+P
fertilization.

To determine the operational dose to
optimize tree growth response at each site.
Methods
Study area
three sites were established for Eucalyptus
nitens RW13 trials in the VIII region of Chile, with
temperatures averaging between 13 and 17
degrees Celsius, rainfall between 1000-1400 mm
per year and soils derived from three geological
formations, recent volcanic ash, ancient
volcanic ash (red clay) and sediment marines
(Table 1).
Trial desing
The
experimental
design
considered
randomized complete blocks with three
replicates. Treatments considered the effect of
a factorial combination of N and P fertilization
on E. nitens early growth considering nine
treatments where a base application of 50 Kgha
-1 de K , 30 Kgha-1 de Mg , 3 Kgha-1 de Zn, 2
Kgha-1 de Cu, 4 Kgha-1 de B, 400Kgha-1 Ca
was applied to the control and other treatments
(Table 2).
Height, diameter and leaf area measurements
for 135411 and 135412 trials were obtained since
treatments application in 2011. For 134905
treatments started in 2012. Trials 135411 and
135412 were established on plantations
established in 2010 (2 year old) and 134905 trial
was established on a plantation 2007 (5 years
old).
Las
Site
Millongue
Bandurrias
Las
Maicas
Code
City
Latitude
135411
San Carlos
36ª26’22.88”
135412
Lebu
37ª33’14.08”
134905
Mulchen
37o49’69.02”
Longitude
71ª41’07.38”
73ª37’08.95”
72º17’96.03”
MAT (ªC)
Rainfall
(mm.year-1)
Soil
14
13
17
1034
1302
1380
Recent
volcanic ash
Sedimentary
marine ter-
Red clays
Table 1. Characteristics of the sites.
T001
T011
T021
0*N + 0P
0N + 30P
0N + 60P
T101
T111
T121
150N + 0P
150N + 30P
150N + 60P
T201
T211
T221
300N + 0P
300N + 30P
300N + 60P
Table 2. Evaluated Regionwide 13 treatments considering
elements combinations in Kg ha-1
Results
After three years since the treatments were
imposed, sites Las Bandurrias and Milongue did
not present significant differences among the
fertilization treatments.
Three year since
treatment application, the productivity for Las
Bandurrias and Millongue sites ranged from 5362 and 14-18 m3/ha, respectively (Figure 1).
Similarly, the site Las Maicas did not show
significant differences among the treatments 2
year since tratment application, with a
productivity ranging from 150-178 m3/ha.
9
Despite of the differences among the
treatments, these did not differ significantly
from the control. However, it seems evident
that these differences are increasing over time
in all the sites.
LAI was measured 2 year since treatment
application in the sites Las Bandurrias and
Millongue. In las Maicas site it was measured 1
year after treatment were imposed.
Significant differences in leaf área index were
observed at all sites. For Las Bandurrias and Las
Maicas sites, LAI reached values close to 7
units. The Millonge site averaged a LAI value of
4.0.
At the Las Bandurrias site the best
treatments (containing 300 Kgha-1 N) had 1
additional unit of LAI relative to the control
treatment. Similar results were observed for the
Millongue site, with the best dose for 150 kgha-1
of N and 60 kgha-1of P with 4.7 units.
Figure 2. Cumulative volume per unit of leaf area index. Blue
lline: Las Bandurrias, Red line: Millongue & Green line: Las
Maicas.
For las Maicas site the differences among
tretament were not more than 0.5. Figure 2
shows the relationship between the cumulative
volumen and LAI measured in 2013, which is a
surrogate measure of growth efficiency. It can
be seen that growth efficiency differed
between sites. Las Bandurrias site presented
the lowest growth efficiency, whereas in the
other sites the growth efficiency was similar.
These results suggest that the highest growth
was achieved at treatments that showed high
LAI responses.
Future Direccion
Trials will be evaluated with anual or bianual
measurements at least for six years based on
lon-term responses to fertilization observed
previously
for
RW13
trials.
Soil
and
environmental variables will be analized to
explain leaf area index responses across sites.
Figure 1. Cumulative volume per treatment for the sites Las
Bandurrias (upper panel), Millongue (middle panel) and Las
Maicas (lower panel).
10
Regionwide 18: Rate and optimal frequency of fertilization at mid-rotation
Eucalyptus grandis stands in Colombia
Trial series objectives
Treatments and experimental design
The RW 18 was designed to investigate the
effects of rate and frequency of N fertilization
application in E. grandis until rotation age. The
major questions under investigation were: 1) Is
there any response to mid-rotation nutrient
additions on these high fertility sites? , 2) What
level of cumulative nutrient additions may be
required to obtain a response?, 3)Does
frequency and rate matters to maximize tree
growth response?, and 4) Is there any
cummulative effect of nutrient additions on
nutrient use efficiency? Considering these
questions the main objectives of these trial series
were to: 1) determine the growth response to
increasing amounts of N availability in midrotation stands, and 2) To determine how
maximum growth response may be obtained by
rate and frequency of application.
The array of treatments for RW18 studies
considered two independent variables: annual
dose of applied fertilizer applied and frequency
of fertilization (Table 2). Studies were established
with four replicates.
A total of 7 studies were initially installed, but a
wildfire destroyed one of the studies during its
second year of development and preliminary
results are not considered in this report. Of the 6
studies successfully established 3 of them are
located in the area of Cauca, 1 in the area of
Valle del Cauca and 1 in the area of Pereira
Quindio considering a range in soil.site
conditions (Table 1).
Table 1 . Climate, geology and soil characteristics
of RW18 trials established in Colombia.
Fertilization Response
Two sites (184604 and 184601) of the six sites
Table 2. Treatments applied at each trial and
accumulated N in kg ha-1 over time.
established
showed
extraordinary
growth
responses in diameter, height and cummulative
volume compared even to experiments with
added nutrient and water resources (Birk and
Turner, 1992). Both sites, having one of the largest
responses observed at FPC trials, were located
at Pubenza soils series. This soil series is widely
distributed in the area with more than 50% of
the soils. These soils are considered of medium
quality and productive potential due their past
intensive agricultural land use despite presenting
high organic matter levels and total nitrogen.
Individual tree growth for the control treatment
had an average of 17 m height and 12.4 cm
DBH for the 4601 site, while for the best treatment
(480 kg N ha-1), for 4604 site, reached 22.6 m in
height and 16.1 cm in DBH (p < 0.05). At 4601 site
cumulative volume after 42 months reached 325
11
m3 ha-1 for the best treatment (960 kg N ha-1)
compared to 172 m3 ha-1 for the control
treatment (p < 0.05). For the 4604 site cumulative
volume for the best treatment (900 kg N ha-1)
was 604 m3 ha-1 after 78 months (Figure 1).
Figure 1. Cumulative volume growth for RW18 trials
that showed response to fertilization.
Figure 2. Leaf area response at 184601.
Optimum response ranged 180-240 Kg N ha-1
every two years (Fig 3). Cumulative doses
maximized response at 800-1000 Kg N ha-1 (Fig
4).
Figure 3. Duration of response for dose and
frequency of application at responsive RW18 trials.
Responses to mid-rotation nutrient additions may
double productivity of E.grandis plantations. The
observed growth rate after fertilization at both
sites is exceptionally high compared to sites in
Brazil and other previous references elsewhere.
Figure 4. Cumulative N dose and cumulative
volume response at RW18 trials.
Figure 5. Stand growth response at 184601.
Cumulative volume growth response observed
at sites with response to fertilization, showed a
type A response post treatment application.
High C/N ratios and low N and P foliar levels
seem to be related to site specific responses.
Nutrient availability mechanisms and plant
uptake in these ecosystems are not well
understood.
Future direction
Nutrient accumulation on aboveground and
belowground biomass analyses will be studied
before harvesting of 184601 site to understand
second rotation carryover of improved nutrition
and nutrient removals of these fast growing E.
grandis stands.
12
Regionwide 21: Eucalyptus Twin-Plot network: the Deforsa Venezuela Trial
Introduction
Objectives
The Twin Plot approach is a paired design used
to capture resource limitations at the landscape
level (Stape et al. 2004). Twin plots have the
advantage of producing knowledge on spatial
distribution of resource limitations (Ferreira and
Stape 2009) and may provide a basis for
parameterization of process based model
(Landsberg, 2011).
To determine the Eucalyptus potential
productivity in San Carlos, Venezuela.
Understanding what limits productivity across a
landscape can help develop site-specific
silvicultural prescriptions. Resource limitations
can be directly verified by the change in forest
current
annual
increment
(CAI) when
confronting CAI of inventory and the twin-plot.
Change in Leaf area index (LAI) (Figure 1) can
be used as a quick surrogate for site resource
limitation due to its direct effect on light
capturing and use, which precedes the final
stemwood increment results.
Methodology
To understand the main factors that constrains
Eucalyptus forest growth across this landscape.
To develop tools to identify managementresponsive Eucalyptus sites.
A total of 76 pairs of twin-plots were established
on Deforsa land (Venezuela) in 2009 (Figures 1,
2). The company has established clonal forest
plantations mainly with E. urophylla and E.
grandis x urophylla. Deforsa’s inventory network
consists of 1 permanent plot per 5 ha of
plantation, being installed when the forests
reaches 2 years-old. Permanent inventory plots
have approximately 60 trees (4 rows X 15 trees/
row). From the existing plots, 8% were selected to
represent Deforsa landscape.
The RW21 Eucalyptus Twin-Plot trial are currently
being used or installed in Deforsa (Venezuela)
and Smurfit Cartón de Colombia. This report
presents the initial results at the Deforsa site in
San Carlos, Venezuela (Figure 1).
Figure 2. Twin Plot: fertilized, no weeds
Figure 1. Pair of Inventory-Twin Plots at Deforsa
13
For every selected inventory plot, a twin-plot
location was carefully chosen in the vicinity,
based on the visual assessment of mortality,
diameter, height, competing vegetation, soil
and topography situation. Twin plots were
established in a way to assure that no run-off
would move from the twin-plot to the inventoryplot carrying applied fertilizers.
In some sites (25% of total) where weeds were
abundant, a third plot was established (called,
the triplet-plot) which received only weed
control in order to isolate the effect of weed
competition in the management response. A
total of twenty triplet plots where established.
For the selected group of plots with high
competing vegetation problems, no statistical
difference was found between Weed Control
and Fertilization + Weed Control treatments.
However both were more productive than the
control plot (Figure 4). These results highlight the
important of weed control in this tropical
environment, which can lead to an increase in
productivity of more than 20% in a very short
time.
Results
Overall, there was a quick and significant
(paired t test p<0.0001) increase in productivity
of Deforsa’s forests in the Twin-Plots due to the
combined weed control and fertilization effects
(Fig 3).
Significant statistical differences were found for
DBH, total height, and biomass increment . At
some specific sites the Management Response
(Twin-Plot growth minus Contorl growth)
reached up to 10 Mg ha-1 yr-1, which represents
an important increase over the productivity
(Figure 3).
Preliminary Conclusions
The Twin-Plot design, with combined effect of
weed control and fertilization, has proven to be
a simple and efficient design to capture
operational
constraints
to
Eucalyptus
productivity in tropical conditions.
The triplet-design, isolating weed control and
fertilization effects, also shown to be effective to
point the relative importance of each
management.
The study, which is also evaluating Leaf Area
Index and Light Use Efficiency, intends to identify
the adequate tool to identify between
responsive and non-responsive sites for a
operational point of view.
Acknowledgements
Currently the same management practice is
applied to all the stands, however it is clear that
in some stands responses are almost zero, and in
others are very high, which shows the possibility
to site-specific optimization. For that purpose, a
tool which allows classifying the sites as
responsive or non-responsive to intensive
treatment need to be established.
Support for this work was provided by Deforsa
Company, the Forest Productivity Cooperative
and the Zobel Grant. We also thanks Deforsa’s
personnel and the Universidad de Los Andes
(Mérida, Venezuela).
14
Regionwide 23: Duration and intensity of weed control of Eucalyptus
grandis in Colombia.
Trial series objectives
Table 1. Characterization establishment sites .
1. To determine the optimal duration and
intensity of weed control to maximize survival
and tree growth of E. grandis.
Farm
Cedral
Suiza
Primavera
RW23
234601
234603
234602
2. To assess biomass and nutrient uptake of
competing vegetation over time and its effect
on resource availability for the target crop.
Area
North
Central
South
Provinces
PereiraRosa
Restrepo
Meseta
Latitude
4° 42' 57'' N
3°50'45'' N
2° 37' 16'' N
Longitud
75° 38' 15'' W
76°29'49'' W
76° 39' 01''
W
Elevation
(m.s.n.m)
1890
1553
1746
Precip.
2673
1256
2616
Previous
use
E.grandis
E. grandis
E.grandis
Plantation
May 2010
May 2010
June 2010
3.- To model the biological and economical
response of E. grandis to duration and intensity
of weed control.
S.
Sites characteristics
Three studies were established in Colombia by
Smurfit Kappa Carton de Colombia (RW234601,
RW234602 and RW234603). Table 1 shows some
characteristic of the sites.
Studies were
established on well drained clay loams and a
moderately well drained loamy soil. The clay
loam sites has nearly twice the precipitation
(2.600mm) than the loamy site (1.250mm). All
soils derived from volcanic p. materials.
Treatments and experimental design
The experimental design considered a complete
randomized block design with 8 to 12 replicates
depending on site and vegetation variability.
Treatments applied considered a factorial
combination of four levels of weed control
intensities (no weed control, 33% , 66% and 100%
of weed control in bands centered in the
planting line) and eight levels of weed control
duration (30, 60, 90, 120, 180, 360, 540 and 720
days free of competing vegetation) (Figure 1).
All sites had a broadcast preplanting chemical
weed control with Glifosate (3,0L/ha) and
Flufenacet (0,45 L/ha) before the application of
weed control duration and intensity treatments.
These trials will be carried until 4 years or shorter
based on canopy closure responses.
Figure 1. Generic block layout of intensity and
duration of weed control treatments.
15
Weed control intensity and duration
The effect of weeds on the nutrient availability
may be very detrimental for tree growth. For
instance, two years after establishment, the
uptake of N varied among the sites from 70 to
330 kg/ha (Figure 2).
After 3 years weed control intensity treatments
effects were clearly observed on tree growth
and showed similar responses for both 66% and
100% weed control intensities at RW234601 and
RW234603. These results suggested that
complete weed control may not be required to
obtain maximum benefits at these sites. In terms
of weed control intensity and duration, the best
response was observed for 66% or 100% intensity
and a period larger than 90 days of duration
allowed to obtain a 40-50 m3ha-1 gain after 3
years of stand development (Figure 3).
Therefore, in sites as 234601 and 234603, the first
90 days of weed control are critical to obtain the
maximum potential growth response to weed
control for these sites. Contrastingly at RW234602
a larger gain was observed when an intensity of
100% weed control independently of the
duration (Figure 3). At this site, the gains of
weed control reached 80 m3ha-1.
Critical results of these trials in all studied sites
suggest that 90 initial days of free weed control
in the line of planting (66% or 100% intensity) are
key to obtain good stand development and
should be key directions for weed control
operations. Advantages of 100% weed control
may be obtained in terms of free stand mobility
but may challenge biodiversity or soil-site
sustainability. The amount of weed at these sites
Figure 3. Mean volume response to the weed
intensity (100% area) and duration for the sites
234601 (upper panel), 234602 (middle panel) and
234603 (lower panel) after 3 years.
were variable, although all the sites had 100%
weed cover a years after establishment.
Future direction
Trials will be carried until rotation age but future
measurements and analyses every 5 years. Additional
soils and nutritional evaluations are under way.
Figure 2. Nitrogen use by weed in the control
treatment over time.
16
Regionwide 24: Eucalyptus Special Project - Identifying cold tolerant
Eucalyptus species and their productivity in the SEUS
Background
Eucalyptus species can have very rapid growth
in other parts of the world. Recent opportunities
in using tree biomass as a replacement for fossil
fuels has generated interest in growing
Eucalyptus in the United States.
Previous
attempts to introduce these species in the 1980s
failed largely due to cold temperatures of the
1983 and 1985 winters. Also, additional entries
that may be more tolerant to cold are now
available from sources around the world. This
new genetic material as well as access to
expertise in growing the species stimulated
interest in this work. Fifteen members of the
Forest Productivity Cooperative are currently
supporting the project (AFM, ArborGen, Boise,
Buckeye, Copener, Deforsa, Fibria, Florida
Grown, IFCO, International Paper. Lykes, MWV,
Rayonier, Suzano and Weyerhaueser).
Study Objectives:
1. To identify cold-hardy Eucalyptus species that
can survive and thrive in the SEUS.
2. To quantify the productivity of these species
under different silvicultural treatments.
Figure 2. E.macarthurii, E.benthamii and E.viminalis in Raleigh
NC (35ª N) after the 2011, 2012 and 2013 winters.
3. T o
investigate
the
environmental factors related
with Eucalyptus cold-risk and
yield zoning; and
4. To provide information for
tree breeding strategies and
hybridization.
Experimental design
One
hundred
and
fifty
Eucalyptus species from 333
entries were planted in 28 sites
across the SE US (Fig.1) in a
biomass or screening design. In
the Biomass design a total of 16,
32 or 64 trees were planted per
plot and the design replicates
Figure1 . Locations of the twenty five RW24 trials installed in 2010, 2011 and 2012
17
are the sites. In the screening one the plot
has 3 plants and many entries.
Current status
Seven trials were planted in 2010, twelve
in 2011, six in 2012 and three in 2013. The
sites were planted between May and
September of each year. Hourly
temperature at each site were recorded
from November to March with a HOBO
datalogger. The 2011, 2012 and 2013
winters can be considered
cooler,
warmer and normal, based on historical
averages (Fig.3). So, they have been ideal
to allow an adequate species screening.
Indeed, for the RW24, a total of 6 winters
will take place prior to the complete
results of the special project. Assessments
of height and survival for the 2010, 2011
and 2012 trials were completed in May of
2013 after the frost free dates in all sites.
Figure 4. Poor growth of E.benthamii (left) when
E.camaldulensis in Palmdale, Florida. Compare with Fig. 2
Results
Entries from seven countries were planted and
survival and growth were very good at all sites
prior to the winter due to careful site prep,
fertilization and weed control. A Canopy
Damage Assessment was developed for visual
evaluation of the frost damage. All 2010, 2011
and 2012 sites experienced subfreezing
temperatures (Fig.3) and, as expected, the
coldest site was in NC and the one with the
mildest temperatures was in South of Florida.
compared
with
result is the G X E interaction as shown, for
instance, in Figs. 2 and 4 where E.benthamii is
adapted to Raleigh NC, but not in Palmdale FL,
and vice-versa with E.camaldulensis, which died
out in Raleigh. Overall, for all sites, five
Eucalyptus species came up as potential
genetic materials for cold-hardy tolerance and
fast-grow: E.benthamii, E.dorrigoensis, E.viminalis,
E.dalrympleana and E. macarthurii. However
other species have site-specific performances.
The weekly evaluation of the E.benthamii growth
in Raleigh using band dendrometers (Fig.5)
confirmed that Eucalyptus has no dormant
period and its photosynthesis is controlled by the
Figure 3. Minimum daily temperature from April 1st 2010 to
March 30th 2013 at Summerville, SC.
However, the daily variation in temperature was
more intense in the flatwoods regions of FL, GA.
In general, across sites the number of species
that survived the winter decreased with the
increasing events of subfreezing temperatures,
except for the flatwoods region where the lack
of cold-hardy acclimatization affected even
cold-hardy species like E.benthamii. Another
Figure 5. Relation between E.benthamii growth and minimum
temperature in Raleigh NC in a weekly-basis.
temperature during the winter time.
Acknowledgements
IPEF
(Brazil),
FCBA
(France),
Deforsa
(Guatemala), Mininco (Chile), University of
Florida (US), CSIRO (Australia) for donating/
selling the seeds. Special thanks to Dr. Thomas
Leidig
18
Regionwide 25: Eucalyptus Nutrient Omission Study
Justification
Table 1. Nutrient Omission trial treatments
Fast-growing Eucalyptus plantations demands
adequate levels of soluble nutrients in the soil
solution to be uptake in order to build a large
and healthy canopy.
The faster the canopy is built, the earlier the
canopy closure is reached, which facilitates the
competing vegetation control and increases the
likelihood of a higher Mean Annual Increment at
the end of the short-rotation.
Traditional areas of Eucalyptus plantations, like
Latina America and Brazil, already have a
calibrated
fertilization
recommendation
protocols based on soil analysis and Eucalyptus
species expected productivity. However, when
new Eucalyptus species are planted in new sites,
with specific soil physical and chemical
attributes, these protocols must be verified.
Some Forest Productivity Cooperative members
are trying or expanding their Eucalyptus
plantations with new species in new sites which
requires a better understanding regarding soil
nutrient availability versus plant nutrient demand
in regions like the SE US, central Brazil and Chile.
infer about the soil natural nutrient availability, a
Complete treatment with the application of all
nutrients to assure plenty availability of them for
adequate tree growth, and a series of
treatments where each nutrient is removed from
the complete balance to characterize its unique
importance in growth (Table 1).
Installations
The
Nutrient
Omission
Design
was
already
The Nutrient Omission design
is therefore proposed by FPC
as an easy and quick way to
capture
four
main
information at the speciessite level: i) Is there a nutrient
limitation for fast canopy
closure?; ii) If so, which
nutrients are limiting the full
species development ?; iii)
Among these nutrients, what
is the ranking in importance
when considering their supply
via fertilization?; and iv) Can
soil analysis be used for
fertilizer recommendation?
Nutrient Omission
Design
The Nutrient Omission design
has a very simple rational. It
has a Control treatment to
Figure 1. E.benthamii in Allendale, SC, at 1 year-old. Control (left), All Nutrients (right)
19
installed in three sites of the SE US (2 in FL and 1 in
SC) and one site in central Brazil (Tocantins
State), in five different FPC members. The species
being studies are E.benthamii, E.grandis and
E.grandis x urophylla clones.
Results
In all sites, the fertilization effect was very clear
(Fig. 1), pointing to the need of an adequate
forest nutrition for Eucalyptus fully express its
growth potential in soils with very low fertility.
Fig.1 and Fig.2 shows the fertilization effect on
the US Coastal Plains. For these very sandy and
poor sites the most limiting nutrients are P, N and
K (Fig.2) which need to be incorporate during
the initial fertilization of the forest to guarantee
fast canopy closure and high growth rates. In all
trials the weed control was imposed.
Figure 3. Eucalyptus grandis stem volume at 1 yr-old in
Palmdale, FL according to the omission treatment (Table 1).
Columns with the same letter do not differ at P=0.05.
Another very important result from these two
trials in the US was the trend to increase coldhardness due to a better nutrition (Fig.3). This
effect is speculated to be related with the K
effect.
Similar to the US results, the omission study in
Brazil also was very effective to picture the
crucial impact of the nutrients for Eucalyptus
clones development, including both macro–
and micronutrients. Notice that 1 year
development in Brazil matches the 2 years
development in the US (Fig.2 and Fig.4).
Next Steps
Figure 2. Eucalyptus grandis volume at 2 years-old at Palmdale,
FL according to the omission treatment (Table 1). Columns with
the same letter do not differ at P=0.05.
Figure 4. Eucalyptus clones volume at 1 yr-old at Tocantins,
Brazil, according to the omission treatment (Table 1).
The Nutrient Omission Design took soil samples
from the sites to evaluate the potential use of soil
analysis for fertilization recommendation. In
addition, at 1 and 2 years-old, leaf samples
(Figure 5) will be analyzed to verify the possibility
to establish critical nutrient levels for the different
Eucalyptus species at each site.
Figure 5. Leaf sampling protocol for Eucalyptus.
20
Regionwide 25: Eucalyptus Nutrient Omission Study in Northern Brazil
JL Stape, Clayton Alvares, Rafaela Carneiro, Thomas Fox, Rafael Rubilar
Justification
Fast-growing Eucalyptus plantations demands
adequate levels of soluble nutrients in the soil
solution to be uptake in order to build a large
and healthy canopy.
The faster the canopy is built, the earlier the
canopy closure is reached, which facilitates the
competing vegetation control and increases the
likelihood of a higher Mean Annual Increment at
the end of a short-rotation.
growth (Table 1).
Installations
The Nutrient Omission Design was installed
Table 1. Nutrient Omission trial treatments
Traditional regions of Eucalyptus plantations in
Brazil, like São Paulo and Minas Gerais states,
already have a calibrated fertilization
recommendation protocols based on soil
analysis and Eucalyptus species expected
productivity. However, when new Eucalyptus
clones are planted in new sites, with specific soil
physical and chemical attributes, these
protocols need to be verified.
Many companies are expanding their
Eucalyptus plantations with new species/clones
in new sites which
requires a better
understanding regarding soil nutrient availability
versus plant nutrient demand in regions like
Tocantins State in Northern Brazil.
The Nutrient Omission design was therefore
proposed by FPC as an easy and quick way to
capture four main information at the species-site
level: i) Is there a nutrient limitation for quick
canopy closure?; ii) If so, which nutrients are
limiting the full species development ?; iii)
Among these nutrients, what is the ranking in
importance when considering their supply via
fertilization?; and iv) Can soil or tissue analysis be
used for fertilizer recommendation?
Nutrient Omission Design
The Nutrient Omission design has a very simple
rational. It has a Control treatment to infer about
the soil natural nutrient availability, a Complete
treatment with the application of all nutrients for
adequate tree growth, and treatments where
each nutrient is removed from the complete
balance to characterize its importance in
Figure 1. E. urograndis in Tocantins-Brazil, at 2.5 year-old. Left
complete treatment (A) and right control treatment (C)
(January 2011) in a Yellow Oxisol in Brejinho de
Nazaré, Tocantins State (Brazil), on site of Valor
Florestal Company. The genotypes being studies
are four hybrid clones of Eucalyptus grandis x
urophylla.
Measurements
Growth is evaluated every six months via DBH
and height measurements, and LAI estimation.
At 3-years old an aboveground biomass
assessment was performed in the control and
21
complete treatments.
Leaf Area Index were estimated for all plots via
the use of the Decagon ceptometer.
Results
Higher LAI are associated with greater
productivity of the forest. Figure 3 highlights the
relationship between LAI and volume. The
control treatment has a lower LAI, and the
complete a higher LAI.
Eucalyptus plantations on soils of low fertility of
the Tocantins state depend on fertilization to
reach their potential growth. The micronutrients
omission revealed to be as limiting as the control
Figure 2. Stem biomass for each block/clones of E.urophylla x E.
grandis at 3 yr-old at Brejinho de Nazaré, Northen Brazil,
according to the omission treatment (Table 1).
treatment (no fertilization).
The content of boron foliar seems to be the most
important indicator, because it showed a
positive correlation with the stem volume across
all treatments (Figure 4).
Figure 4. Relationship between .boron foliar and stem
volume for all 28 plots in 2014.
With the results shown on Figure 4, a new
experiment was installed with the omission of
Figure 3. LAI of the clone A217 at 3 yr-old at Brejinho de
Nazaré, Northern Brazil, according to the omission treatment
(Table 1).
Figure 2 presents stem biomass at 3 years-old
while Figure 3 contains the LAI results.
For the sandy and poor soil of this trial (Yellow
Oxisol, subsoiling site prep), the trial, at just 1 year
-old, already shows the crucial need of fertilizers
to promote forest and growth.
The most limiting nutrients are the micronutrients,
followed by K (Figures 2), clearly showing the
need to incorporate them in the initial
fertilization of the forest to guarantee fast
canopy closure and high growth rates.
In addition, the trial was kept free of weed
competition, pointing that in an operational
scale the fertilization will facilitate the weed
control due to the lower light environment below
the Eucalyptus canopy.
micronutrients, in order to figure out which
elements are most limiting to forest growth in
that region. Figure 5 shows the difference
between the control treatment and complete at
6 months of age on the micronutrient. omission
trial.
Figure 5. Omission of micronutrientes at 6 months in Tocantins
– Brazil. Control plot (left) and complete plot (right).
Next Steps
The site will continue to be measured every six
(DBH, Height, LAI) months for the long term
22
Regionwide 25: Nutrient omission at the establishment and early
development of Eucalyptus grandis in Colombia.
Trial series objectives
Figure 1. Location of RW25 trials in Colombia
1– To determine the critical nutrients for establishment
of E. grandis plantations.
2– To evaluate the clonal response of omission of one
or more nutrients at establishment of each site.
3– To model the biological and economical response
of E. grandis clones to the omission of site specific
nutrients and to determine operational strategies
4.– To update site-specific cooperators models use to
prescribe fertilization of new planted clonal materials.
Treatments and experimental design
Three studies were established by SmurfitKappa
Carton de Colombia. Trials were located at three
representative and contrasting soil conditions of the
south, central and north areas of the Cauca Valley in
Colombia, which are extensively planted with E.
grandis. Trials were established in 2012 considering a
spacing of 3x3 m . The north site has 3.900 mm, the
central site 1.550 mm and the southern site 2558 mm
of rainfall (Figure 1).
The experimental design consisted of a pseudo latin
square
considering
genotypes
and
omission
fertilization treatments with true replicates as sites.
Therefore fertilization omission treatments were
randomly assigned to each genotype being
evaluated. Nutrition omission treatments considered
no fertilization (Control), no P (-P), no K (-K), no N (-N),
no micronutrients (No Micros), no Ca & Mg (-Ca&Mg),
and all nutrients applied (Full Fert. Five genotypes
(clones and seed) were selected to be evaluated
across sites. All plantations had excellent pre-planting
and post-planting full weed control applied during
the first 6 months since establishment.
Annual measurements consider total height, DBH and
mortality and soil and foliage nutrient analyses. Soil
chemical properties will be sampled annually to adjust
soil indicators that may be used as predictors of
potential response to applied nutrients. Environmental
measurements will include temperature, rainfall,
relative humidity, vapour pressure deficit and
photosynthetically active radiation from local weather
stations and periodical sensor measurements
as
available.
Figure 2. Response at 24 months for no fertilized
with P (left) and P fertilized plots (right) at
RW254601.
Preliminary Results
Analysis of variance on basal area 12 months since
establishment showed a significant site by fertilization
treatment interaction (p<0.05). Site 254602 and 254603
had the lower and higher basal area (1.6 versus 3.0 m2/
ha), respectively. Phosphorus was the most limiting
nutrient in all the sites (Figure 3), and its effect on
growth was evident since the evaluation at the 6th
month (Figure 2).
23
Figure 3. Mean basal area per site and fertilization treatment. Bars error correspond to the standard
error.
The sites 254601 and 254602 did not present other
deficiency of any other macro– or micro nutrients. At
the site 254603, N was the second largest nutrient
limitation. Because of the greater productivity of this
site, it may present any other nutrient limitation later in
the rotation, which must be assessed. There were
significant differences in basal area among the
genotypes (p<0.05, Figure 4).
The three best
genotypes doubled the basal area of the genotypes
having the lower productivity.
There were no
interaction among the genotypes neither with the sites
conditions nor with the fertilization treatments.
Figure 5. Clonal response to nutrient omission
treatment at RW254601.
Future direction
Stands will be monitored until canopy closure or until
rotation age given the short rotation period (7 years)
at these sites in order to understand the value of
fertilization until rotation age. New trials are
underway to assess the most efficient P rate for
operational purposes.
Figure 4. Mean basal area per genotype. Bars
error correspond to the standard error.
24
Regionwide 25 Omission of nutritional elements effects on the growth of
young plantations of Eucalyptus nitens in the south-central Chile.
Manuel Yaya M. 1 , Rafael Rubilar P. 2 & Manuel Monsalvez J. 2
1
Forest Science School, University of Concepción, Concepción, Chile.
Objectives
.Table 1. Main characteristics of the study sites.
1. To determine the feasibility of maintaining high
leaf area levels and maximum growth rates in
young Eucalyptus plantations into half or full
rotation stages.
2. To evaluated the importance of macro and
micronutrients on the potential response to N
and P additions across major contrasting soils
and climatic conditions at high nutrient loading
rates.
3. To evaluated leaf area, growth efficiency and
foliar nutrient levels for high rates of N and P
fertilization under a gradient of nutrient
omission at different soil-site conditions.
Methods
Study area: Three sites were established for
Eualyptus nitens (RW25) trials in the VIII region of
Chile. Sites mean annual temperature average 13
-17 °C, rain fall range 1000-1400 mm per year and
soils are considered recent volcanic ashes, marine
sediments and red clay old volcanic ashes( Table
1.)
Site
Code
City
Las
Bandurrias
Millogue
255401
255402
254901
San Carlos
Lebu
Mulchen
Latitude
36ª26’22.88”
37ª33’14.08”
37o49’69.02”
Longitude
71ª41’07.38”
73ª37’08.95”
72º17’96.03”
m.a.t. (ªC)
m.a.p.
(mm.year-1)
Soil
14
13
17
1034
1302
1380
Recent ash
volcanic
Sediment
marines
ancient volcanic ash
Results
After the first 3 years of trials installation, the
cumulative volume in recent volcanic ash was
higher than the other sites. After the first
fertilization treatment in 2011, the sites 255401 and
255402 showed no differences in the cumulated
volume until month 33, afterwards the differences
between the two sites increased (62 m3ha-1 at the
end of the measurement period in January 2014).
Trial design: The experimental design consisted of
a completed randomized block design with three
replicates. Treatments considered
omission of
some nutrients, with low (once a year) and high
(twice a year, autumn & spring) nutrient loading
frequencies focused on extreme additions that
may avoid any potential nutrient limitation at each
site. The whole study considered a total of 18
treatment (Table 2).
Currently there have been 4 applications of
fertilizer treatments in trials 255401 and 255402
which were established in 2011 in 1-year-old
plantations. The trial 254901 was installed in 2012 in
a 4-years-old plantation and treatments have
been applied 3 times.
Las
Maicas
Table 2. RW25 treatments
25
Figure 1 Cumulative volume reached between 12 to 47 months
of 255401 and 255402 and stands between 55 to 73 months in
254901. Red line: 255401, Blue Line: 255402 & Green Line
254901
This indicates a better response to fertilization in
the volcanic ash site. In ancient volcanic ash, at
the end of the measurement period (month 73),
the stand reached 144 m3ha-1 (Figure 1)
Figure 3. Relationship between leaf area index (LAI m2m-2)
and increased cumulative volume for the control and full
fertilization treatments between 2012 and 2014. Red line:
255401, Blue Line: 255402 & Green Line 254901
month 43 at the trial 255402 (4 m3ha-1) and on
month 66 at the trial 254901 (6.7 m3ha-1) (Figure
4).
In the trial 255401, the peak of leaf area index was
reached at the age of 39 months, with 8.3 units. In
the trial 255402 the highest LAI was 6.2 at month
36. The trial 254901 peak LAI at age 60 months,
and reached 7.4 units (Figure 2.)
These results indicate that fertilization produces an
increase in leaf area index, and extend the period
the tree keep te foliage, which is reflected in the
volume reached by the plantation.
Figure 2. Leaf area index for control and full fertilization
between February 2012 and January 2014. Red line: 255401,
Blue Line: 255402 & Green Line 254901
Figura 4. Growth efficiency (m3ha-1 / m2m-2) for treatments
control and full fertilization. Red line: 255401, Blue Line: 255402
& Green Line 254901
In the trial 255401 the growth rate achieved by
the contrast treatments was over 10 m3ha-1 per
unit of LAI. At the trial 255402, the fertilized
treatment produced 1.3 m3ha-1 more than the
control treatment. The growth rates at the trial
254901 was 6.22 m3ha-1 in the fertilized and 5.68
m3ha-1 in the control (Figure 3).
The maximum growth efficiency was achieved
on month 33 at the trial 255401 (13 m3ha-1), on
Future Direccion
Identify environmental and nutritional variables
that explain the growth and leaf area index
response by different nutrient loads to establish
levels of efficiency in the use of resources.
26
Graduate Student Research
28
Effects of intensive silviculture on the productivity of Eucalyptus in
Venezuelan Llanos and a probabilistic analysis of its profitability
Omar Carrero, Jose L Stape, Fred Cubbage, Lee Allen, Joe Roise - NCSU
Background
Increasing forest site productivity is a need in
many areas. High land prices combined with
landowners’ desire for increased profitability
have motivated forest managers to apply more
intensive silvicultural treatments in order to
increase forest site productivity. Even if
Eucalyptus plantations’ growth patterns were
known in a particular area, the gap between
attainable and current productivity is unknown.
This information is important to estimate the
expected responses to intensive silviculture and
will help to decide where it is more convenient
to apply silvicultural treatments to maximize
investment.
Understanding
how
intensive
silviculture practices like weed control and
fertilization + weed control affects productivity
will permit managers to select the best
treatments to increase productivity. Objectives Our objectives were: a) using Twin Plots 1) to
estimate the gap between current and
attainable productivity in Eucalyptus plantations
under extra fertilization + weed control, 2) To
estimate the treatment effect on light use and
light use efficiency, and b) using Triplet Plots: 3)
to estimate the gap between current and
attainable productivity in Eucalyptus plantations
under extra fertilization + weed control and extra
weed control. Figure 1. LAI, nutrient content and biomass determination in the
Twin-Plot Pairs across the Company landscape
Materials and Methods
To estimate the gap between current and
attainable productivity, we established 53 pairs
of plots measured for two years (Fig.1). Each pair
consisted of a control plot
(Control) which received the
management
regime
regularly applied to the
stands and a treated plot
(Twin) which received an
intensive silviculture treatment
(fertilization + weed control)
in addition to the operational
management applied to the
control plots. Figure 2. Initial and Stem Biomass increment after 2 years in twin-plot pairs
In those sites where 75% of
the surface was covered by
weeds a third plot (Triplet)
was established. Those plots
received only an intensive
29
weed
control
besides
operational management. the
Results
After two years, stem biomass
growth in control and twin plots
was 12.4 and 14.8 Mg ha-1 yr-1
respectively (Fig.2). This difference
represents a gain of 2.4 Mg ha-1yr
-1
over 2 years. We found
significant statistical differences
for APAR by twin and control
plots, whose averages where
1406 MJ m-2 yr-1 and 1344 MJ m-2
yr-1 respectively, representing 5%
increase (Fig.3). Figure 3. APAR and LUE of the Twin-Plot pairs showing that the increase in growth (Fig. 2)
was mainly due to the increase in LUE and not APAR.
+ 20%
Figure 4. LAI, N concentration in the canopy and Nitrogen Area Index for the Twin-Plot pairs..
The increase in LUE was higher, reaching 21 % or 1.1 g/MJ (Fig.3). The Nutrient Area Index (LAI x N
concentration) was the best predictor of the
measured yield gain (Fig. 4). We found that the
stem biomass increment averaged 13.4, 15.1,
and 17.2 Mg ha-1 yr-1 for the control, weed
control, and fertilization + weed control (Fig.5). Although weed control (triplet plots) could
increase nutrient availability (beside plant water
availability), it seems the amount of nutrients
released was not enough to cover trees’
demand. No statistical differences in APAR were
found, but they were found in LUE between the
twin plots and the others (control and triplets),
which seem to be a consequence of a higher
foliar nutrient concentration.
The results clearly shows that the Twin-Plot design
was very effective to provide a realistic view of
the forest potential growth at the landscape
level,
allowing
the
financial
analysis
incorporating risk.
Acknowledgements
Figure 5. Stem biomass increment for the control, weed
control only and fertilization plus weed control.
To Universidad de Los Andes, NCSU, PavecaDeforsa, FPC, SOFAC, Zobel Grant and SCC.
30
Eucalyptus forest plantation effects on soil organic carbon across Eastern
Brazil
Rachel L. Cook 1 , Dan Binkley 2 , and Jose Luiz Stape 1
1
Introduction
North Carolina State University, Department of Forestry and Environmental Resources;
2 Colorado State University, Department of Ecosystem Science and Sustainability
Sequestering atmospheric carbon in plant
biomass and soil organic matter can improve soil
quality and potentially mitigate climate change.
However, the effects of tropical plantation
forestry on soil carbon, and therefore on soil
quality and long-term sustainability, requires
further examination to determine what factors
drive changes. Because impacts of intensive
forestry practices on soil organic carbon are
highly
variable,
we
require
a
better
understanding of trends to determine the
sustainability of these operations. The objectives
of this research were to 1) determine trends in
soil carbon stocks and changes over time across
three regions of Brazil, and 2) evaluate certain
factors influencing soil carbon sequestration.
This understanding in tropical plantation forestry
can
help
develop
the
environmental
sustainability of intensive forest management.
Methods
In 2010, we collected soils from 306 Eucalyptus
plantations that had been previously sampled in
2001 and the 1980s and1990s across three
regions of Eastern Brazil in order to measure the
change in soil carbon over time and across
various soils and climates (Figure 1). Eucalyptus
plots were sampled from 0-15 cm and 15-30 cm
for soil carbon concentration and bulk density to
determine soil carbon stocks and evaluate the
long-term effects of continuous Eucalyptus
31
management.
Technical field support was
provided by the forestry companies Copener,
Fíbria, and Suzano.
Results
Analysis of long-term, intensively managed
Eucalyptus plantations in Brazil shows regional
differences in soil carbon changes (Figure 2).
Region 1 had an average value of –0.07 Mg C
ha-1 yr-1 (not significantly different from zero),
Region 2 averaged –0.80 Mg C ha-1 yr-1, and
Region 3 average +0.27 Mg C ha-1 yr-1. Bulk
density had some impact on soil carbon stocks.
Multivariate
models
showed
significant
environmental variables that explained soil
carbon stocks and changes to include: initial
carbon content, clay content, rainfall, and
temperature.
Summary
Results from the 300 Eucalyptus sites show that
even with the same tree species and silvicultural
treatments, soil carbon dynamics vary across
regions. However, the slow, but gradual shift
towards carbon accumulation over the last two
decades may indicated more sustainable
silvicultural practices since about the year 2000,
such as minimum cultivation, improved
fertilization, and elimination of burning practices.
The multivariate model showed variable
relationships depending on the time period, but
the most important factors causing shifts in soil
carbon relate to initial carbon content, clay
content, rainfall, and temperature. Continued
and consistent monitoring of soil quality with soil
organic carbon will help us assess the
sustainability of intensive Eucalyptus production
across a broad geographic gradient.
For more detailed information,
dissertation can be accessed at:
the
full
http://repository.lib.ncsu.edu/ir/
bitstream/1840.16/7759/1/etd.pdf
Acknowledgements
Support for this work was provided by the Forest
Productivity Coop, the Bruce and Barbara Zobel
Travel Grant, and the Brazilian companies:
Suzano, Copener, and Fibria. We also owe a
great deal of thanks to for cooperation from
ESALQ/USP and IPEF in Brazil.
32
Soil carbon dynamics following reforestation of tropical pastures in Brazil
1
Rachel L. Cook 1 , Dan Binkley 2 , Jose Luiz Stape 1 , and Joao Carlos T. Mendez 3
North Carolina State University, Department of Forestry and Environmental Resources;
2 Colorado State University, Department of Ecosystem Science and Sustainability;
3 University of Sao Paulo, ESALQ, Brazil
Introduction
Sequestering atmospheric carbon in plant
biomass and soil organic matter can both
improve soil quality and potentially mitigate
climate change.
However, the effects of
tropical plantation forestry on soil carbon
dynamics, and therefore on soil quality and long
-term sustainability, requires further examination
to determine what factors drive changes.
Because impacts of intensive forestry practices
on soil organic carbon are highly variable, we
require a better understanding of trends to
determine the sustainability of these operations.
The objectives of this research are 1) to
determine the tree species effects on soil carbon
from reforestation of cattle pasture to Conifer or
Broadleaf (primarily pine and Eucalyptus) forest
plantations, and 2) quantify the turnover and soil
carbon dynamics following reforestation.
Methods
We collected soils in 2009 at the Anhembi
Research Station (Figure 1), in the state of Sao
Paulo, Brazil, across 622 ha of reforested pasture.
Transects were sampled across 21 paired plots (1
Reference-Pasture, 10 Pasture-Broadleaf, and 10
Broadleaf-Conifer pairs) to see the effect of
reforestation on soil organic carbon due to tree
species.
The plots were organized as a
chronosequence, substituting space for time
with stand ages of 6-34 years. By examining a
sequence of increasing stand ages, we can
eliminate the need to wait thirty years for
resampling by substituting space for time. Soils
were sampled at the O-horizon and mineral soil
layers of 0-15 cm, 15-30 cm, and 30-45 cm. To
examine soil carbon turnover rates between
vegetation types, we analyzed stable carbon
isotopic signatures to differentiate residual
pasture carbon (C4 plants with higher 13C
accumulation) from new forest-derived
carbon (C3 plants with lower 13C
accumulation).
Results
Analyses of total soil carbon (O-horizon plus
mineral soil to 45 cm) from the Anhembi
Research Station showed an average (±
standard error) of 36.0 ± 1.7 Mg C ha-1 for
pasture paired with 36.8 ± 1.9 Mg C ha-1 for
Broadleaves, and 38.3 ± 1.9 Mg C ha-1 for
Broadleaves paired with 36.0 ± 1.6 Mg C ha1 for Conifers (averaged across all ages of
stands). Soil carbon was not statistically
different between Pasture-Broadleaf paired
plots, but Broadleaves had 6% higher soil
carbon than Conifers (Figure 3) in those
33
pairs. The reduction in soil carbon
from the Reference forest to pasture is
likely due both decomposition and
erosion following original land-use
change.
To estimate annualized soil carbon
change after reforestation, we
divided mean soil carbon by mean
age of Eucalyptus (22.5 yrs) or Pine
(29 yrs) and can be 95% confident
that the change in total soil carbon
falls in the interval of –0.13 to 0.36 Mg
C ha-1 yr-1 for conversion to Broadleaf
plantations, and –0.26 to 0.50 Mg C
ha-1 yr-1 for conversion to Conifers.
While there was no significant soil
carbon accumulation over time
based on the chronosequence
approach, stable carbon isotopes
showed the loss of forest-derived C3
soil carbon decreased with depth.
Particularly interesting was that forestderived carbon gains were offset
exactly with pasture-derived soil
carbon losses in the 0-15 cm layer.
Results showed a significant decrease
in pasture-derived soil carbon in each
soil layer, but slower decomposition as depth
increased (Figure 4).
Summary
Results indicate that tree species of forest
plantations have little effect on soil carbon
stocks up to 34 years following reforestation in
Southeastern Brazil.
Stable carbon isotope
results illustrate slower rates of pasture derived
soil carbon decomposition with increasing depth
and that accumulation of forest-derived soil
carbon is offset by pasture-derived soil carbon
losses. These results improve
our understanding of soil
carbon dynamics following
land-use change.
For
more
detailed
information,
the
full
dissertation
can
be
accessed
at:
http://
repository.lib.ncsu.edu/ir/
bitstream/1840.16/7759/1/
etd.pdf
Acknowledgements
Support for this work was
provided by the Forest
Productivity Coop. We also
owe a great deal of thanks
to for cooperation from
ESALQ and IPEF in Brazil.
34
Effects of Eucalyptus camaldulensis, E. globulus, E. nitens and Acacia
melanoxylon short-rotation forest crops (SRFC) for energy production in
contrasting soils of the Bío-Bío Region, Chile
Edwin Esquivel, Rafael Rubilar, Simón Sandoval, Eduardo Acuña, Jorge Cancino,
Miguel Espinosa y Fernando Muñoz .
Abstract
With the aim of reducing greenhouse gases
emissions, demand for sustainable raw material
in power generation using tree biomass has
increased. However, sustainability of these SRFC
in marginal soils has been questioned due to the
reduction of time between harvests. This project
evaluated Eucalyptus camaldulensis, E. nitens
and Acacia melanoxylon SRFC at three
stockings effect on nitrogen, phosphorus,
potassium and boron use related to biomass
production after four years (harvesting age) at
two contrasting soil conditions. Biomass
production and nutrient use showed major
differences among species and between sites
but little effect from stocking. These responses
contributed to the project´s objective of
obtaining a sustainable protocol for bioenergy
production.
Introduction
Forest biomass is emerging as an alternative for
low emission and renewable energy source
(Richardson and Verwijst, 2007). However,
sustainability of biomass production depends on
its silviculture (Richardson, 2006) and major
limitations for biomass production depend on soil
sustainability
(Mead,
2005),
particularly
considering that SRFC are expected to be
grown on forest or abandoned agricultural land
in order to not compete with food or other
productive land alternatives.
On the other hand, use of SRFC on forest land
can increase nutrient removals when compared
with long-term rotations of traditional timber
production (Stupak et al., 2007). Therefore, the
sustainability of wood energy production may
requires a sustainable nutrient supply (Kauter et
al., 2003). This will also allow for greater above
and below ground carbon sequestration as an
additional objective of SRFC .
Objectives
Evaluate the effect of short rotation species at
contrasting planting densities, in nutrient use per
unit of biomass produced.
Methodology
Experimental Desing: Randomized complete
b
l
o
c
k
(
n
=
3
)
.
Sampling: Soils 2, 4, 11, 16, 23, 28, 36, 40 and 48
months at 0-20 and 20-40 cm depth. Biomass at
11,
23,
36
and
48
months.
Nutritional
and
lab
analyses:
Soils: Available N (N-NO3, N-NH4), P, K and B,
T o t a l
C a r b o n
a n d
p H .
Biomass tissues : Stem, branch and foliage
biomass and N, P, K y B concentrations.
Results
Table 1. Species and
evaluated at each site .
Site
Species
Stockings
trees ha-1
stocking
treatments
Granitic
Sandy
Soils Xeralfs
Soils Psamments
A. melanoxylon,
A. melanoxylon,
E. nitens and E.
camaldulensis
E. nitens and
E. globulus
5.000, 7.500 y 10000
The analyses showed an increase in N availability
for N fixing species at sandy soils (Acacia) at 48
months and interactions with other nutrients.
Effects differed between sites where at granitic
soils ammonium and potassium showed
increased levels at 0-20 cm depth; at sandy soil
nitrate and phosphorus presented higher levels
in concentration at both depths (0-20 20-40 cm).
35
Table 2. P values of analysis of effect of species
(Sp), densities (Den) and interaction (Sp*Den)
evaluated in each site (granitic and sandy) for the
age 48 months.
Site
Depth
20-40 0-20 20-40
0.076
0.005
0.367 0.694 0.375
0.691 0.069 0.407
Sp*Den 0.903
0.701 0.815 0.283
Sp
0.290
0.566 0.702 0.039
Den
0.796
0.405 0.205 0.029
Sp*Den 0.491
0.697 0.823 0.405
Sp
0.773
0.732 0.562 0.609
0.170
0.207 0.007 0.547
Sp*Den 0.561
0.416 0.790 0.367
Sp
0.730 0.586 0.321
Phosphorus Den
Potassium
Boron
Sandy
0-20
Sp
Ammonium Den
Nitrate
Granitic
Den
0.524
Figure 1. Relationship between total nitrogen (kg
ha-1) and total biomass (Mg ha-1). Granitic site.
Figura 2. Relationship between total nitrogen (kg
ha-1) and total biomass (Mg ha-1). Sandy site
<0.001 0.123 0.239 0.891
Sp*Den 0.803
0.472 0.792 0.748
Sp
0.344
0.178 0.531 0.706
Den
0.170
0.578 0.732 0.109
Sp*Den 0.869
0.601 0.843 0.794
Boron did not showed any effects at this age
(Table 2 ).
Most nutrients showed decreased levels after 24
months for all species and stockings suggesting
fast nutrient depletion. However fertilization with
NPKB at 30 months of age allowed soil nutritional
levels to return to their initial status.
Figure 1 and 2 show the relationship between
biomass and aboveground N use. Independent
from site, species showed different nitrogen use
efficiencies and differences in aboveground
biomass production. These results suggest that
A. melanoxylon shows a higher nutrient use
efficiency for N but presented a poor biomass
production at sandy soil sites compared to
granitics. Independent of site E. nitens has
shown interesting results, that suggests a high
potential for this species as a biomass
production alternative.
Conclusions
Differences in nitrogen use among species may
affect sustainability of biomass production
particularly for species with larger biomass
accumulation but less efficiency on nitrogen use
per unit of biomass.
Fertilization is required to maintain soil nutritional
levels and therefore sustained productivity at
these marginal sites.
Acknowledgement
This research was funded by Chilean
government INNOVA BIO BIO Project No. 06-PC
S1-33 and Energía Verde and
MASISA S.A
companies support.
36
Using 3PG to model E. camaldulensis biomass potential for bioenergy
plantations in Chile
Edwin Esquivel, Rafael Rubilar, Simón Sandoval, Eduardo Acuña, Jorge Cancino,
ABSTRACT
Ecophysiological models such as 3PG predict the
productivity of forest plantations using
environmental
parameters
and
species
ecophysiological characteristics. The 3PG model
has been used several times on estimations for
Eucalyptus traditional plantations. There are few
experiences using this model to evaluate
plantations for bioenergy production at high
stockings, and parameterization estimates for
traditional plantations have been not applied to
to estimate productivity from plantations to sitespecific level. To evaluate the prediction
capacity on such plantations, a combination of
3PG parameters from the literature and local
information was used to evaluate Eucalyptus
camaldulensis biomass potential estimates for
bioenergy production using 3PG in the southcentral valley of Chile. Plantations
were
established at high stockings of 5000, 7500 and
10000 trees per hectare at a marginal site for
traditional forest production in granitic soils of the
Bio Bio Region, Chile. Our results suggest that 3PG
successfully estimated biomass production at
7500 and 10000 tph stockings, but failed at 5000
tph stockings due to competing vegetation
caused by poor weed control efficiency but the
model allowed for more real estimates under
better intensive silviculture. Specific allometric
equations calibrations at specific stockings are
key to obtain appropriate biomass estimates.
Our results suggest the need for incorporating a
competing vegetation factor parameter to
calibrate operational 3PG estimations.
INTRODUCTION
Forest biomass can be considered a sustainable
alternative for clean energy production
(Richardson and Verwijst 2007). A source of
clean energy such as biomass from dedicated
energy plantations can both fix carbon as well as
become an important energy source (Berndes et
al., 2003, Sartori et al., 2006). Short rotation crops
for bioenergy may consider 3 to 5 years and a
commercial rotation up to 30 years (Aylott et al.,
2008). High levels of variation on yield depend on
site factors and silvicultural practices, especially
for stands established on marginal sites
(Richardson 2006). Under these conditions, the
benefits that wood energy plantations may have
are: protection of soil erosion (Feikema et al
2010) and organic matter acculation at
impoverished sites.
The aim of this research was to parametrize
and evaluate weaknesses and strengths of 3PG
process based model as an effective tool for
biomass estimation of Eucalyptus camaldulensis
bioenergy plantations established at high
stockings on marginal soils of the Biobío Region,
Chile.
METHODOLOGY
The field study was established as a randomized
complete block experimental design with 3
replicates (n=3). Plantation was fertlized at
establishment and weed control was applied 1
year after planting. Individual tree measurements
of H and DBH were obtained at 2, 4, 11, 16, 23,
28, 36, 40 and 48 months of stand development
in order to evaluate seasonal and climatic
effects on tree growth. Biomass sampling was
estimated at 11, 23, 36 and 48 months, and
nutritional analyses were obtained at each
instance for major biomass components
considering stem, branch and foliage biomass
for N, P, K y B concentrations. 3PG model
parametrization was calibrated with local
biomass equations for each stocking level, and
ecophysiological parameters were obtained
from the literatura. Climatic data was obtained
from the nearest weather station.
RESULTS
Our results suggest that 3PG successfully
estimated biomass production at 7500 and
10000 tph stockings, but failed at 5000 tph. The
model in all cases estimated accurately
plantation survival (data not shown) and major
differences in productivity, at each stocking
level, were explained by individual tree growth.
Unfortunately weed control did not effectively
limited competing vegetation growth at the
lowest stocking and caused a major reduction
Figure 1. 3PG estimates for Eucalyptus camaldulensis stem biomass (Mg ha-1) and DBH (cm) at 5.000 (A),
37
7.500 (B) y 10.000 (C) trees ha-1. Grey squares represent actual biomass estimates at each stcoking level.
Age of the stand (years)
in individual tree growth explaining differences
in model prediction.
As shown in Figure 1, specific allometric
equations calibrations at specific stockings
were key to obtain appropriate stand biomass
estimates and individual tree size parameters
such as diameter and height for each stocking
level. The model estimates resulted in highly
accurate estimates for 10000 trees ha-1 and
had a minor overestimation for the 7500 ha-1.
stocking.
CONCLUSIONS
3PG successfully estimates biomass production
for short rotation bioenergy plantations of E.
camaldulensis at high stockings.
Age of the stand (years)
Accurate biomass estimates are only possible
with 3PG under no competing vegetation
effects.
Stocking
specific
biomass
allometric
relationships are key to obtain accurate
estimates from 3PG.
ACKNOWLEDGMENT S
Funding for this project was provided by
Chilean government INNOVA BIO BIO Agency
Project No. 06-PC S1-33 and Energía Verde and
MASISA S.A companies. the Forest Productivity
Cooperative, the Soil, Nutrition and Forest
Productivity Laboratory at the Faculty of Forest
Sciences at Universidad de Concepción,
38
Using forest inventory as a tool of monitoring silvicultural quality in
Eucalyptus clonal plantations
Rodrigo Eiji Hakamada, José Luiz Stape, Cristiane Camargo Zani de Lemos
Introduction
Forest productivity is defined by the environment
the population is inserted, the implementation of
genetically superior material and the correct
silvicultural management applied to this
material. The correct prescription of technical
recommendations and the execution of those
operations according to acceptable quality
standard are intended to eliminate or minimized
the growth constraints. Despite the silvicultural
improvements that have occurred in recent
decades, it is still possible to obtain gains in
productivity through monitoring, early detection
and correction of deviations in the silviculture.
The objective of this study was to investigate an
index for monitoring the silvicultural quality using
forest inventory networks.
The last step, the indexes and the concept of
ORU were applied on a commercial scale in a
network of forest inventory plots installed in
12.000 hectares of Eucalyptus clonal plantations
in northeastern of São Paulo state, with about 2
years old, and planted within 1995 to 2009.
Results
In the definition phase, the index that represents
the percentage of the total volume of 50% of
smaller trees planted
(which includes the
planting holes) (PV50) was shown as the best
adapted to the purpose of this study because it
has finite limits (50% to 0%), includes the holes of
planting and indirectly represents the distribution
Material and Methods
There were three stages of work: i) Definition of
uniformity indexes of dendrometric variables; ii)
Standardization of so-called optimal range of
uniformity (ORU) and validation of the
methodology through its application in a
network of forest inventory for a single
commercial clone.
In the first stage we used three tests of the
network BEPP (Brazil Eucalyptus Potential
Productivity) with different levels of productivity
to establish the appropriate indexes to
characterize the silvicultural uniformity (Figure 1).
Figure 2: Uniformity at the end of rotation (6 years) (PV50 – 6
years) related with initial uniformity at 5, 9, 12 e 24 months..
The second stage used five clonal test of
Eucalyptus in Sao Paulo state to validate the
concept of Optimum Range of Uniformity.
Figure 1: (left) Uniform treatment at left and heterogeneous
treatment at right and (b) aerial view of BEPP at IP site at 2,5
years (Stape et al., 2010)
Figure 3: – Productivity at the end of rotation (6 years) (MAI – 6
years) related with initial uniformity at 5, 9, 12 e 24 months
39
Figure 5: Relative increment (%) on productivity (MAI)
and uniformity (PV50) based on the year 1
Figure 4 – Productivity (MAI) (up) and uniformity (PV50)
at 2 years at the five clonal tests
of classes of growth. The uniformity at 5, 9, 12
and 24 months were strongly correlated to
uniformity to 6 years (R² > 0.74) showing the
possibility of early monitoring for detection of
quality deviations in forestry (Fig. 2). Furthermore,
the initial rate PV50 was highly correlated with
the final yield (p <0.001) (Fig. 3).
At the stage of standardization, there was no
statistical difference (Tukey, 5%) of PV50 among
clonal tests despite their different yields, showing
that the uniformity index can be generalized,
irrespective of the productivity of the site. The
ORU of PV50 was from 37 to 50%, i.e., sample
plots which have the PV50 within this range can
be considered satisfactory "uniform" (Fig. 4).
In the validation phase, when the concept was
applied on a commercial scale there was a
strong temporal evolution of PV50. In the
plantations made in 1995 the average of PV50
was 29% and increased to 42% in 2009 (Figure 5).
The percentage of plots within the optimal
range of uniformity clearly tended to rise over
this period and could be related to major
improvements in forestry operations and their
monitoring through quality control.
Conclusion
Uniformity was confirmed as a fundamental
variable to monitor operational quality in
Eucalyptus clonal plantation
Uniformity Index (PV50) and the Optimum
Range of Uniformity (ORU) permitted to
capture the evolution of silvicultural quality in
a large scale, being possible to indicate it to
monitor silvicultural quality.
References
Stape et al. 2010. The Brazil Eucalyptus potential
productivity project: influence of water, nutrients
and stand uniformity on
wood production.
Forest Ecology and Management, Amsterdam,
259: 1686–1694.
Table 1. Evolution of MAI and PV50 from 1995 to 2009. Results are shown in average, standard deviation,
maximum, minimum, amplitude and confidence interval.
40
Growth, water use and water use efficiency of Eucalyptus clones under different spacings: from the tree to the watershed scales
Rodrigo Hakamada, Silvio Ferraz, Robert Hubbard, Jose Stape, Cristiane Lemos, Adriano Almeida
Introduction
The definition of planting spacing is a strategic
decision for the establishment of any forest. It directly
affects its productivity, competition for resources
among individuals and the final wood products.
Many studies in planted forests have been done to
define the spacing with the main focus on biomass.
However, there are few studies on the impact of
plant density in relation to water use, water use
efficiency and the interaction among genetics and
spacings (WHITE et al., 2004). Thus, the overall goal of
the study is to determine the effect of spacing on
biomass production, water use and water use
efficiency in Eucalyptus clonal plantation in two
scales: tree and watershed. With this study, we would
like to answer the questions: i) does the interaction
between spacing and genetics have a relationship
with water use and water use efficiency? Why? ii)
Can the effects of water use affect water yield at
watershed scale?
To estimate water use, homemade sap flow probes
were installed in August 2013 at eight trees per
group of spacing in four clones with different
canopy structure and growth (Figure 2) and will be
measured for two years. Precipitation, stemflow and
canopy interception has been measured since
September 2013.
To explain differences in terms of water use (WU)
and water use efficiency (WUE) measurements of
dendrometers (monthly), Leaf Area Index (LAI), leaf
water potencial, soil humidity and fine roots depth
are planned to be done.
Material and Methods
The project encompasses two stages. The first assay
is using a systematic planting density trial of the
TECHS network (Eucalyptus Tolerance of Thermal
and Hydric Stresses) located in Sao Paulo state at
southeast region in Brazil, planted in February 2012.
The objective of this step is to determine water use
and water use efficiency at tree scale (Figure 1).
Measured trees are covering stockings used for the
main wood products and range from 600 to 3,000
trees per hectare.
Figure 2 – Measured clones of Eucalyptus grandis x
E.urophylla (Cenibra, IP22 and GG100) and one clone
of E.grandis x E.camaldulensis (Cop1277)
Results of water balance at tree scale will be
modeled at watershed scale using a watershed 20
kilometers far from the TECHS site which water yield
has been monitored for 10 years. The SWAT
hydrological model will be used (Soil and Water
Assessment Tool - swat.tamu.edu) to spatialize
results.
Current Status
Figure 1 – Layout of systematic spacing design with
the eight spacings (four groups) with water balance
measurements and the main use of wood
International Paper in Brazil has supported this
research through a partnership among the
company and University of São Paulo, USDA Forest
Service and North Carolina State University.
Processed data of twelve months old are already
available and also is the sapflow of one month
measurement (but not validated). Measurements of
wood growth have been done every 6 months and
dendrometer measurements will be monthly.
41
Initial results
At twelve months all clones presented the tendency
of having more wood from wider to tighter spacing
(Figure 3), but with a high variation between clones of
more than 60%. The data of canopy depth showed
the opposite, with more trees per hectare showing a
shallow canopy (Figure 4).
Water use followed the same behavior of wood
growth, with more transpiration with more trees per
hectare. Despite being an initial analysis, the same
water consumption in clones CNB10 and COP1277
can represents a significant difference in terms of
WUE, once Cenibra clone growth 66% more than
COP1277 at 3,000 trees per hectare. It might be an
alternative of management to reduce water use
at watershed scale. Measuring wood growth, LAI
and WU will permit us to have more detailed WUE
in a large range of spacing and clones.
Figure 3 – Wood growth at twelve months at the eight
spacings. Dotted line represents the average of four
clones
Figure 5 – Relative transpiration to 1.500 trees per
hectare (%) of three clones and four groups of
stockings.
Future plans
Figure 4 – Average canopy depth of four clones
These results evidence the beginning of competition
at tightest spacing. Loss of leaves might be
occurring because of light attenuation or because
of water deficit, activating petiole abscission (RIOV,
GOREN, 1979).
We plan to continue measurements of wood growth
every six months and dendrometers monthly. Leaf
water potential and relative water content will be
measured in October 2013, April and October 2014
and April 2015. Soil water content and fine roots
density will be collected in October 2014 and April
2015. Maybe a measurement of non-structural carbohydrate analysis could be done depending on
the initial results of transpiration. Measurements of air
humidity and wind speed to calculate boundary
42
Growth and yield model for Eucalyptus benthamii in the SE United States
Kevin B. Hall, Jose Luiz Stape
Background
The demand for short-rotation hardwood
plantations for pulp and paper as well as the
potential of a biomass market has increased
interests in a cold tolerant Eucalyptus species
suitable for fluctuating winter weather. This has
resulted in eucalypt trials being installed across
the southeastern United States to examine the
cold tolerance and biomass potential.
The FPC cold tolerance eucalypt screening trial
was established with 150 species. After three
years, seven species showed sufficient cold
tolerance through 2010, 2011 and 2012 winters
with E. benthamii being the best performer.
Permanent plots were subsequently installed
across the southeastern states in a range of
stand sizes, from small research plots to
commercial plantations, to be used in
conjunction with the RW24 trials to access the
actual productivity of E. benthamii.
With the support of the IBSS grant and the FPC, a
network of Eucalyptus benthamii permanent
inventory plots have been installed across the
southeast. Standard inventory measurements,
leaf area index estimation, and foliage and soil
samples were taken to provide yield and
nutrition information for each plot.
Growth and yield modeling is necessary for
foresters to make management decisions and
economic evaluations.
Figure 2. FPC RW24 trial in Louisiana.
Site Description
The objective of the permanent inventory plot
network is to encompass all growing conditions
to provide a realistic view of production.
Eucalyptus benthamii plots within FPC RW24
study were used as a foundation to the
permanent inventory plot network. Additional
permanent inventory plots were installed on a
range of sites.
Measurements & Sampling
Twenty-eight permanent plots have been
installed ranging from 113 m2 to 314 m2 pending
on parcel size. Diameter at breast height (DBH)
was measured for all trees, and total height was
measured for the individuals with the four largest
DBH and the first six individuals. Soil samples
were taken between trees within the row and
also between rows. Wood samples has been
collected to estimate wood biomass for the
plots. Foliage samples were taken from four trees
using FPC foliage sampling protocol. When
applicable, leaf area index was estimated using
the FPC LAI Handbook. Table 1 provides a
summary of stand variables used to develop
preliminary site index guide curve and growth
and yield model.
Preliminary Results
Figure 1. FPC/IBSS Inventory Plots of E.benthamii
Dominant height and age were used to develop
initial site index guide curve. Using a base age
43
of six years, Site Index6 was predicted for each
permanent plot to provide an understanding of
site quality. For initial site index guide curve, the
classic relationship of the natural logarithm of
dominant height was regressed against the
inverse of age (Avery and Burkhart, 2002).
An initial growth and yield model was derived
using age (months), basal area (per hectare)
and site index (meters) at base age 72 months
Table 1. Summary of stocking, basal area, site index and total
volume by age.
Standard error of the mean shown in
parenthesis.
Figure 4. 3-year old E. benthamii in Alabama with SI6=18 m.
hectare, a yield table was generated from the
preliminary growth model (Table 2).
Next Steps
To improve site index guide curve by
investigating polymorphic guide curves and to
develop the G&Y models for both volume and
biomass.
Figure 3. Scatter plot of dominant height (meters) for each
sample plot with site index curves with a base age of 6 years.
Work Cited
Avery, Thomas E. and Harold Burkhart.
Measurements. 2002.
Forest
Using a basal area of 20 square meters per
Dougherty, Derek and Jeff Wright.
2012.
“Silviculture and economic evaluation of
eucalypt plantations in the southern US.”
BioResources 7(2): 1994-2001.
Table 2. Preliminary yield table for E. benthamii (m3 ha-1) for SE
US, stocking based on 20 sq. m per hectare.
Acknowledgements
to predict total volume (m3 ha-1). The model was
developed using approximately 71 plots.
This project is supported by the IBSS project and
the FPC and its cooperators. The IBSS project is
supported by the Agriculture and Food
Research Initiative Competitive Grant no. 201168005-30410 from the USDA National Institute of
Food and Agriculture.
44
The radiation-temperature effect on cold tolerant Eucalyptus species
Kevin B. Hall, Jose L. Stape, Thomas R. Fox, Rafael Rubilar, Timothy J. Albaugh
Introduction
Eucalyptus genus has more than 800 species,
which occur in their native range in AustraliaIndonesia from equator to 40º South. In the SE US,
many Eucalyptus initiatives are trying to identify
cold-hardy species that can be used for fiber,
biomass, biofuels or mulch production, with the
ability to coppice after harvesting.
To fully investigate the existence of potential
new species and how their growth rate are
affected by the radiation-temperature, a
screening and detailed measurement trial were
install in a typical Piedmont area of North
Carolina.
The objective of this study is to examine potential
differences in cold tolerant eucalypts rate of
biomass accumulation through the winter
months. Furthermore, this study aims to develop
a robust relationship between average weekly
biomass gain by species and the weekly
average minimum temperature.
Materials & Methods
A frost tolerance eucalypt screening trial was
installed in Raleigh, NC in 2010 consisting of 150
different species on a clayey alfisol soil (fertile).
The first winter was a mild winter and half of the
species were unable to withstand the minimum
temperatures. The 2012 winter proved to be a
harsher winter an another 50 species
experienced severe canopy damage. Of the
remaining species, seven shown to be not only
frost tolerant but had high growth rates (Table 1).
Biomass accumulation for eucalypts can be
measured weekly because their shorter lag time
following the fixation of carbon through
photosynthesis. Unlike pines that store fixed
carbon as carbohydrates during the winter
months, Eucalyptus species immediately use a
big portion of the fixed carbon for growth when
climate factors allow.
Parallel studies within the project include the
assessment of specific leaf area and specific
gravity by age and species. These additional
studies increase the understanding of canopy
biomass and development, as well as provide a
Figure 1. 2.5 years-old E. benthamii (center) and E. viminalis
(right) during snow event February 16th, 2013 in Raleigh, NC.
better understanding of the dynamics of specific
gravity through the duration of the expected
rotation of eucalypts for biomass production in
the SE US. Also, using maximum daily radiation
data and canopy structure, we can begin to
examine the amount of photosynthetically
active radiation that is being absorbed by each
sample tree.
Beginning in October 2012, diameter at breast
height (DBH) was measured and recorded
weekly, using a diameter tape to the 1/100th of
an inch to closely examine changes in DBH each
week. Total height and height to canopy where
measured monthly using a telescopic height
pole. Between height measurements, a linear
Table 1. Average weekly biomass growth by species.
45
Figure 2. Leaf sampling protocol for specific leaf area of cold
tolerant Eucalyptus.
Figure 4. Average weekly stem biomass growth (grams/tree/
week) for seven Eucalyptus species in Raleigh, NC.
Preliminary Results
Figure 3. (left) 8mm increment borer. (right) E. benthamii cores.
relationship was used to estimate change in
height on a weekly basis.
Wood samples were collected using an
increment borer with an 8 mm diameter to
estimate bulk density at breast height for all
species of interest. Sample trees at the Raleigh
Trial consisted of both two– and three-year old
trees.
This methodology was also used in
Ravenel, SC to sample trees with nine– and 13years old to capture change in specific gravity
at breast height for frost tolerance eucalypts
across a range of age classes.
Using the FPC foliage sampling protocol (Figure
3), foliage samples were gathered for all sample
trees to determine average specific leaf area for
each species of interest. Laboratory exercises
include scanning sample leaves to determine
leaf area, drying leaves to determine dry
biomass, as well as grinding foliage for future
nutrient analysis.
All of these works have been implemented with
the intention of increasing the understanding the
limitations of growth for frost tolerant Eucalyptus
species as well as to characterize the wood
properties and biomass allocation.
Average weekly stem biomass accumulation for
each of the seven species of interest can be
seen in Figure 4. The graph shows that frost
tolerant Eucalyptus species will continue to add
stem biomass through the winter months as long
as the weekly average minimum temperature
allows. This gives eucalypts an advantage over
pine species and native hardwoods that store
fixed carbon or remain dormant, respectively,
during winter months. Figure 4 also shows that
during summer months there are climate factors
that are limiting the growth of frost tolerant
eucalypts. Bulk density sampling showed that
there is a significant increase in the specific
gravity of wood at breast height over time. This
suggest that rotation age for frost tolerant
eucalypts should be dependent on the desired
product. For example, forest managers growing
Eucalyptus primarily for biofuel or biomass should
consider extending the rotation from six-years to
eight-years in order to capture the gain in
specific gravity as well as the additional growth
to increase the biomass yield.Next Steps
Radiation-temperature modulators will be
investigated for the 52 weeks of data for the
seven species. We expected that this
information will help on the Cold-hardy
Eucalyptus zoning potential for the SE US.
Acknowledgements
IBSS project (Agriculture and Food Research
Initiative Competitive Grant no. 2011-6800530410 from the USDA National Institute of Food
and Agriculture) has been supporting this study.
46
Individual stem volume, green weight and biomass equations of cold tolerant Eucalypts
Kevin B. Hall, Jose Luiz Stape
Background
A strong research initiative has been focused on
establishing cold tolerant Eucalyptus plantations
across the SEUS in order to expand the region of
potential sites. Freezing winter temperatures
have proven to be the primary climatic variable
that limits the range eucalypts in North America.
As a result, an emphasis has been placed on
frost tolerant Eucalyptus species. The FPC cold
tolerant eucalypt screening trial was established
with 150 species. After three years, six species of
the section Maidenaria have shown sufficient
cold tolerance and fast growth characteristics.
It is necessary to be able to predict individual
stem volume and biomass for the principal
species of interest to accurately assess growth
and yield. Detailed individual stem volume and
biomass equations are necessary when
estimating total yield at the plot or stand level.
With the support of the IBSS grant and the FPC,
forty trees were sampled to determine the stem
volume, green weight and dry weight. The
objectives of this research are to 1) develop
widely applicable total and merchantable
individual stem volume, green weight and
biomass outside– and inside-bark allometric
equations for frost tolerant eucalypts and 2)
quantify the biomass partitioning of E.benthamii
for
both
above–
and
belowground
compartments.
Figure 1. E.benthamii 1-m sections in S.Carolina.
inside-bark was recorded for each disc as well
as the green weight. Discs and bark were dried
to a constant weight to determine moisture
content and overall dry weight of each bolt.
Smalian’s formula was used to estimate the
volume for each bolt and the volume equation
of a cylinder and a cone were used to estimate
the stump and top-section respectively. The
summation of each bolt yields to total stem
volume, green weight and dry weight for each
stem (Table 1).
Table 1. Summary statistics for total stem volume, green weight
and dry weight of frost tolerant Eucalyptus.
Materials & Methods
Study trees ranged from two to nine years old to
capture size classes through the anticipated
rotation age (six years).
Forty trees were
sampled
among
six
species
including
E.badjensis,
E.benthamii,
E.dorrigoensis,
E.macarthurii, E.nitens and E.viminalis. Samples
trees were selected based on the diameter
distribution of a temporary plot at each site to
adequately represent the size classes.
Following felling, dbh, total height and canopy
height were measured. Branches were removed
from the stem, and the stem was separated into
one-meter bolts to a 2.5-cm top. Each bolt was
weighed, and a disc was collected from the
base of each bolt. The diameter outside– and
Variables
Mean
St. error
Min.
Max.
dbh (cm)
13.9
0.98
4.0
32.3
Height (m)
11.8
0.84
6.6
25.5
Bark
2.3
0.23
0.4
7.6
Outside-bark
0.1190
0.0248
0.0051
0.821
Inside-bark
0.0992
0.0202
0.0045
0.666
Outside-bark
113.8
26.2
3.0
852.0
Inside-bark
98.9
22.1
2.6
701.3
Outside-bark
53.2
11.3
2.1
349.9
Inside-bark
46.3
10.0
1.8
310.9
Per tree basis
thickness*
Volume (m3)
Green weight (kg)
Biomass (kg)
47
Six nine-year old E.benthamii trees were selected
to determine the allocation of biomass.
Following the removal of branches, all leaves
were removed and both branches and foliage
weighed separately in the field to determine
total green weight for each compartment. A
one-kilogram sample of each compartment was
dried to determine the moisture content and the
dry weight of each compartment.
A one by one-meter wooden frame was
centered around each E.benthamii stump to
define the belowground sampling area. All soil
and lateral roots were excavated in 20centimeter depth increments.
Roots were
collected to a depth of one-meter and the tap
root was removed. All additional roots from the
tap root beyond one-meter were collected and
labeled to the corresponding sample depth. All
roots were dried at 60 degrees Celsius to a
constant bone dry weight. The total biomass of
the six E.benthamii sample trees can be
estimated using the dry weight estimations of
each compartment (canopy, branch, stem
wood, stem bark, and coarse roots) (Table 2).
Table 2. Descriptive statistics of nine-year old E.benthamii
biomass allocation (dry weight, kg).
Mea
n
8.4
Standard
deviation
7.1
6
30.2
37.0
2.9
98.6
6
162.2
91.7
40.0
310.9
Stem bark
6
21.1
12.0
6.5
39.0
Coarse roots
6
31.0
18.8
7.7
62.3
Compartments
Canopy
n
Branches
Stem wood
6
Min.
Max.
1.6
21.6
Two allometric equations were used to test for
bestness of fit of stem volume, green weight and
dry weight: combined-variable (d2h) (Avery &
Burkhart, 2002) and logarithmic (Schumacher &
Hall, 1933) equation forms. Prior to accepting
the logarithmic equation form, and generalized
logarithmic equation for was examined to
determine the significance of the y-intercept
which was found to not be significant at the
alpha level 0.05.
The logarithmic equation form using dbh as the
only independent variable was used to estimate
dry weight of each compartment.
Preliminary Results
The logarithmic equation form out performed
the estimation of stem volume, green weight
and dry weight compared to the combinedvariable equation form based on the coefficient
of determination and residual plots (Figure 2).
Furthermore, The combined-variable y-intercept
parameter estimate was found to not be
significant at alpha level 0.05 for the green
Figure 2. Residual plot of combined-variable and logarithmic
model forms for stem dry weight outside bark (kg)
weight and dry weight response variables.
The logarithmic equation form showed to be the
Table 3. Parameter estimates for each response variable
outside bark. Model form:
Response variable
Volume (m3)
Green weight
(kg)
Dry weight (kg)
0.00005
3
0.0281
0.0253
1 .7 3 0
7
1 .6 5 2
7
1 .4 8 9
4
Table 4. Parameter estimates
compartment. Model form:
for
1.1236
PseudoR2
0.99566
1.4149
099314
1.3494
0.99295
each
0.7065
0.6174
0.0002
0.0015
0.0747
0.3303
1.7953
1.7993
3.8283
2.7600
1.9432
2.0833
PseudoR2
0.9643
09658
0.9670
0.9791
0.9740
0.9778
0.4002
2.0656
0.9784
Response variable
Stem with bark
Stem without bark
Branch
Foliage
Coarse Roots
Aboveground
woody*
Total woody**
E.benthamii
*Aboveground woody = stem w/ bark + branch
best predictor of stem variables (Table 3) and
E.benthamii compartments (Table 4).
Next Steps
1) Examine merchantable ratio of stem wood
and 2) root:shoot of E.benthamii in SEUS.
48
Development of G&Y hybrid models for clones of Eucalyptus subject to
different environmental and management condition in Brazil
Scolforo, H. F , Stape, J. L.
1. Introduction
The culture of Eucalyptus is expanding in all
regions of the Brazil, however, much of the area
under cultivation has limitations to plant
development, notably with different levels of
water stress. Eucalyptus plantations are primarily
clonal, and these clones are adapted for each
of these regions, due to the strong genotype x
environment (GxE) interaction. This results in a
more complex decision when selecting suitable
genotypes
for
different
environmental
conditions, making it necessary to identify the
most promising materials for each situation
(Stape et al., 2004), ie, those that are best
adapted to abiotic and biotic stresses of each
local.
The success of forest plantations requires the use
of the more adapted genotype to different
climate and soil conditions, but that add
features of economic interest, and generating
social value (Stape et al., 2006; 2010).
In tropical regions of Brazil, the interannual
variability of drought and excessive heat, cause
significant interannual variations of the mean
annual increment of Eucalyptus clonal
plantations that have an accelerated rate of
growth and water use. Therefore, projections
made from different years can lead to
completely different responses which creates
insecurity for those who must do the forest
planning. Thus the development of hybrid or
inclusion of environmental variables in models
describing growth is a concrete form of the
evolution in biometrics area by improving its
general condition and projection accurately.
Brazil in 30 different sites with the insertion of soil
and climate characteristics that most influence
the genotype x environment interaction, in order
to reduce the estimated errors and especially to
provide
a
physiological
sensitivity
to
mathematical models, beyond attempts to
increase the generalizability power of the same.
The following specific objectives:
Develop descriptive models for 11 clones of
Eucalyptus in 30 sites in Brazil;
Develop hybrid models for these same clones
and sites with the incorporation of soil and
climatic variables;
Evaluate the accuracy of the descriptive
models with and without the GE interaction;
Evaluate the accuracy of hybrid models with
and without the GE interaction;
Evaluate from descriptive models if it is feasible
through the mixed effect modeling cluster
two or more clones in ideotypes;
Evaluate from hybrid models if it is feasible
through the mixed effect modeling cluster
two or more clones in ideotypes;
Propose the most appropriate way to model
the GE interaction for Eucalyptus. This form
indicates that modeling is capable of being
generalist, realistic and accurate.
3. Materials and Methods
3.1 Forest inventory data
2. Objective
The goal of this project is to model the growth
and yield of the 11 most important clones in
The dataset from the TECHS project will be
provided. Therefore, the data set will consist of
permanent inventory plots across 30 different
sites of Brazil (Figure 1) evaluating 11 clones of
49
Eucalyptus for each site.
Plantings occurred in 2011 and 2012 and the
forest inventory has completed its first
measurement in 2013, being held semi-annually.
By 2017 there will be approximately 6 years of
semiannual data representing a complete
rotation for each of the 11 clones.
The dendrometric variables to be measured:
diameter at breast height (DBH), height (H),
canopy height (Hc) and survival (N). The
inventory will also contain soil and climatic
variables such as Leaf Area Index (LAI), nutrient
analysis of foliage and soil, soil structure,
hydraulic
analysis,
measurement
of
such as: Sullivan and Clutter model (Sullivan and
Clutter, 1972; Clutter et al., 1983), ChapmanRichards generalization of von Bertalanffy’s
(Pienaar and Turnbull, 1973), as a whole-stand
level approach in addition to others.
Furthermore, it will be tested the efficacy of
physiological process-based (3PG Model, Stape
et al., 2004) and hybrid models, for the same
dataset. Descriptive and hybrid models will also
be considered for the mixed effect modeling.
4. Next Steps
The study was started in August of 2014 and is
supposed to be concluded by the middle of
2018.
References
CLUTTER, J.L., FORTSON, J.C., PIENNAR, L.V., BRISTER, R.G.H.,
BAILEY, R.L. 1983. Timber Management: A Quantitative
approach. Wiley, New York, 333 pp.
PIENAAR, L.V. and TURNBULL, K.J. 1973. The ChapmanRichards generalization of von Bertalanffy’s growth model
for basal area growth and yield in even-aged sands. For.
Sci. 19, 2-22.
STAPE, J.L., BINKLEY, D., JACOB, W. and TAKAHASHI, E. 2006. A
twin-plot approach to determine nutrient limitation and
potential productivity in Eucalyptus plantations at
landscape scales in Brazil. For. Ecol. Manage. 223, 358-362.
evapotranspiration, plus
record of climatic data.
a
complete
daily
Figure 1: Climatic zoning of Brazil and 30 sites of TECHS.
3.2 Modeling
Studies for the proper understanding of the
behavior of clones as well as the interaction of
genotype / environment will be made. Tests will
be performed to compare values and enable
the observation of the trend in behavior
between variables based on the use of Principal
Component
Analysis
(PCA),
correlation
measures and graphical procedures, test of
significance, for example.
Different types of empirical/descriptive models
will be tested and evaluated in terms of fitting,
STAPE, J.L., BINKLEY, D. and RYAN, M.G. 2004. Eucalyptus
production and the supply, use and efficiency of use of
water, light and nitrogen across a geographic gradient in
Brazil. For. Ecol. Manage. 193, 17-31.
STAPE, J.L., BINKLEY, D., RYAN, M.G., FONSECA, R.A., LOOS,
R.A., et al. 2010. The Brazil Eucalyptus Potential Productivity
Project: Influence of water, nutrients and stand uniformity
on wood production. For. Ecol. Manage. 259, 1684-1694.
STAPE, J.L., RYAN, M.G. and BINKLEY, D. 2004. Testing the
utility of the 3-PG model for growth of Eucalyptus grandis x
urophylla with natural and manipulated supplies of water
and nutrients. For. Ecol. Manage. 193, 219-234.
SULLIVAN, A.D. and CLUTTER, J.L., 1972. A simultaneous
growth and yield model for loblolly pine. For. Sci. 18, 76-86.
50
Effects of Intensive Silviculture in the Restoration Process of Northeastern
Atlantic Rainforest in Brazil
Aliisa Harjuniemi UH, Jose Stape NCSU, Jacyr Alves Copener
Study objective
The overall goal of this study is to compare the
effects of intensive management versus
traditional management methods on total
above ground carbon stocks and on stem wood
biomass development for 20 different native
tree species in Atlantic forest restoration in
Northeastern Brazil. I will also determine the
effects of two different spacing and two
different tree species composition for carbon
allocation and long term sustainability. Finally I
will compare the results of this study to the
parallel study site in the Sao Paulo state in
Figure 1. Randomized block design. A means 50:50 ratio of
pioneers and later successional species, B 67:33 ratio, 1
means stocking 3333 tree/ha, 2 means 1667 trees/ha, X is
intensive management methods and U traditional methods.
Southeastern Brazil with similar design looking for
site specific and general patterns.
Study design
The trial was installed in Bahia state, 200 km north
from Salvador. The climate is a typical tropical
climate; annual mean temperature is 24 °C and
annual rainfall 1400 mm. The project has 23
factorial design, where intensive and traditional
treatments (factor 1) were applied to 20 native
species, for two initial stockings (factor 2, 3333
trees ha-1 and 1667 trees ha-1) and two species
composition proportion (factor 3, 50/50 and
67/33 ratio of pioneer : later successional
species). The treatments were applied in a
Figure 2. The picture above is from a plot where
intensive management was applied (A2X) and below a
picture from a plot after traditional management (A1U).
randomized block design (Figure 1). The trial has
40 plots; 4 plots for each treatment (4*8), 4 plots
for control without forest restoration and 4 plot
for destructive sampling. Each plot is 30 m * 39 m
(1170 m2) with an interior plot of 24 m * 32 m (768
m2) on where measurements were made. All
trees were planted on August 2004. Pictures
illustrating the differences between intensive
and traditional managements presented in
Figure 2.
Current status
The height and diameter at 30 cm were
measured for all trees in July 2012. In addition to
this, forest floor and understory were sampled
51
Figure 3. Effects of intensive management to stem wood
development on different tree species.
and their dry weights were determined. Wood
specific gravity was determined for all 20 tree
species four times. The density for bark was
determined four times as well.
Basal areas and stem wood biomass estimates
for each tree and plot were calculated to
compare the effects of different treatments.
Stem wood biomass (dry) was calculated based
on tree cross sectional area, wood specific
gravity and using a taper factor 0,5. Biomass
estimates for bark was calculated separately
because of the lower density of bark and added
to the final stem wood biomass. By summing the
biomass of all stems (including bark and wood),
stem wood biomass estimates per plot were
obtained. The results were presented in Mg/ha.
Statistical analysis to find out the individual
effects of treatments and the possible
interactions was made using ANOVA GLM
procedure with Minitab 15 software.
Next I will calculate the total aboveground
biomass for each plot to get estimates for the
carbon stocks. For this, I will add the canopy dry
mass estimates to obtain the total aboveground
biomass of trees. To get more accurate
estimates of the total amount of carbon in the
system, I will also include understory and forest
floor dry masses. The total amount of carbon per
plot will be calculated by multiplying biomass
estimates with 0,5.
Ceptometer measurements are ongoing and will
be ready approximately in mid October. I’m still
waiting for results from leaf nutrient content
analysis, forest floor furnace and weights from
specific leaf area samples. The results from these
measurements I will use to calculate the amount
of carbon in forest floor, leaf area index and
Figure 4. Effects of different treatments on stem wood stocks.
canopy nutrimental status.
Results
Effects on different tree species.
Intensive treatment had positive effect on the
stem wood biomass development for all tree
species (Figure 3).
Stem wood stocks per treatment
The spacing and the management method had
both statistically significant influence (p < 0,000).
Larger proportion of pioneer species had no
effect on stem wood stocks 8 years after
planting. No significant interactions was
observed at significance level 0,05. Plots with
intensive management method and denser
stockings had bigger stem wood stocks
(respectively 270 % and 70 % bigger).
Comparison to the South site
The responses to the treatments seem to be
following similar patterns so far, at least
regarding the effects of the intensive
management.
Summary
The Intensive management method has a
significant positive effect on stem wood stocks
also in the Northeastern Brazil. However, not all
calculations and analysis are finished yet. Longterm sustainably of larger pioneers species
proportion may be too early to determine.
Acknowledgements
Travel to Brazil and the data analysis were funded by the
Zobel grant and the FPC. Thanks especially to Copener for
supporteding my stay and research in Brazil and for
providing valuable help in the field and in the laboratory. My
studies at NCSU were funded by the Atlantis program.
52
Fertilizer response to Eucalyptus plantations and its correlation with the
soil, climate and silviculture characteristics in different regions of São
Paulo state
Renato Meulman Leite da Silva, José Luiz Stape, Antonio Natal Gonçalves
Introduction
The nutrients are essentials resources to wood
production, therefore a nutritional management
is needful to maintain a sustainable productivity
on forests plantations. These plantations cover
extensive landscapes, over different sites,
climate and soil conditions and the productivity
can be limited by low soils fertility or a
inappropriate fertilization management. This
study aimed to identify nutritional limitation,
evaluate the effect of the soil, climate and
silviculture
characteristics
on fertilization
response and determine the attainable
productivity
in
Eucalyptus
plantations
widespread around 52,700 ha in different regions
through growth estimates to determine the
fertilization response (FR) to each pair of plots.
Experimental blocks were distributed in three
different regions (fig. 1), from north to the south,
increase the clay content, organic matter and
rainfall and decrease the temperature and
drought. To each block has determined the
climate and soil characteristics.
In additional, leaf area index (LAI) was estimated
by hemispherical photos.
Figure 2. Twin (PG-G) and inventory plots (PG-T) volume
at study establishment, 2003. .
Figure 1. Studied Eucalyptus plantations. Distributed
along the state of São Paulo, Brazil.
of São Paulo state, Brazil.
Methods
The Twin Plot design was used in 161 plots which
receive additional fertilizations. These plots
account stands of 2 to 4 years old, seedling or
clonal plantations and establishment or coppice
forests. The plots were paired and the initial
volume was estimated (fig. 2), and compared
for 2.5 years with traditional inventory plots which
receive the usual fertilization management and
Figure 3. Current annual increment (CAI), 2.5 years of
evaluation.
53
to the fertilizations management, while the
northern region was the least responsive, which
was a good correlation with the climate data,
rainfall and water deficit (table 1). Also was
observed the same FR to coppice or established
plantations.
Beyond the increment on wood productivity, the
treatment with extra fertilizations also increased
the leaf area index (LAI) and the light use
efficiency (Fig. 4).
The contents of organic matter, clay, sand and
sum of bases was the soil characteristics more
correlated to the FR, at the studied forest.
Beyond the difference between the regions, FR
was different for soil type, texture, and genetic
materials. The each local analysis presented
good correlations between FR and some soil
characteristics: sand, silt, clay, sum of bases,
Magnesium
(Mg),
Potassium
(K)
and
Phosphorous (P) contents.
Conclusions
Figure 4. Comparison of LAI (A) and LUE (B) between
the treatments. Both were 15% upper in extra fertilized
treatment (PG – G).
Results
The results showed a growth on twin plots
greater than traditional plots (fig. 3) an
attainable productivity of 22 Mg.ha-1.yr-1, 11%
upper than the actual.
Also showed a capacity to reach 28 Mg.ha-1.yr1, at the south of state, more productivity local
(Region 3). The average FR was 5.3 Mg.ha-1.yr1). The south region also was the most responsive
The twin plot approach was effective to
determine the attainable productivity, FR and
showed an opportunity to improve the
fertilization management at studied plantations.
Increased productivity is made by increasing the
light absorption (LAI) and efficiency of its use
(LUE).
The water availability was inversely proportional
to FR, and determine the difference between
regions, in additional the soil nutrients were more
correlated to FR in the regional overview, when
the climate characteristics no differ. The
coppice and established plantations has the
same potential to response the fertilization.
References
STAPE, J.L.; BINKLEY, D.; JACOB, W.S.; TAKAHASHI,
E.N. A twin-plot approach to determine nutrient
limitation and potential productivity in
Eucalyptus plantations as
Table 1: Pearson correlation coefficients – Soil and climate characteristics.
landscapes scales in Brazil.
Forest
Ecology
and
Management 223: 358-362,
2006.
** signiificant (P= 0,05); * siginificant (P= 0,01); ns = not significant.
54
Zoning the productivity of Eucalyptus forest plantation in the northeastern
of São Paulo state in Brazil
Cristiane C.Z. de Lemos, Jose L. Stape
Study objective
In a competitive wood market for land with
other cultures, zoning the productivity can help
the forest managers to make their decision
about land use, based on technical information
of forest productivity. In addition, the correct
forest location on climate condition, possibility to
ensure the wood production and to decrease
the need for land, resulting in a reduce of
pressure on native forest deforestation. The
objective of this study was to zone the actual
productivity of Eucalyptus forest plantation in the
northeastern of São Paulo state and
to
determine constraints for forest growth.
precipitation (Alvares et al., 2012a) and
maximum and minimum temperature estimation
(Alvares et al., 2012b) and a pedological map
(Oliveira et al., 1999). The solar radiation and
numbers of frost day were calculated using
empirical equations (Pereira, Angelloci and
Sentelhas, 2002; Alvares, 2011).
The 3-PG input parameters were calibrated and
validated per type of soil, using 113 pairs of twinplots spread over the study area.
Methods -Site Description
The study was developed in the northeastern of
São Paulo State in Brazil (figure 1), localizated
between 23º 3’S to 18º 58’S and 48º 38’W a 46º
21’W coordinates, totalizing around 5 million of
hectares. The weather condition is Aw, Cwa and
Cwb (Köppen). The predominate soils are
Quartzpsament, red yellow sandy Oxisol, and
red clayey Oxisol.
Figure 2. Eucalyptus plantation in Brazil .
Validation of the estimated productivity
To validate the model, were compared the
mean annual increment at 7 years (MAI 7)
estimated by 3-PG with the projeted production
from 1.884 commercial inventory plot by Clustter
equations (forest with 5 to 8 years, figure 2),
corresponding around 58,000 ha.
Because our map resolution was 400 ha we
compared the forest productivity by stand using
the average weight.
Figure 1.
Study area (brown) in the northeastern of São Paulo
State in Brazil.
Estimation of Eucalyptus plantation
productivity
The estimative of productivity was obtained
using 3-PG model modified by soil water content
associated with ArcGIS tools to spatialize the
analyses. For this purpose, we used maps of
Estimation of Eucalyptus plantation
productivity
The 3-PG modifiers: temperature (ft), frost (ffrost),
soil water content (fѲ),vapor pressure deficit (fvpd)
e soil fertilization (FR) were used to determine
the growth constraints. We randomly selected
10% of our study area to evalute those modifiers
along the Eucalyptus rotation (7 years).
55
Figure 3.
Maps of northeastern of São Paulo State in Brazil - A) Eucalyptus productivity: mean annual increment at 7 years (m3 ha-1
year-1). B) Leaf area index at 7 years (m2 leaf m-2 of soil).
Results
Figure 3 shows us the Eucalyptus productivity
and leaf area índex along the northeastern of
São Paulo. The average of all validated area
was similar between the 3-PG estimation and
the inventory projection, 42 m3 ha-1 year-1.
However, the comparison of the IMA7 between
measured and predicted, stratified by soil type
and farm (figure 4), we found high data
productive areas are located at south. The main
growth constraints were vapor pressure deficit
and soil water content.
Conclusion
The spacial 3-PG model was able to estimate
the forest productivity on northeast of São Paulo
state, and different production based on the
type of soil.
Figure 4. Relationship between mean anual increment (MAI, 7 years) projected using inventory data and mean anual increment at
7 years estimated using 3-PG model per type of soil (Quartzpzament, Red yellow sandy soil and red clayey oxisol), average per
farm.
dispersion, probably because of the average
climatic data used as model input, the soil type
detailing reflecting the soil water content, the
genetic material adaptation and silviculture
management, or the interaction of those factors.
The results suggested that increasing the
analytical scale reduces the precision of the
estimatives. It was possible to make the zoning of
the real productivity of Eucalyptus in the
northeastern state of São Paulo. The most
Acknowledgment
This research was supported by the São Paulo
Research Foundation (FAPESP, 07/57896-5), the
University of São Paulo and International Paper
company. Clayton Alvares helped us on maps
issues. We thank them for their support.
56
Crown architecture, leaf angle distribution, light extinction coefficient
and light interception in Eucalyptus clones
Eduardo Moré de Mattos, José Luiz Stape
Study Objectives
The process of converting light energy from the
sun into wood starts with the interception and
absorption of PAR radiation by the canopy
(APAR). On the other hand, interception of
radiant energy by the canopy depends on
factors related to the structure and spatial
arrangement of the crowns, distribution of leaf
angles,
foliage
amount
and
optical
characteristics of the leaves. There are several
models describing the paths of the radiation
through the canopy. However, simplified models
that assume the canopy as a homogeneous
medium ("Big Leaf") has been successfully
employed to quantify APAR. Based on Beer’s
law, we only need to determine two variables,
one related to the amount of foliage, the leaf
area index (LAI) and the light extinction
coefficient (k), which simplifies the parameters
associated with other characteristics mentioned
above. Our goal is to characterize crown
structure and investigate how it can affect light
absorption.
Site description
This is study is being conduced at the same area
mentioned in the 2-page report entitled
“Patterns of seasonal stem growth of 18 different
Eucalyptus clones” published in this edition.
Methods
Throughout the experiment we have monitored
LAI with a LAI-2000 plant canopy analyzer on a
monthly basis and complemented the
information with periodic PAR readings above
and below the canopy with the LP80
Ceptometer. 20 measurements per plot were
performed with the LAI-2000 and 24 readings (6
points, 4 positions per point) for the LP80.
Destructive sampling was performed in April-May
2014 where 3 to 6 trees were cut down per
clone. Crown was divided in 3 thirds proportional
to its length. We took 30 leaf angles
measurements randomly per third. All the leaves
were removed and weighted, 15 fresh leaves
per third were collected to determine specific
leaf area, and samples were also taken in order
to determine dry weight.
Results - LAI
Figure 1 shows the LAI variation in the period. The
selected clones spanned a large range of LAI
values, from 1.5 to 5.5 m2 m-2. It reinforces the
opportunity with this study to test light
interception across the possible LAI values
attained by Eucalyptus plantations. The
estimates obtained by the LP80 were fairly
correlated to the estimates by the LAI-2000(fig 2)
Figure 1 : Leaf area index (LAI-2000) variation observed through the study period, from 16 to 32 months of age.
57
Figure 4: Relationship between mean inclination angle and k.
Line is a smooth trend and r is the Pearson’s correlation
Figure 2 : Agreement of the LAI estimates by the LP80
Ceptometer and the LAI-2000 plant canopy analyzer. Bars
represent the mean standard error.
However, LAI estimates via destructive sampling
were considerably greater than LAI from
equipments (fig. 3).
We observed a trend in reducing k with
increasing leaf angle inclination. Figure 5 shows
the variation of leaf angle distributions within the
crown. Leaf angle tends to increase with
increasing crown height, this endorses the fact
that k changes within the crown and this may be
taken into account when modeling light
interception.
Figure 3: Examples of two clones with different LAI estimates by
the 3 methods used. Destructive sampling led to higher values.
More investigation is needed, but our
understanding rely on the fact that clones
stacked more leaves into their crowns, which
does not necessarily contributed to light
interception. It reinforces the attention necessary
to what methods to be used when setting up
trials and tests and what is the information
required, sometimes PAR values itself are the
most valuable information.
Results - Light extinction coefficient (k)
Inverting beer’s law equation in conjunction with
the destructive LAI estimates and LP80 above
and below readings, assuming constant k
through the canopy, we could calculate the so
called light extinction coefficient, which ranged
from 0.25 to 0.6 (fig. 4).
Figure 5: Leaf angle distributions for the clones under study and
differences observed at the upper (blue), mid (green) and
lower crown (red).
Acknowledgment
We thank the TECHS project and all its
associated companies in special Duratex that
hosts the experiment. We also thank CAPES for
the scholarship and the Forestry Science and
Research Institute for field support.
58
Patterns of seasonal stem growth of 18 different Eucalyptus clones
Eduardo Moré de Mattos, José Luiz Stape
Introduction
The advances of the silviculture practices and
genetic materials (basically cloning) has led the
plantations in Brazil to a significant increase in
productivity. However, can we keep this
increase? Or can we sustain this productivity
rates? In order to answer some of these
questions we have to understand how and why
patterns of growth may differ from one region to
another but also how clones interact with in-site
weather variability as climate changes and
extreme events are becoming more frequent.
Study Objectives
To better understand the factors regulating
productivity and how growth is subject to soil
and weather conditions we have recorded dbh
every 15 days. Our goals are to:
1.Characterize the patterns of growth of 18
different genotypes (clones) of Eucalyptus,
representative of the variability range found
in genetic materials current in use for
commercial plantation in Brazil.
Experimental design – Methods
The experimental plots consists of 120 trees on a
3x3 m spacing, where 40 trees are reserved for
destructive sampling, giving a 720 m² plot(8 rows
by 10 trees) (fig. 1). The 18 clones were classified
into plastic ones (4), tropicals (7) and
subtropicals (7) according to the region where it
is being planted and prior knowledge about
field
behavior.
Every
fifteen
days
8
representative trees are selected per plot and
their DBH is recorded. Climate information is
provide by public weather stations from the
Brazilian national institute of meteorology which
records
hourly,
radiation,
rainfall,
air
temperature, wind speed, and air relative
humidity data. Destructive sample was
conduced trough April-May 2014 at peak Leaf
area Index. A total of 93 trees were cut down
and measured in order to build allometric
equations. All DBH, heights, and height to the
live crown are measured each 5-6 months.
2.Identify how clones respond to weather
conditions after canopy closure (maximum
growth rates).
3.Develop clone specific models capable to
predict how meteorological variables
weights on growth response.
Site Description
This study is supported by TECHS project
(Tolerance of Eucalyptus clones to hydric and
thermal stresses, www.ipef.br/techs), a research
association led by the Forest Science and
Research Institute (IPEF, www.ipef.br). The
experiment was planted in 2011, December 31st
and now has 33 months old. It is located in Buri
(23°47’57”S and 48°35’15”W ), south region of
the São Paulo state, Brazil (fig 2), on a land held
by Duratex company. The climate is classified as
Cfa, according to Koppen classification with
1200 mm of average annual rainfall and 20.5 °C
of mean annual temperature, the soil is a red
clayey Oxisol (50% clay).
Figure 1 : Experimental design and site location. Only one plot
of each clone at the 3 X 3 spacing is being accessed for the
purposes of this study.
59
Results
The most and least productive clones grew 65,4
and 22,5 m³ ha-1 year-1, respectively. Growth
rates increased in the summer in response to
increase in radiation and air temperature,
ranging from 0 to 0,44 cm2 tree-1 day-1 (fig. 3).
Among all clones the average growth rates was
0.21 cm2 tree-1 day-1, but clones showed great
differences in how sensitive is growth, the
variation coefficient for growth rates ranged
from 25% to 51%. Rainfall is well distributed across
the year, but in late February around 26 months
of age, growth had ceased for most clones. An
interval of 18 days without any rain associated to
high temperatures and low air humidity led
vapor pressure deficit to reach 2.3 kPa (fig. 2) on
a daily basis (including night period). Besides the
soil storage had reached low values (data not
shown) in other periods, but trees only stopped
to grow when the atmospheric demand was
aggressive.
Figure 2: Experimental layout and location.
Figure 2 : Daily rainfall (bars), mean temperature and vapor
preassure deficit (lines) observed at the meteorological station.
Next Steps
Data collection has finished, completing 18
months of monitoring and now we will proceed
to statistical analyses. The data showed here is
subject to validation and refinement. The clonespecific allometric models still need to be
incorporated into the computations, as well as
others approaches shall be used.
Figure 3 : Average sectional area increment for the 3 types of
clones: plastics (top), tropicals (middle) and subtropicals
(bottom).
Acknowledgment
We thank the TECHS project and all its
associated companies in special Duratex that
hosts the experiment. We also thank CAPES for
the scholarship and the Forestry Science and
Research Institute for field support.
60
Leaf anatomical characterization of 16 highly productive Eucalyptus
genotypes from Brazil
Raoni I. Nogueira, Tatiane M. Rodrigues, Jose L. Stape
Justification
Several studies on uptake and use efficiency of
water and light have been conducted looking
for their relations with adaptation and
productivity, however, leaf anatomy and its
relation with plant ecophysiology are still little
explored. In EUCFLUX project studies on carbon,
water, energy and nutrients cycles are
conducted at various scales and for 16 highly
productive and contrasting, in terms of regions
of origin, genotypes. Having such information
available, to characterize this collection of
contrasting genotypes in terms of leaf anatomy,
proved to be an unique possibility, given the
detailed information that has been collected on
the development of clones, to further investigate
whether there is some kind of relationship
between anatomy and ecophysiology of those
that can be observed.
embedded in methacrylate resin, then cross
sections of the material were made with a
rotation microtome. Cross sections were colored
with Toluidine Blue coloring. Measurement and
quantification of cell layers were obtained from
light microscopy images.
Results
Most of the genotypes have presented stomata
only on the abaxial leaf surface. Only two clones
were amphistomatic.
Figure 1. Amphistomatic Clone adaxial leaf surface impression.
Treatments and experimental design
The Eucflux Project has 16 treatments, each one
being a single genotype, to be known, 2 seed
origin E. grandis and 14 hybrids of Eucalyptus
grandis, E. urophylla, E. camaldulensis and E.
saligna. These genotypes are commercially
planted by several companies in Brazil each one
having a unique site specificity. For this Project
they’ve been planted together in Itatinga city,
São Paulo state in November 2009. Considering
their sites specifications, all of them have
phenotypic characteristics inherent to the sites
to which they were genetically improved.
Methodology
Mature, undamaged leaves were collected
from each treatment and taken for analysis.
For stomata studies 3 leaves had an adhesive
applied in both, adaxial and abaxial leaf
surfaces providing leaf surface impressions.
Stomata were quantified and measured at 200 x
magnification.
For leaf anatomy 3 other leaves were
dehydrated in crescent ethanol series,
Figure 2. Amphistomatic Clone abaxial leaf surface impression.
61
Table 1. Leaf Area Index and stomata area.
Contrasts
Unit
Adaxial cuticle lenght
Abaxial cuticle lenght
Adaxial epidermis lenght
Abaxial epidermis lenght
Palisade parenchyma lenght
Spongy parenchyma lenght
Intercelular air spaces
Vascular bundle density
Secretory cavities density
Abaxial stomata density
Individual abaxial stomata size
Stomata area on adaxial leaf surface
Stomata area on abaxial leaf surface
μm
μm
μm
μm
μm
μm
%
# mm-²
# mm-²
# mm-²
μm²
%
%
For mesophyll characterization, 15 genotypes
have one layer of palisade parenchyma and 5
or 6 layers of spongy mesophyll. One genotype,
differered completely from the others, having a
homogeneous mesophyll with 6 layers of
palisade parenchyma.
Data from other studies of the Eucflux project,
three contrasting pairs have been compared,
regarding productivity (least and most
productive), origin (seed and most productive
clone) and light use efficiency (lowest and
highest). No contrasts in leaf anatomy were
observed between the productivity and origin
pairs. Light use efficiency (LUE) genotypes were
clearly different. As seen in figures 3 and 4 the
clone with the highest LUE is the only one that: is
amphistomatic and has a homogeneous
mesophyll, having only palisade parenchyma.
Origin
Seed
Clone
8
7
16
13
73
153
53
10
5
384
207
0
8
6
7
16
13
78
165
61
11
5
362
171
0
6
Productivity
Lowest
Highest
8
6
17
14
72
178
54
11
5
308
226
0
7
6
7
16
13
78
165
61
11
5
362
171
0
6
LUE
Lowest
Highest
6
6
16
13
66
158
49
11
4
375
145
0
5
6
9
15
14
220
0
26
11
9
217
308
3
7
Figure 3. Leaf blade cross section of the clone with lowest LUE.
Figure 4. Leaf blade cross section of the clone with highest LUE.
Considerations
Although being not possible to establish any
cause-effect relation for the differences seen
between the two most contrasting genotypes
regarding LUE, this study suggests a needing for
a best detailing of the effect of this anatomic
distinctions on LUE by this Eucalyptus clone.
Considering the possible relation between
anatomy and physiological behaviors, and
considering a genetic control on leaf anatomy,
it would be possible to predict a system of early
selection for plant breeding between genotypes
with distinct leaf anatomies, identifying in the
hybridized progeny the plants that have
inherited the desired anatomy.
Additional resources
Evans, J.R. and Vogelmann, 2006. The New Phyt.,
171, 771-782.
Tholen, D. et al., 2012. Pl. Science., 197, 92-101.
62
Physiology of tropical Eucalyptus clones under water stress
Marina S. G. Otto, Ricardo F. de Oliveira, Beatriz T. Gonsalez; José L. Stape
Study objective
The
experimental
design
is
completely randomized in a factorial
2 x 8: 2 treatments and 8 clones, with
6 repetitions, totaling 96 experimental
units.
The goal of this study is to evaluate
physiological parameters and
accumulation of γ-aminobutyric acid
(GABA) in Eucalyptus under water
stress. GABA is a four-carbon nonprotein amino acid that has been
linked to stress, signaling and storage
in plants.
We studied two treatments: with and
without water stress. Water stress
began on November 12th when
seedlings were around 3 to 4 meters
tall. Water stress was imposed by non
-irrigation of the plants and pots
coverage to avoid rainfall. These
water stress treatments were
With the advancement of the global
area planted with Eucalyptus to
regions under water stress, it
becomes necessary to increase our
understanding about physiological
processes triggered by the plants
and to identify metabolite signaling
of stress.
Treatments and Experimental design
Eight Eucalyptus clones from different
origins were evaluated (Table 1)
Table 1: Clones and origins Clone
Hibrid
Local
COP1277
E. grandis x
camaldulensis
Bahia
GG100
E.urophylla
Minas Gerais
SUZ2003
E. urophylla x
tereticornis
Maranhão
E. urophylla
VM04
E. urophylla
VER361
CNB10 E. urophylla x grandis
FIB6075 E. urophylla x grandis
JAR2646 E. urophylla x grandis
Minas Gerais
Bahia
Minas Gerais
São Paulo
Pará
Seedlings were planted in 320 liters
pots located in Piracicaba, state of
São Paulo, Brazil in February 2013.
Figure1: Eucalyptus planted in Piracicaba,SP with pot coverage to avoid
rainfall.
maintained for 2 cycles: 3 days
(cycle 1) and 5 days (cycle 2), when
water-stressed plants were rewatered
during 15 days to recovering (Figure
1).
63
This signalization can results in physiology reactions, like stomatal control,
Figure 2 Relation between stomatal conductance and vapour pressure deficit for
Eucalyptus clones in water stress conditions.
Clones from drought areas had more
stomatal sensitivity (COP, VM and SUZ)
than clones from wet areas (JAR and
Figure 3: Relation between water potential and GABA concentration in Eucalyp-
Table 1: Stomatal sensitivity of Eucalyptus
clones during water stressed days
W
e
Figure 4: GABA concentration during the
first water stressed day in Eucalyptus
Future plans
GABA concentration
GABA aminoacid had a high correlation with water potential, showing
that this aminoacid is largely produced in response to water stress
conditions (Figure 3).
Besides during the first day of water
stress, GABA concentration was higher for the clones from drought areas
(COP, VM and SUZ), the same clones
that had more stomatal sensitivity
(Figure 4). This occurrence indicates
We plan develop a new study with
the same clones and on the same
area inducing excess of water, which
is planned for the next semester
(January until July 2015). We will
quantify how these clones respond a
different stress, not without water but
with excess of water.
We will measure water potential,
GABA concentration, leaf anatomic
characteristics and gas exchange
variables, with the supervision of Mike
Ryan (Colorado State University).
64
Biomass evaluation of the Eucalyptus spp. plantations established in Bocaiúva, State
of Minas Gerais, Brazil.
Maria C. Arrevillaga L.; J.L. Stape
Justification
The Brazil Eucalyptus Potential Productivity (BEPP)
Project was designed to evaluate the interaction
between wood production and the use of
natural resources, identifying the influence of
water and nutrients on Eucalyptus spp.
plantations established across a geographic
range of sites and clones in Brazil (Stape et al.
2010).
The BEPP project was developed as a
consortium of eight companies, the USDA Forest
Service, the Forest Science and Research
Institute (IPEF) in Brazil, the University of Sao Paulo
and Colorado State University.
Each company installed the experiment testing
irrigation levels and repeated fertilization at rates
according to its own site index. Although the
BEPP project started in 2001, the establishment of
the experiment in Bocaiúva, State of Minas
Gerais (MG), was initiated in 2005, continuing
the assessments such as forest inventory and
biomass evaluations until July 2014.
Bocaiúva, MG, are highly influenced by the
water availability. The precipitation rates in the
area are approximately of 850 mm per year,
concentrated in the summer season (Table 1).
Table 1. Climate characteristics in Bocaiva, MG.
Average annual
PAR
3500 MJ.m.year
Average precipitation
850 mm.year
Potential evaporation
1220 mm.year
Therefore there was an interest in evaluate the
development of the plantation in conditions
where the water income was even lower than
the average precipitation rate, including an
additional treatment beside the other ones
established in the BEPP project:
water
reduction.
Treatments and experimental design
The BEPP experiment in Bocaiúva was
established following a factorial design (3x2)
evaluating two factors simultaneously, water
availability: control, irrigated and reduced; and
fertilization
rates:
control
and
potential
fertilization. A total of 4 replicates were made
according with the 4 commercial clones of the
company (VM1, VM3, VM4 and MN463).
Figure 1. Localization of Bocaiuva, Minas Gerais
State, Brazil.
BEPP Bocaiúva
The
Eucalyptus
plantations
established
in
In order to evaluate the biomass of every
treatment, destructive sampling was made
selecting 6 trees per plot. The selection of the
trees was based on a forest inventory made
previously. DBH of every treatment was rated in
ascending order and selected
trees
representatives of the average. The block
corresponding to the VM3 clone was not
included in this evaluation because it presented
wilt symptoms of Ceratocystis sp.
65
Measurements
Measurements of height, DBH and tree crown
was made for each tree assessed. Trees were
felled in order to take samples of branches and
wood discs at different heights (0 m, 25%, 50%,
75% and 100% of commercial height).
Canopy was divided into three equal parts,
collecting 30 leafs per part for specific leaf area
Figure 2: MAI for reduced, rainfall and irrigated
plots with 7 years old.
index analysis. It was also evaluated the angle of
insertion of the leafs .
Results
The irrigation increased 30% in the MAI for the
Irrigated
Rainfall
Reduced
Figure 2: MAI for reduced, rainfall and irrigated
plots with 7 years old.
plot that receive only water from the rain and
50% for the plots with reduction rain.
Through the installation of dendrometers can
monitor the effect of treatments monthly,
Figure 3: Pictures of rainfall plot (above) and
redudec plot (below).
especially
October.
during
the
dry
season
April
to
Next Steps
Relate growth variables with the leaf area index
and light use efficiency in order to determinate
the influence of the fertilization and water
availability in the development of the
eucalyptus plantations established in Bocaiúva,
MG.
Based on the results obtained by Stape et al.
(2004) where the production increased by about
2.3 Mg ha–1 . yr–1 for each 100 mm increase in
precipitation, it is expected to obtain a higher
biomass in the irrigated treatment since the area
has a low water availability; is also expected to
observe an influence of the fertilizer treatment
on the growth and carbon allocation of the
forest.
References
66
Effect of soil water availability on PAR capture by Eucalyptus globulus,
Eucalyptus nitens and Eucalyptus camaldulesis hybrids
Matías Pincheira 1,2 , Rafael Rubilar 1,2
Abstract
The effect of contrasting soil water availability on
photosynthetically active radiation capture within
the crown of ten Eucalyptus globulus, nitens and
camaldulensis hybrids was evaluated at 18 months
of development. Light measurements with a Li-Cor
191 ceptometer were made every 25 cm from the
base up to the top of the tree. A logistic model
was adjusted to the rate of light capture (RLE) and
relative crown height to maximum light capture
(MLE). Our results suggest differences among genotypes, irrigated conditions and the interaction effect genotype x irrigation. Three genotypes groups
under irrigation and four groups under no irrigation
were identified with similar RLE and MLE .
planting line of four ramets for 10 selected genotypes Eucalyptus sp. genotypes and hybrids
(Table 1). Each group of genotypes was replicated three times at each irrigation level.
Figure 1. Location map of study area.
Introduction
Little information exist on the crown architecture
and light capture efficiency of Eucalyptus genotypes (Smethurst et al. 2003). This important characteristic reflects the plasticity of each genotype
to the environment, and may be considered an
additional variable to improve the traditional clonal selection criteria since is directly related to biomass productivity (Almeida et al. 2010, Binkley et
al. 2010).
Objectives
To determine and compare the effect of soil water
availability on light capture rates and maximum
light capture along the canopy of Eucalyptus
globules, nitens and camaldulensis genotypes and
hybrids.
Methods
In the dry central valley of the Province of Biobío,
Chile, a clonal trial was established under two contrasting water availability conditions (with and
without irrigation) (Fig. 1).
The experiment was established as a split plot
block design with three replicates under each
irrigation treatment. The experimental unit was a
Table 1. Description of the genetic material evaluated.
Measurements of photosynthetically active radiation (PAR) were made with a Li-Cor 191 ceptometer, along of individual tree crowns in steps of 25
cm from the base up to the top of the tree at 18
months post-establishment. Light extinction along
of the canopy was modeled using a logistic model
[1], and differences among genotypes were evaluated by incorporating dummy variables on the parameters of a nonlinear regression considering genotypes and irrigation treatment.
[1]
Where,
LEr : relative proportion of captured light (%);
Hr : predictive variable (relative height (%));
a, b: parameters of light capture rate and crown
length relative to maximum light extinction (x ≥ a,
b> 0);
e : exponential base.
67
Results
Relative light capture (%)
Differences in the rate of light relative capture (RLE)
along of the canopy was significative between
irrigated conditions for all geneotipes, exept in
EcxEg. Relative crown length to maximum light
capture (MLE) differed except in EgA2, EgB2 and
EnxEc (Figs. 2 y 3).
EgB2 and EnxEc (groups C and D, respectively).
The genotypes EgA1, EgA2, EgM1, EgM2, EnxEg
and EcxEg presented the maximum MLE (group A),
followed by EgB1, EgB2, En and EnxEc (group B).
Figure 3. Effect of soil water availability on light extinction
model parameters (Model 1). (a) parameter of light
capture rate, (b) parameter of relative crown length to
maximum light capture.
Conclusions
Irrigated conditions increased Eucalyptus genotypes relative light capture along the crown. Responses were genotype specific and independent
of the main species tested.
Acknowledgments
Figure 2. Relative light capture along the canopy of Eucalyptus genotypes. Closed dots (irrigated) Inversed triangles (without irrigation ).
Under irrigation, genotypes EgA1, EgA2, EgM2, EgB1, EgB2, En, and EcxEg (group A) presented similar
RLE and MLE, followed by EnxEc (group B), and
EgM1 and EnxEg (group C). EgA1, EgA2, EnxEc
and EcxEg (group A) presented the highest MLE,
followed by EgM2, EgB1, EgB2 and En (group B),
and EgM1 and EnxEg (group C).
Without irrigation, genotype EcxEg presented the
highest RLE (group A), followed by EgA1, EgA2,
EgM1, EgM2, EgB1, En and EnxEg (group B), and
To the Chilean National Fund for Scientific and
Technological Development, FONDECYT 1085093.
To the valuable and continuous support of Forestal
Mininco S.A. to develop this research.
Bibliography
Almeida, A. Siggins, A. Batista, T. Beadle, C. 2010. Mapping the effect of spatial and temporal variation in climate and soils on Eucalyptus plantation production with
3-PG, a process-based growth model. Forest Ecology
and Management 259: 1730-1740.
Binkley, D. Stape, J. Bauerle, W. Ryan, M. 2010. Explaining
growth of individual trees: Light interception and efficiency of light use by Eucalyptus at four sites in Brazil.
Forest Ecology and Management 259: 1704-1713.
Smethurst, P. Baillie, C. Cherry, M. Holz, G. 2003. Fertilizer
effects on LAI and growth of four Eucalyptus nitens plantations. Forest Ecology and Management 176: 531–542.
68
Effect of water availability and temperature on leaf deployment in
genotypes of Eucalyptus globulus, Eucalyptus nitens and
Eucalyptus camaldulensis hybrids
Matías Pincheira 1,2,3* , Rafael Rubilar 1,2 , Jorge Cancino 2 , Manuel Acevedo 1,2,3
1
Forest Productivity Cooperative, 2 Universidad de Concepción - Fac. Cs. Forestales, Chile. 3Instituto Forestal de Chile. *
Abstract
The effect of contrasting water availability
conditions was evaluated in ten genotypes of
Eucalyptus sp. on leaf area growth fortnightly.
Non-destructive leaf measurements were made
every two weeks during the first growing season,
and the effect of soil moisture and temperature
on foliar increment were identified as boundary
conditions of maximum leaf area growth rates.
Under irrigation two optimal growing conditions
were identified where without irrigation only one
condition was observed. The amplitude of
maximum range of foliar growth for genotypes
of E. globulus was directly related its
cummulative volume growth. E. nitens hybrids
showed a greater amplitude
of growth
conditions.
The experiment was established as a split plot
block design with three replicates under each
Figure 1. Location map of study area.
irrigation treatment. The experimental unit
considered a line of four ramets for 10 selected
genotypes of Eucalyptus genotypes (Table 1)
replicated three times for each irriigation level.
Table 1. Description of the genetic material evaluated.
Introduction
In species of the genus Eucalyptus analysis of the
dynamics and evolution of leaf growth is a topic
not completely well understood (Smethurst et al.
2003). Plasticity of genotypes to the environment
may determine common strategies to capture
light but more importantly understanding these
mechanisms may improve traditional clonal
selection criteria (Almeida et al. 2010, Binkley et
al. 2010).
Objectives
To determine the effect of soil water availability
and temperature on leaf area growth in
Eucalyptus genotypes.
Methods
In the dry central valley of the Province of Bio
Bio, Chile, a clonal trial was established under
two contrasting water availability conditions
(with and without irrigation) (Fig. 1).
Two branches (a row and inter-row branch)
were selected from one plant from each
experimental unit. on these non-destructive
measurements of leaf area accumulated
increases (AF) were made every during the first
growing season.
The absolute increases fortnightly branch leaf
area were categorized into deciles and defined
the conditions of soil moisture (%, from 0 to 30 cm
of depth) and average temperature (° C,
fortnightly) limits expression of max growth rates.
We considered two levels of analysis, i. by
condition, and ii. per genotype and condition.
69
Results
Significant differences in the magnitude of
máximum leaf área growth rates were observed
between irrigated and no irrigated conditions
(Fig. 2).
Fig 2. Effect of soil moisture content and fortnightly average
temperature on the average increase leaf area branch
during the fisrt growing season generally for all genotypes.
Colors showing foliar increases deciles (cm2/fortnightly).
Under irrigation, despite the narrower range of
moisture, two optimal boundaries of soil moisture
and temperature were observed for maximum
expression of leaf area growth rates. The first, at
a rate of 30% soil moisture and 21.5 °C, and the
second at a rate of 26% humidity and 25.5 °C
(Figure 2). Under no irrigation only one optimum
area was observed for peak growth rates.
However, this was a broader response surface
(12-24% moisture and 18.5-25 °C) showing that
the magnitude of the highest growth rate decile
under no irrigation was significantly lower than
the first decile under irrigation (fig. 2).
Similar patterns of soil moisture conditions and
high temperatures effects on leaf growth rates
were observed for E. globulus genotypes but
also for E. nitens x E. camaldulensis and
E.camaldulensis x E. globulus genotypes.
The greater range of humidity and temperature
of growth was observed on E.nitens x E. globulus,
E. nitens and E.nitens x E. camaldulensis
genotypes. However, these genotypes also
showed less amplitude for optimal growth (top
decile) under irrigation (fig. 3).
Fig. 3. Range of soil moisture and average temperature of
expression the maximum foliar growth rates fortnightly (top
decile) by genotype and hydric availability condition. Colors
indicate representative genotypes.
Conclusions
Leaf area growth rates of Eucalytpus genotypes
respond positively to favorable conditions of
water availability at this Mediterranean climate
site. However, genotype peak growth rates differ
between irrigated and no irrigated conditions
suggesting as expected larger soil water control
at no-irrigated conditions, but more atmospheric
demand influence at irrigated conditions.
Acknowledgments
Funding for this study was granted by the National
Fund for Scientific and Technological Development,
Project FONDECYT 1085093. Also we would like to
acknowledge the valuable support of Forestal
Mininco S.A., and Conicyt Scholarship for Master's
Studies in Chile of Pincheira, M.
Bibliography

Almeida, A. Siggins, A. Batista, T. Beadle, C. 2010. Mapping the
effect of spatial and temporal variation in climate and soils on
Eucalyptus plantation production with 3-PG, a process-based
growth model. Forest Ecology and Management 259: 1730-1740.
70
The effect of contrasting water availability conditions on water use efficiency
during early development of Eucalyptus globulus, Eucalyptus nitens
genotypes and Eucalyptus camaldulensis hybrids in central Chile.
Manuel Acevedo 1,2,3 *, Rafael Rubilar 1,2, Matías Pincehira 1,2,4,Veronica Emahrt 4 y Oscar Mardones 4
1
Forest Productivity Cooperative, 2 Universidad de Concepción - Fac. Cs. Forestales, Chile. 3Instituto Forestal de Chile. 4 Forestal
Mininco. *[email protected] - [email protected]
ABSTRACT
METHODS
The effect of spring-summer irrigated and no
irrigated conditions were evaluated on water
use efficiency in E. globulus, E. nitens genotypes
and E. camaldulensis hybrids during early
development (3-4 years old). Main results
indicate that monthly water consumption is not
significantly different among genotypes.
However, at species level, irrigation induced a
33% and 28% increase in water use efficiency
(WUE) in E. globulus and E. nitens, respectively,
while WUE in E. camaldulensis was 33% lower.
In the south central valley of the Province of Bio
Bio, Chile, a clonal trial was established under
two contrasting water availability conditions
(with and without irrigation spring-summer
irrigation). The experiment was established as a
split plot block design with three replicates under
each irrigation treatment. The experimental unit
considered a line of four ramets for 10 selected
Eucalyptus genotypes (Table 1) .
Table 1. Description of the genetic material evaluated.
INTRODUCTION
Even though the amount of water used by
Eucalyptus
plantations
is
an
important
ecological and social subject around the world,
the WUE of these forest is also very important,
because wood production is a result from both
efficiency and water consumption (Stape et al.
2004; Binkley et al. 2004). Under the current
scenario of climate change, where an increase
in temperatures and a decrease in precipitation
is expected in many areas of the world (Howe et
al. 2005), it is very important that the
identification of genotypes or clones that are
capable of maintain higher growth rates
associated to lower water consumption values is
recognized (Dye, 2013; Albaugh et al. 2013)
OBJECTIVE
Assess the effect of contrasting water availability
induced via spring-summer irrigation on E.
globulus, E. nitens and E. camaldulensis hybrids
water use efficiency during early development.
Sap flow density was monitored using the
method of constant heat proposed by Granier
(1985, 1987). Thermal dissipation probes were
installed in 60 trees (10 genotypes x 3 ramets x 2
watering conditions). The evaluation period was
between October 2010 and July 2011 (304
days). Monthly height and diameter growth at
10 cm from the ground was measured and
used to calculate estimate individual volume
index. Water use efficiency (WUE) was
determined as the slope of the lineal relationship
between monthly volumetric index vs. water
consumption during the measurement period.
71
(a)
(b)
(c)
volumetric increment at individual tree level was
observed under both irrigation conditions for E.
globulus (egh1, egm1, egm2 y egl1), E. nitens
(en y enxeg) and E. camaldulensis (enxec y
egxec). For E. globulus (Fig. 1a) WUE for no
irrigated conditions (rainfeed) was 0.0225
(R2=0.84) and for
irrigated 0.0299 m3m-3
(R2=0.90). Irrigation significantly increased (P <
0.01) WUE in 0.0074 m3m-3 which represents a
33% increment. For E. nitens (Fig. 1 b) WUE
reached 0.0348 for no irrigated (R2=0.93) and
0.0444 m3m-3 for irrigated conditions (R2=0.97)
for rainfed and irrigated conditions. Irrigation for
E. nitens also increased significantly WUE (P <
0.01) in 0.0096 m3m-3 which represents a 28%
improvement. Contrastingly, WUE for E.
camaldulensis genotypes (Fig. 1 c) reached a
value of 0.0185 (R2=0.94) for no irrigated
(rainfeed) trees and a value of 0.0128 (R2=0.87)
m3m-3 for irrigated conditions. In this case,
irrigation significantly decreased WUE of E.
camaldulensis (P < 0.01) in 0.0057 m3m-3, which
corresponds to a 31% decrease. This result is very
interesting considering that all species showed
larger growth at irrigated conditions, and based
on how usually these species are selected for
dry environments a general understanding
indicates that from higher to lower resistance to
drought these species should rank E.
camaldulensis > E.globulus > E.nitens.
CONCLUSIONS
Fig 1. Linear relationship between water
consumption (m3 tree-1 month-1) and volumetric
increment (m3 of wood tree-1 month-1) at
individual level of E. globulus (a); E. nitens (b)
and E. camaldulensis genotypes by irrigation
treatments. RMSE, root means square error; R2,
determination coefficient. **; Significant for
P<0.01; *** P<0.001.
RESULTS
At this early stage of development, E. nitens
showed the highest WUE from all evaluted
species at this site, and in E. globulus showed a
higher WUE than E. camaldulensis.
A highly significant linear relationship (P < 0.001)
between monthly water consumption and
WUE assessed through individual volume
depends on the species, indicating the E. nitens
is more efficient than E. globulus, and these in
turn are more efficient than E. camaldulensis.
Evaluation of WUE under irrigated and nonirrigated conditions showed that E. globulus and
E. nitens was positively affected by irrigation.
Contrastingly, E. camaldulensis showed a
decrease in WUE at irrigated conditions.
ACKNOWLEDGMENTS
The authors of this research would like to
acknowledge the support of CONICYT graduate
scholarship, FONDECYT Project Nº 1085093,
Forestal Mininco Company SA., the Forest
Productivity Cooperative, the Soil, Nutrition and
Forest Productivity Laboratory at the Faculty of
Forest Sciences at Universidad de Concepción,
72
Physiological response of Eucalyptus globulus, Eucalyptus nitens and
selected E. hibrids to water stress
María Silva F. 12 , Rafael Rubilar P. 12 , Juan Espinosa
1
12
Forest Science School, Universidad de Concepción, Chile.
Introduction
Eucalyptus plantations have expanded around
the world given the diversity of species and
environments in which they may be grow.
According to the National Forestry Institute
(INFOR 2013) in Chile, the area of Eucalyptus
plantations reaches over 700 thousand hectares,
corresponding to 32% of the total area of
planted forests in the country. The main planted
species planted are Eucalyptus globulus and E.
nitens. During the last decade genetic
improvements have focused on increasing
productivity and adaptation for increasing
limiting resources for growth. Over the last years
hybrids of these two species have been also
found to reach high growth rates with desired
wood quality and better performance under
water deficit. However, evaluation of the
physiological behavior that leads to water use
and growth under stress of more productive
genotypes has not been considered, and studies
are needed that include these variables under
conditions of water deficit. A key parameter of
interest in evaluating the response of genotypes
under these conditions is the efficient use of
water, which may be explored at early stages of
development by evaluating leaf level gas
exchange
measurements
considering
transpiration (loss of water by plants) and
photosynthesis (CO2). Thus, the efficient use of
water, especially the transpiration efficiency
may be determined as the assimilation of CO2/
respiration (A / E, mol CO2 / mol H2O).
Objective
To compare Eucalyptus globulus, E. nitens and
E.globulus x nitens genotypes instantaneous gas
exchange variables response to progressive
water deficit.
Methodology
Ten Eucalyptus genotypes including Eucalyptus
globulus, E. nitens and E. globulus x E. nitens
hybrids (Table 1) were grown on 110 cm3
containers on composted pine bark substrate at
the nursery of Forestal Mininco S.A. After four
months of development, 36 plants of each
genotype were selected considering similar
morphological features (height, root collar
diameter & number of leaves) y lack of any
damage. Plants were grown at the nursery of the
Faculty of Forest Sciences at Universidad de
Concepción. Plats were replanted and
acclimated on foam containers of 0,033 m3.
Table 1.– Genotypes included in the study
Especie
N° Genotipos
Códigos
Eucalyptus globulus
6
Eg1, Eg2, Eg3, Eg4, Eg5, Eg6
Eucalyptus nitens
2
En1, En2
E. globulus x nitens
2
Eng1, Eng2
Three replicates of 12 plantas were treated as an
experimental unit. Plants were fertilized and
irrigation to maintain substrate to field capacity
conditions for two weeks before trial
measurements started. After the acclimation
period treatments were established considering
a: Control treatment (C) where substrate was
maintained at as close as possible to field
capacity and a Progressive Irrigation Reduction
(S) treatment were plants were allowed to dry
until a final key –3.0 MPa drop in soil water water
potential. Containers were evaluated daily on
soil water content between 2 y 4 pm by TDR. Soil
water content was evaluated in order to check
critical points of plant assessment at 0 to -0,3
MPa, -1,5 MPa and –3,0 Mpa for gas echange
detailed
measurements. Gas
exchange
measurements considering photosynthesis (A)
and Transpiration (E) were obtained using a
LICOR 6400 XTR equipment with CO2 control.
Water use efficiency was calculated as A/E
(μmol CO2 mol-1 H2O)
73
Results
Progressive water stress reduced stomatal
conductance in all studies genotypes (Figure 1),
this may influence on photosynthesis rates and
genotypes yield. Stomatal conductance and
photosynthesis parameters suggest that stomatal
conductance may play a role as a key variable
predicting plant water stress response to soil
water reduction (Figure 2).
Instantaneous water use efficiency showed a
negative non-linear relationship with leaf
stomatal conductance. This directly affected
reductions in rates of transpiration under
imposed water stress conditions (Figure 3). Better
adapted genotypes under these conditions will
show less variation to these parameters showing
regular (les change) rates of photosynthesis and
will reduce transpiration rates by better
regulation of leaf stomatal control. However, it
may be taken into account that reduced
photosynthesis and transpiration rates will affect
growth affecting genotype performance.
Figure 2. Relation between photosynthetic rate (A: μmol
CO2 m-2 s-1)and stomatal conductance (mol H2O m-2 s-1) on
10 genotypes of Eucalyptus sp under progressive water
stress treatment (S)
Future Steps
A more detailed analyses will be performed
considering genetic interactions to establish
species relationships between gas exchange
parameters, growth rates under stress and
biomass
accumulations,
and
determine
genotypes better or with more potential
adapted to growth and rank in water use
efficiency.
Figure 3. Relation between water use efficiency (A/E: μmol
CO2 mol-1 H2O) and stomatal conductance (mol H2O m-2 s-1)
on 10 genotypes of Eucalyptus sp under progressive water
stress treatment (S)
Acknowledgements
Funding support for this research was provided by
FONDEF D 11I1161: EUCAHYDRO. Project
Logistic and infrastructure support was provided by
Laboratorio de Suelos, Nutrición y Productividad
Forestal de la Facultad de Ciencias Forestales de
la Universidad de Concepción.
Figure 1. Variation of the stomatal conductance (mol H2O m-2
s-1) per genotype under progressive water stress treatment (S)
74
Populus Special Study - Identifying cottonwood and hybrid poplar clones
with the highest productivity on two sites in the SE US.
Nathan Thomas, JL Stape, Steve Kelley
Background
The genus Populus contains one of the
fastest growing tree species in temperate
climates.
Eight species of Populus are
native to the United States with many
regions of natural hybridization. Currently, in
the United States, poplar plantations are
located in the Pacific Northwest, Mississippi
Alluvial Valley, and the Northeast. However,
now there is an increased interest in Populus
across the SE US, because of the effect that
renewable energy goals have created by
increasing the demand for woody biomass.
New Populus clones are being investigated
on sites across the SE US for their productivity
and disease resistance.
This study will
examine 52 clones of cottonwood and
hybrid poplar planted on two sites, North
Carolina and Mississippi.
Figure 2. Fertilizer application.
Study Objectives
1. To identify poplar clones with the highest
growth, and quantify their productivity.
Experimental Design
2. To estimate Light Use Efficiency, Water
Use Efficiency and Nutrient Use Efficiency
for the most productive clones.
The Raleigh site design is composed of 8
reps, with 52 clones per rep, and each rep
randomized.
3. To evaluate any G x E interaction;
The Columbus site design has 10 reps, with
52 clones per rep, and each rep
randomized.
Materials and Sampling
The site in Columbus, MS, and Raleigh, NC,
were both planted in March, 2014 with
genetically identical unrooted cuttings.
The 52 clones are comprised of 47 Populus
deltoides, 4 Populus trichocarpa x deltoides,
and 1 Populus deltoides x maximowiczii.
Figure 1. Location of Field sites in Raleigh, NC, and
Columbus, MS.
Measurements will be taken for two growing
seasons, in Fall 2014 and Fall 2015.
75
The objective of these measurements is to
calculate the volume of hybrid poplar and
its disease resistance.
Volume will be
calculated by measuring the diameter at
breast height (DBH) and total height of
each individual tree. Also, the diameter at
root collar (DRC) will be measured, since not
all trees have developed a DBH by the first
measurement period.
Disease will be
determined by assessing the canopy health.
The trees will be given a numerical value of
1 through 4 depending on the severity of
damage.
Preliminary Results
Average height for the top 15 clones in Raleigh,
Columbus
Raleigh
Figure 4. Clone 7388 in Raleigh and Columbus. Both
trees are over 4.6 meters.
the survival is 100% for the top 10 clones. In
Columbus the average survival per clone is 90%.
The difference in growth between clones is very
obvious in the field, and can be seen in Table 1,
Even in just the top 10 clones the difference in
height and volume is significant.
Figure 3. The 15 tallest clones at the Raleigh,
site
NC
NC, can be seen in Figure 3. The average height
is 3 meters with the best clone averaging 3.7
meters. The tallest clone (7388) at each site is
shown in figure 4. Table 1, shows that In Raleigh
Table 1. Summary measurements comparing the
top 10 clones in Raleigh and Columbus.
Clon Avg. Values (Raleigh) Clon
e
e
Vol Ht (m) Surv. %
In Table 1, 5 of the clones are in the top 10 on
both sites, showing their adaptability across sites.
Whereas some of the clones that were
productive on one site, had very poor growth on
the other site.
Next Steps
Measurements will be taken again following the
2015 growing season in order to calculate the
growth and assess the resistance to disease.
Avg. Values
Vol Ht (m) Surv. %
7388
8019
2.51
2.46
3.76
3.64
100
90
185
229
1.93
1.68
2.64
2.64
100
100
230
2.32
3.55
100
230
1.61
2.68
100
140
339
2.23
2.20
3.06
3.07
100
100
348
187
1.52
1.50
2.18
2.34
100
90
185
116
2.19
2.14
3.18
2.82
100
100
7388
188
1.46
1.45
2.75
1.99
90
90
400
2.13
2.90
100
8019
1.40
2.22
100
7595
188
2.13
2.12
2.92
3.03
100
100
200
176
1.35
1.34
2.11
2.37
80
80
Acknowledgements
This project has been realized by the
support
of
the
Forest
Productivity
Cooperative and its cooperators. The IBSS
76
Special Studies
78
Tolerance of Eucalyptus Clones to Hydric, Thermal and Biotic Stresses - TECHS
Jose Luiz Stape, Dan Binkley, Otávio Campoe, Clayton Alvares
Justification
The two major environmental stressors for
Eucalyptus plantations, both in traditional or new
forest frontiers, are: i) water stress (especially in
regions of the midwest, north, northeast and the
southeast of Brazil), and ii) thermal stresses
related to high temperatures (above 36oC) in
tropical Brazil and low temperatures (below 5oC)
in southern Brazil and Uruguay.
The understanding of the ecophysiological
processes driving different responses and levels
of tolerance of Eucalyptus clones to drought
and extreme temperatures provide tools to
quantify the genotype-environment interactions,
its effects on wood production and improved
zoning of different clones across climatic regions.
The Experimental Design
The experimental design is a combination of
genotype variation, water manipulation and
planting spacing across a wide climatic range
(Figure 1), providing a research platform to
assess genotype-environment interactions.
Experimental sites: The 36 TECHS sites were
planted during 2012 across Brazil and Uruguay,
and are divided in Tropical (above Capricorn
line) and Subtropical (below Capricorn line).
Genotypes: 18 different Eucalyptus clones,
currently planted by companies, were selected
from tropical and humid regions (n=3); tropical
and dry regions (n=4); subtropical from cold
regions (n=7), and intermediary and plastic
(n=4). The clones are pure and hybrids of E.
benthamii, E. camaldulensis, E. dunnii, E.
globules, E. grandis, E. saligna, E. tereticornis, E.
urophylla.
Water manipulation: At least 4 genotypes at
each site have one plot with regular rainfall, and
paired plot with throughfall exclusion to reduce
water reaching the soil to 70% of actual rainfall,
by covering 30% of the soil surface (Figure 2).
Thus, selected genotypes, at each site will be
subject to two levels water availability.
Figure 1. Distribution of the 36 experimental sites across Brazil
and Uruguay. The color scale represents the ratio of historic
annual potential evapotranspiration (ETPo) and annual
precipitation (PPT).
Scientific Questions
The main scientific questions of the TECHS
project concerns about the effect of
environmental stresses (drought and thermal) on
productivity of Eucalyptus plantations:
i) How the productivity of different clones are
affected by climate types ?
ii) How different clones at different planting
spacing respond to water stress ?
Figure 2. Throughfall exclusion system covering 30% of the plot
soil surface, installed across 2013.
79
iii) Which ecophysiological mechanisms allow
some clones to tolerate or avoid water and
thermal stresses?
iv) How much of the clone behavior across sites
is under genetic control (G x E) ? months, at the beginning and at the end of the
dry period provide detailed information about
productivity and its relation with canopy
development (Figure 5).
v) How the clones in different environments and
under different stress levels, behave in terms of
their susceptibility to pests and diseases (biotic
stress)?
Initial results
Wood volume at two years after planting
showed a wide range of clones productivity for
tropical and subtropical sites and also among
clones within each site (Figure 3).
Ranking clones among sites show patterns of
behavior denoting G x E interactions. Some
clones show growth stability across sites, and
some clones are highly responsive to the climatic
gradient (Figure 4).
Figure 4. Ranking of volume for tropical sites and clones.
Clones with similar colors among sites are more stable, and
Acknowledgements
To the 26 companies of the IPEF-TECHS project
for financial and technical support. For more
information: www.ipef.br/techs
Figure 5. Two contrasting clones side by side showing different
canopy structure and development behavior.
Figure 3. Distribution of volume for all sites (upper) and for all
clones (lower) at 2 years after planting . Red dotted line
represents the average
Measurements of crown depth and LAI every six
80
Carbon fluxes on Eucalyptus plantations: genetic and climatic effects
Otávio Campoe, Jose Luiz Stape, Dan Binkley, Mike Ryan
Introduction
Sites and Clones Selection
Wood production is the ultimate goal for
forestry companies and represents a significant
but also variable fraction of gross primary
production (GPP) in Eucalyptus plantations. The
carbon (C) absorbed by photosynthesis is
partitioned to different fluxes above- and
belowground (Figure 1).
Four of the 36 experimental sites of the TECHS
project were selected comprising a wide range
of climatic conditions, with cumulated annual
rainfall ranging from
1000 to 1800 mm,
cumulated annual water deficit from 0 to 700
mm and mean annual temperature from 18 to
28 oC (Figure 2).
Despite the current understanding about
carbon fluxes and partitioning, knowledge
about the effects of climatic seasonality,
genetic variation and their interaction is still
limited. Better knowledge of fluxes and
partitioning to woody and non-woody tissues in
response climatic and genetic variations,
focusing on mechanisms driving the C budget of
these plantations, could provide opportunities to
increase forest productivity.
Figure 1. Diagram of carbon partitioning to above and below
ground components. Adapted from Litton et al. 2007.
The TECHS project (www.ipef.br/techs), a
combination of genotype variation and water
manipulation across a wide climatic range
(Figure 2), provides a research platform to assess
genotype-environment interactions over carbon
fluxes and partitioning.
Our study aim investigating how C fluxes
and partitioning are affected by genotype
variation, climatic seasonality and their
interaction in different regions of Brazil.
Figure 1. Location of the four carbon budget sites (large white
circles) among the 36 TECHS´ experimental sites (small black
circles) across Brazil and Uruguay. The color scale represents
the ratio of historic annual potential evapotranspiration (ETPo)
and annual precipitation (PPT).
Five contrasting clones regarding
productivity and root and canopy development
from different genotypes (pure and hybrids of E.
camaldulensis, E. grandis, E. saligna, E.
tereticornis and E. urophylla) were selected for
complete carbon budget assessment.
The study started at 2 years and will be
performed up to 4 years after planting,
capturing carbon fluxes and partitioning at
maximum current annual increment period of
the rotation.
81
Figure 4. WNPP for clones (gray, average
Figure 3. Monthly soil CO2 efflux (right) and Litterfall (left) for clones (top, average among sites) and for sites (green, average
among clones).
among sites) and for sites (bottom, average among clones)
Methodology
Assessments of annual aboveground net
primary production (ANPP: stem, leaf, and
branch production), total belowground C flux
(TBCF: sum of root production and respiration
and mycorrhizal production and respiration, eq.
2) and GPP (computed as the sum of ANPP,
TBCF and estimated aboveground respiration)
started 2 years after planting (October 2013) at
four sites. The measurements will be performed
for the next 2 years (until October 2015). The
methodology applied was based on Giardina
and Ryan (2002) and Ryan et al. (2010).
Partial Results
Carbon fluxes were more affected by
climatic differences among sites than genetic
differences among clones (Figure 3). The
cumulated soil CO2 efflux for clones over the
period ranged from 1405 (AEC 144) to 1493 g C
m-2 year-1 (CNB 10), and cumulated litterfall
ranged from 232 (CNB 10) 329 g C m-2 year-1
(SUZ MA2003). However, differences among sites
were higher.
The cumulated soil CO2 efflux for sites over the
period ranged from 1159 (20-IPB) to 1901 g C m-2
year-1 (22-KLT), and cumulated litterfall ranged
from 208 (20-IPB) 378 g C m-2 year-1(30-VMT).
Wood production (WNPP, Figure 4) showed a
gradient among clones (from 460 to 1090 g C m2 year-1) and sites (from 540 to 1112 g C m-2 year1).
Conclusions
Studies of C budget in forest plantations
are required to improve our understanding of
the processes driving wood production and
carbon fixation, and develop our capability to
understand forest responses to climatic and
genetic interactions.
Acknowledgements
To the 26 companies of the IPEF-TECHS
project for financial and technical support. For
more information: www.ipef.br/techs
References
Giardina, C.P., Ryan, M.G., 2002. Ecosystems 5, 487–499.
Ryan, M.G. et al. 2010. Forest Ecol. Manag. 259, 1695–1703.
Litton, C.M. et al. 2007. Global Change Biol. 13 , 2089-2109
Figure 5. One of the
TECHS clones showing
different behaviors on
each C budget site. Pictures show a gradient of
water availability, from
no water deficit on the
left to intense water deficit on the right.
82
TECHS Climate: Eucalyptus Tolerance to the Hydric and Thermal Stresses
Clayton A. Alvares, JL Stape, Otavio C. Campoe
Justification
Clonal plantations are the standard for fastgrowing Eucalyptus plantations, in both
operational production areas, and in new
afforested areas. Experience from past rotation
provides a basis for estimating risks and benefits
of silvicultural decisions, but expanding
afforestation in new frontiers poses higher risks of
production due to environmental stresses
different from those where the clones were
selected.
Among the major environmental stresses to
Eucalyptus, two stand out because of their
association with the new forest frontiers, namely:
i) water stress (especially in regions of the
midwest, north, northeast and the southeast of
Brazil), and ii) the thermal stresses related to high
temperatures (above 36º C) in tropical Brazil and
low temperatures (below 5º C) in southern Brazil
and Uruguay.
Even in traditional areas of planting, the effects
of variation in water regime and heat due to
climate
change
will
require
improved
understanding of the sensitivity of clones to these
stresses.
produced (Alvares et al. 2013). Since the
weather information each TECHS be used in
almost all their evaluations, it is necessary to
develop a database of climate for each site.
Methodology
We use meteorological data from Brazilian
Meteorological National Institute (INMET) to
prepare the weather database of TECHS. The
INMET maintains a network of meteorological
stations throughout Brazil, and its data can be
accessed in real time via internet. Based on the
localization of the sites and meteorological
stations we selected the pairs TECHS / INMET
similar in terms of topography and historic
climate (Alvares et al., 2013) (Figure 1). Thus,
from 2012, the meteorological data are
collected in each TECHS at hourly scale. Some
indices such as potential evapotranspiration,
water balance based on Thornthwaite and
Mather method, are calculated at daily scale.
All information was consisted and stored in
webpage
project.
The
rainfall
and
evapotranspiration were plotted in diagrams to
compare them annually. Furthermore, TECHS
sites were classified according Köppen system.
The Cooperative Program on Clonal Eucalyptus
Tolerance to the Hydric and Thermal Stresses
(TECHS) design aims to “understand the variation
of forest adaptability and productivity of a
group of Eucalyptus clones across a wide range
of edaphic-climatic conditions and stocking,
and within each site, the responses to water
deficit.”
General Genetic Material: a group of 18
Eucalyptus clones were selected by a group of
breeders for planting in all TECHS sites. These
clones will represent genetic materials currently
in use in Brazil, but with distinct pedigree
characteristics and susceptibility to hydric and
thermal stresses. With this approach we are
looking for different ecophysiological behaviors
within adequate levels of productivity.
Given the wide range of Brazilian climatic
conditions, each TECHS site was classified into
Tropical and Subtropical, based on detailed
Köppen
climatic
classification,
recently
Figure 1. TECHS sites and weather stations from Brazilian
Meteorological National Institute (INMET).
83
Results
F i
The TECHS sites cover a wide range of climates in
South America, from semi-arid to super humid.
The semi-arid sites present annual rainfall 500 to
700 mm, while the super-humid sites between
1800 and 2000 mm. The year 2013 was more
humid in most of the TECHS, as shown in Figure 2.
gure 4. Sequential water balance for Special TECHS sites.
Figure 4 shows the sequential water balance for
special sites, where it can be noticed the
increasing water deficit in the heading north to
south of Brazil. The first two years of TECHS, 2012
and 2013, showed wide climate variability, as
shown in Figure 6. A few sites keeps the same
climate type in the first two years in relation to
historical climate (Figure 5), while most of them
the change has been observed.
Reference
Alvares, et al. 2013. Köppen's climate
classification map for Brazil. Met. Z., 22:711-728.
Figure 2. TECHS’ climate diagram of 2012 and 2013.
Figure 3. Köppen climate classification and TECHS sites.
Acknowledgements
To IPEF-TECHS and INMET.
Figure 5. Annual climatic variability (2012, 2013) of the TECHS
regarding historic climate.
84
Eucahydro: tools for early assessment of Eucalyptus genotypes for water use
efficiency, water consumption, efficient use of water and drought resistance
Trial objective
Developing a model of water sustainability
analysis for Eucalyptus genotypes (EUCAHYDRO).
EUCAHYDRO will integrate ecophysiological and
genomic tools to evaluate the efficiency, water
consumption
and
drought
resistance
of
genotypes of Eucalyptus globulus, E. nitens and
hybrids in early stages of development (nursery
and/or first year of development).
Treatment and experimental design
The study consider the establishment of two
contrasting climatic site conditions located in the
Maule and Bio-Bio regions. Selected sites differed
mainly on relative humidity and vapor pressure
déficit conditions in Chile (Figure 1). A total of 30
Eucalyptus genotypes of high , medium and low
productivity for E.globulus and Eucalyptus nitens,
and hybrids of Eucalyptus camaldulensis highly
selected genotypes were considered. Two
conditions are being evaluated as study
treatments considering irrigated (1000 mm) and
no irrigated conditions (~700 mm) evaluated at
establishment and until 3 years old.
Figure 1. Geographical
Eucahydro in Chile.
location
of
Project
Studies at each site were established with three
blocks (replicates), each block consisting of two
main plots with both contrasting irrigation system,
where each main plot will contain 30 subplots
each of 5 x 5 tres (25 trees per plot) of each
selected genotype asigned randomly to
experimental units.
Ecophysiological Assessments:
The project involves continuous measurements
of individual tree sap flow , soil moisture and
climatic parameters at each site (Figure 1).
Seasonal measurements on the genotypes
consider tree growth, instantaneous gas
exchange
(photosynthesis
and
evapotranspiration), xylem potential, leaf area
and biomass allocation (Figure 2). Moreover,
collection of samples for genomic analyses will
be obtained during the summer to characterize
genotypes signals.
Figure 2. Summer measurements of intrinsic-water
use
efficiency
(photosynthesis/stomatal
conductance) at the site Yumbel (upper panel)
and Constitucion (lower panel).
85
Figure 4. Measurement campaing of growth (A),
gas Exchange (B) and sap flow (C)
winter measurements in some genotypes.
Figure 3. Summer measurements of collar
diameter (upper panel) and height (lower panel).
Graphs shows the data for the irrigated and water
-stressed treatment at both sites (Yumbel and
Constitucion).
Results
Preliminary results showed differences in water
use efficiency (WUE) among the genotypes
(Figure 2). At the Constitucion site, there was a
general trend that genotypes having higher
WUE in the irrigated treatment also had higher
WUE in the water-stressed treatment. This trend
was not consistent at the Yumbel site. WUE
was higher at Yumbel than Constitucion.
Overall, Photosynthetic rate (Asat) ranged
from 13 to 24 micromole/m2/s in the summer
measurements.
Asat was higher in the
irrigated treatment than the stressed
treatment, but differences did not persist in the
Water stressed decreased the collar diameter
and height at both sites (Figure 3).
The
differences among the genotypes were higher
for collar diameter than tree height. The high
coefficient of correlation (ranging from 0.7 to
0.9) among the growth traits in both irrigated
and water stressed treatment (Figure 3) suggests
a great stability of the genotypes to the water
availability. However, the genotypes changed
in the rank in growth traits between sites
suggesting a significant GxE interaction.
Future direction
New measurements will be obtained annualy to
assess plant development. Other species such as
Eucalyptus will be tested under this scheme.
Moreover, gene expression and genetic isolation
and molecular characterization are also
monitored to understand genotypes genomic
responses to water stress (drought) at these early
stages .
86
Special study: Potential Productivity of Eucalyptus maidenii and E.
globulus in Uruguay.
Objective
To determine the potential response of
Eucalyptus globulus (Eg) and E. maidenii (Em)
plantations established free conditions weed
competition and high levels of nutrient
availability.
To determine the response that validates the
need to incorporate intensive management
Eucalyptus sp. plantation company.
To identify opportunities for increase the
profitability of plantations established in
southeastern Uruguay.
Figure 1. Geographic location of the Potential
Productivity trials in Uruguay.
in the region is 16.5 º C, with an annual rainfall of
1200 mm.
Treatments and experimental design
To determine the potential responses of
Eucalyptus globulus and Eucalyptus maidenii,
the experiment was set up in a number of stand
having different ages (Table 1) The treatments
consisted of two homogeneous plots, one was
considered a control plot. In the case of new
plantations, the control plot was managed using
standard practices by the company.
The
fertilized plots (potential productivity) were kept
free of competing vegetation for 2 years and will
be fertilized annually or seasonally for two years
(Table 2).
Figure 2. Image of the sites Yatay with treatments
control and fertilized+ weed control.
Fertilized+ weed control
Site description
Seven studies were established in Uruguay by
Farm & Forestry Management Services Cattivelli
(Eg) (2007), Esquenet (Em) (2008), Lemosur(Eg)
(2009), Parma (Em)(2008), Singer (Eg)(2009), Villa
Roble (Eg)(2011) and Yatay (Eg)(2007). The trials
were established in 2012 and were located in
the Provinces of Florida, Lavalleja, Maldonado
and Treinta y Tres for E. grandis and E. maidenii
(Figure 1). Studies were established on a well
(sandy loam) and moderately well (clay loam)
drained
soils,
considered
taxonomically
Brunosoles. The previous use of the region was
mostly grazing. The average annual temperature
Control
87
Table 1. Summary of the growth at treatment application in 2012, and soil chemical analysis per site.
Results
Growth responses for the first two growing
season after treatment application were
variable among sites. Similarly, there had not
been a consistent growth response associated
to the Eucalyptus species. In all the sites the
greater growth increment was immediately after
treatment application, except at the Villa Roble
site (Figure 3). There was not clear responses to
fertilization of the Eucalyptus species as some
sites exhibited lower growth in the fertilized
treatment relative to the control (Esquenet,
Yatay). A negative increment in tree growth
was detected in the control plot at the Villa
Roble site, which presented some mortality and
damages problems that explains such decrease.
Responses in DBH were more variable among
sites than height.
Figure 3. Mean increment in DBH (upper panel)
and height (lower panel) for the first two growing
period after treatment application per site
Table . Fertilization treatments for age and
elemental dose (kg ha-1).
Future direction
To increase our understanding about the
potential productivity of plantations of
Eucalyptus globulus and Eucalyptus maidenii in
southeastern Uruguay, and how they are
affected in different instances by environmental
variables that reduces the effectiveness of early
forestry investments. Future assessments may
corroborate these low response found and state
some guideline for future improvements.
88
Integrated Biomass Supply Systems —Comparative water and nutrient use
in Eucalyptus benthamii and Pinus taeda plantations
Timothy J. Albaugh, Chris A. Maier, Kurt H. Johnsen, Jose L. Stape, Thomas Fox, and Jim Vose
Study objective
The Southeast Partnership for Integrated Biomass
Supply Systems (IBSS) is supported by the
Agriculture and Food Research Initiative from
the US Department of Agriculture National
Institute of Food and Agriculture. The overall IBSS
goal is to demonstrate real world solutions to
economically and environmentally sustainable
production and conversion of biomass to biofuel
in the southeast United States.
We are working with the USDA Forest Service to
complete our part of the grant which is focused
on understanding the biology of Eucalyptus to
examine the species as a biofuel crop
alternative in the region. Eucalyptus is grown in
other regions of the world and can be very
productive however it has been associated with
high water consumption. Water requirements for
Eucalyptus plantations in the SE US are largely
unknown. Generally, then, our objective is to
quantify water as well as nutrient use for
Eucalyptus benthamii.
More specifically our objectives are to:
1.Quantify stand water use, water use efficiency,
nutrient use, nutrient use efficiency and tree
growth for nine year old stands of E.
benthamii and Pinus taeda.
2.Define mechanisms for regulating stand water
use.
Mechanisms examined will be
physiological
(photosynthesis, stomatal
conductance), morphological (leaf area,
specific leaf area, crown shape), and
structural (biomass allocation, rooting
depth).
3.Parameterize the 3-PG process model for E.
benthamii in the US.
Treatments and experimental design
We are working at the MWV Ravenel Nursery Site
where adjacent nine-year-old E. benthamii and
P. taeda stands are planted. Twelve trees of
each species will be instrumented with Granierstyle sap flow probes (Figure 2) and automated
dendrometer bands to measure sap flow and
diameter growth for one year. After the sap flow
measurements
are completed we will
destructively harvest a subset of trees to quantify
above and below ground biomass and rooting
depth.
Nutrient determinations will be
completed on tissue samples from the biomass
harvest.
Plots have been established
surrounding the sap flow trees to measure stand
growth and permit scaling to an area basis. The
modeling work requires a full set of
Figure 2. Sap flow probes and sap
flow probes inserted into pine tree
stem.
Figure 1. Nine-year-old Eucalyptus benthamii stand in South
Carolina.
89
meteorological data so temperature, radiation,
precipitation, relative humidity, soil moisture and
wind speed and direction will be measured on
site. with a subset of the plots at the site. .
Current status
The trees were instrumented in January 2013 and
the meteorological equipment was installed in
April 2013. Our site is now one installation of the
RAFES network which is fully described in a
separate two page report. We are monitoring
and maintaining the instrumented trees. We
plan for a March 2014 biomass harvest. .
Results
Daily transpiration on a ground area basis was
typically higher for the E. benthamii than for the
P. taeda from January 20 to May 12, 2013 (Figure
3). The E. benthamii transpired a significantly
greater amount of water (46 mm 122 day-1)than
Figure 5. Daily transpiration per unit leaf area versus average
daily daylight hour vapor pressure deficit for nine year old
Eucalyptus benthamii (black circles) and Pinus taeda (red
circles) for day of year 20 to 132 in 2013.
the P. taeda (Figure 4). Daily transpiration per
unit of leaf area was higher for E benthamii than
P. taeda over the range in vapor pressure deficit
during this time period (Figure 5). Soil moisture
was lower to a deeper depth in the E. benthamii
than it was in the P. taeda. In general then the
E. benthamii appears to be using more water
and depleting soil water faster than the P.
taeda.
Even though water use by the
eucalyptus may be greater than the pine, the
water use efficiency (m3 of wood produced per
m3 of water transpired) has not been
determined at this point.
Future plans
Figure 3. Daily transpiration on a ground area basis for nine year
old Eucalyptus benthamii (black circles) and Pinus taeda (red
circles) for day of year 20 to 132 in 2013.
We plan to continue measurements through
March 2014 to include an entire year of data. At
that time we will complete the destructive
sampling to estimate biomass production,
nutrient use and rooting depth. When this work is
completed these same measurements are
planned for a younger stand on eucalyptus and
pine to help develop how these traits may
change during the rotation.
Acknowledgements
The IBSS Partnership is supported by AFRI
Competitive Grant #2011-68005-30410 from
USDA NIFA.
MWV has provided access to the stands and
additional support to complete the research
The USDA Forest Service provides support for the
RAFES study.
Figure 4. Total transpiration on a ground area basis for nine year
old Eucalyptus benthamii (Euc) and Pinus taeda (Pine) for day of
year 20 to 132 in 2013.
90
Geodatabase System: spatial information
for forestry and environmental applications
Clayton A. Alvares, Jose L. Stape, Thomas Fox, Rafael Rubilar
Introduction
Demand
for
reliable
spatialized
information in forestry is increasing and more
databases are freely accessible on the Internet.
However, these sources are commonly dispersed
and have generalized information that do not
meet specific requirements for forestry and
environmental applications.
All maps were spatialized using geostatistical
tools and GIS that best fit their spatial behavior.
The project to create a system of specialized
maps for use in forestry began in 2010 as a joint
project between IPEF and FPC due to their many
common interests in Brazil and Uruguay.
As a consequence of this effort, the IPEF's FPC’s
Geodatabase System was established in 2012
providing access to Brazilian spatialized edaphobio-climatic information. These maps are
increasingly necessary for research activities and
forest planning of IPEF and FPC members,
besides the forestry community in general.
Development and current status
Under the concept of database management,
the first version of the Geodabatase System was
organized into three hierarchical levels: country,
topic and subtopic.
Figure 1.
IPEF/FPC Geodatabase
(www.ipef.br/geodatabase)
System
Webpage
The relief subtopics contain maps of elevation,
slope and aspect (slope orientation).
On the IPEF’s platform, the two countries of
interest are Brazil and Uruguay.
The geology subtopics are composed by official
Brazilian geological and soil maps, besides some
soil physics and chemical attributes like clay
content and CEC (fertility).
For both countries, the topics comprehend:
climate, topography, geology, administrative
and ecology (Figure 1).
The political maps are the administrative borders
of Brazil, as a region, state, meso- and microregions and municipalities.
The climate subtopics include: mean monthly
temperature, monthly maximum temperature,
minimum temperature monthly (Figure 3),
monthly precipitation (Figure 4), global radiation
(Figure
5),
relative
humidity,
potential
evapotranspiration and vapor pressure deficit.
The ecological subtopics have maps on climatic
zoning, forest diseases, ecological biomes and
estimated productivity of some forest planted
species (Eucalyptus and Pine).
All climatic variables are expressed in monthly or
annual averages, based on historical data from
1950 to 1990.
Maps are available in different levels of access.
There is a level open to the general public,
which guarantees access to maps in PDF format
on A4 size (21 x 24cm).
91
The website restricted level, which requires
access via the login and password, can be
used only by FPC or IPEF members and allow the
download of the raster– and shape-files.
Figure 4. Example ecological map (eucalypt rust)
Figure 2. Example of climatic map (temperature)
Figure 5. Countries in America with FPC members
Figure 3. Example of ecological map (bioclimatic zoning)
Next steps
The geodatabase is an on-going effort and will
start incorporating maps for all the 10 countries
where the FPC has members.
More applied maps with potential and actual
productivity of Eucalyptus, Pine and other
species, together with remote sensing of leaf
area index and vegetation indices are under
development.
For more information:
http://www.ipef.br/geodatabase
http://www.forestproductivitycoop.org
92
Modeling monthly mean air temperature for Brazil
Clayton A. Alvares 1 , José L. Stape 2 , Paulo C. Sentelhas 3 , José L.M. Gonçalves 3
1 IPEF
and FPC; 2 NCSU; 3 ESALQ/USP
Introduction
Air temperature is one of the major effects of
solar radiation in the lower Earth atmosphere. In
agriculture and forestry, air temperature
together with rainfall are the main factors
defining crop zoning, sowing dates, and the
expected yield levels. The Brazilian territory
comprises a wide area, occupies nearly 48% of
South America, and has a wide climate diversity.
However, the density of weather stations in Brazil
is very low and in some areas there is no stations
at all (Figure 1).
descriptive analyzes, multivariate statistics and
geostatistics.
Figure 2. Location of all studied areas used in the evaluation
of the equations obtained in this study
Figure 1. Location of the Brazilian weather stations used for
air temperature modeling
Usually, the normal monthly mean air
temperature is estimated with the use of multiple
linear regressions, and this method can be as
effective as sophisticated local interpolation
methods, especially when dealing with mean
climatic data. In Brazil, coefficients for linear
models for estimating mean air temperature as a
function of latitude, longitude and altitude are
available for approximately 50% of the territory
(Figure 2). Based on this, our study aims to
develop models for estimating monthly and
annual maximum, minimum and mean air
temperatures for the whole Brazilian territory.
Material And Methods
The study was conducted in several steps that
included
the
data
compilation,
data
consistency, exploratory analysis, compilation in
a geodatadase using a geographic information
system, geoprocessing techniques, static
The complete database used in this study comes
from the Brazilian institutions as National Institute
of Meteorology (INMET), National Department of
Works against Droughts (DNOCS) and Northeast
Development Superintendency (SUDENE), and
also the Food and Agriculture Organization of
the United Nations (FAO/ONU). After the phase
of consistency and exploratory analysis we used
2400 weather stations throughout the Brazilian
territory. The models were developed based in
weather stations 1800 and was used for
validating the other 600 (25% of total). Among
other indicators of fit quality and the models
obtained we apply the ME (mean error), MAE
(mean absolute error), RMSE (root-mean-square
error) and MAPE (mean absolute percentage
error). Finally, and as a second validation, we
compare our national models with regional
models of 16 previously published studies in Brazil
(Figure 2).
Results
The correlations between air temperature and
geographical coordinates and altitude were
well defined: higher latitude, lower temperature,
because of seasonal variation of incoming solar
radiation; higher altitude, lower temperature,
due to atmospheric pressure reduction and air
rarefaction of the air. Longitude showed less
93
effect on temperature variation, since its effect
on air temperature amplitude is associated with
the position of the area in relation to the ocean,
which varies with the regions of the country
(Figure 3).
Serrano Plateau (Santa Catarina state) was the
coldest of Brazil; while the valleys of the Amazon,
Araguaia, and Parnaíba rivers were the hottest
regions (Figure 3).
Figure 3.
Pearson’s correlation coefficients for the
relationships (monthly and annual) air temperature and
altitude, longitude, and latitude
All regression models were significant (α ≤ 0.01)
as well as all independent variables at 1 and 5%.
The indicators statistician showed that the
models were of good quality, but were much
better when the residues were added (obtained
by kriging, Table 1). We evaluated a total of 3.67
108 pixels (1 km2), between the results provided
by the proposed models and other studies
(Figure 2). The overall correlation obtained was
0.93, showing that there is a high consistency
between our national models with regional
models.
Table 1. Errors of the temperature multivariate regression
models using database from fitting and test sets considering
kriging effects for the first validation
Figure 4. Annual mean air temperature maps for Brazil,
generated by models, bias by kriging and the final maps
Conclusions
Spatial and temporal variability of monthly and
annual temperatures in Brazil were properly
modeled through the relationship with latitude,
longitude, altitude and their combinations, using
multivariate regression equations, geostatistical
analysis, and GIS. In these kinds of spatial
modeling the bias should be treated properly
and incorporated into the final maps.
94
Stand-level patterns of carbon fluxes and partitioning in a Eucalyptus
grandis plantation across a gradient of productivity, in São Paulo State,
Brazil
Otávio Campoe, José Luiz Stape, Jean-Paul Laclau, Claire Marsden, Yann Nouvellon
Introduction
Wood production represents a large but
variable fraction of gross primary production
(GPP) in
highly
productive
Eucalyptus
plantations. Assessing patterns of carbon (C)
partitioning (C flux as a fraction of GPP)
bet ween
abov eand
belowground
components is essential to understand
mechanisms driving the C budget of these
plantations. Better knowledge of fluxes and
partitioning to woody and non-woody tissues in
response to site characteristics and resource
availability could provide opportunities to
increase forest productivity.
Our study aimed at investigating how C
allocation varied within one apparently
homogeneous 90
ha stand of Eucalyptus
grandis in Southeastern Brazil.
Material and Methods
We assessed annual aboveground net primary
production (ANPP: stem, leaf, and branch
production, eq. 1), total belowground C flux
(TBCF: sum of root production and respiration
and mycorrhizal production and respiration, eq.
2) and GPP (computed as the sum of ANPP,
TBCF and estimated aboveground respiration,
eq. 3) on 12 plots representing the gradient of
productivity found within the stand (Figure 1).
Carbon fluxes were assessed during the last year
of the rotation before harvesting (from 6 to 7 yrs
after planting) and the C partitioning was
studied on a gradient of productivity.
ANPP = FA + Δ(CW + CF) [eq. 1]
FA: literfall, ΔCW: change in carbon content on
aboveground wood biomass (stem, bark and
branches), ΔCF: change in carbon content on
foliage in the canopy.
TBCF = FS - FA + Δ(CR + CL + CT) [eq. 2]
FS: soil respiration, FA: literfall, Δ (CR, CL, CT):
change of carbon content in coarse roots, litter
layer and stumps.
Figure 1. Experimental area topography and plot
location. The arrow indicates the trend of increase
in soil clay content from ≈20 to ≈ 40% across the
site. Plots are classified by ascending order of
elevation.
GPP = ANPP + TBCF + RP [eq. 3]
RP: aboveground autotrophic respiration was
estimated as a fraction of ANPP (53%)
The methodology applied was based on
Giardina and Ryan (2002) and Ryan et al. (2010).
Results
The spatial heterogeneity of topography and
associated soil attributes (Figure 1) strongly
affected the component fluxes of GPP and C
partitioning across the area. Stand GPP ranged
from 2971 g C m-2 yr-1 to 4132 g C m-2 yr-1 (+39%).
These contrasting plots were also the most
contrasted for ANPP, which attained 1208 g C
m-2 yr-1 versus 1791 g C m-2 yr-1 (+48%), and for
stem wood NPP (554 g C m-2 yr-1 versus 923 g C
m-2 yr-1 (+ 67%).
95
Total belowground carbon flux ranged from 497
g C m-2 yr-1 (low elevation and median GPP) to
1235 g C m-2 yr-1 (higher elevation and high GPP)
and showed no significant relationship with GPP
(P=0.33) or ANPP (P=0.36). Carbon partitioning to
stem production ranged from 0.19 to 0.23,
presenting a trend of linear increase with GPP
(Figure 2). In contrast, the fraction of primary
production partitioned belowground was not
correlated with GPP, while C partitioning to leaf
production showed a significant negative
correlation with GPP (Figure 3).
Figure 3. Linear increase in leaf NPP (A) and
decrease in C partitioning to foliage NPP (B) as a
function of increasing GPP. Numbers represent
plots.
Conclusion
Figure 2. Linear increase in stem NPP (A) and C
partitioning to stem NPP (B) as a function of
increasing GPP. Numbers represent plots.
Comparing the two most contrasting plots in
terms of GPP, the 67% difference in stem wood
NPP was a combined result of higher
photosynthesis and higher C flux.
Stem production across the gradient of
productivity observed at our experimental site
was a combined result of the variability in GPP,
and carbon partitioned to stem NPP.
Comprehensive studies focusing on the
components of C budgets in forest plantations
are required to improve our understanding of
the processes driving wood production and
carbon fixation, and develop our capability to
predict forest responses to environmental
changes.
Acknowledgements
We thank to FAPESP-Brazil, the EUCFLUX ProjectIPEF, Itatinga Research Station – ESALQ/USP and
Floragro Company for field work and laboratory
support.
References
Giardina, C.P., Ryan, M.G., 2002. Ecosystems 5,
487–499.
Partial view of the 7 years old Eucalyptus grandis
plantation studied in São Paulo state/Brazil
Ryan, M.G., et al. 2010. Forest Ecology and
Management. 259, 9, 1695–1703.
96
Stem production, light absorption and light use efficiency between
dominant and non-dominant trees of Eucalyptus grandis across a
productivity gradient in Brazil
Otavio Campoe, Jose L. Stape, Yann Nouvellon, Jean-Paul Laclau, William Bauerle, Dan
Binkley, Guerric Le Maire
Introduction
Brazilian Eucalyptus plantations are some of the
most productive forest plantations in the world,
sustaining mean growth rates of 25 Mg ha-1 year1 (50 m3 ha-1 year-1) over the 4.7 million hectares
planted across the country. To better
understand forest productivity, studies at the
stand scale need to be coupled with tree level
evaluations of the production ecology.
Study design
We measured (from 6 to 7 years after planting)
stem wood dry biomass growth and estimated
light absorption and light use efficiency at the
tree level with a three-dimensional array model
(MAESTRA) in 12 plots within a seed-origin
Eucalyptus grandis plantation. The soil clay
content (from ~20% to ~40%) and topography of
the experimental site generated a natural
gradient in productivity (Campoe et al. 2012).
Radiation absorption for each tree was
simulated with MAESTRA (Medlyn, 2004), a three
dimensional ecophysiological model that
estimates radiation absorption, photosynthesis
and transpiration at the individual tree level.
MAESTRA accounts for the influence of shading
from leaves within a crown and by those of
neighboring crowns. From destructive sampling,
we
measured
tree
crown
structural
characteristics to parameterize MAESTRA (Figure
1).
We investigated the hypothesis that dominant
trees (the 20% largest) are more productive than
non-dominant trees (the 20% smallest) as a result
of greater light absorption and light use
efficiency; and that with increasing productivity
across plots, dominant trees would show larger
increases in light use and light use efficiency in
comparison to non-dominant trees.
Figure 1. Schematic representation (in scale) of the positioning
and size of trees present in plot 1 with elevation of 726 m (A:
frontal view, B: orthogonal view). The red crown trees are the
focal trees used in MAESTRA simulations, and the outer five rows
of green crown trees were included in the simulation only for
their potential shading on focal trees. This figure was made
using the Maeswrap package in R.
Results
The 20% smallest of the trees averaged 10.6 kg of
stem wood dry biomass (1.6 kg of stem wood
growth during the last year of the rotation),
compared with 185 kg per stem wood in the 20%
largest of trees (34 kg of stem wood growth over
the same period). The smallest trees contained
7.2% of the leaf area as compared to the largest
trees (3.0 m2 versus 41.7 m2), and they absorbed
only 6.7% as much light (2.2 versus 32.8 GJ year1, Figure 2). The smallest trees grew at about
4.7% of the rate of the largest trees, which is a
smaller percentage than the difference in
absorbed photosynthetically active radiation;
97
Figure 2. Stem wood dry biomass growth and
absorbed photosynthetically active radiation
(APAR) as a function of tree stem wood dry
biomass (A and B), and as a function of tree
ranking (C and D) for all trees and plots. The
pattern of increasing stem growth and APAR
with larger or high ranking trees was consistent
among the 12 plots.
therefore the light use efficiency was lower for
the smallest trees (0.75 kg GJ-1 versus 1.03 kg GJ1).
Averaging by plot, higher growth rates showed
higher light absorption and stem wood growth
increased with increasing light interception
across all size classes of trees. Higher
growth rates also showed greater LUE, and the
effect of LUE on growth rate was higher for small
trees than for large trees (Figure 3).
Canopy view of the 7 years old Eucalyptus
grandis plantation studied in São Paulo-Brazil.
Conclusion
The significant contribution of dominant trees to
stand productivity and the importance of
evaluating production ecology at the individual
tree scale show the relevance of further studies
should focus on understanding the impact of
tree dominance on planted forest productivity.
Figure 3. Averaging by plot, higher growth rates showed
higher light absorption (A). Higher growth rates also showed
greater efficiency of light use (B), and the difference in
efficiency was much stronger for small trees than for medium
and large trees.
Acknowledgements
We thank FAPESP-Brazil, the EUCFLUX Project-IPEF,
Itatinga Research Station - ESALQ/USP and Floragro
Company.
References
Campoe O.C., et al. 2012. Tree Physiol. 32, 696–706.
Medlyn, B.E., 2004. Forests at the land–atmosphere
interface, 105–122.
98
Fondecyt project 1140482: Water and nutrient use efficiency in
dendroenergetic plantations.
Trial series objectives
To provide strategies for managing short rotation
woody crops to provide a sustainable source of
fiber and bioenergy feed stocks while
maintaining a positive water and nutrient
balance.
de 1.400 mm. El área presenta un déficit hídrico
que se extiende entre 3 y 4 meses. La
temperatura media anual varía entre 12.5 y 13.9°
C. Another potential study may be added near
Trial specific objectives
To quantify water use for Acacia dealbata and
E. globulus over a range of stocking levels,
nutrient availability (sites) and soils.
To determine site water balance for Acacia
dealbata and E. globulus over a range of
stocking levels, nutrient availability (sites) and
soils and estimate water that may be available
for use off site.
To quantify nutrient use for Acacia dealbata and
E. globulus over a range of stocking levels,
nutrient availability (sites) and soils during a
bioenergy rotation to determine what nutrients
will likely become limiting in a short rotation
silviculture regime.
Experimental design
Two studies were established on a sandy
volcanic soils and recent volcanic ash soils,
located between the Biobío and Araucanía
Regions, Chile (Figure 1). Both trials were
established in September 2010, using 3 blocks at
each site, 2 species (Acacia dealbata,
Eucalyptus globulus), 2 stocking levels (5000,
15000 stems ha-1) with a 1.41 x 1.41 m (324 tree
ha-1) and 0.81 x 0.81 m (900 tree ha-1) spacing
respectively.
Description of study area
El clima en el área de estudio es mediterráneo
templado con una precipitación media anual
Figure 1. Localization studies, on a sandy volcanic
soils (property “la Aguada”) and recent volcanic
ash soils (property “Parcelas Collipulli”).
Micrometeorological measurements and
soil moisture status
Micrometeorological measurements will be
collected from an automatic weather station
located at each site. Data from sensors will
include rainfall, anemometer, leaf wetness sensor,
air temperature and relative humidity and
photosynthetically active sensor. Volumetric soil
water content and temperature data will be
recorded on a daily basis.
99
Soil probes will be installed at 0-20, 20-40 and 4060 cm depths; a block was selected at each trial
where three sensors were installed per plot.
Sap flux measurements
Stem sap flux measurements will be made on 10
trees per plot across different size classes, at breast
height diameter using 2 cm heat dissipation
probes. Averages of temperature difference data
will be recorded daily and stored in data loggers.
Sap flux density data will be scaled and converted
to a tree transpiration rate per unit leaf area (E,
mmol m-2 s-1) using tree sapwood area, stand LAI
(projected) and stand tree density. While trees
with sap flow sensor will have a record of their
diameters by means of electronic dendrometers
attached to the steam of the trees.
Xylem pressure
exchange
potential
and
gas
Pre-dawn and mid-day plant water potential will
be measured with a PSM model 1000 pressure
bomb on one leaf or twing, from the upper half to
one-third of the crown, from each of six trees per
plot. Leaf-level gas exchange rates and stomatal
conductance will be measured using an openpath portalephotosynthetic system. Leaf areas
enclosed in the cuvette will be scanned using an
laser leaf area meter, and the subsequent digital
image will be analyzed with Image software.
These measurements will be made quarterly.
Ecosystem
balance
water
use
and
water
Ecosystem water use will be estimated for each
species and stocking by integrating individual
tree sapflow to the stand level using inventory
data. A simplified water balance equation for
calculating stand water use, is:
P = T + I + D + ETu + Δθ.
Where P = precipitation, T = tree transpiration, I
= tree canopy interception, D = drainage, and
ET = understory evapotranspiration. Interception
will be calculated as a function of throughfall
and stemflow.
Water use efficiency
Integrated water use efficiency will be
calculated over the growing season as total
useable tree biomass produced per total water
transpired per unit land area.
Nutrient use efficiency
Each year in the dormant season 20 trees per
trial will be harvested. The components of the
trees will be separated, and samples will be
obtained, these will be dried at 65°C until a
constant weight and weighed. A subsample of
each component will be obtained to
determine elemental nutrient concentrations
for nitrogen, phosphorus, potassium, calcium,
magnesium, and boron. Nutrient use for each
Figure 2. View of one of the potential sites containing A.dealbata, Eucaliptus species and hybrid poplars
planted at different spacing.
100
Nutrient omission at establishment and doses and efficiency of P
fertilizers trials in Eucalyptus urophylla and Eucayptus pellita in
Colombia.
Background
Eucapyptus urophyllla and Eucalyptus pellita are
species with great potential to be developed
extensively in some Colombian regions. Only with
gains in silvicultural management (and not
genetic improvement), E.pellita and E.urophylla
could reach productivities ranging from 20-35
and 30-40 m3/ha/year, respectively. Ultisoals and
Oxisols are typical soils found in the Orinoquia
region in Colombia, presenting high nutrient
limitations, low pH (4.5-5.5) and consequently high
aluminum saturation. In the last five years, CONIF
(Corporacion Nacional de Investigacion y
Fomento Forestal de Colombia) has developed
some research in these species under different
fertilizations regimes, and it has been recognized
the positive responses to the fertilization with Ca
and P. In this context, and in order to improve
the knowledge about the fertility management of
these species in these areas, it was created an
alliance between CONIF and the company
Cambium (both
members of the Forest
Productivity Cooperative). In this initiative, two
studies were planned for the Orinoquia region,
one nutrient omission (RW25) trial that was
already installed in 2014. , and one P fertilization
strategy (RW26) trial that will be installed in 2015.
Trial series objectives of RW25
1– To determine the critical nutrients
establishment of E.urophylla and E.pellita.
for
2– To evaluate the omission response of one or
more nutrients at establishment on the soils of the
region.
3– To model the biological and economical
response of the species to the omission of site
specific nutrients and to determine operational
strategies
4.– To model the biological and economical
responses of these species to the nutrient omission
to prescribe fertilization of new planted clonal
materials.
Trial series objectives of RW26
1.– To determine a minimal dose of P that
increases the growth response of E.urophylla
and E.pellita to the most limiting nutrient.
2.– To evaluate the effectiveness of controlledrelease fertilizers during the first years of
development.
3.- To model the biological and economical
response to doses and sources of fertilizers
during the first 4 years of development.
Figure 1. Site view and soil profile of the study
area.
101
Treatments and experimental design
The study area is located in the San Martin
province, with an area of 15.597 hectares, and
highly weathered soils (Ultisol and Oxisol), typical
of the Orinoquia region (Figure 1). Soils are
classified as Plinthic Hapludos, with loamy
textures. The specific treatments for the RW25
and RW26 are shown in table 1 and 2,
respectively.
Table 1. Treatment description for the RW25 trial.
Table 2.
trial.
Treatment description for the RW26
Future direction
Annual growth measurement will be taken in
subsequent years in the RW25. The RW26 will be
established in 2014. Pre– and post– sampling of
foliage and soil will allow the effectiveness of
the fertilization treatment on the species under
study.
102
Special study: Nelder spacing trial to evaluateAcacia mangium
development in Northeast Llanos Area of Venezuela.
Trial objective
To assess the
survival and
introduced in
availability in
Venezuela.
effect of planting density on
growth of Acacia mangium
the low fertility and low water
the Llanos Area of northeast
Figure 1. Nelder trial installations in Venezuela
farm (Figure 1,2). The site considered for
establishment was previously occupied by a
Pinus caribaea plantation and was burned
before manual planting with spade. Weed
control was applied mechanically to all plots
after planting .
Table1 . Treatments on spacing (Nelder method)
for Acacia mangium in Venezuela and dasometric
features the establishment of the trial.
Spacing
Distance
from the center
Distance
between a
radio and
others
Spacing1
11.4
Spacing2
13.8
2.4
1545
Spacing3
16.7
2.9
1052
Spacing4
20.2
3.5
717
Spacing5
24.5
4.2
489
Spacing6
29.7
5.1
333
Spacing7
36
6.2
227
Spacing8
43.6
7.6
200
Tree /Ha
2267
Figure 2. Nelder trial for Acacia mangiun.
Treatment and experimental design
The study was established as a circular Nelder
trial design considering planting densities of 227
to 1545 trees per ha and treatment with and
without pruning (Table 1). The circular planting
was established within a square plot of 90 m by
90 m and had a radio from the center of 11.4 m
for the first planting density.
The Nelder plot Seedlings of 20 to 25 cm height
with collar diameter above 3 mm of SSOLanercost
PNG-N-QLD
provenance
were
established in July 2011 at Chamaraguas Norte
103
Measurements
After 25 months since establishment, the mean
cumulative diameter was 10.2 cm and 8.8 m for
the treatment with and without pruning,
respectively (Figure 3). The mean cumulative
height did not vary among the treatments, with
an average height 10 m for both treatments at
month 25th.
In both pruning treatments
diameter consistently declined over the planting
densities of 333 trees/ha, and remained without
variation at higher planting densities.
The
pruning treatment had higher diameter in all the
planting densities than the treatment without
pruning. These differences were greater at
lower planting densities (Figure 3).
As was
expected, the effect of planting density was less
on tree height. Micro-spatial variation in soil
compaction has decreased tree growth (Figure
5).
Figure 4. Accumulated increment for DBH and
height 25 month since establishment by pruning
treatment and planting density.
Future direction
New measurements will be obtained annualy to
assess plant development. Other species such as
Eucalyptus will be tested under this scheme.
Figure 3. Accumulated increment for DBH and
height over time.
Figure 5. Accumulated increment for DBH and
height 25 moths since establishment by pruning
treatment and planting density.
104
Special study: Deep positioning of fertilizers on fast-growing plantations
under intensive management.
Background
Fertilizers responses have been reported
extensively in forest research. However, theses
responses vary enormously among species and
sites. Moreover, there is a little understanding
about how to optimize the nutrient capture
efficiency by trees. In this context, the use of
controlled-release fertilizers is becoming more
common in plantations.
These have the
advantages of increasing the nutrient capture
and diminishing the losses. Other way to increase
the uptake may be optimizing the positioning of
the fertilizer in depth.
Trial objectives
purposes, it will be considered only the interior
plot of 5 by 5 trees. The treatments consider the
testing alone or in combination of:
A)a controlled-release fertilizer, which may be
applied in contact with the root system and in
depth.
B)Boron applied superficially.
C)A blended-fast released fertilizer that may be
applied in different depths.
The experiment was established in two sites of
South-central valley of Chile (Figure 1) in August
2014. The stand was planted at a spacing of 3 x 2
m. A total of 10 fertilization treatment are tested
at each site (Table 1, Figure 2).
This study attempts to assess the growth responses
in fast-growing plantations to the positioning
strategy of the fertilizers relative to the root system.
The specific objectives are:
1) To assess the responses in survival, growth and
stand uniformity to the addition of controlledreleased fertilizers in contact with the root
system versus fast-released fertilizers.
2) To assess the positioning of fertilizers at
different depths and their interaction with the
water availability and
3) To assess the interaction of controlled-release
fertilizers used in contact with the root system
(starters) with traditional fertilization.
Treatments and experimental design
The experimental design is a randomized
complete block design, with 3 replicates. The
plots have 49 trees, but for measurements
Figure 1. Location of the trials near Nacimiento (site
1) and Mulchen (site 2).
105
Figure 2. Diagram describing the way in which the fertilizers are applied.
Table 1.
Description of the fertilization
treatments. LC1 and LC2 are two controlledrelease fertilizers, whereas LR1 and LR2 are two
fast-release fertilizers. Dpending og the type of
the fertilizer, these were applied in contact with
the root, superficially or at 50-60 cm in depth.
Future direction
Three months after stablishing, the distribution of the root system will be assessed through pits excavations and using soil cores.
The study will also be
monitored in soil moisture at different
depth to evaluate the fertilizer by water
availability as well as the interaction between the types of fertilizers. Growth responses will be measured annually.
106
Special study: Deep positioning of fertilizers on fast-growing plantations
under intensive management.
Background
Fertilizers responses have been reported
extensively in forest research. However, theses
responses vary enormously among species and
sites. Moreover, there is a little understanding
about how to optimize the nutrient capture
efficiency by trees. In this context, the use of
controlled-release fertilizers is becoming more
common in plantations.
These have the
advantages of increasing the nutrient capture
and diminishing the losses. Other way to increase
the uptake may be optimizing the positioning of
the fertilizer in depth.
Trial objectives
purposes, it will be considered only the interior
plot of 5 by 5 trees. The treatments consider the
testing alone or in combination of:
A)a controlled-release fertilizer, which may be
applied in contact with the root system and in
depth.
B)Boron applied superficially.
C)A blended-fast released fertilizer that may be
applied in different depths.
The experiment was established in two sites of
South-central valley of Chile (Figure 1) in August
2014. The stand was planted at a spacing of 3 x 2
m. A total of 10 fertilization treatment are tested
at each site (Table 1, Figure 2).
This study attempts to assess the growth responses
in fast-growing plantations to the positioning
strategy of the fertilizers relative to the root system.
The specific objectives are:
1) To assess the responses in survival, growth and
stand uniformity to the addition of controlledreleased fertilizers in contact with the root
system versus fast-released fertilizers.
2) To assess the positioning of fertilizers at
different depths and their interaction with the
water availability and
3) To assess the interaction of controlled-release
fertilizers used in contact with the root system
(starters) with traditional fertilization.
Treatments and experimental design
The experimental design is a randomized
complete block design, with 3 replicates. The
plots have 49 trees, but for measurements
Figure 1. Location of the trials near Nacimiento (site
1) and Mulchen (site 2).
107
Figure 2. Diagram describing the way in which the fertilizers are applied.
Table 1.
Description of the fertilization
treatments. LC1 and LC2 are two controlledrelease fertilizers, whereas LR1 and LR2 are two
fast-release fertilizers. Dpending og the type of
the fertilizer, these were applied in contact with
the root, superficially or at 50-60 cm in depth.
Future direction
Three months after stablishing, the distribution of the root system will be assessed through pits excavations and using soil cores.
The study will also be
monitored in soil moisture at different
depth to evaluate the fertilizer by water
availability as well as the interaction between the types of fertilizers. Growth responses will be measured annually.
108
Fertilizer Survey—Summary of Operational Forest Fertilization for Forest
Productivity Cooperative Members: 2013 Update
Timothy J. Albaugh, Thomas R. Fox, Jose L. Stape, Rafael A. Rubilar
Study objectives
Beginning in 1969, the Forest Productivity
Cooperative surveyed its members annually
regarding forest fertilization practices in the
Southeastern United States. In 2009, With over
half of our full members controlling land outside
of the United States, we began surveying
members about their fertilizer applications on
land inside and outside the United States.
Treatments and experimental design
This report summarizes the annual survey of forest
fertilization practices by the Forest Productivity
Cooperative member companies. This survey
has been conducted for 43 years and provides
a long-term record of forest fertilization practices
in the Southeast United States. This is the fourth
year we have collected and reported
fertilization by Forest Productivity Cooperative
members on land outside the United States.
Members were asked to voluntarily report the
area fertilized in the previous year. The data
reported by the members are then pooled for
the Southeast US and for
land outside the United
States. Information provided
by individual members is not
revealed.
3.Nitrogen and phosphorus applied in
established stands. These applications
may include other elements.
4.Other elements applied in established
stands.
These applications do not
include nitrogen or phosphorus.
5.All applications made on land outside the
southeastern US.
Results
Fertilization in the United States
The total area fertilized in the United States in
2013 was 725404 acres (293561 ha). This was a
4% decrease from 2012 applications (Figure 1).
Applications at planting in 2013 were made on
67824 ac (27447 ha); a 20% decrease from 2012.
Phosphorus only applications in established
stands decreased 45% from 2012 levels to 36524
ac (14781 ha) in 2013. Nitrogen and phosphorus
applications in established stands increased 4%
Many different fertilization
treatments are used by FPC
members. To simplify the
reporting
process,
fertilization treatments are
combined into five major
categories with the first four
incorporating only data from
applications in the southeast
US:
1.Applications made at
the time of planting.
2.Phosphorus
applied
alone in established
stands.
Figure 1. Area fertilized by Forest Productivity Cooperative members from 1990 to 2013.
109
2014. In 2014 the cost for fertilizer material
needed for a 200N and 25P lbs ac-1 elemental
rate would be $147, a 4% decrease from 2013.
Other elements applied in 2013 were potassium,
boron, and calcium along with micronutrient
mixes.
These elements were applied in
combination with nitrogen and phosphorus.
Fertilization outside the United States
This is the fifth year we have surveyed Forest
Productivity members with land outside the
United States about their fertilization practices.
As in previous years, the area fertilized shown
represents only a portion of the area fertilized
outside the United States because all members
surveyed did not respond (Figure 3). Members
who did report their data applied fertilizer to
102040 ac (41294 ha) a 67% drop from the
previous year. About half the applications were
at planting and half were nitrogen and
phosphorus in established stands.
Common
fertilizers applied include NPK blends, limestone,
diammonium phosphate, triple super phosphate
and rock phosphate.
Additional Resources
Albaugh et al. 2007. South. J. Appl. For. 31(3):
129-137.
Albaugh et al. 2012. For. Sci. 58(5): 419-429 or
http://dx.doi.org/10.5849/forsci.11-050.
Figure 2. United States farm price for diammonium phosphate
(DAP) and urea (http://www.ers.usda.gov/data-products/
fertilizer-use-and-price.aspx#26727) (top) and fertilizer material
cost to apply 200 N and 25 P elemental lbs./ac (bottom).
Material cost does not include application.
from 2012 levels to 621056 ac (251333 ha) in
2013. Fertilized area is likely an underestimate
this year because not all companies responded
to the survey.
Fluctuations in area fertilized have been linked to
fluctuations in fertilizer costs (Figure 2).
Compared to 2013 prices, diammonium
phosphate prices dropped 5% and urea prices
dropped 4% in 2014. Given the relatively small
fluctuation in price from 2013 to 2014 fertilizer
applications are likely to remain flat from 2013 to
Figure 3. Response rate from members receiving the fertilizer
survey..
110
Forest Productivity Cooperative Team
North Carolina State University
Universidad de Concepción
Staff
Staff
Jose Luiz Stape Co--Director
Timothy Albaugh Senior Research Associate
Graduate Students
Yuan Fang (PhD)
Kevin Hall (MSc)
April Meeks (MSc)
Aliisa Harjuniemi (MSc)
Virginia Polytechnic Institute and State Univ.
Staff
Thomas Fox Co-Director
Andy Laviner Research Associate
Colleen Carlson Data Manager and Analyst*
Matthew Summall Research Associate
Kathryn Hollandsworth Administrative Assistant
Christine Blinn Research Associate
Graduate Students
Marco Yanez (PhD)
Andy Laviner (PhD)
Kevan Minnick (PhD)
Santosh Subedi (PhD)
Jay Raymond (PhD)
Laura Skeith (PhD)
Eric Carbaugh (MSc)
Antonio Tacilla (MSc)
Amy Werner (MSc)
Lara Nichols (MSc)
(* In memoriam)
Rafael Rubilar Co-Director
Manuel Monsalvez Research Assistant
Diego Escoban Research Assistant
Luis Castro Lab-technician
Graduate Students
Mailing Vanessa (PhD)
Manuel Acevedo (PhD)
Edwin Esquivel (PhD)
Andres Rodriguez (PhD)
Eduardo Cartes (PhD)
Manuel Yaya (MSc)
Matias Pincheira (MSc)
Instituto de Pesquisas e Estudos Florestais
Staff
Clayton Alvares Research Associate
Rafaela Carneiro Research Assistant
USP and Unesp - Sao Paulo Brazil
Juliana Munhoz (PhD)
Marina Otto (PhD)
Rodrigo Hakamada (PhD)
Cecila Arrevillaga (MSc)
Eduardo Mattos (MSc)
Raoni Nogueira (MSc)
Isabel Deliberali (MSc)
112