16.06.2013 Views

Ground-based Validation of the MODIS Leaf Area Index Product for ...

Ground-based Validation of the MODIS Leaf Area Index Product for ...

Ground-based Validation of the MODIS Leaf Area Index Product for ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Ground</strong>-<strong>based</strong> <strong>Validation</strong><br />

<strong>of</strong> <strong>the</strong> <strong>MODIS</strong> <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> <strong>Product</strong><br />

<strong>for</strong> East African Rain Forest Ecosystems<br />

Der Naturwissenschaftlichen Fakultät<br />

der Friedrich-Alexander-Universität Erlangen-Nürnberg<br />

zur Erlangung des Doktorgrades<br />

vorgelegt von<br />

Tanja Kraus<br />

aus Nürnberg


Als Dissertation genehmigt<br />

von der Naturwissenschaftlichen Fakultät<br />

der Friedrich-Alexander-Universität Erlangen-Nürnberg.<br />

Tag der mündlichen Prüfung: 10. Oktober 2008<br />

Vorsitzender der Promotionskommission Pr<strong>of</strong>. Dr. Eberhard Bänsch<br />

Erstberichterstatter Pr<strong>of</strong>. Dr. Cyrus Samimi<br />

Zweitberichterstatter Pr<strong>of</strong>. Dr. Roland Baumhauer


i<br />

Summary<br />

In <strong>the</strong> context <strong>of</strong> current global changes, tropical rain <strong>for</strong>ests are undergoing significant alterations. De<strong>for</strong>estation<br />

and degradation processes caused by human exploitation have already destroyed between 8 and 12 million km2 (i.e. 35-50%) <strong>of</strong> <strong>the</strong> original <strong>for</strong>est cover. Large amounts <strong>of</strong> CO <strong>for</strong>merly stored in vegetation and soil are thus<br />

2<br />

released and add to globally rising atmospheric CO rates. Regional and global climate change is en<strong>for</strong>ced<br />

2<br />

as precipitation patterns, surface temperature and carbon uptake are modified. Several international initiatives<br />

(mostly under <strong>the</strong> framework <strong>of</strong> <strong>the</strong> United Nations Framework Convention on Climate Change, UNFCCC) call<br />

<strong>for</strong> action to stop or at least reduce de<strong>for</strong>estation and degradation in <strong>the</strong> tropics and thus emissions. Concrete<br />

mechanisms will probably be defined in a post-2012 climate agreement, <strong>the</strong> details <strong>of</strong> which are currently being<br />

discussed and negotiated among nations.<br />

One <strong>of</strong> <strong>the</strong> biggest challenges to <strong>the</strong> estimation <strong>of</strong> changes in <strong>for</strong>est cover is <strong>the</strong> monitoring <strong>of</strong> <strong>for</strong>est areas on a<br />

reliable, fast, cost effective and area-wide basis. Here, operational remote sensing techniques <strong>for</strong>m a valuable<br />

data source <strong>for</strong> <strong>the</strong> scientific and political community since <strong>the</strong>y permit repetitive and synoptic observations <strong>of</strong><br />

vegetation cover. Changes in <strong>for</strong>est structure and dynamics can, <strong>for</strong> instance, be monitored through repeated<br />

measurement <strong>of</strong> remotely measured biophysical attributes. In contrast to discrete representations <strong>of</strong> land<br />

cover, biophysical variables alter continuously over space and time and may <strong>the</strong>reby reveal early ecosystem<br />

modifications.<br />

In this context, leaf area index (LAI) is one <strong>of</strong> <strong>the</strong> key biophysical variables. It characterises canopy structure<br />

and accounts <strong>for</strong> differences in phenological development, assimilation and biomass growth among plant<br />

species. Operational standard products <strong>of</strong> LAI derived from satellite data, as e.g. <strong>the</strong> <strong>MODIS</strong> LAI product, can<br />

thus contribute to <strong>the</strong> constant and repetitive monitoring <strong>of</strong> <strong>for</strong>est areas.<br />

However, a prerequisite <strong>for</strong> <strong>the</strong> use <strong>of</strong> operational satellite products is <strong>the</strong> evaluation <strong>of</strong> <strong>the</strong>ir accuracy in a<br />

process called validation. <strong>Validation</strong> is defined as <strong>the</strong> process <strong>of</strong> assessing by independent means <strong>the</strong> quality <strong>of</strong><br />

data products. So far, <strong>the</strong>re has been a clear lack <strong>of</strong> validation sites <strong>for</strong> <strong>the</strong> <strong>MODIS</strong> LAI product in tropical rain<br />

<strong>for</strong>est environments. Especially in Africa, no test sites have been available.<br />

Consequently, this <strong>the</strong>sis deals with <strong>the</strong> validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product <strong>for</strong> two test sites in East Africa<br />

that were accessible within <strong>the</strong> framework <strong>of</strong> <strong>the</strong> BIOTA East Africa project. <strong>Ground</strong>-<strong>based</strong> measurements were<br />

per<strong>for</strong>med in Budongo Forest (Uganda) and Kakamega Forest (Kenya) and upscaled to <strong>the</strong> spatial resolution<br />

<strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product <strong>based</strong> on high resolution satellite data <strong>of</strong> ASTER and SPOT-4. Finally <strong>the</strong> <strong>MODIS</strong><br />

LAI product was validated with respect to its accuracy in representing landscape LAI and to its temporal<br />

consistency.<br />

First <strong>of</strong> all, existing methods <strong>of</strong> in situ measurements were reviewed concerning <strong>the</strong>ir applicability to tropical<br />

rain <strong>for</strong>ests. Additionally, a representative and valid sampling scheme was elaborated <strong>based</strong> on recommendations<br />

<strong>of</strong> <strong>the</strong> CEOS-LPV and VALERI networks. LAI-2000 PCA and digital hemispherical photography were found to<br />

be suitable <strong>for</strong> ground measurements as <strong>the</strong>y meet <strong>the</strong> demand <strong>of</strong> fast and area wide sampling. Yet environmental<br />

conditions in <strong>the</strong> tropics (especially with respect to prevailing radiation regimes) are not ideally suited to <strong>the</strong>se


instruments, so that <strong>the</strong> measurements had to be analysed thoroughly with respect to different error sources. A<br />

correction method <strong>for</strong> LAI-2000 PCA data was developed <strong>for</strong> measurements made under direct radiation. O<strong>the</strong>r<br />

errors, influencing both LAI-2000 PCA and digital hemispherical photography, could not be corrected, but <strong>the</strong>ir<br />

influence was quantified in terms <strong>of</strong> measurement precision.<br />

Theil-Sen regression was found to be suitable <strong>for</strong> <strong>the</strong> upscaling process, as its per<strong>for</strong>mance is robust even in <strong>the</strong><br />

presence <strong>of</strong> measurement errors in field and satellite data. For Budongo Forest, where a better quality <strong>of</strong> in situ<br />

data could be retrieved than <strong>for</strong> Kakamega Forest, a thorough analysis revealed that different transfer functions<br />

had to be established <strong>for</strong> different <strong>for</strong>est stages. Interestingly, mean in situ LAI retrieved <strong>for</strong> intermediate and<br />

late <strong>for</strong>est stages was relatively similar, but structural differences mainly in <strong>the</strong> upper canopy lead to significantly<br />

different surface reflectances. Whereas Simple Ratio was found to per<strong>for</strong>m best <strong>for</strong> early and intermediate <strong>for</strong>est<br />

stages (R2 <strong>of</strong> 0.94) a texture measurement (GLCM variance <strong>of</strong> ASTER band 4) retrieved better results <strong>for</strong> late<br />

<strong>for</strong>est stages (R2 <strong>of</strong> 0.71). Consequently, a high resolution LAI map could be produced <strong>for</strong> Budongo Forest<br />

e<br />

with a relative accuracy <strong>of</strong> 9%. This signifies only a slight degradation over field measurement precision. For<br />

Kakamega Forest inferior in situ data quality led to reduced quality <strong>of</strong> <strong>the</strong> high resolution LAI map. A regression<br />

model <strong>based</strong> on Reduced Simple Ratio was here used to estimate LAI <strong>for</strong> <strong>the</strong> whole test site. Overall this model<br />

yielded an R2 <strong>of</strong> 0.53 with a relative accuracy <strong>of</strong> 16% (RMSE <strong>of</strong> 0.8).<br />

Based on <strong>the</strong> high resolution LAI maps, <strong>the</strong> <strong>MODIS</strong> LAI product was <strong>the</strong>n validated <strong>for</strong> <strong>the</strong> two test sites. The<br />

spatial validation <strong>for</strong> Budongo Forest revealed that it represented <strong>the</strong> up-scaled in situ LAI with an accuracy <strong>of</strong><br />

0.53. This corresponds to a relative accuracy <strong>of</strong> 9%, which is identical to <strong>the</strong> accuracy <strong>of</strong> <strong>the</strong> high resolution<br />

LAI maps. For Kakamega Forest validation led to a comparatively low accuracy <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product <strong>of</strong><br />

1.5 (relative accuracy <strong>of</strong> 25%). This is most likely <strong>the</strong> result <strong>of</strong> inferior field data quality <strong>for</strong> this test site and <strong>the</strong><br />

resulting degradation <strong>of</strong> accuracy in <strong>the</strong> upscaling process.<br />

The investigation <strong>of</strong> <strong>the</strong> temporal consistency <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product was <strong>based</strong> on a time series analysis<br />

<strong>for</strong> <strong>the</strong> years 2000-2005. Results showed that <strong>the</strong> data was reliable and stable, but only if temporal interpolation<br />

was applied to bad data quality pixels. The observed variability in LAI (0.4 <strong>for</strong> intermediate and late <strong>for</strong>est<br />

stages) fur<strong>the</strong>r corresponds to in situ measured seasonal LAI trajectories found by o<strong>the</strong>r studies <strong>for</strong> comparable<br />

semi-deciduous rain <strong>for</strong>ests. Maximum LAI values are associated with <strong>the</strong> end <strong>of</strong> <strong>the</strong> rainy season, minimum<br />

LAI values with <strong>the</strong> dry season. It can thus be assumed that <strong>the</strong> <strong>MODIS</strong> LAI product responds correctly to<br />

biome-level LAI changes associated with interannual climate variability.<br />

This <strong>the</strong>sis shows that <strong>the</strong> <strong>MODIS</strong> LAI product represents <strong>the</strong> LAI <strong>of</strong> <strong>the</strong> two East African test sites with an<br />

accuracy that is comparable to <strong>the</strong> accuracy <strong>of</strong> field measurements. In addition, seasonal variations are captured<br />

correctly, but only if quality in<strong>for</strong>mation <strong>for</strong> <strong>the</strong> <strong>MODIS</strong> LAI product is included in <strong>the</strong> analyses.<br />

Although structural changes in rain <strong>for</strong>ests can only be monitored if absolute LAI values are affected (differences<br />

in in situ LAI <strong>of</strong> intermediate and late <strong>for</strong>est changes were not found to be significant <strong>for</strong> <strong>the</strong> test sites), <strong>the</strong><br />

outcomes <strong>of</strong> this <strong>the</strong>sis can never<strong>the</strong>less help to improve knowledge <strong>of</strong> tropical rain <strong>for</strong>ests and <strong>the</strong>ir response<br />

to climatic changes and human disturbances. The results may e.g. help to reduce predictive uncertainties in<br />

biophysical process models and serve as supplementary data <strong>for</strong> model calibration.<br />

ii


iii<br />

Zusammenfassung<br />

Im Kontext des globalen Wandels unterliegen tropische Regenwälder derzeit starken Veränderungen.<br />

Schätzungen zufolge haben Entwaldung und Degradationsprozesse bereits 8 bis 12 Millionen km2 (35-50%) der<br />

ehemaligen Waldfläche zerstört. Die Freisetzung großer, ehemals in der Vegetation und im Boden gespeicherter<br />

Mengen an Kohlenst<strong>of</strong>f trägt wiederum zum Anstieg der atmosphärischen CO -Konzentration bei und verstärkt<br />

2<br />

so den Treibhauseffekt. Modifizierte Niederschlagsmuster, Oberflächentemperaturen und eine veränderte<br />

Kohlenst<strong>of</strong>faufnahme durch die Vegetation haben ihrerseits Einfluss auf das regionale und globale Klima.<br />

Mehrere internationale Initiativen haben daher im Zusammenhang mit der Klimarahmenkonvention der Vereinten<br />

Nationen, UNFCCC (United Nations Framework Convention on Climate Change), zum Handeln aufgerufen.<br />

Ziel ist es, die Abholzung und Degradation von Waldflächen in den Tropen zu stoppen oder wenigstens zu<br />

reduzieren. Konkrete Mechanismen werden wahrscheinlich in einer Klimavereinbarung nach 2012 („Post-<br />

Kyoto-Prozess“) definiert, deren Details momentan zwischen den Nationen diskutiert und verhandelt werden.<br />

Eine der größten methodischen Heraus<strong>for</strong>derungen um die Veränderungen in der Waldfläche abzuschätzen,<br />

ist derzeit ein verlässliches, schnelles, günstiges und qualitatives Monitoring der Waldbereiche. Hier stellen<br />

operationelle Fernerkundungstechniken eine wertvolle Quelle für Wissenschaft und Politik dar, da sie<br />

flächendeckende Aufnahmen der Vegetationsbedeckung liefern. Veränderungen in der Waldstruktur und in<br />

der Dynamik bestimmter Waldflächen können zudem durch wiederholte Abschätzung biophysikalischer<br />

Eigenschaften gemessen werden. Im Gegensatz zu einer diskreten Darstellung der Landbedeckung verändern<br />

sich biophysikalische Variablen kontinuierlich in Zeit und Raum und haben dadurch das Potential, frühe<br />

Veränderungen in Ökosystemen darzustellen.<br />

In diesem Zusammenhang ist der Blattflächenindex (leaf area index, LAI) eine der Schlüsselvariablen. Er<br />

charakterisiert die Struktur von Waldbeständen und spiegelt deren phänologische Veränderung wider. Auch<br />

die Unterschiede in den Assimilationsraten verschiedener Spezies und Biomasseveränderungen werden in<br />

der Blattfläche sichtbar. Operationelle fernerkundliche Standardprodukte zum LAI, wie z.B. das <strong>MODIS</strong> LAI<br />

Produkt, können so zu einem konstanten und wiederholten Monitoring von Waldflächen beitragen.<br />

Eine Voraussetzung für den Gebrauch operationeller Satellitenprodukte ist jedoch die Überprüfung ihrer<br />

Genauigkeit durch eine sogenannte Validierung. Validierung ist in diesem Zusammenhang definiert als ein<br />

Prozess, in dem anhand unabhängiger (Feld-)Daten die Qualität von Fernerkundungsprodukten überprüft wird.<br />

Bis zum heutigen Zeitpunkt sind Validierungsflächen für das <strong>MODIS</strong> LAI Produkt in tropischen Regenwäldern<br />

unterrepräsentiert. Besonders in Afrika steht bislang keine einzige derartige Validierungsfläche zur Verfügung<br />

Dementsprechend beschäftigt sich diese Dissertation mit der Validierung des <strong>MODIS</strong> LAI Produkts für zwei<br />

Testflächen in Ostafrika, die im Rahmen des BIOTA-Ost-Projekts zur Verfügung standen. Feldmessungen<br />

zum Blattflächenindex wurden im Budongo Forest (Uganda) und im Kakamega Forest (Kenia) durchgeführt<br />

und mithilfe von hochaufgelösten Fernerkundungsdaten (ASTER, SPOT-4) auf die räumliche Auflösung des<br />

<strong>MODIS</strong> LAI Produkts skaliert. Schließlich wurde das <strong>MODIS</strong> LAI Produkt auf Basis dieser Datengrundlage<br />

validiert und auf seine räumliche Genauigkeit und zeitliche Konsistenz hin überprüft.


Im ersten Schritt wurden die existierenden Methoden zur in situ Messung von LAI im Hinblick auf ihre<br />

Anwendbarkeit für tropische Regenwälder begutachtet. Zusätzlich wurde ein repräsentatives und gültiges<br />

Aufnahmeschema für die Feldmessungen erarbeitet. Dieses basiert im Wesentlichen auf Empfehlungen der<br />

Netzwerke VALERI und CEOS-LPV. LAI-2000 PCA und digitale hemisphärische Bilder wurden schließlich<br />

als geeignet erachtet, um LAI im Feld abzuschätzen. Beide Methoden erlauben eine relative schnelle Aufnahme<br />

des Blattflächenindex für große Flächen. Allerdings sind die Messbedingungen für beide Instrumente in den<br />

Tropen nicht ideal. So stellt zum Beispiel die Tatsache, dass indirekte Einstrahlung kaum und meist nicht für<br />

ausreichend lange Zeiträume vorherrscht, ein Problem für die LAI-2000 PCA Messungen dar. Aus diesem<br />

Grund wurden die Messergebnisse beider Instrumente im Hinblick auf verschiedene Fehlerquellen genau<br />

analysiert. Für LAI-2000 PCA Messungen, die unter direkten Strahlungsbedingungen aufgenommen wurden,<br />

wurde eine Korrekturmethode entwickelt. Andere Fehlerquellen, die sowohl LAI-2000 PCA Messungen und<br />

hemisphärische Fotos beeinflussten, konnten nicht korrigiert werden. Allerdings wurde ihr Einfluss auf die<br />

Messungen zumindest indirekt durch die Feststellung der Messgenauigkeit quantifiziert.<br />

Für die Hochskalierung der in situ Messungen auf Basis der hochauflösenden ASTER- und SPOT-4-<br />

Daten erwies sich die Theil-Sen Regression als geeignet. Sie zeigte sich als robuste Methode, selbst wenn<br />

Messungenauigkeiten in Feld- und Fernerkundungsdaten vorliegen. Für Budongo Forest, wo eine bessere<br />

Datenqualität der in situ Messungen erzielt werden konnte als für Kakamega Forest, ergab eine genaue<br />

Datenanalyse, dass verschiedene Regressionsmodelle für frühe und mittlere (d.h. gestörte/degradierte)<br />

Waldstadien etabliert werden mussten. Interessanterweise waren die mittleren LAI-Werte für mittlere und späte<br />

(d.h. ungestörte) Waldstadien fast gleich, allerdings führten strukturelle Unterschiede im Kronenbereich zu<br />

signifikanten Unterschieden in der Oberflächenreflexion. Während das Theil-Sen Regressionsmodell basierend<br />

auf dem Vegetationsindex „Simple Ratio“ die besten Ergebnisse für frühe und mittlere Waldstadien erzielte<br />

(R2 =0,94), konnte der LAI für späte Waldstadien besser durch ein Texturmaß (GLCM Varianz von ASTER,<br />

Band 4) modelliert werden (R2 =0,71). Dementsprechend konnte eine hochauflösende LAI-Karte mit einer<br />

relativen Genauigkeit von 9% für Budongo Forest erstellt werden. Dies stellt nur einen leichten Qualitätsverlust<br />

gegenüber der Messgenauigkeit der Feldmessungen dar. Für Kakamega Forest führte die mindere Datenqualität<br />

der in situ Messungen konsequenterweise zu einer niedrigeren Genauigkeit der hochauflösenden LAI-Karte.<br />

Ein Regressionsmodell basierend auf dem Vegetationsindex „Reduced Simple Ratio“ wurde hier verwendet, um<br />

LAI für das gesamte Studiengebiet abzuschätzen. Dieses Modell erreichte ein R2 von 0,53 mit einer relativen<br />

Genauigkeit von 16% (RMSE=0,8).<br />

Basierend auf den hochaufgelösten LAI-Karten wurde das <strong>MODIS</strong> LAI Produkt schließlich für die beiden<br />

Studiengebiete validiert. Die räumliche Validierung für Budongo Forest zeigte, dass das <strong>MODIS</strong> LAI Produkt<br />

die hochskalierten Felddaten mit einer Genauigkeit von 0,53 repräsentiert. Dies entspricht einer relativen<br />

Genauigkeit von 9% und ist somit identisch mit der relativen Genauigkeit der hochaufgelösten LAI-Karten.<br />

Für Kakamega Forest zeigte die Validierung des <strong>MODIS</strong> LAI Produkts dagegen eine vergleichsweise geringe<br />

Genauigkeit von 1,5 (relative Genauigkeit von 25%). Dies ist auf die mindere Felddatenqualität für dieses<br />

Studiengebiet zurückzuführen, die durch das Hochskalieren entsprechend auch eine mindere Qualität der<br />

hochaufgelösten LAI-Karte nach sich zieht.<br />

iv


v<br />

Die Untersuchungen der zeitlichen Konsistenz des <strong>MODIS</strong> LAI Produkts basierte auf der Analyse von Zeitserien<br />

der Jahre 2000-2005. Die Ergebnisse zeigten, dass die Daten verlässlich und stabil sind, allerdings nur, wenn<br />

Pixel mit niedriger Qualität zeitlich interpoliert wurden. Die beobachtete saisonale Variabilität im LAI (0,4 für<br />

mittlere und späte Waldstadien) entspricht Feldmessungen anderer Studien für vergleichbare halbimmergrüne<br />

Regenwälder. Maximale LAI-Werte wurden am Ende der Regenzeit registriert, minimale LAI-Werte in der<br />

Trockenzeit. Somit darf angenommen werden, dass das <strong>MODIS</strong> LAI Produkt die saisonale LAI-Variabilität<br />

korrekt abbildet.<br />

Diese Dissertation zeigt, dass das <strong>MODIS</strong> LAI Produkt den Blattflächenindex der beiden ostafrikanischen<br />

Studiengebiete mit einer Genauigkeit repräsentiert, die innerhalb der Messfehler der Feldmessungen liegt.<br />

Saisonale Veränderungen werden außerdem korrekt wiedergegeben, wenn die Qualitätsin<strong>for</strong>mation des <strong>MODIS</strong><br />

LAI Produkts zur Maskierung und zeitlichen Interpolation der entsprechenden Pixel herangezogen wird.<br />

Obwohl strukturelle Veränderungen in Regenwäldern mit Hilfe des <strong>MODIS</strong> LAI Produkts nur beobachtet werden<br />

können, wenn sie auch die absoluten LAI-Werte beeinflussen (signifikante Unterschiede im in situ gemessenen<br />

LAI der mittleren und späten Waldstadien konnten nicht ermittelt werden), zeigen die Ergebnisse dieser<br />

Dissertation dennoch, dass das <strong>MODIS</strong> LAI Produkt einen Beitrag zur Abschätzung der Reaktion tropischer<br />

Regenwälder auf klimatische Veränderungen und anthropogene Störungen liefern kann. Die Ergebnisse können<br />

z.B. dazu beitragen, die Vorhersagegenauigkeit biophysikalischer Prozessmodelle zu erhöhen und als zusätzliche<br />

Daten in die Modellkalibrierung einfließen.


Acknowledgements<br />

Many people have supported me over <strong>the</strong> last few years during <strong>the</strong> completion <strong>of</strong> this <strong>the</strong>sis. Although <strong>the</strong> list <strong>of</strong><br />

individuals I wish to thank extends beyond <strong>the</strong> limit <strong>of</strong> this page, I would like to show gratitude to <strong>the</strong> following<br />

individuals.<br />

First <strong>of</strong> all my thanks go to my supervisor Pr<strong>of</strong>. Dr. Cyrus Samimi at <strong>the</strong> University <strong>of</strong> Erlangen. He not<br />

only introduced me to <strong>the</strong> field <strong>of</strong> remote sensing during my studies, but also supported this work from <strong>the</strong><br />

very beginning with helpful advice and critical discussion. I am also grateful to Pr<strong>of</strong>. Dr. Roland Baumhauer<br />

(University <strong>of</strong> Würzburg) <strong>for</strong> his role as co-reviewer.<br />

This <strong>the</strong>sis was written during my time at DFD-DLR and at <strong>the</strong> Chair <strong>for</strong> Remote Sensing at <strong>the</strong> University <strong>of</strong><br />

Würzburg. My sincere gratitude goes to Pr<strong>of</strong>. Dr. Stefan Dech, who made it possible <strong>for</strong> me to work in both<br />

places and to benefit from a stimulating atmosphere. Despite his manifold responsibilities he was always ready<br />

to lend an ear to problems and to make helpful comments.<br />

Writing a <strong>the</strong>sis while being involved with o<strong>the</strong>r projects apart from BIOTA East has not always been easy. Yet<br />

this <strong>the</strong>sis has greatly benefited from <strong>the</strong> constant support and steady encouragement <strong>of</strong> Dr. Michael Schmidt.<br />

Thanks <strong>for</strong> never giving up on me. I am fur<strong>the</strong>r grateful to my <strong>for</strong>mer team leader, Pr<strong>of</strong>. Dr. Günter Strunz, who<br />

initially <strong>of</strong>fered me <strong>the</strong> opportunity to work at DFD. His door was always open, even at stressful times.<br />

Carrying out fieldwork in Kakamega and Budongo Forest was one <strong>of</strong> <strong>the</strong> most exciting experiences <strong>for</strong> me.<br />

The people I encountered and <strong>the</strong> experiences I have had have continued to influence me, both personally<br />

and academically. All this would not have been possible without <strong>the</strong> framework <strong>of</strong> <strong>the</strong> BIOTA East Africa<br />

project funded by BMBF. Yet a network only thrives thanks to <strong>the</strong> individuals who support it and thus I thank<br />

my colleagues, especially those from <strong>the</strong> subproject E02. The subproject leader Pr<strong>of</strong>. Dr. Gertrud Schaab, and<br />

PhD students Tobias Lung and Nick Mitchell have helped with valuable discussions and with sharing data and<br />

knowledge concerning <strong>the</strong> two test sites.<br />

Without <strong>the</strong> assistance <strong>of</strong> Jaqueline Kennedy Ayuka, Benson Bwibo Chituyi, Kennedy Andama, and Afeku<br />

Alfred fieldwork would not have been possible. I greatly appreciate that you shared your valuable knowledge<br />

concerning <strong>the</strong> flora <strong>of</strong> Kakamega and Budongo Forests with me. Thanks go also to Mark Broich and Susanne<br />

Zenzinger <strong>for</strong> <strong>the</strong>ir help with in situ measurements.<br />

Several international cooperations and contacts helped advance my work. Thanks go especially to <strong>the</strong> VALERI<br />

network and to Frederic Baret and Marie Weiss <strong>for</strong> practical support with field equipment and s<strong>of</strong>tware issues,<br />

as well as <strong>the</strong>oretical considerations with respect to field measurements. Thanks also to CEOS-LPV and to Jeff<br />

Morisette, Sebastien Garrigues and Jaime Nickeson <strong>for</strong> recommendations and provision <strong>of</strong> ASTER data.<br />

At DFD and <strong>the</strong> University <strong>of</strong> Würzburg my <strong>the</strong>sis has benefited from various discussions with and <strong>the</strong> technical<br />

support <strong>of</strong> several colleagues and students: I would especially like to thank Christopher Conrad, Rolf Richter,<br />

Gunter Schorcht, Rene Colditz and Johannes Hetzel. I am also very grateful to Martin Bachmann, Christopher<br />

vi


vii<br />

Conrad, Wouter Dorigo, Susan Giegerich, Rosslynne Grandinger, Miriam Machwitz and Katharina Winterroth<br />

<strong>for</strong> pro<strong>of</strong>reading and commenting on various parts <strong>of</strong> this <strong>the</strong>sis. Nadine Hauff was a great help in completing<br />

some <strong>of</strong> <strong>the</strong> figures, Conny Schödl with <strong>the</strong> layout.<br />

Finally, this dissertation would not have been possible without <strong>the</strong> support and understanding <strong>of</strong> <strong>the</strong> most<br />

important people in my life: my family. Especially my husband Marcus was not only patient when I spent more<br />

time with my computer than with him, but also provided me with help and encouragement when it was most<br />

needed. This <strong>the</strong>sis is dedicated to my parents who have taught me never to give up.<br />

Tanja Kraus München, July 2008


Table <strong>of</strong> Contents<br />

1 Introduction 1<br />

1.1 Status <strong>of</strong> research 3<br />

1.2 Thesis objectives 6<br />

2 The study areas 9<br />

2.1 Budongo Forest 10<br />

2.1.1 Physiogeographic aspects 10<br />

2.1.2 Vegetation characteristics 13<br />

2.1.3 Forest management 18<br />

2.2 Kakamega Forest 20<br />

2.2.1 Physiogeographic aspects 20<br />

2.2.2 Vegetation characteristics 22<br />

2.2.3 Forest management 25<br />

3 Theoretical background 27<br />

3.1 Definition <strong>of</strong> LAI 27<br />

3.2 Characteristics <strong>of</strong> LAI in tropical rain <strong>for</strong>ests 28<br />

3.3 <strong>Ground</strong>-<strong>based</strong> measurements <strong>of</strong> LAI 29<br />

3.3.1 Methods 31<br />

3.3.2 Instruments 34<br />

3.3.3 Limitations <strong>of</strong> indirect optical methods 37<br />

3.3.4 Spatial sampling strategies 40<br />

3.4 Satellite-<strong>based</strong> derivation <strong>of</strong> LAI 41<br />

3.4.1 Empirical approaches 42<br />

3.4.2 Physical models 43<br />

3.4.3 Spatial and spectral scale <strong>of</strong> sensors 45<br />

3.4.4 Suitability and limitations <strong>of</strong> empirical methods 46<br />

3.4.5 <strong>Validation</strong> approaches 51<br />

4 Remote sensing data and preprocessing 55<br />

4.1 Available high resolution satellite data 55<br />

4.1.1 ASTER 55<br />

4.1.2 SPOT-HRVIR 56<br />

4.2 Preprocessing <strong>of</strong> high resolution satellite data 56<br />

4.2.1 Geometric correction 56<br />

4.2.2 Atmospheric correction 57<br />

4.3 The <strong>MODIS</strong> LAI product 58<br />

4.3.1 The <strong>MODIS</strong> instrument 59<br />

4.3.2 Data set characteristics 60<br />

viii


ix<br />

4.3.3 Algorithm description 61<br />

4.3.4 Algorithm refinements in C5 63<br />

4.3.5 <strong>Validation</strong> status <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product 64<br />

5 <strong>Ground</strong>-<strong>based</strong> LAI measurements 67<br />

5.1 Spatial sampling strategy 67<br />

5.1.1 Modifications made to <strong>the</strong> CEOS-LPV/VALERI methodology 67<br />

5.1.2 Field campaigns 69<br />

5.1.3 Measurement set-up 71<br />

5.2 Processing, correction and error analysis 73<br />

5.2.1 LAI-2000 PCA 73<br />

5.2.2 DHP 80<br />

5.2.3 Comparison between LAI-2000 PCA and DHPs 85<br />

5.3 Results <strong>of</strong> in situ measurements 87<br />

5.3.1 Budongo Forest 88<br />

5.3.2 Kakamega Forest 93<br />

5.4 Conclusion 98<br />

6 Derivation <strong>of</strong> high resolution LAI maps 101<br />

6.1 Calculation <strong>of</strong> spectral vegetation indices and texture measures 101<br />

6.2 Establishment <strong>of</strong> transfer functions 101<br />

6.2.1 Quantification <strong>of</strong> measurement errors 102<br />

6.2.2 Regression analysis 107<br />

6.3 Results 109<br />

6.3.1 Budongo Forest 109<br />

6.3.2 Kakamega Forest 121<br />

6.4 Conclusion 124<br />

7 <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product 127<br />

7.1 Upscaling <strong>of</strong> <strong>the</strong> high resolution LAI e map 127<br />

7.2 Correction <strong>for</strong> foliage clumping 127<br />

7.3 Results 128<br />

7.3.1 Budongo Forest 128<br />

7.3.2 Kakamega Forest 138<br />

7.4 Conclusion 143<br />

8 Discussion and Outlook 145<br />

References 151<br />

Appendix <strong>of</strong> Figures 173<br />

Appendix <strong>of</strong> Tables 181<br />

Appendix <strong>of</strong> Equations 191<br />

Curriculum Vitae 192


<strong>Index</strong> <strong>of</strong> Figures<br />

Figure 1-1 <strong>Validation</strong> sites <strong>of</strong> CEOS LPV 5<br />

Figure 2-1 Map <strong>of</strong> East Africa with <strong>the</strong> study sites in Kenya and Uganda 9<br />

Figure 2-2 Budongo Forest 10<br />

Figure 2-3 Climatic characteristics <strong>of</strong> Budongo Forest 12<br />

Figure 2-4 Assumed movement <strong>of</strong> <strong>for</strong>est species after 10,000 B.P. across Uganda 13<br />

Figure 2-5 Present distribution <strong>of</strong> <strong>for</strong>est in Uganda and estimated distribution <strong>of</strong> <strong>for</strong>est be<strong>for</strong>e clearance 13<br />

Figure 2-6 Forest types in Budongo Forest in 1990 16<br />

Figure 2-7 Vegetation types in Budongo Forest 16<br />

Figure 2-8 Characteristic features <strong>of</strong> trees belonging to <strong>the</strong> upper tree layer 17<br />

Figure 2-9 Logging compartments in Budongo Forest 19<br />

Figure 2-10 Illegal logging activities in <strong>the</strong> N15 nature reserve 19<br />

Figure 2-11 Kakamega Forest and its associated <strong>for</strong>est areas 20<br />

Figure 2-12 Surroundings <strong>of</strong> Kakamega Forest 21<br />

Figure 2-13 Climatic characteristics <strong>of</strong> Kakamega Forest 22<br />

Figure 2-14 Vegetation in Kakamega Forest 23<br />

Figure 2-15 Illegal activities in Kakamega Forest 26<br />

Figure 3-1 Vertical LAI variation in a semi-evergreen tropical rain <strong>for</strong>est in Panama 29<br />

Figure 3-2 LI-COR LAI-2000 PCA 35<br />

Figure 3-3 Illustration <strong>of</strong> <strong>the</strong> different viewing angles <strong>of</strong> <strong>the</strong> LAI-2000 PCA 36<br />

Figure 3-4 Digital camera with fish-eye lens 36<br />

Figure 3-5 Hemispherical photograph taken in Budongo Forest, Nature Reserve on 11 October, 2005 36<br />

Figure 3-6 Effect <strong>of</strong> foliage clustering on gap fraction 38<br />

Figure 3-7 Sampling schemes 41<br />

Figure 3-8 Reflectance from one to six layers <strong>of</strong> cotton leaves 42<br />

Figure 3-9 Simulation <strong>of</strong> vegetation leaf canopy as one- dimensional turbid medium 44<br />

Figure 3-10 Simulation <strong>of</strong> a 3D vegetation canopy with a Monte Carlo Ray Tracing model 45<br />

Figure 3-11 Relationship <strong>of</strong> NDVI versus SR derived from ASTER data 48<br />

Figure 3-12 Structural differences in primary rain <strong>for</strong>est and selectively logged <strong>for</strong>est 51<br />

Figure 3-13 Schematic illustration <strong>of</strong> validation strategy 52<br />

Figure 3-14 VALERI validation scheme <strong>of</strong> medium resolution satellite data 53<br />

Figure 4-1 <strong>MODIS</strong> tiles <strong>for</strong> processing levels 2G, 3 and 4 60<br />

Figure 5-1 Location <strong>of</strong> ESUs in Kakamega Forest 69<br />

Figure 5-2 Location <strong>of</strong> ESUs in Budongo Forest 70<br />

Figure 5-3 Modified sampling scheme 71<br />

Figure 5-4 Transects on ESUs in Kakamega Forest and Budongo Forest 71<br />

Figure 5-5 Schematic illustration <strong>of</strong> measurement set-up with three LAI-2000 PCA devices 72<br />

Figure 5-6 Measurement set-up <strong>of</strong> LAI-2000 PCA 73<br />

Figure 5-7 LAI e (LAI2000) calculated from rings 1-4 versus LAI e (LAI2000) calculated from ring 4 only 74<br />

Figure 5-8 Test measurements with LAI-2000 PCA 75<br />

x


xi<br />

Figure 5-9 Hemispherical photograph 75<br />

Figure 5-10 LAI e (LAI2000) calculated from continuous B1 readings on ESU 7 in Budongo Forest 77<br />

Figure 5-11 LAI e (LAI2000) derived from continuous B1 measurements on ESUs 16, 18 and 30 78<br />

Figure 5-12 Scatter plot <strong>of</strong> LAI e (LAI2000, cen) and θ sun and correction <strong>for</strong> e θsun 79<br />

Figure 5-13 Optical centre <strong>of</strong> Nikon Coolpix 4300 camera used in Budongo Forest 80<br />

Figure 5-14 Illustration <strong>of</strong> DHP area used <strong>for</strong> LAI derivation 81<br />

Figure 5-15 Subset <strong>of</strong> hemispherical photograph as original and classified image 81<br />

Figure 5-16 Variation <strong>of</strong> <strong>the</strong> G-function with <strong>the</strong> average leaf inclination angle θ 83<br />

Figure 5-17 Equidistant projection 83<br />

Figure 5-18 Illustration <strong>of</strong> calculated projection function <strong>for</strong> NIKON coolpix 4300 and FC-E8 fisheye 84<br />

Figure 5-19 LAI e (DHP) derived from classification <strong>of</strong> gap fraction by two different operators 85<br />

Figure 5-20 LAI e (DHP) plotted against θ sun 85<br />

Figure 5-21 Illustration <strong>of</strong> field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA with 45° view cap and DHP 86<br />

Figure 5-22 Boxplots showing light variables (P 0 and τ 0 ) 87<br />

Figure 5-23 LAI e (LAI2000) plotted against LAI e (DHP) 87<br />

Figure 5-24 Thinned upper canopy and illegal logging activities on ESU 28 in Budongo Forest 88<br />

Figure 5-25 Comparison <strong>of</strong> percentage <strong>of</strong> <strong>for</strong>est stages <strong>of</strong> Budongo Forest and <strong>of</strong> sampled ESUs 89<br />

Figure 5-26 Frequency distribution <strong>of</strong> LAI e (LAI2000) on sample level 89<br />

Figure 5-27 ESU 21 and ESU 15 in Budongo Forest 89<br />

Figure 5-28 Frequency distributions <strong>of</strong> LAI e (LAI2000) , LAI e (DHP) and LAI true (DHP) on ESU level 90<br />

Figure 5-29 DHP <strong>of</strong> understorey on ESU 4 and ESU 28 91<br />

Figure 5-30 Mean LAI per <strong>for</strong>est stage in Budongo Forest 91<br />

Figure 5-31 ESU 19 in Kakamega Forest 93<br />

Figure 5-32 Comparison <strong>of</strong> percentage <strong>of</strong> <strong>for</strong>est stages <strong>of</strong> Kakamega Forest and sampled ESUs 94<br />

Figure 5-33 Frequency distribution <strong>of</strong> LAI e (LAI2000) on sample level in Kakamega Forest 95<br />

Figure 5-34 ESU 25 and ESU 27 in Kakamega Forest 95<br />

Figure 5-35 Frequency distribution <strong>of</strong> LAI e (LAI2000) , LAI e (DHP) and LAI (DHP) on ESU level 96<br />

Figure 6-1 Illustration <strong>of</strong> bias, precision and accuracy 103<br />

Figure 6-2 Relation between c v and LAI e (2000) <strong>for</strong> all ESUs in Budongo Forest 105<br />

Figure 6-3 Derivation <strong>of</strong> reference values <strong>for</strong> regression analysis 108<br />

Figure 6-4 Bivariate plot <strong>of</strong> LAI e (DHP) with ρ red derived from ASTER 110<br />

Figure 6-5 Bivariate plot <strong>of</strong> LAI e (DHP) and SR derived from ASTER data 110<br />

Figure 6-6 Bivariate plots <strong>of</strong> LAI e (DHP) and SR (ASTER) and LAI (DHP) and NDVI c (SPOT) 111<br />

Figure 6-7 Bivariate plots <strong>of</strong> LAI e (DHP) and NDVI (ASTER) and both variables in log trans<strong>for</strong>mation 112<br />

Figure 6-8 Bivariate plot <strong>of</strong> LAI e (DHP) and variance (band 4, 9x9 kernel) derived from ASTER 113<br />

Figure 6-9 Relationship between observed LAI e (DHP) and mean LAI e (DHP) values over associated groups<br />

and observed surface reflectances and mean values over associated groups in <strong>the</strong> red<br />

and NIR spectral bands <strong>of</strong> ASTER, taking into account observation errors in field LAI 114<br />

Figure 6-10 Relationship between observed surface reflectance and mean values over associated groups<br />

in <strong>the</strong> red and NIR spectral bands <strong>of</strong> ASTER and observed LAI e (DHP) and mean LAI e (DHP)<br />

values over associated groups, taking into account observation errors in ASTER data 115


Figure 6-11 Relationship between observed LAI e (DHP) and mean LAI e (DHP) values over associated groups<br />

and observed surface reflectance and mean values over associated groups in <strong>the</strong> red<br />

and NIR spectral bands <strong>of</strong> ASTER, taking into account observation errors in both<br />

field LAI and ASTER data 116<br />

Figure 6-12 Bivariate plots <strong>of</strong> log trans<strong>for</strong>med LAI e (DHP) and NDVI and LAI e (DHP) and SR 116<br />

Figure 6-13 Bivariate plot <strong>of</strong> LAI e (DHP) and SR 117<br />

Figure 6-14 Relation between LAI e estimated from ASTER data and in situ measured LAI 118<br />

Figure 6-15 High resolution LAI e map <strong>of</strong> Budongo Forest estimated from ASTER data 119<br />

Figure 6-16 Focus areas <strong>of</strong> Budongo Forest displayed in Figure 6-15. 120<br />

Figure 6-17 Bivariate plot <strong>of</strong> LAI e (LAI2000) and RSR derived from SPOT 121<br />

Figure 6-18 Bivariate plot <strong>of</strong> LAI e (LAI2000) and RSR derived from SPOT <strong>for</strong> Kakamega Forest 122<br />

Figure 6-19 Bivariate plot <strong>of</strong> LAI e (DHP) and homogeneity (band 4, 7x7 kernel) derived from SPOT 122<br />

Figure 6-20 High resolution LAI e map <strong>of</strong> Kakamega Forest modelled from SPOT data. 123<br />

Figure 6-21 Focus areas <strong>of</strong> Kakamega Forest displayed in Figure 6-22 124<br />

Figure 7-1 Quality in<strong>for</strong>mation on cloud cover etc. <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> data sets JD 273-353 2005 129<br />

Figure 7-2 Percentage <strong>of</strong> invalid (i.e. cloud contaminated) pixels <strong>for</strong> JD 273 to 353 in 2005 129<br />

Figure 7-3 Type 3 <strong>of</strong> MOD12Q1 data sets 2001-2004 130<br />

Figure 7-4 Quality in<strong>for</strong>mation on <strong>the</strong> applied <strong>MODIS</strong> LAI algorithm <strong>for</strong> JD 289 and 329 131<br />

Figure 7-5 (Effective) LAI maps <strong>for</strong> <strong>the</strong> Budongo Forest test site 132<br />

Figure 7-6 Frequency distribution <strong>of</strong> (effective) LAI values derived from maps <strong>of</strong> Figure 7-6 133<br />

Figure 7-7 Regression between ASTER LAI e (1000 m) and C5 <strong>MODIS</strong> LAI product 134<br />

Figure 7-8 Quality in<strong>for</strong>mation on <strong>the</strong> applied <strong>MODIS</strong> LAI algorithm (C4 data) <strong>for</strong> JD 329 135<br />

Figure 7-9 Relative frequencies <strong>of</strong> C5 <strong>MODIS</strong> LAI algorithm usage <strong>for</strong> <strong>the</strong> years 2000-2005 136<br />

Figure 7-10 Mean <strong>MODIS</strong> LAI retrieved <strong>for</strong> <strong>the</strong> centre <strong>of</strong> <strong>the</strong> Budongo Forest test site 137<br />

Figure 7-11 Annual <strong>MODIS</strong> LAI trajectories <strong>for</strong> <strong>the</strong> different <strong>for</strong>est stages <strong>of</strong> Budongo Forest 138<br />

Figure 7-12 Quality in<strong>for</strong>mation <strong>of</strong> <strong>the</strong> MOD15A2 data sets JD 281-361 2004 138<br />

Figure 7-13 Percentage <strong>of</strong> invalid (i.e. cloud contaminated) pixels <strong>for</strong> JD 281 to 361 in 2004 139<br />

Figure 7-14 Type 3 <strong>of</strong> <strong>MODIS</strong> land cover <strong>of</strong> 2004 and <strong>MODIS</strong> LAI algorithm applied to data set JD 281 139<br />

Figure 7-15 (Effective) LAI maps <strong>for</strong> <strong>the</strong> Kakamega Forest test site 140<br />

Figure 7-16 Frequency distribution <strong>of</strong> (effective) LAI values 141<br />

Figure 7-17 Regression between SPOT LAI e (1000 m) and <strong>the</strong> C5 <strong>MODIS</strong> LAI product<br />

and SPOT LAI (1000 m) and <strong>the</strong> <strong>MODIS</strong> LAI product 142<br />

Figure 7-18 Annual <strong>MODIS</strong> LAI trajectory <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> Kakamega Forest 143<br />

xii


xiii<br />

<strong>Index</strong> <strong>of</strong> Tables<br />

Table 2-1 Formations <strong>of</strong> tropical moist <strong>for</strong>ests 14<br />

Table 3-1 Review <strong>of</strong> LAI values reported <strong>for</strong> tropical rain <strong>for</strong>ests in <strong>the</strong> literature 30<br />

Table 3-2 Nominal angular coverage <strong>of</strong> <strong>the</strong> different rings <strong>of</strong> LAI-2000 PCA 35<br />

Table 3-3 cos θ and W i values <strong>for</strong> LAI-2000 PCA rings 1-5 and rings 1-4 35<br />

Table 3-4 Optical remote sensing data types, mainly used LAI derivation methods 47<br />

Table 3-5 Overview <strong>of</strong> most important SVIs with respect to LAI derivation <strong>for</strong> tropical rain <strong>for</strong>ests 49<br />

Table 4-1 Spectral and spatial properties <strong>of</strong> <strong>the</strong> ASTER VNIR and SWIR instruments 55<br />

Table 4-2 Spectral and spatial properties <strong>of</strong> SPOT-4 HRVIR 56<br />

Table 4-3 Reference data <strong>for</strong> geometric correction 57<br />

Table 4-4 Acquisition and preprocessing in<strong>for</strong>mation <strong>for</strong> <strong>the</strong> high resolution satellite data 58<br />

Table 4-5 Spectral and spatial properties <strong>of</strong> <strong>the</strong> first seven bands <strong>of</strong> <strong>MODIS</strong> 59<br />

Table 4-6 MOD15A2 data set characteristics 61<br />

Table 4-7 Biome classes in MOD12 land cover product 62<br />

Table 4-8 Canopy structural attributes <strong>of</strong> <strong>the</strong> six biomes used in <strong>the</strong> C4 LAI algorithm 63<br />

Table 4-9 Overview on validation studies <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product (C4) 66<br />

Table 5-1 Methodology proposed by VALERI <strong>for</strong> in situ sampling and modifications made 68<br />

Table 5-2 Statistical measures <strong>for</strong> light intensity recorded by rings 1-4 <strong>of</strong> LAI-2000 PCA 74<br />

Table 5-3 Descriptive statistics <strong>of</strong> continuous LAI-2000 PCA measurements 76<br />

Table 5-4 Allocation <strong>of</strong> ESUs to <strong>the</strong> respective radiation regimes and necessary processing steps 76<br />

Table 5-5 Descriptive statistics <strong>for</strong> LAI e (LAI2000) derived from continuous B1 measurements 78<br />

Table 5-6 Mean and standard deviation <strong>of</strong> LAI e (DHP) per ESU processed by two different operators 84<br />

Table 5-7 Spearman Rho correlation matrix between τ 0 (LAI-2000 PCA) and P 0 (DHP) 86<br />

Table 5-8 Forest stage, logging status, NFA zone, comp. no. and LAI measures <strong>of</strong> 30 ESUs<br />

in Budongo Forest 92<br />

Table 5-9 Differences in LAI sampling between <strong>the</strong> first and <strong>the</strong> second field campaign 93<br />

Table 5-10 Forest stage, logging status, NFA zone, compartment number and LAI measures<br />

<strong>of</strong> <strong>the</strong> 30 ESUs in Kakamega Forest 97<br />

Table 5-11 Comparison between <strong>the</strong> properties <strong>of</strong> LAI-2000 PCA and DHP 99<br />

Table 6-1 Second-order texture measures used in this <strong>the</strong>sis 102<br />

Table 6-2 Descriptive statistics <strong>of</strong> LAI e (LAI2000) per ESU derived from B1 measurements 104<br />

Table 6-3 Descriptive statistics <strong>for</strong> LAI e (DHP) derived from DHP time series 106<br />

Table 6-4 Relative precision α <strong>for</strong> red, NIR and SWIR spectral bands 107<br />

Table 6-5 Theil-Sen and OLS regression models <strong>based</strong> on LAI e (DHP) and SVIs <strong>for</strong> early<br />

and intermediate <strong>for</strong>est stages in Budongo Forest 111<br />

Table 6-6 Theil-Sen and linear OLS regression models <strong>based</strong> on LAI variables and<br />

texture measures <strong>for</strong> late <strong>for</strong>est stages in Budongo Forest 113<br />

Table 6-7 Relative bias, precision and accuracies <strong>of</strong> relationships between LAI e (DHP)<br />

and surface reflectance derived from ASTER data 117<br />

Table 6-8 Theil-Sen and linear OLS regression models <strong>based</strong> on LAI e (DHP) and SVIs 117


Table 6-9 Theil-Sen and linear OLS regression models <strong>based</strong> on LAI e (LAI2000) and SVIs<br />

<strong>for</strong> early and intermediate <strong>for</strong>est stages in Kakamega Forest 122<br />

Table 6-10 Theil-Sen and linear OLS regression models <strong>based</strong> on LAI variables and texture<br />

measures <strong>for</strong> late <strong>for</strong>est stages in Kakamega Forest 122<br />

Table 7-1 Comparison <strong>of</strong> λ derived from DHP with λ given in Chen et al. (2005) 128<br />

Table 7-2 Mean (and standard deviation) <strong>of</strong> (effective) LAI derived from ASTER<br />

and <strong>MODIS</strong> <strong>for</strong> different <strong>for</strong>est stages 134<br />

Table 7-3 Relative frequency <strong>of</strong> C4 and C5 LAI algorithm usage <strong>for</strong> <strong>the</strong> Budongo Forest test site 135<br />

Table 7-4 Mean (and standard deviation) <strong>of</strong> (effective) LAI derived from SPOT<br />

and <strong>MODIS</strong> <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> Kakamega Forest 142<br />

xiv


xv<br />

Glossary<br />

6S Second Simulation <strong>of</strong> Satellite Signal in <strong>the</strong> Solar Spectrum<br />

AATSR Advanced Along Track Scanning Radiometer<br />

ALA Average <strong>Leaf</strong> Inclination Angle<br />

ANN Artificial Neural Network<br />

ASTER Advanced Spaceborne Thermal Emission and Reflection Radiometer<br />

AVHRR Advanced Very High Resolution Radiometer<br />

BRF Bi-directional Reflectance Factors<br />

C4 Collection 4 (version <strong>of</strong> <strong>MODIS</strong> products and respective algorithms)<br />

C5 Collection 5 (version <strong>of</strong> <strong>MODIS</strong> products and respective algorithms)<br />

CAR Central African Republic<br />

CCD Charge Coupled Device<br />

CCRS Canada Centre <strong>for</strong> Remote Sensing<br />

CEOS Committee on Earth Observing Satellites<br />

CNES Centre National d’Etudes Spatiales<br />

COI Circle <strong>of</strong> Interest<br />

CYCLOPES Carbon Cycle and Change in Land Observational <strong>Product</strong>s from an Ensemble <strong>of</strong> Satellites<br />

DAAC Distributed Active Archive Center<br />

DART Discrete Anisotropic Radiative Transfer<br />

DEM Digital Elevation Model<br />

DFD Deutsches Fernerkundungsdatenzentrum (German Remote Sensing Data Center)<br />

DHP Digital Hemispherical Photograph<br />

DLR Deutsches Zentrum für Luft- und Raumfahrt e. V. (German Aerospace Center)<br />

DRC Democratic Republic <strong>of</strong> <strong>the</strong> Congo<br />

EC European Commission<br />

EDC EROS Data Center<br />

EDOS EOS Data and Operations System<br />

EOS Earth Observing System<br />

ERS Earth Resources Satellite<br />

EROS Center <strong>for</strong> Earth Resources Observation and Science<br />

ESA European Space Agency<br />

ESU Elementary Sampling Unit<br />

ETM+ Enhanced Thematic Mapper Plus<br />

EVI Enhanced Vegetation <strong>Index</strong><br />

FAO Food and Agricultural Organization<br />

FPAR Fraction <strong>of</strong> Photosyn<strong>the</strong>tically Active Radiation absorbed by vegetation<br />

FRA Global Forest Resources Assessment<br />

GCP <strong>Ground</strong> Control Point<br />

GLCM Grey Level Co-Occurrence Matrix<br />

GPP Gross Primary <strong>Product</strong>ion


GPS Global Positioning System<br />

GSFC Goddard Space Flight Center<br />

HDF Hierarchical Data Format<br />

IDL Interactive Data Language<br />

IFOV Instantaneous Field Of View<br />

INRA Institut National de la Recherche Agronomique<br />

IMF Institut für Methodik der Fernerkundung (Remote Sensing Technology Institute)<br />

IPCC Intergovernmental Panel <strong>of</strong> Climate Change<br />

ITCZ Intertropical Convergence Zone<br />

JD Julian Day<br />

KFS Kenya Forest Service<br />

KWS Kenyan Wildlife Service<br />

LAI <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong><br />

LAI-2000 PCA LAI-2000 Plant Canopy Analyzer<br />

LP DAAC Land Processes Distributed Active Archive Center<br />

LPV Land <strong>Product</strong> <strong>Validation</strong><br />

LUT Look-Up-Table<br />

MERIS Medium Resolution Image Spectrometer Instrument<br />

MODAPS <strong>MODIS</strong> Adaptive Processing System<br />

<strong>MODIS</strong> Moderate Resolution Imaging Spectroradiometer<br />

MODTRAN Moderate Resolution Atmospheric Transmission<br />

MSR Modified Simple Ratio<br />

NASA National Aeronautics and Space Administration<br />

NDMI Normalized Difference Moisture <strong>Index</strong><br />

NDVI Normalized Difference Vegetation <strong>Index</strong><br />

NDVI c<br />

Corrected Normalized Difference Vegetation <strong>Index</strong><br />

NFA National Forestry Authority <strong>of</strong> Uganda<br />

NIR Near Infrared (part <strong>of</strong> <strong>the</strong> electromagnetic spectrum between 0.78 and 1.40 µm)<br />

NOAA National Oceanic and Atmospheric Administration<br />

NSIDC National Snow and Ice Data Center<br />

OASIS Optimising access to SPOT infrastructure <strong>for</strong> Science<br />

OLS Ordinary Least Squares<br />

PAI Plant <strong>Area</strong> <strong>Index</strong><br />

PAR Photosys<strong>the</strong>tically Active Radiation<br />

POLDER Polarization and Directionality <strong>of</strong> <strong>the</strong> Earth’s Reflectances<br />

(view imaging radiometer developed by CNES)<br />

PROSPECT <strong>Leaf</strong> Optical Properties Spectra<br />

REDD Reducing Emissions from De<strong>for</strong>estation and Degradation<br />

RMA Reduced Major Axis<br />

RMSE Root Mean Square Error<br />

RSR Reduced Simple Ratio<br />

SAIL Scattering by Arbitrarily Inclined Leaves<br />

xvi


xvii<br />

SDS Scientific Data Set (<strong>MODIS</strong>)<br />

SLA Specific <strong>Leaf</strong> <strong>Area</strong><br />

SPOT Satellite Pour l’Observation de la Terre<br />

SR Simple Ratio<br />

SRTM Shuttle Radar Topography Mission-<br />

SVI Spectral Vegetation <strong>Index</strong><br />

SWIR Shortwave Infrared (part <strong>of</strong> <strong>the</strong> electromagnetic spectrum between 1.4 and 3.0 µm)<br />

TiSeG Time Series Generator<br />

TIR Thernal Infrared (part <strong>of</strong> <strong>the</strong> electromagnetic spectrum between 7.0 and 15.0 µm)<br />

TM Thematic Mapper (sensor carried on Landsat 3, 4, and 5)<br />

UNECSO United Nations Educational, Scientific and Cultural Organization, Organisation<br />

UNFCCC United Nations Framework Convention <strong>of</strong> Climate Change<br />

USGS United States Geological Survey<br />

VALERI <strong>Validation</strong> <strong>of</strong> Land European Remote Sensing Instruments<br />

VGT VEGETATION (sensor carried on SPOT-4)<br />

VIS Visible (part <strong>of</strong> <strong>the</strong> electromagnetic spectrum between 0.38 and 0.78 µm)<br />

WAI Woody <strong>Area</strong> <strong>Index</strong><br />

WGCV Working Group on Calibration and <strong>Validation</strong><br />

ZEF Zentrum für Entwicklungs<strong>for</strong>schung


xviii


1<br />

Introduction<br />

Tropical rain <strong>for</strong>ests are <strong>the</strong> most important habitat<br />

type <strong>for</strong> biodiversity conservation worldwide (Myers<br />

et al. 2000). Although <strong>the</strong>y only cover about 7% <strong>of</strong><br />

<strong>the</strong> global land surface (Hansen & DeFries 2004),<br />

<strong>the</strong>y shelter a large variety <strong>of</strong> life. At least 44% <strong>of</strong><br />

<strong>the</strong> world’s vascular plants and 35% <strong>of</strong> terrestrial<br />

vertebrate species are endemic to 25 global<br />

biodiversity hotspots, 15 <strong>of</strong> which are tropical rain<br />

<strong>for</strong>ests (Brooks et al. 2002, Myers et al. 2000).<br />

Additionally, tropical rain <strong>for</strong>ests are playing an<br />

important role in <strong>the</strong> world’s carbon cycle. Vast<br />

amounts <strong>of</strong> carbon are stored in <strong>the</strong>ir vegetation<br />

and soils and are processed in photosyn<strong>the</strong>sis<br />

and respiration processes (Clark 2007). Although<br />

<strong>the</strong> net-carbon balance <strong>of</strong> mature <strong>for</strong>ests is still<br />

<strong>the</strong> subject <strong>of</strong> various scientific discussions, <strong>the</strong>re<br />

is no doubt that carbon is released if degradation<br />

and de<strong>for</strong>estation processes take place, thus adding<br />

to <strong>the</strong> concentration <strong>of</strong> atmospheric greenhouse gases<br />

(Gullison et al. 2007).<br />

In <strong>the</strong> context <strong>of</strong> current global changes pan-tropical<br />

de<strong>for</strong>estation has already destroyed between 8 and<br />

12 million km² (i.e. 35-50%) <strong>of</strong> <strong>the</strong> original primary<br />

<strong>for</strong>est cover (Wright & Muller-Landau 2006).<br />

According to <strong>the</strong> latest global Forest Resources<br />

Assessment (FRA) <strong>of</strong> <strong>the</strong> Food and Agricultural<br />

Organization (FAO) <strong>the</strong> annual net loss in tropical<br />

<strong>for</strong>est area during <strong>the</strong> years 1990-2005 was highest<br />

in Africa with approximately 43,000 km²/a (FAO<br />

2006). Whereas in Europe net <strong>for</strong>est cover is actually<br />

increasing, growing human population numbers in<br />

Africa and a continuing high amount <strong>of</strong> people living<br />

in rural areas are predicted to maintain high levels <strong>of</strong><br />

net de<strong>for</strong>estation on <strong>the</strong> African continent through<br />

2030 (Wright & Muller-Landau 2006). Additionally,<br />

an estimated two-thirds <strong>of</strong> <strong>the</strong> remaining <strong>for</strong>est cover<br />

in <strong>the</strong> tropics show clear signs <strong>of</strong> human disturbance<br />

and degradation (FAO 2006). Habitat loss, habitat<br />

modification and fragmentation lead to altered<br />

vegetation structure and species composition and,<br />

as a result, disturbed natural processes (such as seed<br />

dispersal or pollination), lower species abundance<br />

and extinction (Cuarón 2000). Fur<strong>the</strong>r, <strong>the</strong> possible<br />

carbon sink capacity is reduced and additional<br />

greenhouse gases are emitted. Gullison et al. (2007)<br />

state in this context that de<strong>for</strong>estation and degradation<br />

processes contributed to almost 20% <strong>of</strong> anthropogenic<br />

greenhouse gas emissions during <strong>the</strong> 1990s.<br />

Rising rates <strong>of</strong> atmospheric CO 2 and temperature<br />

will in turn affect various ecosystem processes in<br />

tropical rain <strong>for</strong>ests, <strong>for</strong> example, photosyn<strong>the</strong>tic<br />

rates, net carbon balance, or carbon sequestration.<br />

In turn, declining productivity or even <strong>for</strong>est die-<strong>of</strong>f<br />

could en<strong>for</strong>ce regional and global climate change<br />

as precipitation patterns, surface temperature<br />

and carbon uptake are modified (Clark 2007,<br />

Jin & Zhang 2002). Several studies have also<br />

revealed latitudinal or altitudinal shifts <strong>of</strong> species<br />

ranges due to changing climatic conditions in <strong>the</strong><br />

last 30 years (e.g. Parmesan & Yohe 2003, Root<br />

et al. 2003). According to <strong>the</strong> latest report <strong>of</strong> <strong>the</strong><br />

Intergovernmental Panel on Climate Change (IPCC)<br />

increasing global mean temperatures are likely to<br />

aggravate this effect with <strong>the</strong> risk <strong>of</strong> ecosystem<br />

disruption and species extinction (Schneider et al.<br />

2007).<br />

1


2<br />

Within this context, <strong>the</strong> prospects <strong>of</strong> jointly addressing<br />

concerns about climate change, biodiversity loss and<br />

poverty by Reducing Emissions from De<strong>for</strong>estation<br />

and Degradation (REDD) have attracted growing<br />

attention from <strong>the</strong> international environment and<br />

development communities (Gullison et al. 2007).<br />

Especially tropical countries that are severely<br />

affected by de<strong>for</strong>estation (e.g. Costa Rica and Papua<br />

New Guinea), but <strong>of</strong>ten lack <strong>the</strong> financial power to<br />

stop ongoing processes, have submitted proposals<br />

to <strong>the</strong> United Nations Framework Convention on<br />

Climate Change (UNFCCC) on how to share <strong>the</strong><br />

responsibility <strong>for</strong> <strong>the</strong>se processes among nations.<br />

The objective is to reduce emissions by <strong>the</strong><br />

implementation <strong>of</strong> trading mechanisms, as<br />

compensation <strong>for</strong> protection <strong>of</strong> existing <strong>for</strong>ests and<br />

funding <strong>for</strong> af<strong>for</strong>estation and re<strong>for</strong>estation. While<br />

<strong>the</strong> political implementation <strong>of</strong> <strong>the</strong> so-called REDD<br />

mechanism is still subject to various discussions,<br />

<strong>the</strong>re are also methodological challenges that need<br />

to be addressed. They mainly relate to <strong>the</strong> definition<br />

<strong>of</strong> a baseline to which future (and past) de<strong>for</strong>estation<br />

rates can be compared, as well as to <strong>the</strong> establishment<br />

<strong>of</strong> reliable monitoring systems <strong>for</strong> both de<strong>for</strong>estation<br />

and degradation (Ernsting & Rughani 2007).<br />

In this context, whenever surveying, monitoring and<br />

modelling issues are to be addressed on a regional to<br />

global scale, operational remote sensing techniques<br />

<strong>for</strong>m a valuable data source <strong>for</strong> <strong>the</strong> scientific (and<br />

political) community since such techniques permit<br />

repetitive and synoptic observations <strong>of</strong> vegetation<br />

cover.<br />

Subtle changes in <strong>for</strong>est structure and dynamics<br />

can, <strong>for</strong> instance, be monitored through repeated<br />

measurement <strong>of</strong> remotely derived biophysical<br />

attributes. In contrast to discrete representations<br />

<strong>of</strong> land cover and land cover change, biophysical<br />

variables alter continuously over space and time. They<br />

may <strong>the</strong>reby reveal starting ecosystem modifications<br />

that do not necessarily lead to a land cover conversion<br />

in <strong>the</strong> observed time frame (Lambin 1999).<br />

One <strong>of</strong> <strong>the</strong> key biophysical variables in this context<br />

is <strong>the</strong> leaf area index (LAI), which refers to <strong>the</strong> onesided<br />

foliage area per unit ground area. It characterises<br />

canopy structure, quantifies <strong>the</strong> size <strong>of</strong> canopyatmosphere<br />

interface and accounts <strong>for</strong> differences in<br />

phenological development, assimilation and biomass<br />

growth among plant species (Weiss et al. 2004,<br />

Wilson et al. 2007). As <strong>for</strong>est degradation also affects<br />

phenological cycles in semi-evergreen rain <strong>for</strong>ests<br />

(Lambin 1999), operational standard products <strong>of</strong> LAI<br />

derived from satellite data are here a valuable data<br />

source.<br />

One <strong>of</strong> <strong>the</strong> most widely used operational standard<br />

LAI products is derived from optical remote sensing<br />

data acquired by <strong>the</strong> Moderate Resolution Imaging<br />

Spectroradiometer (<strong>MODIS</strong>) on board <strong>the</strong> Terra and<br />

Aqua plat<strong>for</strong>ms. The <strong>MODIS</strong> LAI product provides<br />

global in<strong>for</strong>mation on LAI at 1 km resolution <strong>based</strong><br />

on a compositing period <strong>of</strong> 8 days (Knyazikhin et<br />

al. 1999). Modelled relationships between spectral<br />

reflectances <strong>of</strong> vegetation canopies and <strong>the</strong>ir<br />

structural characteristics are stored in a look-uptable<br />

(LUT) that is used to achieve inversion <strong>of</strong> <strong>the</strong><br />

radiative transfer problem. If <strong>the</strong> main algorithm fails,<br />

a backup method <strong>based</strong> on <strong>the</strong> normalized difference<br />

vegetation index (NDVI) is used <strong>for</strong> LAI retrieval<br />

(Myneni et al. 2002). The product is distributed free <strong>of</strong><br />

charge toge<strong>the</strong>r with additional per pixel in<strong>for</strong>mation<br />

on product quality by <strong>the</strong> National Aeronautics and<br />

Space Administration (NASA).<br />

Yet a key issue in working with operational satellite<br />

products is product accuracy. Though <strong>the</strong> <strong>MODIS</strong><br />

LAI algorithm is constantly refined, various errors<br />

still persist. Comparison <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

with comparable products derived from <strong>the</strong> POLDER,<br />

MERIS, and VGT sensors (<strong>for</strong> explanation see<br />

Glossary) revealed significant discrepancies (Baret et


1 Introduction<br />

al. 2006). An important step towards <strong>the</strong> evaluation<br />

<strong>of</strong> satellite product accuracy - and in consequence<br />

<strong>of</strong> <strong>the</strong> scientific results <strong>based</strong> on <strong>the</strong>se products - is<br />

<strong>the</strong>re<strong>for</strong>e <strong>the</strong> validation <strong>of</strong> algorithm outputs.<br />

1.1<br />

Status <strong>of</strong> research<br />

According to <strong>the</strong> Working Group on Calibration and<br />

<strong>Validation</strong> (WGCV) within <strong>the</strong> Committee on Earth<br />

Observing Satellites (CEOS), validation is defined<br />

as “<strong>the</strong> process <strong>of</strong> assessing by independent means<br />

<strong>the</strong> quality <strong>of</strong> <strong>the</strong> data products derived from <strong>the</strong><br />

system outputs” (Justice et al. 2000, see also WGCV<br />

2007). <strong>Validation</strong> refers in this context to <strong>the</strong> detailed<br />

examination <strong>of</strong> satellite data products with respect to<br />

<strong>the</strong>ir accuracy in depicting actual surface conditions.<br />

Based on analytical comparison to independent<br />

reference data, products must be analyzed <strong>for</strong><br />

different ecosystems, varying atmospheric conditions<br />

and temporal consistency.<br />

Independent reference data can only be ga<strong>the</strong>red<br />

through <strong>the</strong> sampling <strong>of</strong> accurate and representative<br />

field measurements <strong>for</strong> a certain test site. These in situ<br />

methods <strong>for</strong> LAI assessment have been developed<br />

<strong>for</strong> more than 70 years. Early studies focused on<br />

<strong>the</strong> direct determination <strong>of</strong> LAI to calculate growth<br />

rates <strong>of</strong> crops and grasslands (Rozhnyatovsky 1954,<br />

Warren-Wilson 1959, Watson 1947), but soon novel<br />

methods were developed that also allowed semidirect<br />

and indirect LAI estimation <strong>for</strong> more complex<br />

vegetation types such as <strong>for</strong>ests (e.g. Anderson<br />

1964, Ogawa et al. 1961, Rogers & Hinckley 1979).<br />

As direct and semi-direct methods are usually very<br />

elaborate in terms <strong>of</strong> time and labour, <strong>the</strong> focus has<br />

fur<strong>the</strong>r shifted towards <strong>the</strong> development <strong>of</strong> new and<br />

faster indirect methods (Chen & Black 1991, Lang &<br />

Yuequin 1986, Nilson 1971, Welles 1990, Welles &<br />

Norman 1991). Today, direct and indirect approaches<br />

have been thoroughly tested <strong>for</strong> crops, grasslands<br />

as well as temperate and boreal <strong>for</strong>est ecosystems.<br />

Discussions on error sources have been <strong>based</strong> on<br />

<strong>the</strong>oretical and practical considerations (Bréda 2003,<br />

Fassnacht et al. 1994, Ferment et al. 2003, Weiss et<br />

al. 2004, among o<strong>the</strong>rs).<br />

In tropical rain <strong>for</strong>ests however <strong>the</strong> in situ assessment<br />

<strong>of</strong> LAI remains challenging due to various constraints.<br />

Difficulties with site accessibility, complexity <strong>of</strong> <strong>for</strong>est<br />

structure and canopy height need to be taken into<br />

account, as well as <strong>the</strong> fact that many conventional<br />

direct field methods, that were developed to estimate<br />

LAI in <strong>for</strong>est stands <strong>of</strong> higher latitudes, do not work<br />

in evergreen broadleaf <strong>for</strong>ests owing to <strong>the</strong> high<br />

amount <strong>of</strong> foliage, species richness and <strong>the</strong> lack <strong>of</strong><br />

seasonal litter fall (e.g. Bréda 2003, Dufrêne & Bréda<br />

1995). Optical methods that have been evolving<br />

since 1990 and are widely applied today, in turn,<br />

show constraints related to illumination conditions<br />

under which <strong>the</strong> measurements should be per<strong>for</strong>med.<br />

Although some studies have been conducted with<br />

optical devices in rain <strong>for</strong>est ecosystems (e.g. Aragão<br />

et al. 2005 in Eastern Amazonia, De Wasseige et al.<br />

2003 in <strong>the</strong> Central African Republic [CAR], and<br />

Wirth et al. 2001 in Panama), <strong>the</strong> effects <strong>of</strong> non-ideal<br />

measurement conditions are not yet well understood.<br />

Parallel to <strong>the</strong> improvement <strong>of</strong> in situ <strong>based</strong> LAI<br />

estimation <strong>the</strong> derivation <strong>of</strong> LAI <strong>based</strong> on remote<br />

sensing data has also been subject to significant<br />

progress over <strong>the</strong> past 30 years. Especially through<br />

fur<strong>the</strong>r development <strong>of</strong> <strong>the</strong> physical and ma<strong>the</strong>matical<br />

foundation, techniques have evolved from simple<br />

empirical approaches to complex physically-<strong>based</strong><br />

canopy reflectance models (Liang 2004).<br />

Empirical methods ei<strong>the</strong>r rely on regression analyses<br />

or neural networks <strong>based</strong> on field measurements and<br />

canopy reflectance data. Recent examples <strong>for</strong> LAI<br />

estimation with optical remote sensing data can be<br />

found in case studies by Kalácska et al. (2004) <strong>for</strong><br />

tropical rain <strong>for</strong>ests (with Landsat ETM+), Fassnacht<br />

et al. (1997) and Jensen & Bin<strong>for</strong>d (2004) <strong>for</strong><br />

3


4<br />

temperate <strong>for</strong>ests (both with Landsat TM), or Chen<br />

& Cihlar (1996) and Cohen et al. (2003) <strong>for</strong> boreal<br />

<strong>for</strong>ests (with Landsat TM and ETM+ respectively).<br />

Fernandes et al. (2003) produced Canada-wide LAI<br />

maps <strong>based</strong> on VGT data with a semi-empirical<br />

method, taking into account field data, vegetation<br />

indices and land use in<strong>for</strong>mation. For global<br />

applications and operational processing empirical<br />

methods are, however, not suitable as <strong>the</strong> established<br />

relationships are very time and site specific.<br />

Constraints are thus put on <strong>the</strong> transferability <strong>of</strong><br />

empirical approaches. However empirical approaches<br />

are still <strong>of</strong> importance when it comes to <strong>the</strong> upscaling<br />

<strong>of</strong> field data <strong>for</strong> validation purposes.<br />

By contrast, physical models, as e.g. applied in <strong>the</strong><br />

<strong>MODIS</strong> LAI algorithm, use an inversion <strong>of</strong> canopy<br />

radiation models to estimate biophysical variables.<br />

Radiative transfer models describe <strong>the</strong> interaction<br />

between sunlight and vegetation elements within<br />

<strong>the</strong> canopy on a <strong>the</strong>oretical basis and are thus ideal<br />

<strong>for</strong> global applications. They can be run <strong>for</strong> given<br />

patterns <strong>of</strong> soil, canopy structure, view-illumination<br />

conditions etc. The major challenge lies here in<br />

solving <strong>the</strong> inverse problem, i.e. <strong>the</strong> derivation <strong>of</strong><br />

biophysical variables (such as LAI) given canopy<br />

reflectance, as due to various error sources and a<br />

large number <strong>of</strong> possible combinations <strong>of</strong> canopy<br />

properties <strong>the</strong> inversion <strong>of</strong> a physical model does not<br />

have a unique solution (Knyazikhin et al. 1998a).<br />

Apart from <strong>the</strong> <strong>MODIS</strong> LAI algorithm, physical<br />

models have also been applied to VGT data in <strong>the</strong><br />

European Commission (EC) funded CYCLOPES<br />

programme (Baret et al. 2007) as well as to combined<br />

VGT, AATSR and MERIS data in <strong>the</strong> European<br />

Space Agency’s GLOBCARBON project (Plummer<br />

et al. 2005).<br />

In order to validate <strong>the</strong> above-mentioned LAI<br />

products, CEOS established a subgroup on Land<br />

<strong>Product</strong> <strong>Validation</strong> (LPV) in 2000. The objective is<br />

to define standard guidelines and common protocols<br />

<strong>for</strong> validating satellite products characterising<br />

<strong>the</strong> land surface. A special ef<strong>for</strong>t has fur<strong>the</strong>r been<br />

put into international collaboration in validation<br />

and intercomparison <strong>of</strong> land biophysical products<br />

(Morisette et al. 2006a). Several international<br />

programmes and o<strong>the</strong>r individual initiatives<br />

contributed validation exercises <strong>for</strong> LAI products<br />

derived from different medium spatial resolution<br />

sensors, e.g., BigFoot (funded by NASA’s Terrestrial<br />

Ecology Program), <strong>Validation</strong> <strong>of</strong> Land European<br />

Remote Sensing Instruments (VALERI, mainly<br />

supported by <strong>the</strong> French Centre National d’Etudes<br />

Spatiales, CNES, and <strong>the</strong> Institut National de la<br />

Recherche Agronomique, INRA), or <strong>the</strong> Canada<br />

Centre <strong>for</strong> Remote Sensing (CCRS).<br />

An overview <strong>of</strong> <strong>the</strong> CEOS-LPV test sites is given<br />

in Figure 1-1 (cf. Morisette et al. 2006a). Although<br />

fur<strong>the</strong>r validation sites <strong>of</strong> o<strong>the</strong>r initiatives and<br />

networks do exist, <strong>the</strong> geographical site distribution<br />

is quite representative. <strong>Validation</strong> sites are mainly<br />

situated in Europe and North America with most<br />

studies being conducted in agricultural (e.g., Cohen<br />

et al. 2003, Pisek & Chen 2007, Tan et al. 2005) or<br />

<strong>for</strong>est areas (e.g., Abuelgasim et al. 2006, Cohen et<br />

al. 2003, Pisek & Chen 2007, Tian et al. 2002b, Wang<br />

et al. 2004). Only a few studies have so far dealt<br />

with test sites in Africa, and all <strong>of</strong> <strong>the</strong>m validated<br />

<strong>the</strong> <strong>MODIS</strong> LAI product <strong>for</strong> savannah ecosystems<br />

(Privette et al. 2002, Tian et al. 2002a) in Sou<strong>the</strong>rn<br />

Africa (not included in Figure 1-1).<br />

The studies mentioned above have contributed to <strong>the</strong><br />

detection <strong>of</strong> anomalies in previous versions <strong>of</strong> <strong>the</strong><br />

<strong>MODIS</strong> LAI product and quality assessment <strong>of</strong> LAI<br />

retrieval over different biomes. However <strong>the</strong>re is a<br />

clear lack <strong>of</strong> adequate validation data over broadleaf<br />

evergreen <strong>for</strong>ests (Baret et al. 2006), especially<br />

in Africa. So far, three validation ef<strong>for</strong>ts dealing<br />

with <strong>MODIS</strong> LAI <strong>of</strong> tropical rain <strong>for</strong>ests have been<br />

published, all exhibiting major or minor difficulties<br />

with in situ LAI assessment and geolocation,


1 Introduction<br />

upscaling <strong>of</strong> in situ measurements or algorithm<br />

per<strong>for</strong>mance due to cloud cover. Whereas two <strong>of</strong><br />

<strong>the</strong> studies were dealing with <strong>the</strong> same tropical rain<br />

<strong>for</strong>est site in Brazil (Aragão et al. 2006 and Cohen<br />

et al. 2006), <strong>the</strong> third was part <strong>of</strong> a larger nationwide<br />

assessment in Australia (Hill et al. 2006). Among<br />

<strong>the</strong> reported problems was a general overestimation<br />

<strong>for</strong> tropical <strong>for</strong>ests and a general weak correlation<br />

between in situ data and <strong>the</strong> <strong>MODIS</strong> LAI product.<br />

It must be mentioned, however, that all three studies<br />

were dealing with a previous version <strong>of</strong> <strong>the</strong> <strong>MODIS</strong><br />

LAI product. Improvements are supposed to be<br />

integrated into <strong>the</strong> current version.<br />

Although fur<strong>the</strong>r field measurements were made in<br />

a tropical rain <strong>for</strong>est in Counami, French Guiana, in<br />

<strong>the</strong> context <strong>of</strong> VALERI (Baret & Rosello 2007), no<br />

fur<strong>the</strong>r <strong>MODIS</strong> LAI validation has been published<br />

<strong>for</strong> this site.<br />

Figure 1-1<br />

An important part <strong>of</strong> validation is <strong>the</strong> establishment<br />

<strong>of</strong> standardised sampling schemes. Morisette et<br />

al. (2006a) give a good overview <strong>of</strong> <strong>the</strong> validation<br />

methodologies used by different LPV groups.<br />

Differences lie especially in <strong>the</strong> instruments used <strong>for</strong><br />

field measurements and subsequent LAI calculation,<br />

sampling schemes, site extents and upscaling<br />

approaches. In order to establish common standards<br />

and achieve comparability <strong>of</strong> validation exercises,<br />

VALERI has focused on developing effective standard<br />

methodologies <strong>for</strong> field measurements and scaling to<br />

higher resolution remotely sensed imagery (Baret et<br />

al. submitted). High spatial resolution maps <strong>of</strong> LAI<br />

are generated from ancillary satellite data and used<br />

<strong>for</strong> <strong>the</strong> validation <strong>of</strong> moderate-resolution biophysical<br />

products, such as <strong>MODIS</strong> (Morisette et al. 2006a).<br />

So far most validation exercises have focused on <strong>the</strong><br />

spatial domain, i.e., upscaling <strong>of</strong> in situ measurements<br />

<strong>Validation</strong> sites <strong>of</strong> CEOS-LPV. Classes represent bare soil, water bodies, deciduous broadleaf <strong>for</strong>est, evergreen<br />

needleleaf <strong>for</strong>est, evergreen broadleaf <strong>for</strong>est, crops, and grassland (Morisette et al. 2006a).<br />

5


6<br />

and comparing <strong>the</strong> results with <strong>MODIS</strong> LAI maps in<br />

order to assess algorithm correctness (e.g. Abuelgasim<br />

et al. 2006, Cohen et al. 2003, Pisek & Chen 2007,<br />

Tan et al. 2005, Tian et al. 2002a, Tian et al. 2002b,<br />

Wang et al. 2004). Only a few publications deal with<br />

<strong>the</strong> temporal dimension <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product in<br />

terms <strong>of</strong> <strong>the</strong> correct retrieval <strong>of</strong> phenological patterns<br />

(Fensholt et al. 2004, Huemmrich et al. 2005, Kang<br />

et al. 2003, Wang et al. 2005).<br />

There are several reasons <strong>for</strong> focusing this <strong>the</strong>sis<br />

on East African rain <strong>for</strong>ests. First <strong>of</strong> all, <strong>the</strong>se<br />

ecosystems belong to a globally outstanding<br />

ecoregion with respect to <strong>the</strong>ir biological diversity.<br />

Although endemism rates are low, both flora and<br />

fauna are characterised by high levels <strong>of</strong> species<br />

richness (Burgess et al. 2004). However, <strong>the</strong> <strong>for</strong>mer<br />

mosaic <strong>of</strong> tropical <strong>for</strong>ests mixed with savanna<br />

woodlands has been significantly modified. Severe<br />

de<strong>for</strong>estation has taken place during <strong>the</strong> last century.<br />

Most <strong>of</strong> <strong>the</strong> remaining <strong>for</strong>est areas are now protected,<br />

but never<strong>the</strong>less <strong>the</strong> diverse flora is imperilled by<br />

human disturbance and possible future climatic<br />

impact. Second, <strong>the</strong> close link to <strong>the</strong> VALERI project<br />

encouraged <strong>the</strong> establishment <strong>of</strong> rain <strong>for</strong>est study<br />

sites in Africa <strong>for</strong> <strong>the</strong> mentioned reasons.<br />

With respect to <strong>the</strong> above-described current state <strong>of</strong><br />

research, <strong>the</strong> objectives <strong>of</strong> this <strong>the</strong>sis will now be<br />

outlined.<br />

1.2<br />

Thesis objectives<br />

As documented in <strong>the</strong> previous chapter, <strong>the</strong> <strong>MODIS</strong><br />

LAI product holds a large amount <strong>of</strong> in<strong>for</strong>mation on<br />

<strong>for</strong>est structural and phenological properties. It can<br />

thus contribute to monitoring and modelling purposes<br />

in <strong>the</strong> context <strong>of</strong> current scientific and political<br />

discussions. Never<strong>the</strong>less – and especially if financial<br />

support to developing countries is involved – it can<br />

only be applied if <strong>the</strong> product is reliable and accurate<br />

<strong>for</strong> different geographical regions and land surfaces.<br />

To date <strong>the</strong>re is a clear lack <strong>of</strong> validation sites <strong>for</strong><br />

<strong>the</strong> <strong>MODIS</strong> LAI product in tropical rain <strong>for</strong>ests.<br />

Especially <strong>for</strong> African rain <strong>for</strong>ests no in<strong>for</strong>mation<br />

on <strong>the</strong> quality <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product is thus far<br />

available. Fur<strong>the</strong>r, <strong>the</strong> current version <strong>of</strong> <strong>the</strong> <strong>MODIS</strong><br />

LAI product has not been validated yet <strong>for</strong> tropical<br />

rain <strong>for</strong>ests.<br />

The main objective <strong>of</strong> this <strong>the</strong>sis is <strong>the</strong>re<strong>for</strong>e <strong>the</strong><br />

validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product <strong>for</strong> rain<br />

<strong>for</strong>est ecosystems in East Africa <strong>based</strong> on ground<br />

measurements <strong>of</strong> LAI. For that reason adequate in<br />

situ sampling strategies are necessary. As explained<br />

earlier, <strong>the</strong> available optical measurement techniques<br />

have not been thoroughly tested in tropical rain<br />

<strong>for</strong>ests. There<strong>for</strong>e a detailed comparison <strong>of</strong> available<br />

and suitable instruments as well as <strong>the</strong>ir advantages<br />

and constraints is required. In addition, an in-depth<br />

analysis <strong>of</strong> <strong>the</strong> results is planned in order to reveal<br />

errors introduced by non-ideal sampling conditions.<br />

With respect to spatial sampling strategies <strong>the</strong><br />

recommendations proposed by <strong>the</strong> CEOS-LPV and<br />

VALERI networks will be reviewed with respect to<br />

<strong>the</strong>ir suitability <strong>for</strong> test sites in tropical rain <strong>for</strong>ests.<br />

If necessary, <strong>the</strong> strategies will be adapted and<br />

optimized.<br />

Direct upscaling from in situ measurements to <strong>MODIS</strong><br />

data bears multiple uncertainties because a) field<br />

samplings <strong>of</strong> LAI representing an adequate area would<br />

be very time intensive and b) smaller sampling sizes<br />

are unable to cover <strong>the</strong> entire heterogeneity within<br />

one <strong>MODIS</strong> pixel. Additionally geometry problems<br />

can lead to errors in <strong>the</strong> direct upscaling <strong>of</strong> in situ<br />

measurements. There<strong>for</strong>e <strong>the</strong> intention is to integrate<br />

high resolution satellite data in order to accurately<br />

scale <strong>the</strong> ground-<strong>based</strong> measurements to <strong>MODIS</strong><br />

resolution. If possible, measurement precision <strong>of</strong> both<br />

field and satellite data will be included to develop<br />

robust regression models that lead to <strong>the</strong> production<br />

<strong>of</strong> high resolution LAI maps with known accuracy.


1 Introduction<br />

In <strong>the</strong> last step, <strong>the</strong> <strong>MODIS</strong> LAI product will be<br />

compared with this high resolution reference data in<br />

a detailed analysis. Here <strong>the</strong> analysis <strong>of</strong> additional<br />

in<strong>for</strong>mation on <strong>the</strong> quality <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI<br />

product will shed light on <strong>the</strong> per<strong>for</strong>mance <strong>of</strong> <strong>the</strong><br />

LAI algorithm <strong>for</strong> tropical rain <strong>for</strong>ests. Owing to <strong>the</strong><br />

observation <strong>of</strong> seasonal variability <strong>of</strong> LAI due to <strong>the</strong><br />

presence <strong>of</strong> some deciduous species in <strong>the</strong> two study<br />

sites, <strong>the</strong> temporal consistency <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI<br />

product will also be investigated. The generation and<br />

analysis <strong>of</strong> time series will help to reveal product<br />

smoothness over time and address <strong>the</strong> feasibility<br />

<strong>of</strong> using <strong>the</strong> <strong>MODIS</strong> LAI product as a proxy <strong>for</strong><br />

structural differences in <strong>for</strong>est stages.<br />

In summary, <strong>the</strong> following research objectives will be<br />

addressed within this <strong>the</strong>sis<br />

1. Sound review <strong>of</strong> in situ methods <strong>of</strong> LAI<br />

assessment toge<strong>the</strong>r with a judgement on<br />

transferability and applicability <strong>of</strong> <strong>the</strong>se<br />

methods in tropical rain <strong>for</strong>ests.<br />

2. Establishment <strong>of</strong> a representative and valid<br />

sampling scheme (<strong>based</strong> on CEOS-LPV<br />

and VALERI recommendations) <strong>for</strong> in situ<br />

measurements <strong>of</strong> LAI on <strong>the</strong> test sites.<br />

3. Per<strong>for</strong>mance <strong>of</strong> in situ LAI measurements and<br />

analysis <strong>of</strong> ga<strong>the</strong>red field data (also with respect<br />

to different <strong>for</strong>est stages).<br />

4. Upscaling <strong>of</strong> field measurements <strong>based</strong> on high<br />

resolution satellite data and derivation <strong>of</strong> high<br />

resolution LAI maps <strong>for</strong> <strong>the</strong> test sites.<br />

5. <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product <strong>based</strong> on<br />

derived high resolution LAI maps.<br />

6. Check <strong>of</strong> temporal consistency <strong>of</strong> <strong>the</strong> <strong>of</strong> <strong>the</strong><br />

<strong>MODIS</strong> LAI product.<br />

This research was carried out under <strong>the</strong> framework<br />

<strong>of</strong> <strong>the</strong> international and interdisciplinary research<br />

network BIOTA AFRICA, funded by <strong>the</strong> German<br />

Ministry <strong>of</strong> Education and Research (BMBF). As part<br />

<strong>of</strong> <strong>the</strong> regional initiative BIOTA East Africa, which<br />

focuses on biodiversity changes in Kenyan and<br />

Ugandan rain <strong>for</strong>ests with different degradation and<br />

disturbance grades, this study <strong>of</strong>fered <strong>the</strong> opportunity<br />

to get access to two <strong>of</strong> <strong>the</strong>se <strong>for</strong>ests <strong>for</strong> detailed<br />

field and validation campaigns. Close cooperation<br />

with <strong>the</strong> BIOTA West and South Africa projects<br />

led to comparable validation exercises in Namibia<br />

and Burkina Faso and will serve as comparison<br />

<strong>for</strong> <strong>the</strong> methodological aspects <strong>of</strong> this study in<br />

later publications. Additionally, field data and high<br />

resolution LAI maps will be made available within<br />

<strong>the</strong> VALERI and LPV networks <strong>for</strong> <strong>the</strong> validation <strong>of</strong><br />

biophysical products derived from o<strong>the</strong>r sensors.<br />

7


2<br />

Figure 2-1<br />

The study areas<br />

Map <strong>of</strong> East Africa with <strong>the</strong> study sites in Kenya and Uganda.<br />

Research in <strong>the</strong> BIOTA East Africa project focuses<br />

on three tropical rain <strong>for</strong>ests, namely Mabira and<br />

Budongo Forests in Uganda and Kakamega Forest<br />

in Kenya. Each study site shows different grades<br />

<strong>of</strong> anthropogenic interference, with Kakamega and<br />

Budongo Forests including areas that are considered<br />

relatively undisturbed (Mitchell & Schaab, in print).<br />

By contrast, not a single part <strong>of</strong> Mabira Forest<br />

can be identified as unlogged (Mitchell, personal<br />

communication). To cover as many different <strong>for</strong>est<br />

types with fieldwork as possible (with respect to<br />

stand age, logging history and canopy structure)<br />

field campaigns were thus conducted in Kakamega<br />

and Budongo Forests (cf. Figure 2-1). Both study<br />

sites will be described in <strong>the</strong> following text with<br />

respect to <strong>the</strong>ir physiogeographic aspects, vegetation<br />

characteristics and <strong>for</strong>est management.<br />

9


10<br />

2.1<br />

2.1.1<br />

Budongo Forest<br />

Physiogeographic aspects<br />

Budongo Forest is one <strong>of</strong> Uganda’s largest remaining<br />

rain <strong>for</strong>ests, ranking third in species diversity<br />

nationwide (Howard et al. 1997, Reynolds 2005). It is<br />

situated in <strong>the</strong> northwest <strong>of</strong> <strong>the</strong> country in <strong>the</strong> Hoima<br />

and Masindi districts, close to <strong>the</strong> eastern shoreline <strong>of</strong><br />

Lake Albert (cf. Figure 2-1). Including all remaining<br />

<strong>for</strong>est fragments, its geographic location ranges<br />

approximately from N 169,500 to N 218,800 and E<br />

317,100 to E 369,100 (Projection: UTM, Zone 36 N,<br />

Datum: WGS 84).<br />

Figure 2-2<br />

Figure 2-2 shows a subset <strong>of</strong> a Landsat ETM+ image<br />

acquired on 17 February, 2000. Budongo Forest lies<br />

in <strong>the</strong> centre with <strong>the</strong> <strong>for</strong>ested areas appearing in<br />

bright green colours. Lake Albert is clearly visible<br />

on <strong>the</strong> western side, with <strong>the</strong> Albertine Rift stretching<br />

along its shoreline from southwest to nor<strong>the</strong>ast. As<br />

<strong>the</strong> scene was acquired at <strong>the</strong> end <strong>of</strong> <strong>the</strong> dry season<br />

and severe fires had occurred in <strong>the</strong> woodlands and<br />

grasslands <strong>of</strong> Bugungu Game Reserve, burnt areas<br />

are perceptible and represented by purple colours.<br />

According to Howard et al. (1996) <strong>the</strong> reserve covers<br />

an area <strong>of</strong> 793 km² and comprises four smaller<br />

blocks, namely Budongo itself (364 km²), Siba in<br />

Landsat ETM+ scene acquired on 17 February, 2000, showing Budongo Forest (band combination 5-4-3, Projection:<br />

UTM 36 N, WGS84). The protected areas <strong>of</strong> Budongo Forest Reserve (including <strong>the</strong> main <strong>for</strong>est blocks), Bugungu<br />

and Karuma Game Reserves as well as Murchinson Falls National Park are outlined in yellow (<strong>based</strong> on data <strong>of</strong> <strong>the</strong><br />

National Forestry Authority, NFA, Uganda).


2 Study areas<br />

<strong>the</strong> southwest (66 km²), Kitigo in <strong>the</strong> northwest (95<br />

km²) and Kanyo-Pabidi in <strong>the</strong> nor<strong>the</strong>ast (268 km²).<br />

An estimated 47% <strong>of</strong> <strong>the</strong> reserve comprises partly<br />

wooded grassland communities (Howard 1991). They<br />

are mainly found in Kanyo-Pabidi and Kitigo and are<br />

indicated by light green and reddish tones that are<br />

clearly distinguishable from <strong>the</strong> lush green colours <strong>of</strong><br />

<strong>the</strong> <strong>for</strong>ested areas.<br />

The <strong>for</strong>est itself lies approximately between 870 m<br />

and 1170 m a.s.l. with <strong>the</strong> whole area gently sloping<br />

down towards <strong>the</strong> escarpment <strong>of</strong> <strong>the</strong> Albertine Rift<br />

in <strong>the</strong> northwest. It is separated by several partly<br />

ephemeral rivers draining ei<strong>the</strong>r through Weiga,<br />

Waisoke, Sonso, or Bubwe Rivers towards Lake<br />

Albert.<br />

Budongo Forest is surrounded by populated areas<br />

in <strong>the</strong> south, southwest and sou<strong>the</strong>ast. Subsistence<br />

farming is widely spread and indicated by <strong>the</strong> smallscale<br />

mosaic <strong>of</strong> lighter and darker reddish tones. In<br />

<strong>the</strong> very south <strong>of</strong> Budongo Forest large sugarcane<br />

fields are visible in pink (harvested) and light green<br />

(standing crops); <strong>the</strong>y belong to <strong>the</strong> second leading<br />

sugar producer in Uganda, Kinyara Sugar Works<br />

Limited. In <strong>the</strong> north and nor<strong>the</strong>ast, <strong>the</strong> <strong>for</strong>est<br />

reserve is contiguous with Bugungu Game Reserve<br />

and Murchinson Falls National Park. Towards<br />

<strong>the</strong> nor<strong>the</strong>ast Karuma Game Reserve extends <strong>the</strong><br />

protected area.<br />

Geology and Soils<br />

The largest part <strong>of</strong> Budongo Forest lies on an<br />

Archean gneissic-granulitic basement complex<br />

that makes up large parts <strong>of</strong> nor<strong>the</strong>rn, eastern<br />

and northwestern Uganda. It was created during<br />

several orogenic processes along <strong>the</strong> margin <strong>of</strong><br />

<strong>the</strong> Congo and Tanzania Cratons thus creating a<br />

polyphase mobile belt (Schlüter 1997). Rocks in <strong>the</strong><br />

Budongo region were affected by Watian (2.88 Ga,<br />

11<br />

granulite facies) and Aruan (2.25 Ga, amphibolite<br />

facies) tectono-metamorphic events and consist <strong>of</strong><br />

highly metamorphic granulites, paragneisses and<br />

orthogneisses toge<strong>the</strong>r with various schists, quarzites<br />

and marbles (Schlüter 1997).<br />

The sou<strong>the</strong>astern part <strong>of</strong> Budongo Forest is occupied<br />

by <strong>the</strong> low metamorphic sedimentary rocks <strong>of</strong> <strong>the</strong><br />

Bunyoro Series (see MacDonald 1966). The series<br />

is <strong>of</strong> late Precambrian age and rests uncon<strong>for</strong>mably<br />

on <strong>the</strong> crystalline basement complex, probably<br />

representing a shallow remnant <strong>of</strong> a <strong>for</strong>merly more<br />

extensive sedimentary <strong>for</strong>mation (Bjørlykke 1973,<br />

MacDonald 1966). The sequence is described<br />

as “sandstones and pellitic sediments” with<br />

“conglomerates and pebbly mudstones in <strong>the</strong> lower<br />

parts” <strong>of</strong> glacial origin (Bjørlykke 1973).<br />

To <strong>the</strong> west, Budongo Forest stretches towards <strong>the</strong><br />

above-mentioned escarpment <strong>of</strong> <strong>the</strong> Albertine Rift<br />

that is situated only about 10 kilometres away. It is<br />

<strong>the</strong> result <strong>of</strong> tectonic activities that started during<br />

<strong>the</strong> Miocene and continue today. The uplift <strong>of</strong> <strong>the</strong><br />

rift walls led to a modification <strong>of</strong> regional drainage<br />

patterns, as <strong>the</strong> area <strong>of</strong> Uganda and western Kenya<br />

was draining towards <strong>the</strong> west into <strong>the</strong> Atlantic in<br />

<strong>the</strong> late Miocene-Pliocene, but later changed towards<br />

<strong>the</strong> north through <strong>the</strong> Nile River, thus influencing<br />

regional and interrift climates (Schlüter 1997).<br />

Eggeling (1947) states that <strong>for</strong>merly existing lateritic<br />

layers, having <strong>the</strong>ir origin in <strong>the</strong> peneplanation<br />

following <strong>the</strong> uplift <strong>of</strong> <strong>the</strong> rift shoulders, are mainly<br />

eroded today and only remain on a few hills<br />

surrounding <strong>the</strong> <strong>for</strong>est. The soils in <strong>the</strong> Budongo<br />

region are heavily wea<strong>the</strong>red and mainly consist<br />

<strong>of</strong> ferralitic clay soils (Sheil 1999). According to<br />

Eggeling (1947) <strong>the</strong>y vary from heavy loam or sandy<br />

clay to very sandy loams. Immediately below <strong>the</strong><br />

humus horizon this soil is reported to be less clayey<br />

and sandier than at greater depth. Soils in <strong>the</strong> Budongo<br />

region are moderately fertile (Howard 1991).


12<br />

Climate<br />

As in most parts <strong>of</strong> Uganda, <strong>the</strong> climate <strong>of</strong> Budongo<br />

Forest is governed by <strong>the</strong> country’s latitudinal<br />

position, <strong>the</strong> resulting movements <strong>of</strong> <strong>the</strong> Intertropical<br />

Convergence Zone (ITCZ), and altitudinal and<br />

topographical characteristics. In consequence<br />

Budongo Forest’s precipitation pattern is bimodal<br />

with more or less pronounced dry seasons.<br />

Figure 2-3a shows <strong>the</strong> mean monthly rainfall as<br />

recorded at Sonso Camp between 1993 and 2004<br />

(source: Budongo Forest Project, courtesy <strong>of</strong> Ge<strong>of</strong>frey<br />

Muhanguzi). Though <strong>the</strong> observed time frame is<br />

short and climate stations with records over several<br />

decades do exist outside <strong>the</strong> <strong>for</strong>est (<strong>for</strong> instance at<br />

Busingiro or Nyabyeya, cited in Reynolds 2005), data<br />

from Sonso is preferred <strong>for</strong> <strong>the</strong> illustration below, as<br />

precipitation can vary significantly between <strong>the</strong> <strong>for</strong>est<br />

and surrounding areas. Un<strong>for</strong>tunately, data from 2004<br />

to 2007 was not available.<br />

The two rainfall peaks between March to May and<br />

September to November are clearly visible as are<br />

<strong>the</strong> dry seasons, which are experienced from mid-<br />

December to mid-February and less pronounced<br />

between June and July (see also Reynolds 2005).<br />

Figure 2-3<br />

However, it has to be noted that following an El-Niño-<br />

Sou<strong>the</strong>rn-Oscillation event, December 1997 was<br />

exceptionally wet with <strong>the</strong> precipitation summing up<br />

to 520 mm. The mean value <strong>for</strong> that month as shown<br />

in Figure 2-3a was thus strongly influenced (mean<br />

monthly rainfall <strong>for</strong> December drops from 84 mm<br />

to 40 mm when excluding data from 1997). Yearly<br />

rainfall at Sonso varied between 1,240 mm and<br />

2,187 mm in <strong>the</strong> observed period, with a mean <strong>of</strong><br />

1,642 mm (±232 mm).<br />

Temperature records from <strong>the</strong> Sonso site (same<br />

source) only exist <strong>for</strong> an even shorter period due to<br />

inconsistent data sets from April 1999 on. Figure 2-3b<br />

shows <strong>the</strong> mean monthly maximum and minimum<br />

temperatures from 1993 to 1999. As expected, mean<br />

maximum temperatures are highest during <strong>the</strong> dry<br />

season from December to February with more than<br />

30°C on average. The short dry season in June/July<br />

has in turn no effect on mean maximum temperatures.<br />

Mean minimum temperatures show little variance<br />

throughout <strong>the</strong> year, with slightly higher values<br />

during <strong>the</strong> rainy seasons. Mean monthly values<br />

range from 14.3–15.7°C minimum and 24.3–31.0°C<br />

maximum temperatures with an annual mean <strong>of</strong><br />

21.1°C (±1.1°C).<br />

Climatic characteristics <strong>of</strong> Budongo Forest. a) Mean monthly rainfall at Sonso Camp from 1993-2004 and b) mean<br />

monthly temperatures at Sonso Camp from 1993-1999 (source: Budongo Forest Project).


2 Study areas<br />

2.1.2<br />

Vegetation characteristics<br />

Biogeographic position<br />

Apart from recent anthropogenic influences, climatic<br />

fluctuations since <strong>the</strong> Upper Quaternary have been<br />

a major driver <strong>for</strong> <strong>the</strong> distribution <strong>of</strong> <strong>for</strong>est species<br />

in East Africa (Hamilton 1981). Based on pollen<br />

diagrams and <strong>the</strong> analysis <strong>of</strong> floristic disjunctions,<br />

assumptions can be made about <strong>the</strong> <strong>for</strong>mer<br />

distribution <strong>of</strong> <strong>for</strong>ested areas as well as changes in<br />

<strong>the</strong>ir extent during <strong>the</strong> late Pleistocene and Holocene.<br />

Especially water availability greatly influenced<br />

rain <strong>for</strong>est distribution. Whereas rain <strong>for</strong>ests were<br />

restricted to a small number <strong>of</strong> refuges (e.g., in <strong>the</strong><br />

eastern Democratic Republic <strong>of</strong> <strong>the</strong> Congo, DRC)<br />

during a very dry period be<strong>for</strong>e 12,000 B.P., <strong>the</strong>y<br />

greatly increased during <strong>the</strong> following warmer and<br />

wetter stages (Hamilton 1981, Reynolds 2005).<br />

Figure 2-4 shows <strong>the</strong> assumed movement <strong>of</strong> <strong>for</strong>est<br />

species after 10,000 B.P. across <strong>the</strong> country with<br />

<strong>the</strong> <strong>for</strong>mer refuges in DRC and a probable minor<br />

Figure 2-4<br />

Assumed movement <strong>of</strong> <strong>for</strong>est species after<br />

10,000 B.P. across Uganda (Hamilton 1982,<br />

modified).<br />

13<br />

one in <strong>the</strong> western Lake Victoria region as starting<br />

points. It has to be pointed out that <strong>the</strong> extent <strong>of</strong> <strong>the</strong><br />

<strong>for</strong>mer <strong>for</strong>est cover is still a question <strong>of</strong> scientific<br />

discussion. Whereas generic similarities as well as<br />

common species <strong>of</strong> coastal <strong>for</strong>ests in East Africa and<br />

<strong>the</strong> <strong>for</strong>ested regions in <strong>the</strong> DRC are evident, <strong>the</strong>re<br />

is no pro<strong>of</strong> yet whe<strong>the</strong>r both areas were linked by a<br />

vast continuous <strong>for</strong>est cover, as assumed by Schmidt<br />

(1992), that would have included both Budongo<br />

and Kakamega Forests. Alternatively, small isolated<br />

<strong>for</strong>ests that expanded and maybe even connected<br />

during wetter periods and shrank again during dry<br />

conditions could have contributed to seed dispersal<br />

<strong>of</strong> <strong>for</strong>est species towards <strong>the</strong> East African coastline<br />

(Hamilton 1981, Lind & Morrison 1974).<br />

Lowland <strong>for</strong>est cover in East Africa probably reached<br />

its maximum in a relatively warm and wet period<br />

between 7,000 and 5,000 B.P. (Kendall 1969 and<br />

Reynolds 2005, derived from pollen analysis). After<br />

that time trees characteristic <strong>for</strong> semideciduous<br />

<strong>for</strong>ests became more abundant, whereas pollen from<br />

trees found in both evergreen and semideciduous<br />

Figure 2-5<br />

Present distribution <strong>of</strong> <strong>for</strong>est and estimated<br />

distribution <strong>of</strong> <strong>for</strong>est be<strong>for</strong>e clearance (Hamilton<br />

1974, modified).


14<br />

<strong>for</strong>ests decreased in numbers according to a study<br />

from Pilkington Bay, Lake Victoria (Kendall 1969).<br />

Today, one <strong>of</strong> <strong>the</strong> major factors affecting <strong>for</strong>est cover<br />

and species diversity is anthropogenic influence.<br />

According to Hamilton (1974) anthropogenic<br />

<strong>for</strong>est destruction began already around 1,000 B.P.,<br />

but human pressure on remaining <strong>for</strong>est patches<br />

tremendously increased within <strong>the</strong> last 100 years.<br />

As a result, tropical <strong>for</strong>ests in Uganda have suffered<br />

massive loss. Figure 2-5 shows <strong>the</strong> <strong>for</strong>est distribution<br />

in 1974 (dark green) in relation to <strong>the</strong> assumed <strong>for</strong>mer<br />

limits <strong>of</strong> <strong>for</strong>est cover <strong>based</strong> on distribution <strong>of</strong> <strong>for</strong>est<br />

remnants, precipitation and altitude (light green).<br />

According to Andrua (2003) Uganda’s tropical <strong>for</strong>ests<br />

decreased from 12.7% <strong>of</strong> <strong>the</strong> land area in 1900 to<br />

3.7% in 2002. This corresponds to 8,847 km² <strong>of</strong> rain<br />

<strong>for</strong>est cover in <strong>the</strong> country. An alarming 32% <strong>of</strong> <strong>the</strong>se<br />

<strong>for</strong>est areas are degraded.<br />

Vegetation classification and floristics<br />

The distinction <strong>of</strong> <strong>for</strong>est <strong>for</strong>mations in tropical Africa<br />

is usually <strong>based</strong> on floristic elements as well as<br />

environmental factors such as altitude and moisture.<br />

With respect to floristic elements, White’s descriptive<br />

memoir on <strong>the</strong> UNESCO vegetation map <strong>of</strong> Africa<br />

(White 1983), classifies Budongo Forest as belonging<br />

to <strong>the</strong> Lake Victoria Regional Mosaic floristic region.<br />

This transition zone is influenced by <strong>the</strong> phytochoria<br />

<strong>of</strong> <strong>the</strong> Guineo-Congolian, Sudanian, Zambezian,<br />

Somaila-Masai and Afromontane floristic regions.<br />

In consequence only very few endemic species and<br />

probably no endemic genera are being recorded <strong>for</strong><br />

this zone (White 1983). In Budongo Forest however,<br />

elements <strong>of</strong> <strong>the</strong> Guineo-Congolian phytochorion<br />

prevail.<br />

Whitmore (2003) fur<strong>the</strong>r distinguishes five types<br />

<strong>of</strong> rain <strong>for</strong>est <strong>for</strong>mations according to elevation and<br />

climate (cf. Table 2-1). As Budongo Forest is situated<br />

in a seasonally dry climate with less than 100 mm<br />

Table 2-1<br />

Formations <strong>of</strong> tropical moist <strong>for</strong>ests<br />

(modified according to Whitmore 2003).<br />

Climate Elevation a.s.l. Forest <strong>for</strong>mation<br />

Seasonally dry Semi-evergreen<br />

rain <strong>for</strong>est<br />

Perhumid < 1200 m Lowland evergreen<br />

rain <strong>for</strong>est<br />

Perhumid 1200-1500 m Lower montane<br />

rain <strong>for</strong>est<br />

Perhumid 1500-3000 m Upper montane<br />

rain <strong>for</strong>est<br />

Perhumid > 3000 m Subalpine <strong>for</strong>est<br />

to tree line<br />

<strong>of</strong> rainfall during some months, it is categorized as<br />

semi-evergreen rain <strong>for</strong>est. Langdale-Brown et al.<br />

(1964) mention, however, that <strong>the</strong> transition between<br />

evergreen and semi-evergreen rain <strong>for</strong>ests is fluid and<br />

in consequence somewhat arbitrary. The latter shows<br />

a higher amount <strong>of</strong> leaf-shedding trees <strong>for</strong> a slightly<br />

longer period due to climatic conditions. It should be<br />

noted that in semi-evergreen rain <strong>for</strong>ests only some<br />

species shed <strong>the</strong>ir leaves, and not necessarily all at<br />

<strong>the</strong> same time (White 1983), resulting in slightly<br />

lower LAI values during <strong>the</strong> dry seasons.<br />

With respect to floristics Budongo Forest is one <strong>of</strong><br />

<strong>the</strong> best described <strong>for</strong>ests in Uganda (Howard et al.<br />

1996). Since 1930 many researchers have conducted<br />

vegetation studies <strong>of</strong> different scopes. Among <strong>the</strong> most<br />

important publications are <strong>the</strong> <strong>of</strong>ten cited description<br />

<strong>of</strong> vegetation types from Eggeling (1947), Plumptre’s<br />

studies on <strong>the</strong> effects <strong>of</strong> selective logging on species<br />

distribution (Plumptre 1995, and 1996) and some<br />

more recent vegetation surveys that can partly be<br />

seen as a continuation <strong>of</strong> Eggeling’s work. A critical<br />

review <strong>of</strong> Eggeling’s hypo<strong>the</strong>ses on succession was<br />

conducted by Sheil (e.g. 1995, 1998, and 1999). Apart<br />

from that, inventories were carried out on behalf <strong>of</strong> <strong>the</strong><br />

Uganda Forest Department in 1986 (<strong>based</strong> on a threeday<br />

reconnaissance and aerial imagery, published


2 Study areas<br />

by Howard 1991) and, more thorough, in 1993 and<br />

1994 (<strong>based</strong> on 42 days <strong>of</strong> fieldwork and sampling<br />

1,053 km <strong>of</strong> transects throughout <strong>the</strong> <strong>for</strong>est, published<br />

in <strong>the</strong> Budongo Forest Reserve Biodiversity Report<br />

by Howard et al. 1996). As a result <strong>of</strong> <strong>the</strong> latter,<br />

465 tree and shrub species are now known to exist<br />

<strong>the</strong>re, 93 <strong>of</strong> which are restricted-range species that<br />

only occur in five o<strong>the</strong>r <strong>for</strong>ests in Uganda. In <strong>the</strong><br />

following, <strong>the</strong> data published in <strong>the</strong> above-mentioned<br />

sources is summarized.<br />

According to Eggeling (1947) four <strong>for</strong>est types can<br />

be distinguished in Budongo Forest, three <strong>of</strong> which<br />

follow an ecological succession series. Apart from<br />

swamp <strong>for</strong>est, that <strong>for</strong>ms an edaphic climax stadium<br />

and occupied only 2% <strong>of</strong> <strong>the</strong> total <strong>for</strong>est area at <strong>the</strong><br />

time <strong>of</strong> Eggeling’s survey, mainly colonising <strong>for</strong>est,<br />

mixed <strong>for</strong>est and a so-called ironwood <strong>for</strong>est can be<br />

distinguished. Colonising <strong>for</strong>est is <strong>the</strong> youngest stage,<br />

occurring mainly at <strong>the</strong> <strong>for</strong>est edge and covering 6%<br />

<strong>of</strong> <strong>the</strong> <strong>for</strong>ested land. It is described as moderately<br />

species rich and comprises Maesopsis eminii stands<br />

on deeper and better soils and a species mixture <strong>of</strong><br />

Olea welwitschii, Sapium ellipticum and Phyllanthus<br />

discoideus on shallow poor soils. Most <strong>of</strong> <strong>the</strong>se<br />

species are light loving and <strong>the</strong>re<strong>for</strong>e replaced very<br />

quickly by a mixed <strong>for</strong>est with high species richness,<br />

including several mahogany and sapotaceous species<br />

(Sheil 1999). Eggeling (1947) estimated that mixed<br />

<strong>for</strong>est covered <strong>the</strong> largest part <strong>of</strong> <strong>the</strong> <strong>for</strong>est reserve,<br />

approximately 60% <strong>of</strong> <strong>the</strong> area. The most abundant<br />

species include Khaya anthoteca, Entandrophragma<br />

spp. and Cynometra alexandri. Mixed <strong>for</strong>est species<br />

regenerate in gaps, but gradually and over a long time<br />

period only <strong>the</strong> more shade-tolerant trees survive<br />

and ironwood <strong>for</strong>est develops, with Cynometra<br />

alexandri being <strong>the</strong> dominant species in <strong>the</strong> canopy<br />

and Lasiodiscus mildbraedii in <strong>the</strong> understorey. This<br />

is also supported by Synott (personal communication<br />

in Reynolds 2005), who discovered that Cynometra<br />

spp. are less <strong>of</strong>ten eaten by rodents and grow more<br />

successfully in shady conditions. According to<br />

15<br />

Eggeling (1947), ironwood <strong>for</strong>est covered around<br />

32% <strong>of</strong> Budongo Forest at <strong>the</strong> time <strong>of</strong> his survey.<br />

Whereas <strong>the</strong> above description <strong>of</strong> colonising <strong>for</strong>est<br />

has found wide acceptance, Eggeling’s assumptions<br />

<strong>of</strong> later succession stages have prompted controversial<br />

discussion (Langdale-Brown et al. 1964, Lieth &<br />

Werger 1989, Sheil 1999). Especially his <strong>the</strong>ory<br />

about late-successional species decline, resulting in<br />

<strong>the</strong> dominance <strong>of</strong> Cynometra alexandri in <strong>the</strong> upper<br />

tree layer, has been a subject <strong>of</strong> ongoing discussion.<br />

Whereas Langdale-Brown et al. (1964) agree in<br />

most parts with Eggeling’s description, <strong>the</strong>y also<br />

mention that Cynometra <strong>for</strong>est might only represent<br />

<strong>the</strong> climax stage on poor soils, relying on evidence<br />

from observations in Semliki and Bugoma Forests. In<br />

<strong>the</strong>ir opinion mixed <strong>for</strong>est can also develop to Celtisor<br />

Chrysophyllum-dominated types. This <strong>the</strong>ory is<br />

in line with Sheil (1999), who evaluated Eggeling’s<br />

successional interpretation and found inconsistencies<br />

<strong>of</strong> <strong>the</strong> data with time series collected on Eggeling’s<br />

original plots <strong>for</strong> <strong>the</strong> following 60 years. Sheil (1999)<br />

proposes a possible change <strong>of</strong> successional processes<br />

as <strong>the</strong> reason.<br />

It is indisputable that since Eggeling’s study <strong>the</strong> <strong>for</strong>est<br />

and its vegetation composition has been influenced<br />

strongly by anthropogenic processes. Many parts <strong>of</strong><br />

<strong>the</strong> <strong>for</strong>est have been disturbed by <strong>for</strong>mer or present<br />

legal or illegal logging activities, thus annulling<br />

any <strong>the</strong>ory <strong>of</strong> climax vegetation (though <strong>the</strong>re are<br />

protected areas that are still pretty well intact). A very<br />

detailed study <strong>of</strong> Plumptre (1996) <strong>based</strong> on aerial<br />

imagery <strong>of</strong> 1951 and 1990, <strong>for</strong>est inventory data<br />

and fieldwork showed a decrease in Cynometra and<br />

Cynometra-mixed <strong>for</strong>est <strong>of</strong> more than 205 km² during<br />

<strong>the</strong> above-mentioned time span. Mixed <strong>for</strong>est in turn<br />

increased more than 350 km² also at <strong>the</strong> expense<br />

<strong>of</strong> colonizing <strong>for</strong>est. The state <strong>of</strong> <strong>for</strong>est types in<br />

Budongo Forest <strong>based</strong> on Plumptre’s work is shown<br />

in Figure 2-6. Cynometra-mixed <strong>for</strong>est is pictured in<br />

Figure 2-7a.


16<br />

Figure 2-6<br />

Forest types in Budongo Forest in 1990 (Plumptre 1996).<br />

Outside <strong>the</strong> <strong>for</strong>est reserve natural vegetation cover<br />

almost disappeared due to uninhibited exploitation <strong>of</strong><br />

natural resources by <strong>the</strong> local population. The protected<br />

regions in <strong>the</strong> northwest towards Kanyo-Pabidi and<br />

along <strong>the</strong> escarpment line in <strong>the</strong> north and nor<strong>the</strong>ast,<br />

however, still support moist Combretum-Savanna<br />

(Langdale-Brown et al. 1964) with a grass layer up<br />

to 1.5 to 2 m thick and characterized by Hyparrhenia<br />

rufa (cf. Figure 2-7b). The light to moderate tree<br />

cover consists mainly <strong>of</strong> deciduous trees <strong>of</strong> <strong>the</strong><br />

Combretaceae family with abundant Combretum<br />

molle, Terminalia glaucescens and Albizia zygia 3 to<br />

12 m in height. Patches <strong>of</strong> dry Combretum savannas<br />

also exist with varying combinations <strong>of</strong> Combretum-<br />

Terminalia-Loudetia and Combretum-Hyparrhenia<br />

types (Langdale-Brown et al. 1964).<br />

Figure 2-7<br />

Vegetation types in Budongo Forest.<br />

a) Cynometra-mixed and b) moist Combretum-<br />

Savanna close to Kanyo-Pabidi.


2 Study areas<br />

Vegetation structure<br />

Structural differences in rain <strong>for</strong>ests are mainly<br />

associated with different succession or disturbance<br />

stages. According to FAO definitions (FAO 2006),<br />

primary <strong>for</strong>ests are <strong>for</strong>ests that have never been<br />

logged and have developed under natural processes<br />

and following natural disturbances. They are thus<br />

<strong>the</strong> end stage <strong>of</strong> natural succession processes.<br />

As already mentioned, primary <strong>for</strong>est can hardly<br />

be found in Budongo Forest. Solely areas that<br />

were under protection since <strong>the</strong> 1930s and only<br />

experienced selective illegal logging activities can<br />

still be regarded as near primary <strong>for</strong>est. Most o<strong>the</strong>r<br />

parts are ra<strong>the</strong>r characterized as secondary <strong>for</strong>ests<br />

that according to FAO (2006) have been logged and<br />

recovered or have lost, through human disturbance,<br />

<strong>the</strong> structure, function, species composition or<br />

productivity normally associated with primary<br />

<strong>for</strong>ests. Primary and secondary rain <strong>for</strong>ests differ<br />

significantly in vegetation composition and structure.<br />

In order to describe <strong>the</strong> latter, a strata or canopy layer<br />

concept is usually applied. Though this is usually<br />

a simplification <strong>of</strong> reality, it may be a useful aid to<br />

description or analysis (Whitmore 2003).<br />

For Budongo Forest Eggeling (1947) proposed a three<br />

layer concept <strong>for</strong> undisturbed parts which is probably<br />

applicable to many o<strong>the</strong>r lowland and medium<br />

Figure 2-8<br />

Characteristic features <strong>of</strong> trees belonging to <strong>the</strong> upper tree layer. a) Umbrella shaped crowns and b) buttresses.<br />

17<br />

altitude <strong>for</strong>ests in East Africa. In <strong>the</strong> upper layer,<br />

trees are approximately 21-36 m high, have umbrella<br />

shaped crowns and are <strong>of</strong>ten heavily buttressed (cf.<br />

Figure 2-8). Some trees emerge through this stratum,<br />

growing to heights above 40 m. The second layer<br />

is <strong>for</strong>med by individuals ca. 11 m to 12 m high.<br />

They usually have oblong crowns in lateral contact.<br />

Toge<strong>the</strong>r with smaller trees up to 11 m high <strong>the</strong>y <strong>for</strong>m<br />

a closed canopy. Due to <strong>the</strong>se dense strata, <strong>the</strong> ground<br />

is intensively shaded and understorey vegetation in<br />

<strong>the</strong> <strong>for</strong>m <strong>of</strong> shrub and herb cover is hardly present<br />

(Schulz & Wagner 2002). Whereas this concept is<br />

still applicable to only slightly disturbed parts <strong>of</strong><br />

Budongo Forest, it has to be altered <strong>for</strong> those areas<br />

where more or less intense logging took place in <strong>the</strong><br />

20th century.<br />

In secondary <strong>for</strong>est around Sonso Camp, <strong>for</strong> example,<br />

where large amounts <strong>of</strong> timber had been selectively<br />

logged until <strong>the</strong> 1970s, no distinct canopy stories<br />

are present today. The upper canopy is more even as<br />

emergent trees are hardly present. Greater gaps are<br />

still perceptible and as a result <strong>of</strong> <strong>the</strong> higher amount<br />

<strong>of</strong> light reaching <strong>the</strong> <strong>for</strong>est floor a dense cover <strong>of</strong><br />

young shrubs and trees is growing in <strong>the</strong> understorey,<br />

frequently comprising pioneer species. Climbers<br />

(lianas, woody and herbaceous climbers) are more<br />

abundant in this <strong>for</strong>est type than in <strong>the</strong> near primary<br />

<strong>for</strong>est areas (Schulz & Wagner 2002). Vascular


18<br />

epiphytes are also well represented throughout<br />

mature <strong>for</strong>est stages, especially in <strong>the</strong> upper canopy<br />

(Eggeling 1947).<br />

In addition to vertical differences in <strong>for</strong>est structure,<br />

horizontal variability is also present. Apart from a<br />

small-scale mosaic <strong>of</strong> natural growth cycles driven<br />

by gap dynamics in <strong>the</strong> undisturbed <strong>for</strong>est areas,<br />

human interference in terms <strong>of</strong> timber extraction<br />

patterns had a great influence on horizontal structural<br />

variability in Budongo Forest. Plumptre (1996)<br />

found major differences between areas that had been<br />

logged since 1941 and areas that have been protected<br />

<strong>for</strong> more than 80 years. The latter are characterized<br />

by a significantly denser canopy, more trees in <strong>the</strong><br />

lower canopy, fewer lianas and less light reaching <strong>the</strong><br />

<strong>for</strong>est floor. In order to understand spatial patterns <strong>of</strong><br />

<strong>for</strong>est structure, <strong>the</strong> <strong>for</strong>est management and logging<br />

history over <strong>the</strong> last 70 years will be described in <strong>the</strong><br />

following chapter.<br />

2.1.3<br />

Forest management<br />

According to Howard (1991) Budongo Forest,<br />

being famous <strong>for</strong> its precious mahogany trees, has<br />

been <strong>the</strong> largest and most valuable timber <strong>for</strong>est<br />

in Uganda. The first sawmill was established in<br />

1926 (Eggeling 1947). Be<strong>for</strong>e that time only minor<br />

quantities <strong>of</strong> timber had been cut by pit sawyers,<br />

mainly in <strong>the</strong> sou<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> <strong>for</strong>est. After 1926<br />

more extensive exploitation began by concession<br />

holders. Between 1932 and 1939 Budongo Forest was<br />

gazetted as a Forest Reserve by <strong>the</strong> British Colonial<br />

Administration (Howard 1991). At <strong>the</strong> same time,<br />

<strong>the</strong> first <strong>of</strong> several detailed 10-year working plans<br />

was elaborated in order to control exploitation. The<br />

<strong>for</strong>est was divided into different compartments with<br />

individual logging schemes. Additionally, a nature<br />

reserve <strong>of</strong> approximately 100 ha was established<br />

in <strong>the</strong> southwestern part <strong>of</strong> <strong>the</strong> <strong>for</strong>est (N15), where<br />

vegetation was to remain untouched.<br />

The working plans prescribed a rotating system <strong>of</strong><br />

selective logging with species-dependent rotating<br />

times <strong>of</strong> 40-, 60- and 80-years and minimum girths<br />

(Reynolds 2005). Mechanical logging activities<br />

increased until <strong>the</strong> 1960s and decreased in <strong>the</strong> 1970s,<br />

with <strong>the</strong> most valuable timber species being <strong>the</strong><br />

mahoganies Khaya anthoteca, Entandrophragma<br />

angolense, E. cylindricum and E. utile as well as <strong>the</strong><br />

ironwood Cynometra alexandri that was used <strong>for</strong><br />

flooring, amongst o<strong>the</strong>rs (Eggeling 1947, Plumptre<br />

1996, Reynolds 2005).<br />

Ef<strong>for</strong>ts were also made to plant mahogany trees<br />

to increase <strong>the</strong> stock inside <strong>the</strong> <strong>for</strong>est, but without<br />

success (Reynolds 2005). Plumptre (1996) reports<br />

arboricide treatment in <strong>the</strong> 1950s to 70s <strong>for</strong> tree<br />

species with little or no market value, such as Ficus<br />

spp, Celtis spp. and Lasiodiscus spp. A mixture <strong>of</strong><br />

2,4,5-T and 2,4-D toge<strong>the</strong>r with diesel was applied (a<br />

similar mixture became famous during <strong>the</strong> Vietnam<br />

war under <strong>the</strong> name Agent Orange) and was supposed<br />

to prevent <strong>the</strong>se species from competing with <strong>the</strong><br />

mahoganies. As consequence, structural changes<br />

and changes in vegetation types can still be observed<br />

in Budongo Forest after 60 years <strong>of</strong> commercial<br />

and illegal timber extraction as well as arboricide<br />

treatment (Plumptre 1996).<br />

Today <strong>the</strong> <strong>for</strong>est is managed by <strong>the</strong> Uganda Forest<br />

Department through <strong>the</strong> Masindi District Forest<br />

Office with two local stations at Nyakafunjo and<br />

Biiso. The current Management Plan (1997-2007)<br />

<strong>for</strong>esees <strong>the</strong> expansion <strong>of</strong> strict nature reserves,<br />

where no human activity is allowed to take place,<br />

surrounded by buffer zones (Figure 2-9). It should be<br />

noted however, that with respect to logging history,<br />

only compartments N15, W17, W31-34 and KP11-13<br />

can be considered as unlogged <strong>for</strong> <strong>the</strong> past 80 years<br />

(Plumptre 1996). Though <strong>the</strong> creation <strong>of</strong> protected<br />

zones is definitely a step in <strong>the</strong> right direction, human<br />

interference in <strong>the</strong> <strong>for</strong>est remains. Even parts that<br />

should stay totally undisturbed, such as <strong>the</strong> nature


2 Study areas<br />

Figure 2-9<br />

Logging compartments in Budongo Forest and <strong>for</strong>est management (made available by NFA through BIOTA-E02).<br />

reserve in compartment N 15, showed clear signs<br />

<strong>of</strong> illegal logging during <strong>the</strong> time <strong>of</strong> <strong>the</strong> fieldwork<br />

(Figure 2-10). This is also confirmed by Reynolds<br />

(2005) <strong>for</strong> o<strong>the</strong>r parts <strong>of</strong> <strong>the</strong> <strong>for</strong>est.<br />

Apart from logging, fur<strong>the</strong>r anthropogenic threats<br />

to Budongo Forest include charcoal burning, illegal<br />

encroachment (crop planting on <strong>for</strong>est meadows),<br />

firewood collection and illegal grazing activities<br />

(Aryal 2002, Reynolds 2005). Fur<strong>the</strong>r in<strong>for</strong>mation on<br />

<strong>the</strong> <strong>for</strong>est compartments can be found in Table A-1<br />

(cf. Appendix <strong>of</strong> Tables).<br />

Figure 2-10<br />

19<br />

Illegal logging activities in <strong>the</strong> nature reserve N15.


20<br />

2.2<br />

2.2.1<br />

Kakamega Forest<br />

Physiogeographic aspects<br />

Kakamega Forest is situated in Kenya’s Western<br />

Province between N 41,236 and N 15,984 and<br />

E 696,777 and E 717,761 E (Projection: UTM, 36 N,<br />

WGS 84) at an altitude <strong>of</strong> about 1,460-1,760 m a.s.l.<br />

Figure 2-11 shows a subset <strong>of</strong> a Landsat ETM+ scene<br />

that was acquired on 10 January, 2003. Forested areas<br />

are shown in dark green colours, whereas bushland,<br />

tea cultivation zones and grasslands appear in brighter<br />

green tones. It is obvious that Kakamega Forest and<br />

its associated <strong>for</strong>est areas are much more fragmented<br />

than Budongo Forest.<br />

Figure 2-11<br />

Within its <strong>of</strong>ficial boundaries <strong>the</strong> <strong>for</strong>est covers an area<br />

<strong>of</strong> 236 km². However only 123 km² comprise natural<br />

<strong>for</strong>est (Mitchell et al. 2006). The rest is composed<br />

<strong>of</strong> bushland, grassland and plantations. Kakamega<br />

Forest is fur<strong>the</strong>rmore surrounded by <strong>for</strong>est patches<br />

<strong>of</strong> various sizes, altoge<strong>the</strong>r adding up to 257 km² <strong>of</strong><br />

natural <strong>for</strong>est cover. Apart from <strong>the</strong> South and North<br />

Nandi Forests with which Kakamega Forest was once<br />

connected via South Nandi, <strong>the</strong> largest remaining<br />

<strong>for</strong>est patch is Kisere Forest in <strong>the</strong> north with an<br />

approximate area <strong>of</strong> 4 km² (Mitchell et al. 2006).<br />

Two major rivers run through <strong>the</strong> <strong>for</strong>est in a westerly<br />

direction, <strong>the</strong> Isiukhu in <strong>the</strong> north and <strong>the</strong> Yala in <strong>the</strong><br />

south, both draining towards Lake Victoria.<br />

Landsat ETM+ scene acquired on 10 January 2003, showing Kakamega Forest and its associated <strong>for</strong>est areas<br />

(band combination 5-4-3, Projection: UTM 36 N, WGS8). The <strong>for</strong>est coverage as marked in topographic maps<br />

1:50,000 published in 1970 is outlined in yellow (<strong>based</strong> on aerial photography from 1967).


2 Study areas<br />

Kakamega Forest is surrounded by densely populated<br />

areas (Blackett 1994). Agricultural areas around <strong>the</strong><br />

<strong>for</strong>est are shown in Figure 2-11 in pink and bluish<br />

colours and consist <strong>of</strong> small fields where sugarcane,<br />

maize and beans are grown, amongst o<strong>the</strong>r crops (cf.<br />

Figure 2-12a). In <strong>the</strong> south <strong>of</strong> Kakamega Forest tea is<br />

cultivated along <strong>the</strong> <strong>for</strong>est edges (Figure 2-12b).<br />

Figure 2-12<br />

Geology and Soils<br />

Surroundings <strong>of</strong> Kakamega Forest.<br />

a) Agricultural land between Buyangu and Kisere<br />

(photo: C. Brueckl), b) tea zone at <strong>the</strong> <strong>for</strong>est<br />

edge close to Isecheno Forest Station.<br />

The rocks beneath <strong>the</strong> Kakamega Forest region<br />

belong to <strong>the</strong> Nyanzian and <strong>the</strong> overlying Kavirondian<br />

systems (Cahen et al. 1984, Schlüter 1997). They<br />

were <strong>for</strong>med in <strong>the</strong> late Archean, approximately 3.0<br />

to 2.5 Ga years ago. The Nyanzian System comprises<br />

21<br />

predominantly volcanic rocks <strong>of</strong> basic to acidic<br />

composition as well as poorly sorted sediments<br />

occurring as greenstone belts within <strong>the</strong> Basement<br />

Complex <strong>of</strong> western Kenya (Hester & Boberg 2006).<br />

The Kavirondian system uncon<strong>for</strong>mably overlays <strong>the</strong><br />

Nyanzian system and consists <strong>of</strong> sediments derived<br />

from <strong>the</strong> erosion <strong>of</strong> supracrustal material, following<br />

<strong>the</strong> <strong>for</strong>mation <strong>of</strong> late granites (Schlüter 1997).<br />

Kavirondian sediments occur in an east-west oriented<br />

syncline between <strong>the</strong> western Nandi slopes and<br />

Yala Town in <strong>the</strong> southwest <strong>of</strong> Kakamega Forest.<br />

The nor<strong>the</strong>rn and sou<strong>the</strong>astern flanks <strong>of</strong> this basin<br />

structure are framed by palaeoproterozoic granitic<br />

intrusions (Schlüter 1997). The Kavirondian system<br />

can be subdivided in four lithographic <strong>for</strong>mations –<br />

from bottom to top called Shivakala (conglomerates<br />

<strong>of</strong> granitic and basaltic, andesitic and rhyolitic<br />

volcanic sources), Igukhu (greywacke), Mroda<br />

(cross-stratified sandstones and minor shales and<br />

greywackes) and Mudaa (shales).<br />

As in Budongo <strong>the</strong> soils <strong>of</strong> Kakamega Forest are<br />

mainly deeply wea<strong>the</strong>red ferralitic soils that are poor<br />

to moderate in <strong>the</strong>ir nutrient content, as occurs very<br />

<strong>of</strong>ten under topical rain <strong>for</strong>est vegetation (Musila et al.<br />

2004, Schultka 1974). To a lesser extent <strong>the</strong>re are also<br />

lixisols and cambisols, in some places phaeozems can<br />

also be found (Musila et al. 2004). Schultka (1974)<br />

reported dark reddish to dark brownish-red crumbly,<br />

sandy mudstones and sandy, loamy mudstones,<br />

mainly occurring on granite. In lower horizons he<br />

even found laterite crusts and compressed layers.<br />

Climate<br />

Analogous to Budongo Forest <strong>the</strong> climate <strong>of</strong> Western<br />

Kenya is influenced by its latitudinal position and<br />

<strong>the</strong> resulting seasonal ITCZ movements. Kakamega<br />

Forest also experiences a bimodal rainfall pattern<br />

with a long dry season from December to February


22<br />

and a short, less pronounced dry spell in June and<br />

July. Most precipitation falls in <strong>the</strong> rainy seasons<br />

between March and May and August to October. The<br />

average monthly precipitation recorded by <strong>the</strong> <strong>for</strong>mer<br />

Forest Department (now Kenya Forest Service, KFS)<br />

at Isecheno Forest Station between 1982 and 2006<br />

is shown in Figure 2-13a. Mean yearly rainfall is<br />

1,972 mm (±336 mm).<br />

Temperature recorded by a meteorological station<br />

set up by <strong>the</strong> BIOTA project in <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong><br />

<strong>the</strong> <strong>for</strong>est close to <strong>the</strong> <strong>of</strong>fice <strong>of</strong> <strong>the</strong> Kenya Wildlife<br />

Service (KWS) showed daytime maximum values<br />

up to 32°C in February and a minimum nighttime<br />

temperature <strong>of</strong> 9°C recorded in October (Steinbrecher<br />

et al. 2004). As no full year <strong>of</strong> data records was<br />

available, temperature data as cited in Mutangah<br />

(1996) is displayed in Figure 2-13b. Minimum mean<br />

monthly temperatures are again recorded during <strong>the</strong><br />

dry seasons.<br />

Figure 2-13<br />

2.2.2<br />

Vegetation characteristics<br />

Biogeographic position<br />

The main aspects concerning <strong>the</strong> history <strong>of</strong> <strong>for</strong>est<br />

development in East Africa were described in<br />

Chapter 2.2.2. It is again emphasized that according<br />

to Hamilton (1982) <strong>for</strong>est species diversity in this<br />

region decreases with increasing distance to <strong>the</strong><br />

Guineo-Congolian Quaternary refuge. Alth<strong>of</strong> (2005)<br />

mentions, that Kakamega Forest probably developed<br />

in <strong>the</strong> postglacial period after 10,000 B.P., an era<br />

favourable to <strong>for</strong>est expansion. Accordingly, this<br />

relatively young age toge<strong>the</strong>r with <strong>the</strong> distance to <strong>the</strong><br />

Guineo-Congolian centre <strong>of</strong> endemism results in low<br />

species numbers and low endemism in Kakamega<br />

Forest. Fischer (personal communication) states that<br />

only about 400 different vascular plant species are<br />

known from Kakamega, which is equivalent to some<br />

mixed deciduous <strong>for</strong>ests in Germany.<br />

Kakamega Forest is generally referred to as <strong>the</strong><br />

easternmost remnant <strong>of</strong> <strong>the</strong> <strong>for</strong>mer Guineo-Congolian<br />

rain <strong>for</strong>est belt (Kowkaro 1988). On <strong>the</strong> o<strong>the</strong>r hand<br />

Lind & Morrison (1974) claim, “botanical evidence<br />

suggests that <strong>the</strong>re have been at least two periods<br />

Climatic characteristics <strong>of</strong> Kakamega Forest. a) Mean monthly rainfall at Isecheno Forest Station from 1982-2006<br />

(source: Forest Department) and b) mean monthly temperatures at Kakamega Agricultural Station from 1980-1992<br />

(source: Mutangah 1996).


2 Study areas<br />

<strong>of</strong> connection between <strong>the</strong> coastal rain <strong>for</strong>ests [<strong>of</strong><br />

Kenya and Tanzania] and <strong>the</strong> Zaire rain <strong>for</strong>est” (p.<br />

203). This would imply that some still existing <strong>for</strong>est<br />

patches along <strong>the</strong> coastline <strong>of</strong> East Africa could also<br />

be regarded as relicts with strong Guineo-Congolian<br />

influence. Alth<strong>of</strong> (2005), however, argues that <strong>the</strong><br />

influence <strong>of</strong> o<strong>the</strong>r floristic elements in <strong>the</strong>se <strong>for</strong>ests<br />

is higher, whereas in Kakamega Forest species <strong>of</strong><br />

Guineo-Congolian origin prevail.<br />

On a regional scale Mitchell (2004) analysed <strong>the</strong><br />

exploitation and disturbance history <strong>of</strong> Kakamega<br />

Forest and its surrounding <strong>for</strong>est patches <strong>based</strong> on<br />

published data, oral records, and place name evidence.<br />

According to his study, <strong>the</strong> <strong>for</strong>est was still linked to<br />

South Nandi Forest in 1959, which in turn <strong>for</strong>med an<br />

23<br />

entity with North Nandi Forest (cf. Figure 2-11). No<br />

evidence has been found that Kakamega Forest had<br />

ever been fully joined with <strong>the</strong> <strong>for</strong>est patches (<strong>of</strong>ten<br />

referred to as “fragments”) in <strong>the</strong> north and west.<br />

Vegetation classification and floristics<br />

White (1983) also classifies Kakamega Forest as<br />

belonging to <strong>the</strong> floristic region <strong>of</strong> <strong>the</strong> Lake Victoria<br />

Regional Mosaic. Kakamega Forest, however,<br />

harbours a much higher number <strong>of</strong> afromontane<br />

species than Budongo. These might have survived<br />

arid conditions in close-by mountainous regions in<br />

Kenya (e.g., around Mount Elgon, as reported by<br />

Hamilton 1982). According to Alth<strong>of</strong> (2005) about<br />

Figure 2-14 Vegetation in Kakamega Forest. a) Dracena fragrans, b) Deinbollia kilimandscharica-Markhamia lutea alliance in<br />

Kisere Forest, c) Harungana madagascariensis-Desmodium adscendens alliance close to Buyangu Camp and<br />

d) <strong>for</strong>est glade close to Isecheno Nature Reserve.


24<br />

33% <strong>of</strong> <strong>the</strong> woody species recorded in Kakamega<br />

Forest are afromontane floristic elements. Ano<strong>the</strong>r<br />

26% are so-called transition species and 41% are<br />

<strong>of</strong> Guineo-Congolian origin. Of <strong>the</strong> latter, several<br />

species such as Aningeria altissima, Cordia milennii,<br />

Entandrophragma angolsense, Maesopsis eminii<br />

and Moodora myristica reach <strong>the</strong>ir easternmost limit<br />

here. Just as Budongo Forest, Kakamega Forest is<br />

classified as semi-evergreen rain <strong>for</strong>est following<br />

Whitmore’s <strong>for</strong>est <strong>for</strong>mations (see Table 2-1).<br />

Compared to Budongo Forest, relatively few studies<br />

on floristics variation have been conducted in<br />

Kakamega Forest. Though Beentje (1990), Mutangah<br />

(1996), and Fashing & Mwangi Gathua (2004)<br />

analysed <strong>the</strong> tree species composition on certain<br />

plots in Kakamega Forest and Schultka (1974)<br />

studied <strong>for</strong>est regeneration in different surrounding<br />

grasslands and shrublands, Alth<strong>of</strong> (2005) was <strong>the</strong> first<br />

to describe plant communities in more detail. Her<br />

study was <strong>based</strong> on 200 phytosociological relevés that<br />

were distributed throughout <strong>the</strong> <strong>for</strong>est and resulted in<br />

<strong>the</strong> differentiation <strong>of</strong> two vegetation alliances (I and<br />

II) with thirteen communities and subcommunities.<br />

Though a detailed and spatially explicit vegetation<br />

map <strong>of</strong> <strong>the</strong> whole <strong>for</strong>est is still missing, her results<br />

will be briefly described below.<br />

Alliance I mainly refers to more or less disturbed<br />

mature <strong>for</strong>est sites. The dominant tree species in this<br />

alliance are Antiaris toxicaria, which is <strong>of</strong> guineocongolian<br />

origin, and Diospyros abyssinica, an<br />

afromontane element. In <strong>the</strong> lower canopy layers<br />

Bequaertiodendron oblanceolatum, Trilepisium<br />

madagascariense as well as Trichilia emetica,<br />

Heinsenia diervilleoides, Cassipourea ruwensorensis,<br />

Morus mesozygia and Chrysophyllum albidum are<br />

frequent. Dracena fragrans is highly abundant in<br />

<strong>the</strong> shrub layer (cf. Figure 2-14a). Lianas also occur<br />

under <strong>the</strong> dense canopy cover.<br />

Alliance I can be fur<strong>the</strong>r split up into two major<br />

groups, <strong>the</strong> Deinbollia kilimandscharica-Markhamia<br />

lutea (cf. Figure 2-14b) and <strong>the</strong> Celtis mildbraedii-<br />

Craibia brownii alliances. Whereas <strong>the</strong> first group<br />

mainly occurs in <strong>the</strong> nor<strong>the</strong>rn parts <strong>of</strong> Kakamega<br />

(Kisere Forest and Buyangu Nature Reserve) and<br />

comprises different stages <strong>of</strong> middle-aged, old<br />

secondary and near primary <strong>for</strong>ests, <strong>the</strong> second one<br />

mainly occurs in <strong>the</strong> middle and sou<strong>the</strong>rn <strong>for</strong>est parts<br />

and reflects a combination <strong>of</strong> different anthropogenic<br />

treatment and natural variation.<br />

Alliance II indicates young initial <strong>for</strong>est stages<br />

that occur after clear felling when bushland and<br />

subsequent young secondary <strong>for</strong>est develop. Two<br />

species are characteristic <strong>for</strong> this alliance: Harungana<br />

madagascariensis as dominant tree and Desmodium<br />

adscendens in <strong>the</strong> herb layer (cf. Figure 2-14c).<br />

For both major alliances, different communities and<br />

subcommunities can be distinguished. As <strong>the</strong>y are,<br />

however, less important <strong>for</strong> this <strong>the</strong>sis, <strong>the</strong> reader is<br />

referred to Alth<strong>of</strong> (2005) <strong>for</strong> fur<strong>the</strong>r in<strong>for</strong>mation.<br />

Within its <strong>of</strong>ficial <strong>for</strong>est boundary, Kakamega<br />

Forest also contains various glades where grassland<br />

is present (cf. Figure 2-14d). Whe<strong>the</strong>r <strong>the</strong>ir origin<br />

is natural or anthropogenic remains unclear.<br />

Mitchell (2004) mentions that some glades might<br />

possibly be remnants <strong>of</strong> <strong>for</strong>mer un<strong>for</strong>ested areas.<br />

Due to <strong>the</strong>ir shallow soils <strong>for</strong>est tree species<br />

could not establish. Fur<strong>the</strong>rmore, human induced<br />

fire and grazing animals might have kept <strong>the</strong><br />

glades open. However, Mitchell (2004) reports<br />

that <strong>for</strong>est succession is taking place, resulting<br />

in <strong>the</strong> shrinking <strong>of</strong> most glades with some<br />

even disappearing altoge<strong>the</strong>r, a process that<br />

has been especially marked in <strong>the</strong> nor<strong>the</strong>rn<br />

part <strong>of</strong> <strong>the</strong> <strong>for</strong>est since <strong>the</strong> establishment <strong>the</strong>re<br />

<strong>of</strong> <strong>the</strong> Kakamega National Reserve in 1986.<br />

Alth<strong>of</strong> (2005) fur<strong>the</strong>r mentions several plantations<br />

that were mainly established in <strong>the</strong> sou<strong>the</strong>rn


2 Study areas<br />

parts <strong>of</strong> <strong>the</strong> <strong>for</strong>est primarily consisting <strong>of</strong> Maesopsis<br />

eminii, Eucalyptus saligna, Pinus patula or Cupressus<br />

lusitanica.<br />

Vegetation structure<br />

Concerning vertical <strong>for</strong>est structure Alth<strong>of</strong> (2005)<br />

describes <strong>the</strong> less disturbed <strong>for</strong>est stages as having<br />

canopy heights <strong>of</strong> 20 m to 35 m in <strong>the</strong> upper tree<br />

layer with some emergents reaching up to 40 m. A<br />

more or less developed second layer exists with tree<br />

heights <strong>of</strong> 10 m to 12 m. Understorey vegetation,<br />

mainly comprising <strong>of</strong> Dracena fragrans, is dense in<br />

most parts. Younger <strong>for</strong>est stages <strong>of</strong> <strong>the</strong> Harungana<br />

madagascariensis-Desmodium adscendens alliance<br />

are in turn characterized by a single tree layer up<br />

to 12 m high with more or less dense herb cover.<br />

Understorey vegetation in <strong>the</strong> <strong>for</strong>m <strong>of</strong> larger shrubs<br />

is hardly present.<br />

Just like in Budongo Forest, horizontal variability<br />

<strong>of</strong> <strong>for</strong>est structure is mainly driven by human<br />

exploitation <strong>of</strong> <strong>the</strong> rain <strong>for</strong>est ecosystem. Valuable<br />

timber species once abundant in Kakamega, such as<br />

<strong>the</strong> mahogany Entandrophragma angolense, are now<br />

very rare and only occur in restricted areas, though<br />

seedlings and saplings were recorded in o<strong>the</strong>r areas<br />

by Alth<strong>of</strong> (2005). Thus it is not astonishing that a true<br />

primary <strong>for</strong>est no longer exists in Kakamega Forest<br />

(Alth<strong>of</strong> 2005). Today <strong>the</strong> <strong>for</strong>est is ra<strong>the</strong>r a mixture<br />

<strong>of</strong> different disturbance stages, such as near primary<br />

<strong>for</strong>est, secondary <strong>for</strong>est and clearings mixed with tea<br />

and timber plantations. The different <strong>for</strong>est stages are<br />

however not spatially grouped in compartments as in<br />

Bodongo Forest, but ra<strong>the</strong>r irregularly distributed.<br />

Based on Mitchell (2004), Kisere Forest, as well as <strong>the</strong><br />

Yala and Isecheno Nature Reserves can be identified<br />

as <strong>the</strong> most undisturbed parts <strong>of</strong> <strong>the</strong> Kakamega Forest<br />

area. Buyangu Nature Reserve in <strong>the</strong> nor<strong>the</strong>rn part<br />

<strong>of</strong> Kakamega Forest has in turn undergone extensive<br />

25<br />

exploitation. Natural succession can now be studied<br />

on areas that are now protected from fire and grazing,<br />

which <strong>for</strong>merly maintained <strong>the</strong>m as grassland.<br />

2.2.3<br />

Forest management<br />

According to Mitchell (2004) Kakamega Forest was<br />

gazetted as a Forest Reserve in 1933 after <strong>the</strong> <strong>for</strong>mer<br />

Forest Department took over <strong>the</strong> management in<br />

1931. Commercial timber extraction started during<br />

<strong>the</strong> 1930s, triggered by <strong>the</strong> need <strong>of</strong> <strong>the</strong> <strong>the</strong>n existing<br />

gold mines (Mitchell 2004). The number <strong>of</strong> logged<br />

trees continually rose until <strong>the</strong> 1980s, supported by<br />

numerous sawmills (Bleher et al. 2006, Mitchell<br />

2004), and had a major influence on <strong>for</strong>est structural<br />

properties and <strong>for</strong>est species composition. Clear<br />

felling <strong>of</strong> indigenous <strong>for</strong>est took place until <strong>the</strong> mid-<br />

1970s when it was partly replaced with fast growing<br />

s<strong>of</strong>twood or tea plantations, especially in <strong>the</strong> sou<strong>the</strong>rn<br />

<strong>for</strong>est parts. In <strong>the</strong> north, more selective logging took<br />

place. Olea capensis had been identified as <strong>the</strong> best<br />

timber tree, but o<strong>the</strong>r species were also cut, such as<br />

Zanthoxylum gillettii, Aningeria altissima, Funtumia<br />

africana, Prunus africana and Cordia africana,<br />

after 1948 also Celtis mildbraedii amongst o<strong>the</strong>rs<br />

(Mitchell 2004).<br />

Until <strong>the</strong> nature reserves <strong>of</strong> Yala, Isecheno and<br />

Kisere (cf. Figure 2-11) were established in 1967,<br />

local people were allowed to use <strong>the</strong> whole <strong>for</strong>est<br />

<strong>for</strong> <strong>the</strong>ir own purposes (collection <strong>of</strong> firewood, poles<br />

<strong>for</strong> construction material, cattle grazing). In 1986<br />

Kisere Forest and <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> Kakamega<br />

Forest, Buyangu, were <strong>of</strong>ficially handed over to<br />

<strong>the</strong> management <strong>of</strong> <strong>the</strong> KWS. Since <strong>the</strong> 1980s <strong>the</strong><br />

conversion <strong>of</strong> indigenous <strong>for</strong>est to plantation, <strong>the</strong><br />

cutting <strong>of</strong> indigenous trees, as well as cultivation<br />

have been banned from all parts <strong>of</strong> <strong>the</strong> <strong>for</strong>est. Cattle<br />

grazing and <strong>the</strong> collection <strong>of</strong> firewood and grass are<br />

still allowed under permit outside <strong>the</strong> nature reserves<br />

(Mitchell 2004). Though legislation is clear, KFS and


26<br />

KWS do not en<strong>for</strong>ce <strong>the</strong>se laws with equal strictness.<br />

Whereas <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> <strong>for</strong>est managed by<br />

KWS shows less anthropogenic disturbance today, <strong>the</strong><br />

sou<strong>the</strong>rn part under KFS direction still experiences<br />

high levels <strong>of</strong> illegal logging activity as well as<br />

frequent charcoal burning, firewood collection (cf.<br />

Figure 2-15) and cattle grazing (Bleher et al. 2006,<br />

Mitchell & Schaab, in print).<br />

Figure 2-15<br />

Illegal activities in Kakamega Forest. a) Firewood<br />

collection at Isecheno and b) charcoal burning<br />

close to Shandarema.


3 Theoretical background<br />

The following chapter provides <strong>the</strong> <strong>the</strong>oretical<br />

background <strong>for</strong> <strong>the</strong> <strong>MODIS</strong> LAI product validation.<br />

LAI estimation from both field and satellite data<br />

will be described with special emphasis on <strong>the</strong><br />

derivation <strong>for</strong> tropical rain <strong>for</strong>est canopies. Starting<br />

with <strong>the</strong> definition <strong>of</strong> LAI and an overview <strong>of</strong> LAI<br />

characteristics in tropical rain <strong>for</strong>ests, existing in<br />

situ methodologies, instruments, constraints and<br />

sampling strategies are to be reviewed. The state<br />

<strong>of</strong> research in satellite-<strong>based</strong> LAI derivation will<br />

be fur<strong>the</strong>r discussed with respect to physical and<br />

empirical approaches, as well as constraints related to<br />

sensor resolution. Focus will <strong>the</strong>n be put on spectral<br />

vegetation indices (SVIs) and texture measures as<br />

<strong>the</strong>y <strong>for</strong>m <strong>the</strong> basis <strong>for</strong> <strong>the</strong> upscaling <strong>of</strong> field data in<br />

this <strong>the</strong>sis. Last but not least validation approaches<br />

<strong>for</strong> satellite-<strong>based</strong> LAI products will be appraised.<br />

3.1<br />

Definition <strong>of</strong> LAI<br />

27<br />

LAI is a dimensionless variable that was first<br />

mentioned by Watson (1947) following earlier<br />

studies <strong>of</strong> foliage amount estimation. According to<br />

his definition, it refers to <strong>the</strong> total one-sided area <strong>of</strong><br />

photosyn<strong>the</strong>tic tissue per unit ground surface area.<br />

Whereas this definition is adequate <strong>for</strong> broad leaves,<br />

problems occur with needle leaves as here <strong>the</strong> onesided<br />

area is not clearly defined (Jonckheere et al.<br />

2004). There<strong>for</strong>e several authors (<strong>for</strong> instance, Ross<br />

1981, Smith et al. 1991) proposed a projected leaf<br />

area to take into account asymmetrical <strong>for</strong>ms <strong>of</strong><br />

certain foliage types. Yet if <strong>the</strong> projection angle is<br />

not clearly specified, <strong>the</strong> chosen projection does not<br />

necessarily result in <strong>the</strong> highest possible LAI values<br />

<strong>of</strong> <strong>the</strong> examined vegetation. Consequently Myneni<br />

et al. (1997) defined LAI as <strong>the</strong> maximum projected<br />

leaf area per unit ground surface area.<br />

Whereas <strong>the</strong> projected leaf area is more relevant<br />

to radiometric assessments <strong>of</strong> LAI, where <strong>the</strong><br />

parameter is estimated indirectly from solar radiation<br />

interception, Chen & Black 1992 in agreement with<br />

Lang (1991) suggested using half <strong>the</strong> total green leaf<br />

area per unit ground surface as definition <strong>of</strong> LAI.<br />

The basis <strong>for</strong> <strong>the</strong>ir <strong>the</strong>oretical reasoning was that <strong>the</strong><br />

hemisurface leaf area is a simple definition that also<br />

allows <strong>for</strong> fur<strong>the</strong>r calculation <strong>of</strong> projected leaf area<br />

with a few assumptions (Hyer & Goetz 2004). This<br />

adheres to <strong>the</strong> original concept <strong>of</strong> Watson (1947), as<br />

at least <strong>for</strong> broad-leaved trees <strong>the</strong> one-sided leaf area<br />

is equivalent to <strong>the</strong> hemisurface leaf area.<br />

This last definition <strong>for</strong> LAI is now widely used in<br />

literature (Casa & Jones 2005, Eriksson et al. 2006,<br />

Leblanc et al. 2005, Nilson & Kuusk 2005, Weiss<br />

et al. 2004, among o<strong>the</strong>rs). The latter also described<br />

optical methods <strong>of</strong> in situ LAI estimation as well as<br />

<strong>the</strong> <strong>MODIS</strong> LAI algorithm, which are <strong>based</strong> on this<br />

definition (Privette et al. 2002). Consequently it will<br />

be <strong>the</strong> one used throughout this study. However, it is


28<br />

important to note that different definitions result in<br />

significant discrepancies <strong>of</strong> calculated LAI values<br />

and <strong>for</strong> comparison purposes <strong>the</strong> respective definition<br />

should be checked carefully.<br />

3.2<br />

Characteristics <strong>of</strong> LAI<br />

in tropical rain <strong>for</strong>ests<br />

As mentioned in Chapter 2, <strong>the</strong> leaf area <strong>of</strong> tropical<br />

rain <strong>for</strong>est stands varies according to species<br />

composition, phenological stage (in <strong>the</strong> case <strong>of</strong><br />

semi-evergreen rain <strong>for</strong>ests with a certain percentage<br />

<strong>of</strong> deciduous species), disturbance and climatic<br />

conditions. Additionally local changes in LAI can be<br />

observed due to <strong>for</strong>est dynamics (Welles 1990).<br />

Compared to higher latitudes, field measurements <strong>of</strong><br />

LAI in tropical environments are rare. Doubtlessly <strong>the</strong><br />

reason <strong>for</strong> that is <strong>the</strong> higher complexity <strong>of</strong> vegetation<br />

structure and species diversity in rain <strong>for</strong>ests that<br />

limits <strong>the</strong> transferability <strong>of</strong> well-established in situ<br />

methods <strong>of</strong>, <strong>for</strong> instance, agricultural lands or boreal<br />

<strong>for</strong>ests. In <strong>the</strong> following a literature review will shed<br />

light on available in<strong>for</strong>mation on LAI <strong>of</strong> tropical rain<br />

<strong>for</strong>ests.<br />

One <strong>of</strong> <strong>the</strong> most extensive and global LAI data sets<br />

was ga<strong>the</strong>red by Scurlock et al. (2001) and is fur<strong>the</strong>r<br />

summarized by Asner et al. (2003). Consisting <strong>of</strong><br />

almost 1000 LAI measurements from nearly 400<br />

unique and globally distributed field sites, it is <strong>based</strong><br />

on published in situ data sampled between 1932 and<br />

2000. However, LAI was estimated with different<br />

methods and no in<strong>for</strong>mation on measurement<br />

precision is available. Un<strong>for</strong>tunately <strong>the</strong> database<br />

does not comprise a single test site <strong>for</strong> East Africa<br />

and only 6% <strong>of</strong> all valid observations refer to tropical<br />

evergreen broadleaf <strong>for</strong>est sites. For <strong>the</strong>se sites <strong>the</strong><br />

mean reported LAI is 4.78 with a standard deviation<br />

<strong>of</strong> 1.70 and minimum and maximum values <strong>of</strong> 1.48<br />

and 8.0 respectively. Interestingly, Asner et al. (2003)<br />

mention that an analysis <strong>of</strong> <strong>the</strong> collected LAI values<br />

over time revealed a decline in mean measured LAI<br />

values over <strong>the</strong> decades. This is most noticeable in<br />

<strong>the</strong> 1990s and is probably related to <strong>the</strong> methods<br />

predominantly used. Be<strong>for</strong>e 1990 mainly direct<br />

and semi-direct methods were applied, whereas <strong>for</strong><br />

optical approaches that were developed after 1990<br />

underestimations are frequently reported (cf. Chapter<br />

3.3.1).<br />

One <strong>of</strong> <strong>the</strong> first studies on LAI <strong>of</strong> tropical rain <strong>for</strong>ests<br />

is <strong>the</strong> one by Ogawa et al. (1961). It was per<strong>for</strong>med<br />

using destructive measurements and reports LAI<br />

values between 7 and 11 <strong>for</strong> evergreen tropical rain<br />

<strong>for</strong>ests in Thailand. Slightly lower LAI values are<br />

reported by Alexandre (1981), who reviewed several<br />

studies, mostly from tropical rain <strong>for</strong>ests in South<br />

America and Asia. He comes to <strong>the</strong> conclusion that<br />

a maximum annual LAI <strong>of</strong> 8.2 seems to be a reliable<br />

value <strong>for</strong> tropical rain <strong>for</strong>ests in general.<br />

By contrast, more recent studies show <strong>the</strong> abovementioned<br />

tendency <strong>for</strong> lower LAI values. Aragão<br />

et al. (2005) measured LAI in evergreen primary<br />

and secondary rain <strong>for</strong>ests in Eastern Amazonia<br />

with optical devices. Their very detailed statistical<br />

analysis reveals a spatial variability <strong>of</strong> LAI at<br />

different landscape units (individual measurements to<br />

plot level). For <strong>the</strong> plot level, which is probably <strong>the</strong><br />

most comparable to o<strong>the</strong>r studies, <strong>the</strong>y recorded LAI<br />

values between 3.06 (±0.31) <strong>for</strong> secondary <strong>for</strong>est<br />

and 5.85 (±1.19) <strong>for</strong> primary rain <strong>for</strong>est. Comparable<br />

values were retrieved by Wirth et al. (2001), who<br />

examined <strong>the</strong> vertical and horizontal variability <strong>of</strong><br />

canopy structures and seasonal variability <strong>of</strong> LAI<br />

in a semi-evergreen tropical rain <strong>for</strong>est in Panama.<br />

Although <strong>the</strong> mean decrease in LAI from wet to dry<br />

season was low with 5.5% from 5.41 (±1.27) to 5.11<br />

(±0.68), Wirth et al. (2001) report considerable smallscale<br />

heterogeneities in horizontal LAI distribution.<br />

The results <strong>of</strong> Wirth et al. (2001) concerning <strong>the</strong><br />

vertical variation <strong>of</strong> LAI are displayed in Figure 3-1.


3 Theoretical background<br />

Summing up to a mean total LAI <strong>of</strong> 6.04,<br />

measurements along five vertical pr<strong>of</strong>iles revealed<br />

that more than 50% <strong>of</strong> canopy leaf area is found<br />

in <strong>the</strong> uppermost 5 m due to <strong>the</strong> clustering <strong>of</strong> tree<br />

crowns in <strong>the</strong> upper canopy. Below 30 m <strong>the</strong> canopy<br />

is dominated by almost leafless trunks until a second<br />

maximum <strong>of</strong> leaf area arises at <strong>the</strong> height <strong>of</strong> trees<br />

in <strong>the</strong> second tree layer or understorey (Wirth et<br />

al. 2001). In ano<strong>the</strong>r study from Central America<br />

Kalácska et al. (2004) estimated in situ LAI values<br />

<strong>of</strong> 4.80 (±0.82) <strong>for</strong> intermediate and 6.90 (±1.96) <strong>for</strong><br />

late <strong>for</strong>est stages in a tropical moist <strong>for</strong>est in Costa<br />

Rica.<br />

The mean LAI values given by Wirth et al. (2001)<br />

correspond closely to one <strong>of</strong> <strong>the</strong> very few examples<br />

<strong>for</strong> LAI measurements in African rain <strong>for</strong>ests. De<br />

Wasseige et al. (2003) investigated seasonal changes<br />

in leaf area <strong>of</strong> a nearly undisturbed semi-evergreen<br />

tropical rain <strong>for</strong>est in <strong>the</strong> CAR, which is comparable<br />

to Budongo and Kakamega Forests in terms <strong>of</strong><br />

vegetation type, average yearly rainfall and length <strong>of</strong><br />

dry season. At <strong>the</strong> end <strong>of</strong> <strong>the</strong> rainy season an average<br />

LAI <strong>of</strong> 5.47 was measured at plot level. This value<br />

changed to 5.13 on <strong>the</strong> same site at <strong>the</strong> end <strong>of</strong> <strong>the</strong><br />

dry season. Slightly higher values were derived by<br />

Laumonier et al. (1994) <strong>for</strong> an evergreen tropical rain<br />

<strong>for</strong>est in Cameroon. A mean value <strong>of</strong> 5.6 is reported<br />

here, with minimum and maximum values <strong>of</strong> 3 and<br />

7 at sample level. So far, no published LAI values<br />

could be found <strong>for</strong> East African rain <strong>for</strong>ests.<br />

Table 3-1 summarizes <strong>the</strong> above-mentioned<br />

publications. The examples show that <strong>the</strong> variability<br />

<strong>of</strong> in situ LAI <strong>for</strong> tropical rain <strong>for</strong>ests can be relatively<br />

high. Whereas a minimum value <strong>of</strong> 3.06 (mean LAI<br />

at plot level) is reported by Aragão et al. (2005) <strong>for</strong> a<br />

secondary evergreen rain <strong>for</strong>est in Panama, Ogawa et<br />

al. (1961) derived maximum LAI values <strong>of</strong> 11 with<br />

destructive measurements in an evergreen tropical<br />

rain<strong>for</strong>est in Thailand. These values do not only reflect<br />

differences in <strong>for</strong>est <strong>for</strong>mation and structure due to<br />

Figure 3-1<br />

29<br />

natural environmental parameters or anthropogenic<br />

influences (mainly represented by different <strong>for</strong>est<br />

stages); <strong>the</strong> variation <strong>of</strong> <strong>the</strong> results might also be partly<br />

attributed to <strong>the</strong> instrumentation and methodologies<br />

used (e.g. <strong>the</strong> destructive measurements employed<br />

by Ogawa et al. [1961] versus studies using LAI-<br />

2000 PCA). For a better understanding <strong>of</strong> <strong>the</strong>se<br />

discrepancies, <strong>the</strong> most important methods <strong>for</strong> in situ<br />

leaf area index determination will be described in <strong>the</strong><br />

following.<br />

3.3<br />

Vertical LAI variation in a semi-evergreen tropical<br />

rain <strong>for</strong>est in Panama (Wirth et al. 2001).<br />

<strong>Ground</strong>-<strong>based</strong> measurements<br />

<strong>of</strong> LAI<br />

Per<strong>for</strong>ming ground-<strong>based</strong> measurements <strong>of</strong> LAI is<br />

quite challenging in tropical rain <strong>for</strong>ests. Besides<br />

problems in transferability or applicability <strong>of</strong> methods<br />

to tropical rain <strong>for</strong>ests, in situ sampling is complicated<br />

by species diversity and stand complexity. Fur<strong>the</strong>r, in<br />

situ data should be representative <strong>for</strong> <strong>the</strong> study sites,<br />

which can be very difficult in areas where <strong>the</strong> spatial<br />

distribution <strong>of</strong> sampling sites is restricted by <strong>the</strong><br />

accessibility <strong>of</strong> remote <strong>for</strong>est regions. There<strong>for</strong>e an<br />

in-depth understanding <strong>of</strong> LAI, its definition and its<br />

characteristics in tropical rain <strong>for</strong>est stands, as well as<br />

<strong>the</strong> knowledge about state-<strong>of</strong>-<strong>the</strong>-art methodologies<br />

<strong>of</strong> in situ LAI estimation is crucial. After an overview<br />

on <strong>the</strong> available field methods, <strong>the</strong> instruments used<br />

in this study will be described and limitations and


30<br />

5.85 (±1.19) <strong>for</strong> primary rain<br />

<strong>for</strong>est (plot level)<br />

3.06 (±0.31) <strong>for</strong> secondary rain<br />

<strong>for</strong>est (plot level)<br />

Alto Tapajós, Eastern<br />

Amazonia, Brazil<br />

(2°24’S/4°01’1S;<br />

55°30’W/54°29’W)<br />

Table 3-1<br />

Evergreen rain <strong>for</strong>est Primary and secondary<br />

rain <strong>for</strong>est<br />

LAI-2000 PCA One-sided leaf area<br />

per unit ground area<br />

Aragão et al. (2005)<br />

<strong>for</strong>est (plot level)<br />

6.90 (±1.96) <strong>for</strong> primary rain<br />

<strong>for</strong>est (plot level)<br />

3.20 (±0.82) <strong>for</strong> secondary rain<br />

Los Innocentes, Costa<br />

Rica<br />

(11°1’N, 85°30’W)<br />

Evergreen rain <strong>for</strong>est Primary and secondary<br />

rain <strong>for</strong>est<br />

17°00’E/17°30’E)<br />

(wrong definition <strong>for</strong><br />

LAI-2000 PCA)<br />

LAI-2000 PCA One-sided leaf area<br />

per unit ground area<br />

Kalácska et al. (2004)<br />

(plot level, seasonal changes)<br />

5.13 to 5.47<br />

Ngotto Forest, Central<br />

African Republic<br />

(3°40’N/4°10’N;<br />

Semi-evergreen<br />

rain <strong>for</strong>est<br />

specified)<br />

<strong>based</strong> on radiative<br />

measurements<br />

Primary <strong>for</strong>est LAI-2000 PCA Total leaf area per unit<br />

ground area<br />

De Wasseige et al.<br />

(2003)<br />

4.78 (±1.70) <strong>for</strong> evergreen rain<br />

<strong>for</strong>est<br />

3.92 (±2.53) <strong>for</strong> semi-evergreen<br />

rain <strong>for</strong>est<br />

5.11 to 5.41<br />

(seasonal changes)<br />

Barro Colorado Island,<br />

Panama<br />

(9°10’N, ° W not<br />

Semi-evergreen rain<br />

<strong>for</strong>est<br />

Semi-evergreen<br />

rain <strong>for</strong>est<br />

Primary <strong>for</strong>est Inversion <strong>of</strong> light<br />

interception model<br />

Not mentioned Wirth et al. (2001)<br />

Worldwide (no East<br />

African rain <strong>for</strong>est<br />

included)<br />

Evergreen rain <strong>for</strong>est<br />

n/a Various methods<br />

applied<br />

Not mentioned Asner et al. (2003)/<br />

Scurlock et al. (2001)<br />

(5°15’N/ 52°55’W)<br />

4.51 to 5.44 Paracou, French<br />

Guiana<br />

Evergreen rain <strong>for</strong>est Primary <strong>for</strong>est LAI-2000 PCA One-sided leaf area<br />

per unit ground area<br />

Review <strong>of</strong> LAI values reported <strong>for</strong> tropical rain <strong>for</strong>ests in <strong>the</strong> literature.<br />

Ferment et al. (2001)<br />

5.6 Campo Nature<br />

Reserve, South<br />

Cameroon<br />

Evergreen rain <strong>for</strong>est Not mentioned, primary<br />

<strong>for</strong>est assumed<br />

LAI-2000 PCA Not mentioned Laumonier et al. (1994)<br />

Maximum value <strong>of</strong> 8.2 Worldwide Evergreen rain <strong>for</strong>est n/a <strong>Leaf</strong> litter fall method Total leaf surface area<br />

per unit ground area<br />

Alexandre (1981)<br />

7 to 11 Thailand Evergreen rain <strong>for</strong>est Not mentioned, primary<br />

<strong>for</strong>est assumed<br />

Destructive<br />

measurements<br />

One-sided leaf area<br />

per unit ground area<br />

Ogawa et al. (1961)<br />

In situ LAI Study area Forest <strong>for</strong>mation Forest stage Methodology/<br />

instruments<br />

LAI definition Source


3 Theoretical background<br />

caveats <strong>of</strong> in situ LAI assessment will be reviewed.<br />

The subchapter closes with <strong>the</strong> description <strong>of</strong><br />

adequate ground sampling strategies.<br />

3.3.1<br />

Methods<br />

Methods <strong>for</strong> LAI determination have been developed<br />

since <strong>the</strong> early 1930s. Most <strong>of</strong> <strong>the</strong> literature since<br />

published in this context is related to methodology<br />

and constraints, validation <strong>of</strong> methods and<br />

instruments, <strong>the</strong> development <strong>of</strong> new instrumentation<br />

and <strong>the</strong> comparison <strong>of</strong> results with reference to leaf<br />

area measurements both in stands and in laboratory<br />

conditions.<br />

Ross (1981) gives a good overview <strong>of</strong> <strong>the</strong> historical<br />

methodological development and <strong>the</strong>ir <strong>the</strong>oretical<br />

background until 1980. He separates <strong>the</strong> methods<br />

into direct and indirect approaches, as well as several<br />

o<strong>the</strong>rs that have lost importance since <strong>the</strong> 1980s and<br />

will not be mentioned here. More recently Bréda<br />

(2003) published a useful review <strong>of</strong> methods, giving<br />

fur<strong>the</strong>r detail to newer optical instruments and<br />

current controversies such as error analysis, crosscalibration,<br />

sampling strategy, spatial validation or<br />

scaling problems. Jonckheere et al. (2004) discuss<br />

different methods <strong>for</strong> in situ LAI determination with<br />

special regard to indirect measurements. The most<br />

important methods <strong>for</strong> tropical rain <strong>for</strong>est stands will<br />

be described in <strong>the</strong> following.<br />

Direct and semi-direct measurements<br />

Generally, direct methods measure <strong>the</strong> leaves<br />

<strong>the</strong>mselves. In consequence <strong>the</strong>y are <strong>the</strong> most<br />

accurate method <strong>of</strong> retrieving LAI <strong>for</strong> a certain<br />

sample, yet <strong>the</strong> direct approach is usually very time<br />

consuming and labour intensive and is <strong>the</strong>re<strong>for</strong>e only<br />

feasible <strong>for</strong> small areas. According to Ross (1981)<br />

<strong>the</strong> earliest method <strong>of</strong> leaf area determination was<br />

31<br />

drawing <strong>the</strong> contours <strong>of</strong> a leaf on a piece <strong>of</strong> paper<br />

and relating it to a reference grid. Though this is very<br />

accurate (Rozhnyatovsky [1954] employing a very<br />

similar method found an error in <strong>the</strong> order <strong>of</strong> 1%),<br />

it is not very efficient and practically inapplicable to<br />

small compound and crimpy leaves. Today optically<strong>based</strong><br />

automatic area measurement systems exist,<br />

such as <strong>the</strong> LI-3000 or <strong>the</strong> LI-3100 <strong>Area</strong> Meters (LI-<br />

COR, Lincoln, Nebraska). Single leaves are scanned<br />

(attached or unattached) and <strong>the</strong> respective leaf area<br />

is automatically calculated.<br />

However, when scaling direct LAI measurements<br />

up to a whole stand, errors can be introduced that<br />

might lead to a large over- or underestimation <strong>of</strong> <strong>the</strong><br />

parameter. This also puts <strong>the</strong> above-mentioned early<br />

work <strong>of</strong> Ogawa et al. (1961) into perspective (cf.<br />

Table 3-1).<br />

In general, direct methods are not compatible with <strong>the</strong><br />

long-term monitoring <strong>of</strong> spatial and temporal patterns<br />

<strong>of</strong> LAI that can be done with optical devices. Yet <strong>the</strong>y<br />

are still <strong>of</strong> relevance when it comes to <strong>the</strong> validation<br />

<strong>of</strong> indirect measurement methods (Jonckheere et al.<br />

2004).<br />

Semi-direct methods are <strong>based</strong> on direct<br />

measurements <strong>of</strong> o<strong>the</strong>r vegetation parameters that<br />

are closely related to LAI, e.g. sapwood area. Here,<br />

measurement errors mainly result from <strong>the</strong> empirical<br />

relationship established between LAI and <strong>the</strong> variable<br />

taken into account, as well as <strong>the</strong> applied upscaling<br />

approach. A common example is <strong>the</strong> collection <strong>of</strong><br />

leaf litter fall in temperate deciduous <strong>for</strong>ests during<br />

<strong>the</strong> autumn exfoliation period (e.g. Dufrêne & Bréda<br />

1995, Eriksson et al. 2005). A litter trap with a known<br />

surface area is emptied on a regular basis to avoid<br />

weight loss through decomposition. The collected<br />

leaves are usually weighed and converted to leaf<br />

area by determining <strong>the</strong> specific leaf area (leaf area/<br />

leaf mass) <strong>for</strong> sub-samples (Eriksson et al. 2005).<br />

The method is believed to give reliable results and is


32<br />

<strong>the</strong>re<strong>for</strong>e widely used as reference <strong>for</strong> o<strong>the</strong>r methods<br />

<strong>of</strong> LAI determination and <strong>the</strong>ir validation. However,<br />

it should be noted that <strong>the</strong> LAI assessment using litter<br />

traps is also <strong>based</strong> on an estimation technique, as it<br />

depends on <strong>the</strong> specific leaf area (SLA) relationship<br />

which is species dependent and also varies according<br />

to tree vertical position. Eriksson et al. (2005) e.g.<br />

found that <strong>the</strong> top leaves <strong>of</strong> beech and oak stands<br />

had a higher SLA than lower leaves. If an SLA<br />

relationship is only <strong>based</strong> on low leaves, <strong>the</strong> resulting<br />

LAI from litter trap measurements is likely to be<br />

overestimated.<br />

In tropical rain <strong>for</strong>ests <strong>the</strong> use <strong>of</strong> leaf litter traps is<br />

ra<strong>the</strong>r difficult. The correlation <strong>of</strong> seasonal litter fall<br />

shows a correlation to <strong>the</strong> average life span <strong>of</strong> leaves<br />

or climatic variability ra<strong>the</strong>r than to real leaf area.<br />

Scurlock et al. (2001) e.g. mention leaf turnover times<br />

<strong>for</strong> tropical rain <strong>for</strong>ests <strong>of</strong> up to 6 years. However,<br />

Kalácska et al. (2005) used leaf litter data from a<br />

tropical dry <strong>for</strong>est to determine LAI. As <strong>the</strong>y also<br />

mention that especially in late succession stages not<br />

all trees on <strong>the</strong> sample plots shed <strong>the</strong>ir entire leaves,<br />

<strong>the</strong> LAI determination with litter traps is questionable<br />

in this case. Moreover, as SLA is species dependent,<br />

representative relationships with LAI are not easy to<br />

establish in a species rich environment.<br />

LAI can also be retrieved by semi-direct allometric<br />

methods, which describe <strong>the</strong> relation between leaf area<br />

and an independent variable (Gower et al. 1999). These<br />

independent variables are commonly characterized by<br />

easy-to-measure physical dimensions, as e.g. stem<br />

diameter at breast height (DBH). Species- or standspecific<br />

relationships are <strong>the</strong>n applied on <strong>the</strong> basis <strong>of</strong><br />

detailed destructive measurements <strong>of</strong> sub-samples.<br />

Good correlation coefficients have been found<br />

between leaf area and sapwood area (Kaufmann &<br />

Troendle 1981, O’Hara & Valappil 1995, Rogers &<br />

Hinckley 1979, Vertessy et al. 1995), stem basal area<br />

(Bartelink 1997) and DBH (e.g. Le Dantec et al. 2000)<br />

<strong>for</strong> different species or vegetation types.<br />

The disadvantage is that allometric relationships<br />

are very stand-, species- and time-specific. Nei<strong>the</strong>r<br />

phenological change over <strong>the</strong> year nor LAI recovery<br />

after a natural or human-induced canopy opening can<br />

be described (Bréda 2003). In addition, considerable<br />

fieldwork is required to establish a stable relationship<br />

<strong>for</strong> a whole tree stand, especially when it consists <strong>of</strong><br />

a high number <strong>of</strong> different species. There<strong>for</strong>e it is<br />

hardly – if at all – applicable to tropical rain <strong>for</strong>ests.<br />

Indirect methods<br />

In recent years indirect measurements, in particular,<br />

have grown in importance as <strong>the</strong>y allow rapid LAI<br />

assessment <strong>of</strong> large areas. A good overview <strong>of</strong> state<strong>of</strong>-<strong>the</strong>-art<br />

methodologies is given by Weiss et al.<br />

(2004) and Jonckheere et al. (2004). Indirect methods<br />

usually refer to optical measurements <strong>of</strong> radiation<br />

with photosensitive instruments. The trans<strong>for</strong>mation<br />

into LAI is <strong>the</strong>n per<strong>for</strong>med ei<strong>the</strong>r by a calibrated<br />

relationship or an inversion model that takes statistic<br />

and probabilistic models <strong>of</strong> canopy geometry into<br />

account (see e.g. Bréda 2003, Eriksson et al. 2005,<br />

Frazer et al. 2001, Jonckheere et al. 2004, Leblanc et<br />

al. 2005, Ross 1981, Welles 1990, Welles & Norman<br />

1991, Zhang et al. 2005, among o<strong>the</strong>rs). While <strong>the</strong>se<br />

methods have <strong>the</strong> advantage <strong>of</strong> convenience and<br />

low labour costs, many difficulties arise from <strong>the</strong><br />

complexity <strong>of</strong> radiation transfer in <strong>for</strong>est canopies.<br />

Major problems include unknown foliage angle<br />

distribution, improper model assumptions and<br />

<strong>the</strong> contribution <strong>of</strong> woody material to radiation<br />

interception. The <strong>the</strong>oretical background <strong>of</strong> indirect<br />

optical measurements will be described briefly in <strong>the</strong><br />

following.<br />

As it penetrates <strong>the</strong> canopy, incident solar radiation<br />

is intercepted by plant organs. Consequently light<br />

attenuation corresponds to <strong>the</strong> vertical depth <strong>of</strong> <strong>the</strong><br />

canopy. Though this process can be quite variable<br />

in space and time, mean horizontal flux densities


l that 3, that ERIKSSON takes statistic et al. and 2005, probabilistic FRAZER models et models al. 2001, <strong>of</strong> <strong>of</strong> canopy JONCKHEERE geometry et into<br />

al. 2004,<br />

If we consider G ( , )<br />

constant along <strong>the</strong> path and LAI bein<br />

If we consider G ( , )<br />

constant along <strong>the</strong> path and LAI being de<br />

981, RIKSSON IKSSON WELLES et et al. al. 1990, 2005, 3 Theoretical WELLES FRAZER background & et et NORMAN al. al. 2001, 1991, JONCKHEERE ZHANG et et et al. al. al. 2005, 2004,<br />

among Z<br />

Z<br />

33<br />

WELLES have ELLES <strong>the</strong> 1990, advantage WELLES <strong>of</strong> & convenience & NORMAN THEORETICAL 1991, and 1991, low ZHANG labour et BACKGROUND<br />

et costs, al. al. 2005, many among<br />

difficulties LAI u dz ,<br />

LAI <br />

f <strong>the</strong> <strong>the</strong> radiation advantage transfer <strong>of</strong> <strong>of</strong><br />

7 u dz ,<br />

0 THEORETICAL BACKGROUND<br />

convenience in <strong>for</strong>est canopies. and low labour Major labour costs, problems costs, many include difficulties<br />

unknown 0<br />

ndirect methods usually refer to optical measurements <strong>of</strong> radiation<br />

iation tion proper transfer model in in assumptions <strong>for</strong>est canopies. and Major <strong>the</strong> Major contribution problems include <strong>of</strong> include woody unknown<br />

material to<br />

z<br />

rans<strong>for</strong>mation into (2004) usually LAI and decrease is JONCKHEERE <strong>the</strong>n per<strong>for</strong>med resulting et in al. ei<strong>the</strong>r (2004). a downward by a Indirect calibrated pattern methods usually where where refer z is z is <strong>the</strong> to <strong>the</strong> optical total total canopy measurements canopy height <strong>of</strong> and radiation ε z , , it follows aft<br />

heoretical r er model assumptions background assumptions <strong>of</strong> and indirect <strong>the</strong> <strong>the</strong> contribution optical measurements <strong>of</strong> <strong>of</strong> woody material will material be described to to where z is <strong>the</strong> total canopy height and cos , it follows after L<br />

takes statistic and with <strong>of</strong> light probabilistic photosensitive reduction models (Saeki instruments. <strong>of</strong> 1975). canopy The The geometry trans<strong>for</strong>mation relationship into into it follows LAI is after <strong>the</strong>n Lang per<strong>for</strong>med et al (1985) ei<strong>the</strong>r that by <strong>the</strong> a calibrated gap cosfraction<br />

<br />

tical cal background <strong>of</strong> <strong>of</strong> indirect optical measurements will be be described<br />

ON et al. 2005,<br />

7<br />

described by<br />

relationship between FRAZER light et or al. reduction an 2001, inversion JONCKHEERE and model canopy that et geometry takes al. 2004, statistic can and described is described probabilistic by by models THEORETICAL <strong>of</strong> canopy BACKGROUND<br />

geometry into<br />

LES incident 1990, solar WELLES radiation account be & characterized NORMAN (see is intercepted e.g. 1991, by BRÉDA statistical ZHANG by 2003, plant models et organs. ERIKSSON al. 2005, that Consequently among describe et al. 2005, light FRAZER et al. 2001, LAI<br />

GLAI<br />

, JONCKHEERE et al. 2004,<br />

(2004) and JONCKHEERE et al. (2004). Indirect methods usually refer to optical cos measurements nt e ent vertical solar radiation depth radiation <strong>the</strong> <strong>of</strong> is <strong>the</strong> is probability intercepted canopy. Though <strong>of</strong> by by radiation plant this organs. process interception Consequently can be (or quite non- light variable light P ( ,<br />

)<br />

<br />

<strong>of</strong> radiation<br />

advantage <strong>of</strong> convenience LEBLANC et and al. low 2005, labour ROSS costs, 1981, many WELLES difficulties 1990, WELLES in & NORMAN o Ge<br />

, <br />

1991, cos<br />

.<br />

P<br />

ZHANG<br />

et al. 2005, among<br />

o ( ,<br />

)<br />

e .<br />

(3.3)<br />

with photosensitive instruments. The trans<strong>for</strong>mation into LAI is <strong>the</strong>n per<strong>for</strong>med ei<strong>the</strong>r by a calibrated<br />

ntal tical cal transfer depth flux depth densities <strong>of</strong> in <strong>of</strong> <strong>the</strong> <strong>for</strong>est <strong>the</strong> o<strong>the</strong>rs). interception) canopy. usually canopies. While Though decrease Though within Major <strong>the</strong>se this resulting <strong>the</strong> methods problems process canopy. in can have a can Usually include downward be be <strong>the</strong> quite advantage <strong>the</strong>se unknown variable pattern models <strong>of</strong> in <strong>of</strong> in convenience light and low labour costs, many difficulties<br />

relationship or an inversion model that takes statistic and probabilistic models <strong>of</strong> canopy geometry into<br />

lux x odel e densities relationship densities assumptions usually arise are between decrease and <strong>based</strong> decrease from <strong>the</strong> light <strong>the</strong> on resulting contribution resulting complexity simplified reduction in in a a downward <strong>of</strong> assumptions and <strong>of</strong> downward woody radiation canopy pattern material pattern transfer concerning geometry <strong>of</strong> <strong>of</strong> to light in light can <strong>for</strong>est Lang be canopies. (1986) showed Major that problems <strong>the</strong> gap include fraction unknown is equivalent<br />

account (see e.g. BRÉDA 2003, ERIKSSON et al. 2005, FRAZER et al. 2001, JONCKHEERE et al. 2004,<br />

lationship background tionship odels that between describe <strong>of</strong> foliage <strong>the</strong> indirect light radiative light <strong>the</strong> angle optical reduction probability reduction and distribution, measurements spatial and <strong>of</strong> canopy characteristics radiation improper canopy will geometry be interception model described <strong>of</strong> canopy assumptions can (or be be non- to and radiation <strong>the</strong> contribution transmittance <strong>of</strong> τ woody through material <strong>the</strong> canopy, to<br />

LEBLANC et al. 2005, ROSS 1981, WELLES 1990, WELLES & NORMAN 1991, ZHANG et al. 2005, among<br />

ls y. that Usually describe <strong>the</strong>se radiation structural <strong>the</strong> models <strong>the</strong> probability are interception. properties, <strong>based</strong> <strong>of</strong> on <strong>of</strong> radiation simplified as radiation The e.g. <strong>the</strong>oretical leaf interception assumptions interception area density, background (or concerning (or leaf non-<br />

<strong>of</strong> indirect <strong>the</strong> if <strong>the</strong> latter optical is measured measurements at wavelengths will be described where <strong>the</strong><br />

o<strong>the</strong>rs). While <strong>the</strong>se methods have <strong>the</strong> advantage <strong>of</strong> convenience and low labour costs, many difficulties<br />

istics allyually <strong>the</strong>se <strong>of</strong> <strong>the</strong>se canopy models briefly angle structural are are <strong>based</strong> and in <strong>based</strong> <strong>the</strong> properties, foliage on following. on simplified distribution as e.g. assumptions leaf among area density, concerning o<strong>the</strong>rs. leaf The angle <strong>the</strong> <strong>the</strong> and assumption <strong>of</strong> opaque leaves (i.e. non-reflecting and<br />

olar radiation is arise intercepted from <strong>the</strong> by complexity plant organs. <strong>of</strong> Consequently radiation transfer light in <strong>for</strong>est canopies. Major problems include unknown<br />

<strong>the</strong>rs. f <strong>of</strong> canopy The structural key key variable properties, variable properties, in in those as those as e.g. models leaf area is is gap density, gap density, fraction leaf angle P angle<br />

o ( ,<br />

and )<br />

and ,<br />

which<br />

depth <strong>of</strong> <strong>the</strong> canopy. As it Though penetrates this <strong>the</strong> process canopy, can incident be quite solar variable radiation in is intercepted non-transmitting) by plant is valid.<br />

foliage angle distribution, improper model assumptions and <strong>the</strong> contribution organs. <strong>of</strong> woody Consequently material light to<br />

ky . The vs.<br />

key<br />

foliage key variable seen which in<br />

at in point describes those 0<br />

models<br />

in models <strong>the</strong> zenith proportion is is gap<br />

gap and<br />

fraction <strong>of</strong> fraction azimuth sky vs. P oP<br />

( <strong>of</strong>oliage<br />

( <br />

<br />

, ,<br />

<br />

directions<br />

) ,<br />

) , which seen which<br />

ensities usually radiation attenuation decrease resulting interception. corresponds in a downward The to <strong>the</strong> <strong>the</strong>oretical vertical pattern depth background <strong>of</strong> light <strong>of</strong> <strong>the</strong> canopy. <strong>of</strong><br />

from<br />

indirect Though optical this measurements process can be will quite be variable described in<br />

. foliage<br />

ERE foliage et<br />

seen<br />

al. seen 2004).<br />

at at point at point 0 0 in in zenith and azimuth directions from<br />

ship between space light and reduction time, mean and horizontal canopy geometry flux densities can be usually decrease A good resulting overview in <strong>of</strong> a downward o<strong>the</strong>r <strong>the</strong>oretical pattern <strong>of</strong> models light <strong>of</strong><br />

briefly in <strong>the</strong> following.<br />

t et al. al. 2004).<br />

THEORETICAL beneath <strong>the</strong> canopy (Jonckheere et al. 2004).<br />

different at describe statistical <strong>the</strong> reduction<br />

BACKGROUND<br />

probability (SAEKI<br />

models exist <strong>of</strong> radiation 1975).<br />

(each characterizing interception The relationship<br />

a (or different non- between light canopy reduction geometry and and canopy <strong>the</strong>ir relation geometry to gap can fraction be<br />

As it penetrates <strong>the</strong> canopy, incident solar radiation statistical is intercepted<br />

nt ent ts), <strong>the</strong>se statistical one statistical models <strong>of</strong> <strong>the</strong> models are models characterized<br />

best <strong>based</strong> known exist on (each and simplified (each by<br />

most characterizing statistical characterizing frequently assumptions models<br />

applied a a different concerning different that describe<br />

in <strong>the</strong> statistical<br />

estimation <strong>the</strong> <strong>the</strong> is probability given by by<br />

<strong>of</strong> Nilson plant<br />

radiation (1971), organs.<br />

interception who, Consequently in addition light<br />

(or non- to <strong>the</strong><br />

er to optical measurements attenuation <strong>of</strong> <strong>of</strong><br />

Though a large corresponds radiation<br />

number <strong>of</strong> different statistical models<br />

e ne e nopy e.g. <strong>of</strong> <strong>of</strong> <strong>the</strong> MONSI <strong>the</strong> structural best known & known interception)<br />

SAEKI properties, and 1953, most as within<br />

THEORETICAL to <strong>the</strong> vertical<br />

MUSSCHE e.g. frequently leaf <strong>the</strong> area canopy.<br />

et applied al. applied density, Usually<br />

BACKGROUND depth <strong>of</strong> <strong>the</strong> canopy.<br />

2001, in in NILSON leaf <strong>the</strong> <strong>the</strong> angle <strong>the</strong>se estimation models<br />

1999, and WEISS <strong>of</strong> <strong>of</strong> are Poisson Though<br />

<strong>based</strong> on model, this process<br />

simplified describes can<br />

assumptions binomial be quite variable<br />

concerning models in<br />

<strong>the</strong> (<strong>the</strong><br />

<strong>the</strong>n per<strong>for</strong>med space ei<strong>the</strong>r et<br />

exist and by<br />

(each time, a calibrated<br />

characterizing mean horizontal a different statistical<br />

canopy e MONSI ONSI key variable & is & SAEKI divided SAEKI radiative in 1953, into 1953, those N MUSSCHE and<br />

statistically models spatial et is et al. characteristics THEORETICAL flux densities<br />

independent gap al. 2001, fraction NILSON layers, P <strong>of</strong><br />

o ( canopy<br />

BACKGROUND usually decrease<br />

1999, , 1999, )<br />

in , which WEISS structural WEISS leaves et et properties, stand is resulting divided as e.g. into in a<br />

leaf area a downward finite density, number pattern<br />

leaf <strong>of</strong> angle statistically <strong>of</strong> light<br />

and<br />

irect ilistic methods models <strong>of</strong> usually reduction canopy refer geometry to are<br />

distribution (SAEKI optical into<br />

<strong>of</strong> canopy 1975). measurements<br />

elements), The relationship <strong>of</strong> radiation<br />

one <strong>of</strong> <strong>the</strong> between best light<br />

iage endently y is is divided seen divided at <strong>of</strong> into point each into foliage N o<strong>the</strong>r 0 statistically in zenith distribution<br />

(NILSON independent independent and among<br />

1999). azimuth If layers, o<strong>the</strong>rs.<br />

N layers, <br />

directions<br />

in , in The<br />

<strong>the</strong> which key<br />

number leaves from leaves variable<br />

<strong>of</strong> are overlaps are in independent reduction<br />

those models layers, and<br />

is gap a canopy positive geometry<br />

fraction binomial can<br />

P o ( ,<br />

)<br />

, model be<br />

which <strong>for</strong><br />

et 04). s<strong>for</strong>mation al. Indirect 2001, into JONCKHEERE methods characterized LAI is usually <strong>the</strong>n et refer per<strong>for</strong>med al. 2004, to optical ei<strong>the</strong>r measurements by a calibrated <strong>of</strong> radiation<br />

known THEORETICAL and most by BACKGROUND<br />

frequently statistical applied models in that <strong>the</strong> estimation describe <strong>the</strong> regular probability dispersion <strong>of</strong> radiation <strong>of</strong> foliage, interception a negative (or binomial non-<br />

. son The MAN kes tly ntly 2004). <strong>of</strong> statistic distribution. <strong>of</strong> trans<strong>for</strong>mation 1991, each o<strong>the</strong>r ZHANG and o<strong>the</strong>r describes The (NILSON gap <strong>the</strong> fraction 1999). proportion If is If N <strong>the</strong>n N <br />

<strong>of</strong> <strong>the</strong><br />

,<br />

sky , <strong>the</strong> probability <strong>the</strong> vs. number foliage <strong>of</strong> <strong>for</strong> <strong>of</strong> seen overlaps zero overlaps at overlaps point 0 in zenith and azimuth directions from<br />

interception) probabilistic et into al. LAI 2005,<br />

<strong>of</strong> LAI is <strong>the</strong> within models is among <strong>the</strong>n<br />

Poisson <strong>the</strong> <strong>of</strong> per<strong>for</strong>med<br />

model canopy. canopy<br />

(see Usually geometry ei<strong>the</strong>r<br />

e.g. Monsi <strong>the</strong>se by into a calibrated<br />

& models Saeki are model <strong>based</strong> on <strong>for</strong> simplified clumped assumptions dispersion concerning <strong>of</strong> foliage) <strong>the</strong> and<br />

ually stribution. and istribution. l al. that et (1985) low refer al. takes 2005, labour to be The statistic optical described FRAZER gap beneath gap fraction as <strong>the</strong> canopy is is <strong>the</strong>n (JONCKHEERE <strong>the</strong> <strong>the</strong> probability et <strong>for</strong> <strong>for</strong> al. zero 2004). zero overlaps<br />

tatistical models radiative costs, measurements and many et<br />

1953, exist Mussche (each and probabilistic al. difficulties spatial 2001, <strong>of</strong><br />

characterizing et al. characteristics JONCKHEERE radiation models <strong>of</strong> canopy<br />

2001, Nilson a different 1999, <strong>of</strong> et canopy al. geometry 2004,<br />

statistical Weiss structural into<br />

et al. properties, Markov models as e.g. (layers leaf area are density, dependent, leaf i.e. angle probability and<br />

S es. 85) LAI 985) RIKSSON 1990, Major be be is described WELLES <strong>the</strong>n described et problems per<strong>for</strong>med al. as as<br />

f <strong>the</strong> best known foliage Though & 2005, NORMAN include FRAZER<br />

2004). and most It distribution a ei<strong>the</strong>r<br />

assumes large unknown 1991, by et<br />

frequently number a al. ZHANG<br />

that among calibrated 2001,<br />

applied <strong>the</strong> <strong>of</strong> canopy o<strong>the</strong>rs. et JONCKHEERE different al. 2005,<br />

in <strong>the</strong> is The estimation divided statistical among key et<br />

into variable al.<br />

<strong>of</strong> models 2004,<br />

N in <strong>of</strong> exist those transmission (each models characterizing differs is gap according fraction a different P to o ( contact<br />

, )<br />

statistical , which or non-<br />

vantage e probabilistic WELLES contribution <strong>of</strong> 1990, convenience models<br />

SI & SAEKI 1953, describes distribution WELLES <strong>of</strong> woody <strong>of</strong> and canopy & material<br />

statistically MUSSCHE <strong>the</strong><br />

NORMAN <strong>of</strong> low<br />

independent proportion et canopy geometry labour to 1991,<br />

al. 2001, elements), costs,<br />

layers, <strong>of</strong> into ZHANG<br />

NILSON sky<br />

many<br />

in vs. one 1999, which foliage<br />

difficulties et <strong>of</strong> al. <strong>the</strong> 2005,<br />

WEISS leaves seen best among<br />

et at are known point (3.1) contact and 0 in most zenith in frequently adjacent and azimuth layers). applied in The <strong>the</strong> directions Poisson estimation model from <strong>of</strong> is<br />

ptical ransfer e AZER <strong>the</strong> advantage measurements et in al. <strong>for</strong>est 2001, (3.1)<br />

ough divided <strong>the</strong> into canopy N statistically LAI <strong>of</strong> canopies. convenience JONCKHEERE<br />

dispersed is will <strong>the</strong> be<br />

in direction randomly independent Poisson Major described and et problems model low al. 2004, labour<br />

( , )<br />

, and u is layers, independently (see include<br />

<strong>the</strong> leaf in e.g. costs,<br />

area which MONSI unknown many<br />

density <strong>of</strong> leaves each & difficulties<br />

beneath <strong>the</strong> canopy (JONCKHEERE et al. 2004). SAEKI<br />

(defined are o<strong>the</strong>r 1953, MUSSCHE et al. 2001, NILSON 1999, WEISS et<br />

as <strong>the</strong> simplest <strong>of</strong> <strong>the</strong> above-mentioned models, as no<br />

& el iation NORMAN assumptions transfer 1991,<br />

<strong>of</strong><br />

photosyn<strong>the</strong>tic<br />

he <strong>the</strong> each canopy o<strong>the</strong>r in in direction (NILSON direction al. in and<br />

(Nilson 2004). <strong>for</strong>est ZHANG <strong>the</strong><br />

( 1999). tissue per<br />

( , It canopies. contribution et<br />

, )<br />

assumes al. 2005, Major<br />

unit<br />

,<br />

) u , If u<br />

canopy<br />

is is N <strong>the</strong> <strong>the</strong> that<br />

among <strong>of</strong><br />

leaf leaf<br />

, <strong>the</strong> problems woody<br />

<strong>the</strong> volume),<br />

area number canopy material<br />

number density is include<br />

<strong>of</strong> <strong>of</strong><br />

and G<br />

(defined divided to unknown<br />

overlaps (defined overlaps into<br />

( , )<br />

is<br />

as can as<br />

<strong>the</strong><br />

<strong>the</strong> <strong>the</strong> N statistically additional independent parameters layers, (as e.g. in layer which thickness) leaves are are<br />

per ckground nience model and <strong>of</strong> assumptions low indirect Though mean<br />

<strong>the</strong><br />

yn<strong>the</strong>tic<br />

d ution.<br />

syn<strong>the</strong>tic<br />

by plant The<br />

direction<br />

tissue organs. gap dispersed labour optical costs, a and large<br />

per<br />

be fraction<br />

( , <br />

per<br />

described Consequently<br />

) .<br />

unit is randomly many measurements <strong>the</strong> number contribution difficulties <strong>of</strong> different<br />

canopy <strong>the</strong>n by <strong>the</strong><br />

volume),<br />

<strong>the</strong> light and<br />

volume),<br />

probability Poisson independently will <strong>of</strong> be woody described statistical<br />

and distribution.<br />

G<br />

<strong>for</strong><br />

G ( ( <br />

zero <strong>of</strong><br />

, , )<br />

)<br />

is<br />

overlaps each material models<br />

is<br />

The o<strong>the</strong>r to exist (each characterizing a different statistical<br />

<strong>the</strong> <strong>the</strong> mean<br />

gap (NILSON 1999). If N , <strong>the</strong> number <strong>of</strong> overlaps<br />

mean<br />

needed. For fur<strong>the</strong>r reading see also Weiss et al.<br />

etical canopies. background Major distribution<br />

gh be this described process as can problems <strong>of</strong><br />

fraction be indirect <strong>of</strong><br />

quite described include canopy optical<br />

is <strong>the</strong>n variable <strong>the</strong> by probability <strong>the</strong> unknown elements), measurements one <strong>of</strong><br />

in Poisson <strong>for</strong> distribution. will <strong>the</strong> be best described known and most frequently applied in <strong>the</strong> estimation <strong>of</strong><br />

zero overlaps The and gap fraction (2004). is <strong>the</strong>n <strong>the</strong> probability <strong>for</strong> zero overlaps<br />

irection ection and <strong>the</strong> ( ( , contribution , )<br />

.<br />

) . LAI is <strong>the</strong><br />

nt sulting r radiation along in <strong>the</strong> a is downward path intercepted and can and according can <strong>of</strong> Poisson<br />

LAI pattern according woody<br />

by being plant to <strong>of</strong> Lang defined to material model (see<br />

light organs. LANG et al. as (1985) et to e.g. MONSI & SAEKI 1953, MUSSCHE et al. 2001, NILSON 1999, WEISS et<br />

Consequently al. (1985) be described be light described as as<br />

irect<br />

dent ction g ng th <strong>the</strong><br />

optical<br />

<strong>of</strong> <strong>the</strong> solar path <strong>the</strong> and path canopy. radiation canopy and<br />

measurements al. 2004). It assumes<br />

and LAI Though is geometry being intercepted being this defined<br />

will be that<br />

defined can process by as<br />

described <strong>the</strong> canopy is divided into N statistically independent layers, in which leaves are<br />

be as plant can be organs. quite Consequently variable<br />

(3.1) in light Based on <strong>the</strong> above-described models, measurements<br />

dispersed randomly ( <br />

uG<br />

, dand<br />

) independently <strong>of</strong> each o<strong>the</strong>r (NILSON 1999). If N , <strong>the</strong> number <strong>of</strong> overlaps<br />

0<br />

rtical sities y <strong>of</strong> depth usually radiation <strong>of</strong> decrease <strong>the</strong> Pinterception<br />

o ( canopy. ,<br />

resulting<br />

) eThough<br />

(or in non- a this downward , process pattern can be quite <strong>of</strong> light variable (3.1) in (3.2) <strong>of</strong> gap fraction (or transmittance) at ground (3.1) level can<br />

anopy in direction can ( be , described<br />

)<br />

tercepted flux ip simplified between densities by assumptions plant light usually organs. reduction decrease concerning<br />

, u<br />

Consequently<br />

is <strong>the</strong> by<br />

resulting and<br />

leaf <strong>the</strong><br />

<strong>the</strong> canopy<br />

area Poisson<br />

in light a<br />

density distribution.<br />

downward geometry<br />

(defined<br />

can pattern<br />

as<br />

be<br />

<strong>the</strong> The gap fraction is <strong>the</strong>n <strong>the</strong> probability <strong>for</strong> zero overlaps<br />

(3.2) <strong>of</strong> (3.2) light be used to derive LAI. Several detailed reviews, as<br />

and<br />

hetic tissue per<br />

where can<br />

unit canopy<br />

according is <strong>the</strong><br />

volume),<br />

path to length LANG<br />

and<br />

through et al. (1985)<br />

G ( , <br />

<strong>the</strong><br />

) is<br />

canopy be described<br />

<strong>the</strong> mean<br />

in direction as ( , )<br />

, u is <strong>the</strong> leaf area density (defined as <strong>the</strong><br />

y. elationship describe as Though e.g. leaf <strong>the</strong> this area between probability process where density, light can ε leaf is <strong>of</strong> be <strong>the</strong> reduction angle quite radiation path and variable length and interception in canopy through geometry <strong>the</strong> (or canopy non- can in be e.g. Jonckheere et al. (2004) or Bréda (2003), describe<br />

z<br />

on rease ight ese ls ( and , )<br />

, it follows after LANG et al (1985) that <strong>the</strong> gap fraction is<br />

z.<br />

total one-sided leaf area <strong>of</strong> photosyn<strong>the</strong>tic tissue per unit canopy volume), and G ( , )<br />

odels that models resulting is describe<br />

is <strong>the</strong> mean<br />

zgap<br />

are cos fraction in <strong>based</strong> direction a <strong>the</strong> downward on probability P osimplified<br />

( (<br />

, <br />

u)<br />

Gpattern<br />

, u which , is<br />

<strong>of</strong> assumptions d<strong>the</strong><br />

) radiation <strong>of</strong> leaf light area concerning interception density (defined <strong>the</strong> (or as non- <strong>the</strong>se respective sensors toge<strong>the</strong>r with <strong>the</strong>ir advantages<br />

0<br />

d nd , it , it follows after LANG et et al al (1985) that <strong>the</strong> <strong>the</strong> gap fraction is is<br />

ith sually py t reduction structural <strong>the</strong>se cos and cos azimuth models<br />

properties, and projection<br />

Po<br />

( ,<br />

)<br />

e<br />

<strong>the</strong> canopy total are directions <strong>based</strong> one-sided as <strong>of</strong> e.g. geometry unit on leaf simplified leaf<br />

from leaf area area<br />

,<br />

can area density, in assumptions be <strong>of</strong> <strong>the</strong> photosyn<strong>the</strong>tic direction leaf angle concerning ( and , tissue<br />

) .<br />

(3.1)<br />

<strong>the</strong> and problems. Instruments <strong>for</strong> transmittance<br />

e path and LAI being defined as<br />

key obability <strong>of</strong> canopy variable <strong>of</strong> structural in radiation where per those unit properties, models interception canopy is volume), as gap e.g. (or fraction leaf non- and area P Go<br />

density, ( ,<br />

)<br />

, is which leaf <strong>the</strong> angle mean and<br />

If we consider<br />

is <strong>the</strong> path<br />

G ( ,<br />

length<br />

)<br />

constant<br />

through<br />

along<br />

<strong>the</strong><br />

<strong>the</strong><br />

canopy<br />

path<br />

in<br />

and<br />

direction measurements<br />

LAI being<br />

( , )<br />

defined<br />

, u is <strong>the</strong> include<br />

as<br />

leaf area e.g. density <strong>the</strong> Sunfleck (defined or AccuPAR as <strong>the</strong><br />

e ased s. seen The on at key simplified point variable total 0 projection in assumptions zenith one-sided in those <strong>of</strong> unit and models concerning leaf leaf azimuth area area is <strong>of</strong> in gap <strong>the</strong> <strong>the</strong> photosyn<strong>the</strong>tic directions fraction direction P o ( from ,<br />

tissue )<br />

. , which per<br />

ch characterizing a different (3.2) (3.3) unit Ceptometers canopy volume), (Decagon and Devices G ( , )<br />

is Inc., <strong>the</strong> mean Pullman,<br />

Z statistical<br />

s. perties,<br />

04). foliage as seen e.g. leaf at point area 0 density, in<br />

LAI <br />

(3.3)<br />

frequently applied in <strong>the</strong> u dz estimation ,<br />

zenith leaf angle and and<br />

projection <strong>of</strong> unit leaf azimuth directions from WA, USA), <strong>the</strong> Demon device (CSIRO, Canberra,<br />

<strong>of</strong> area in <strong>the</strong> direction ( , )<br />

.<br />

(3.2)<br />

those et al. models 2004). is If gap we consider fraction 0 P Go<br />

( ,<br />

)<br />

, constant which along <strong>the</strong> path ε and Australia), or <strong>the</strong> LAI-2000 Plant Canopy Analyzer<br />

HE istical et zal.<br />

models 2001, NILSON exist (each 1999, characterizing WEISS et a different statistical<br />

in zenith , it follows and If LAI we azimuth after being consider LANG defined G directions et ( al as , (1985) )<br />

constant from that <strong>the</strong> along gap <strong>the</strong> fraction path is and LAI (LI-COR being defined Inc., as NE, USA). A study from Dufrêne &<br />

rent independent e best cosstatistical<br />

known<br />

layers, and models most in which exist frequently (each leaves applied characterizing are in <strong>the</strong> estimation a different z<strong>of</strong><br />

statistical<br />

where z is <strong>the</strong> total canopy height and , it follows Bréda after (1995) LANG shows et al (1985) good agreement that <strong>the</strong> gap between fraction Demon is<br />

Z<br />

one 9). & SAEKI If <strong>of</strong> N <strong>the</strong> 1953,<br />

best , <strong>the</strong> known MUSSCHE number and et most <strong>of</strong> al. overlaps 2001, frequently NILSON applied 1999, in WEISS <strong>the</strong> estimation coset<br />

<strong>of</strong><br />

LAI <br />

xist . ided <strong>the</strong>n MONSI (each into <strong>the</strong> & probability N characterizing SAEKI statistically described u dz , (3.2) and LAI-2000 Plant Canopy Analyzer (PCA). (3.2) With<br />

1953, <strong>for</strong> independent MUSSCHE zero a by 0 different overlaps et layers, statistical al. 2001, in which NILSON leaves 1999, are WEISS et<br />

respect to this <strong>the</strong>sis <strong>the</strong> main disadvantage <strong>of</strong> <strong>the</strong><br />

each py d most is o<strong>the</strong>r divided frequently (NILSON into N applied 1999). statistically in If <strong>the</strong> N independent estimation LAI<br />

, <strong>the</strong> number <strong>of</strong> layers, <strong>of</strong> in overlaps which<br />

<br />

(3.3) leaves are<br />

G<br />

, <br />

cos<br />

ion. ently MUSSCHE The <strong>of</strong> each gap et al. o<strong>the</strong>r fraction where P2001,<br />

( ,<br />

)<br />

<br />

z<br />

o (NILSON is NILSON z is <strong>the</strong>n e<strong>the</strong><br />

1999). <strong>the</strong> total 1999, probability canopy If WEISS . N height<br />

et<br />

<strong>for</strong> , <strong>the</strong> zero and number overlaps <strong>of</strong> overlaps , it follows after LANG et al (1985) that <strong>the</strong> gap fraction (3.3) is<br />

cos<br />

istically distribution. described independent as The gap layers, fraction in which is <strong>the</strong>n (3.1) leaves <strong>the</strong> probability are <strong>for</strong> zero overlaps<br />

described by<br />

1985) ON 1999). be described If N as<br />

, <strong>the</strong> number <strong>of</strong> overlaps<br />

u<br />

ction<br />

is <strong>the</strong><br />

is<br />

leaf<br />

<strong>the</strong>n<br />

area<br />

<strong>the</strong><br />

density<br />

probability<br />

(defined<br />

<strong>for</strong> zero<br />

as LAI <strong>the</strong><br />

G<br />

, <br />

<br />

overlaps<br />

cos <br />

(3.1)<br />

py volume), and PoG<br />

(<br />

(<br />

,<br />

,<br />

)<br />

<br />

<br />

)<br />

eis<br />

<strong>the</strong> mean . (3.3)<br />

(3.1)


quadrants <strong>of</strong> 0°, 45°, 90°, 180°, or 270° (cf. Figure 3-2). Due to a<br />

34<br />

wavelengths between 0.32 µm and 0.49 µm is recorded (LI-COR 1992<br />

Measurements are usually taken above and below <strong>the</strong> canopy to coll<br />

sky radiation (no direct radiation present). Transmittance can subseq<br />

Demon and <strong>the</strong> AccuPAR Ceptometer is that data<br />

I trans ( )<br />

must be collected over several hours over <strong>the</strong> course 9 ( ) ,<br />

(3.4)<br />

I o ( )<br />

<strong>of</strong> <strong>the</strong> day to capture an adequate range <strong>of</strong> sun angles.<br />

Gap fraction can fur<strong>the</strong>r be calculated directly from where where I trans ( ) is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted light as measured<br />

fish eye photographs (e.g. Jonckheere et al. 2005, intensity <strong>of</strong> <strong>the</strong> hemispheric light source (i.e. <strong>the</strong> sun) as measured<br />

Leblanc et al. 2005, Zhang et al. 2005), which are function <strong>of</strong> zenith angle .<br />

independent <strong>of</strong> illumination conditions (Baret 2005,<br />

personal communication).<br />

3.3.2 Instruments<br />

In order to determine LAI in <strong>the</strong> field, LAI-2000 PCA<br />

and digital hemispherical photographs were used in<br />

this <strong>the</strong>sis. Both are standard methods applied in <strong>the</strong><br />

CEOS, VALERI and BigFoot validation ef<strong>for</strong>ts and<br />

will thus be described in more detail.<br />

The standard method to<br />

<strong>of</strong> <strong>the</strong> above-mentioned Poisson model. In wavelengths where<br />

radiation, leaves are assumed to be non-reflecting and non-transmitti<br />

LAI-2000 Plant Canopy Analyzer<br />

fraction and is fur<strong>the</strong>r averaged over azimuth, Equation (3.3) become<br />

LAI<br />

The LAI-2000 Plant Canopy Analyzer (LI-COR,<br />

G<br />

cos<br />

o e<br />

Lincoln, Nebraska, cf. Figure 3-2) records light<br />

intensity with an optical sensor (LAI-2050) consisting<br />

<strong>of</strong> five concentric silicon detectors with fields <strong>of</strong> view<br />

between 0° and 74.1°. The nominal coverage <strong>of</strong> each<br />

detector is shown in Table 3-2. Incoming radiation is<br />

projected through a fisheye lens onto <strong>the</strong> photoelectric<br />

sensor rings, where it is integrated over <strong>the</strong> complete<br />

azimuth range and stored in <strong>the</strong> associated control<br />

unit (LAI-2070). In order to avoid unwanted effects,<br />

such as operator shadow, <strong>the</strong> 360° azimuthal field <strong>of</strong><br />

view may be restricted with view caps into quadrants<br />

<strong>of</strong> 0°, 45°, 90°, 180°, or 270° (cf. Figure 3-2). Due<br />

to an optical filter, only radiation with wavelengths<br />

between 0.32 µm and 0.49 µm is recorded (LI-COR<br />

1992).<br />

Measurements are usually taken above and below <strong>the</strong><br />

canopy to collect above and below canopy diffuse sky<br />

radiation (no direct radiation present). Transmittance<br />

τ can subsequently be calculated <strong>for</strong> each ring as<br />

<br />

,<br />

which can also be written as<br />

G ( ) LAI ln<br />

( ) cos<br />

.<br />

For flat leaves with a random distribution <strong>of</strong> foliage and leaf azimuth<br />

/ 2<br />

G ( ) sind<br />

0.<br />

5 .<br />

0<br />

Inserting Equation 3.6 in 3.7 leads to<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

<br />

9 THEORETICAL BACKGROUND<br />

is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted<br />

where I trans ( ) is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted light as as measured below <strong>the</strong> canopy and I o ( ) is <strong>the</strong><br />

intensity <strong>of</strong> <strong>the</strong> hemispheric light source (i.e. <strong>the</strong> sun) as measured above <strong>the</strong> canopy, both varying as a<br />

function <strong>of</strong> zenith angle .<br />

Table 3-2: Nom<br />

rings <strong>of</strong> LAI-2000<br />

LAI-2000 PCA<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Figure 3-2: LI-CO<br />

The standard method to calculate LAI is <strong>based</strong> on inversion<br />

<strong>of</strong> <strong>the</strong> above-mentioned Poisson model. In wavelengths where <strong>the</strong> LAI-2000 PCA sensor records<br />

radiation, leaves are assumed to be non-reflecting and non-transmitting. As ( ) is thus equivalent to gap<br />

fraction and is fur<strong>the</strong>r averaged over azimuth, Equation (3.3) becomes<br />

LAI<br />

G<br />

cos<br />

o e<br />

.<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical solution <strong>for</strong> Equation 3<br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance in different s<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

,<br />

i1<br />

where i stands <strong>for</strong> rings 1-5 and Wi are normalized weighting factors a<br />

with <strong>the</strong> different rings. Whereas cos values are constant <strong>for</strong> each<br />

viewing angle, weighting factors can vary according to <strong>the</strong> number <strong>of</strong><br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings are employe<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly fo<br />

<br />

, (3.5)<br />

which can also be written as<br />

G ( ) LAI ln<br />

( ) cos<br />

. (3.6)<br />

For flat leaves with a random distribution <strong>of</strong> foliage and leaf azimuth angles, MILLER (1967) showed that<br />

/ 2<br />

G ( ) sind<br />

0.<br />

5 . (3.7)<br />

0<br />

Inserting Equation 3.6 in 3.7 leads to<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

<br />

foliage angle distribution, improper model assumptions and <strong>the</strong> contribution <strong>of</strong> woody material to<br />

9 THEORETICAL BACKGROUND<br />

radiation interception. The <strong>the</strong>oretical background <strong>of</strong> indirect optical measurements will be described<br />

where<br />

briefly<br />

I<br />

in <strong>the</strong> following.<br />

trans ( ) is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted light as measured below <strong>the</strong> canopy and I o ( ) is <strong>the</strong><br />

intensity As it <strong>of</strong> penetrates <strong>the</strong> hemispheric <strong>the</strong> canopy, light incident source (i.e. solar <strong>the</strong> radiation sun) as is measured intercepted above by <strong>the</strong> plant canopy, organs. both Consequently varying as a light<br />

function attenuation <strong>of</strong> zenith corresponds angle . to <strong>the</strong> vertical depth <strong>of</strong> <strong>the</strong> canopy. Though this process can be quite variable in<br />

space and time, mean horizontal flux densities usually decrease resulting in a downward pattern <strong>of</strong> light<br />

reduction (SAEKI 1975). The relationship between light reduction and canopy geometry can be<br />

characterized by statistical models that describe <strong>the</strong> probability <strong>of</strong> radiation interception (or noninterception)<br />

within <strong>the</strong> canopy. Usually <strong>the</strong>se models are <strong>based</strong> on simplified assumptions concerning is <strong>the</strong><br />

radiative and spatial characteristics <strong>of</strong> canopy structural <strong>the</strong> intensity properties, <strong>of</strong> <strong>the</strong> hemispheric as e.g. leaf area light density, source (i.e. leaf <strong>the</strong> angle and<br />

foliage distribution among o<strong>the</strong>rs. The key sun) variable as measured in those above models <strong>the</strong> canopy, is gap fraction both varying P o ( ,<br />

as )<br />

a , which<br />

describes <strong>the</strong> proportion <strong>of</strong> sky vs. foliage seen function Table at point <strong>of</strong> 3-2: zenith 0 in Nominal zenith angle angular . and azimuth coverage directions <strong>of</strong> <strong>the</strong> different from<br />

beneath <strong>the</strong> canopy (JONCKHEERE et al. 2004). rings <strong>of</strong> LAI-2000 PCA (LI-COR 1992)<br />

The standard<br />

LAI-2000<br />

method<br />

PCA ring<br />

to calculate<br />

Aangular<br />

LAI is<br />

coverage<br />

<strong>based</strong> on<br />

Though a large number <strong>of</strong> different statistical models exist (each characterizing a different statistical [°]<br />

inversion <strong>of</strong> <strong>the</strong> above-mentioned Poisson model. In<br />

distribution <strong>of</strong> canopy elements), one <strong>of</strong> <strong>the</strong> best known and 1 most frequently applied 0.0°-12.3° in <strong>the</strong> estimation <strong>of</strong><br />

The standard wavelengths method where to <strong>the</strong> calculate LAI-2000 LAI PCA is <strong>based</strong> sensor on records inversion<br />

LAI is <strong>the</strong> Poisson model (see e.g. MONSI & SAEKI 1953, MUSSCHE 2 et al. 2001, NILSON 16.7°-28.6° 1999, WEISS et<br />

<strong>of</strong> <strong>the</strong> above-mentioned Poisson model. In wavelengths radiation, leaves where are <strong>the</strong> assumed LAI-2000 to PCA be non-reflecting sensor records<br />

al. 2004). It assumes that <strong>the</strong> canopy is divided into N statistically 3 independent layers, 32.4°-43.4° in which leaves are<br />

radiation, leaves are assumed to be non-reflecting and non-transmitting. As ( ) is thus equivalent to gap<br />

dispersed randomly and independently <strong>of</strong> each o<strong>the</strong>r (NILSON 4 1999). If N , <strong>the</strong> 47.3°-58.1° number <strong>of</strong> overlaps<br />

fraction and is fur<strong>the</strong>r averaged over azimuth, Equation (3.3) becomes<br />

can be described by <strong>the</strong> Poisson distribution. The gap fraction 5 is <strong>the</strong>n <strong>the</strong> probability 62.3°-74.1° <strong>for</strong> zero overlaps<br />

and can according LAI<br />

G<br />

to LANG et al. (1985) be described as<br />

cos<br />

Figure 3-2: LI-COR LAI-2000 PCA (LI-COR 2007)<br />

o e<br />

<br />

( <br />

uG,<br />

d<br />

)<br />

0<br />

Po<br />

( ,<br />

)<br />

e , (3.1)<br />

where is <strong>the</strong> path length through <strong>the</strong> canopy in direction ( , )<br />

, u is <strong>the</strong> leaf area density (defined as <strong>the</strong><br />

total one-sided leaf area <strong>of</strong> photosyn<strong>the</strong>tic tissue per unit canopy volume), and G ( , )<br />

is <strong>the</strong> mean<br />

projection <strong>of</strong> unit leaf area in <strong>the</strong> direction ( , )<br />

.<br />

If we consider G ( , )<br />

constant along <strong>the</strong> path and LAI being defined as<br />

Z<br />

LAI u dz , (3.2)<br />

0<br />

z<br />

where z is <strong>the</strong> total canopy height and , it follows after LANG et al (1985) that <strong>the</strong> gap fraction is<br />

cos<br />

described by<br />

LAI<br />

G,<br />

<br />

cos<br />

P<br />

<br />

o ( ,<br />

)<br />

e . (3.3)<br />

. (3.8)<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical solution <strong>for</strong> Equation 3.8. It is <strong>based</strong> on LANG et al. (1985),<br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance in different sun angles, and includes <strong>the</strong> measured<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

, (3.9)<br />

i1<br />

where i stands <strong>for</strong> rings 1-5 and Wi are normalized weighting factors associated to <strong>the</strong> relative area covered<br />

with <strong>the</strong> different rings. Whereas cos values are constant <strong>for</strong> each ring as <strong>the</strong>y depend on <strong>the</strong> zenithal<br />

viewing angle, weighting factors can vary according to <strong>the</strong> number <strong>of</strong> rings used <strong>for</strong> LAI calculation. Table<br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings are employed, as well as <strong>the</strong> use <strong>of</strong> four rings (as<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly <strong>for</strong> both cases, but <strong>the</strong>y always sum up<br />

<br />

, (3.5)<br />

which can also be written as<br />

G ( ) LAI ln<br />

( ) cos<br />

. (3.6)<br />

For flat leaves with a random distribution <strong>of</strong> foliage and leaf azimuth angles, MILLER (1967) showed that<br />

/ 2<br />

G ( ) sind<br />

0.<br />

5 . (3.7)<br />

0<br />

Inserting Equation 3.6 in 3.7 leads to<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

<br />

9<br />

where I trans ( ) is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted light as measured<br />

intensity <strong>of</strong> <strong>the</strong> hemispheric light source (i.e. <strong>the</strong> sun) as measured a<br />

function <strong>of</strong> zenith angle .<br />

Table 3-2: Nominal angular coverage <strong>of</strong> <strong>the</strong> different<br />

rings <strong>of</strong> LAI-2000 PCA (LI-COR 1992)<br />

LAI-2000 PCA ring Aangular coverage [°]<br />

1 0.0°-12.3°<br />

2 16.7°-28.6°<br />

3 32.4°-43.4°<br />

4 47.3°-58.1°<br />

5 62.3°-74.1°<br />

Figure 3-2: LI-COR LAI-2000 PCA (LI-COR 2007)<br />

The standard method to<br />

<strong>of</strong> <strong>the</strong> above-mentioned Poisson is thus model. equivalent In wavelengths to where t<br />

radiation, gap fraction leaves and are is assumed fur<strong>the</strong>r averaged to be non-reflecting over azimuth, and non-transmittin<br />

fraction Equation and (3.3) is fur<strong>the</strong>r becomes averaged over azimuth, Equation (3.3) becomes<br />

LAI<br />

G<br />

cos<br />

o e<br />

. (3.8)<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical solution <strong>for</strong> Equation 3.8. It is <strong>based</strong> on LANG et al. (1985),<br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance in different sun angles, and includes <strong>the</strong> measured<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

, (3.9)<br />

i1<br />

where i stands <strong>for</strong> rings 1-5 and Wi are normalized weighting factors associated to <strong>the</strong> relative area covered<br />

with <strong>the</strong> different rings. Whereas cos values are constant <strong>for</strong> each ring as <strong>the</strong>y depend on <strong>the</strong> zenithal<br />

viewing angle, weighting factors can vary according to <strong>the</strong> number <strong>of</strong> rings used <strong>for</strong> LAI calculation. Table<br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings are employed, as well as <strong>the</strong> use <strong>of</strong> four rings (as<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly <strong>for</strong> both cases, but <strong>the</strong>y always sum up<br />

<br />

,<br />

which can also be written as<br />

G ( ) LAI ln<br />

( ) cos<br />

.<br />

For flat leaves with a random distribution <strong>of</strong> foliage and leaf azimuth a<br />

/ 2<br />

G ( ) sind<br />

0.<br />

5 .<br />

0<br />

Inserting Equation 3.6 in 3.7 leads to<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

<br />

9<br />

where I trans ( ) is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted light as measured<br />

intensity <strong>of</strong> <strong>the</strong> hemispheric light source (i.e. <strong>the</strong> sun) as measured a<br />

function <strong>of</strong> zenith angle .<br />

Table 3-2: Nom<br />

rings <strong>of</strong> LAI-2000<br />

LAI-2000 PCA<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Figure 3-2: LI-CO<br />

The standard method to<br />

<strong>of</strong> <strong>the</strong> above-mentioned Poisson model. In wavelengths where t<br />

radiation, leaves are assumed to be non-reflecting and non-transmittin<br />

fraction and is fur<strong>the</strong>r averaged over azimuth, Equation (3.3) becomes<br />

(3.5)<br />

LAI<br />

G<br />

cos<br />

o e<br />

which can also be written as<br />

.<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical solution <strong>for</strong> Equation 3<br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance in different su<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

,<br />

i1<br />

where i stands <strong>for</strong> rings 1-5 and Wi are normalized weighting factors a<br />

with <strong>the</strong> different rings. Whereas cos values are constant <strong>for</strong> each<br />

viewing angle, weighting factors can vary according to <strong>the</strong> number <strong>of</strong><br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings are employed<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly <strong>for</strong><br />

<br />

,<br />

which can also be written as<br />

G ( ) LAI ln<br />

( ) cos<br />

.<br />

For flat leaves with a random distribution <strong>of</strong> foliage and leaf azimuth a<br />

/ 2<br />

G ( ) sind<br />

0.<br />

5 .<br />

0<br />

Inserting Equation 3.6 in 3.7 leads to<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

<br />

9<br />

where I trans ( ) is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted light as measured<br />

intensity <strong>of</strong> <strong>the</strong> hemispheric light source (i.e. <strong>the</strong> sun) as measured a<br />

Table 3-2: Nom<br />

function <strong>of</strong> zenith angle .<br />

rings <strong>of</strong> LAI-2000<br />

LAI-2000 PCA<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Figure 3-2: LI-CO<br />

The standard method to<br />

<strong>of</strong> <strong>the</strong> above-mentioned Poisson model. In wavelengths where t<br />

radiation, leaves are assumed to be non-reflecting and non-transmittin<br />

fraction and is fur<strong>the</strong>r averaged over azimuth, Equation (3.3) becomes<br />

LAI<br />

G<br />

cos<br />

o e<br />

(3.6)<br />

For flat leaves with a random distribution <strong>of</strong> foliage<br />

and leaf azimuth angles, Miller (1967) showed that<br />

.<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical solution <strong>for</strong> Equation 3<br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance in different su<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

,<br />

i1<br />

where i stands <strong>for</strong> rings 1-5 and Wi are normalized weighting factors a<br />

with <strong>the</strong> different rings. Whereas cos values are constant <strong>for</strong> each<br />

viewing angle, weighting factors can vary according to <strong>the</strong> number <strong>of</strong><br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings are employed<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly <strong>for</strong><br />

<br />

,<br />

which can also be written as<br />

G ( ) LAI ln<br />

( ) cos<br />

.<br />

For flat leaves with a random distribution <strong>of</strong> foliage and leaf azimuth a<br />

/ 2<br />

G ( ) sind<br />

0.<br />

5 .<br />

0<br />

Inserting Equation 3.6 in 3.7 leads to<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

<br />

9<br />

where I trans ( ) is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted light as measured<br />

intensity <strong>of</strong> <strong>the</strong> hemispheric light source (i.e. <strong>the</strong> sun) as measured a<br />

function <strong>of</strong> zenith angle .<br />

Table 3-2: Nom<br />

rings <strong>of</strong> LAI-2000<br />

LAI-2000 PCA<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Figure 3-2: LI-CO<br />

The standard method to<br />

<strong>of</strong> <strong>the</strong> above-mentioned Poisson model. In wavelengths where t<br />

radiation, leaves are assumed to be non-reflecting and non-transmittin<br />

fraction and is fur<strong>the</strong>r averaged over azimuth, Equation (3.3) becomes<br />

LAI<br />

G<br />

cos<br />

o e<br />

(3.7)<br />

Inserting Equation 3.6 in 3.7 leads to<br />

.<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical solution <strong>for</strong> Equation 3<br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance in different su<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

,<br />

i1<br />

where i stands <strong>for</strong> rings 1-5 and Wi are normalized weighting factors a<br />

with <strong>the</strong> different rings. Whereas cos values are constant <strong>for</strong> each<br />

viewing angle, weighting factors can vary according to <strong>the</strong> number <strong>of</strong><br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings are employed<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly <strong>for</strong><br />

<br />

,<br />

which can also be written as<br />

G ( ) LAI ln<br />

( ) cos<br />

.<br />

For flat leaves with a random distribution <strong>of</strong> foliage and leaf azimuth a<br />

/ 2<br />

G ( ) sind<br />

0.<br />

5 .<br />

0<br />

Inserting Equation 3.6 in 3.7 leads to<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

<br />

9<br />

where I trans ( ) is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted light as measured<br />

intensity <strong>of</strong> <strong>the</strong> hemispheric light source (i.e. <strong>the</strong> sun) Table as measured 3-2: Noma<br />

function <strong>of</strong> zenith angle .<br />

rings <strong>of</strong> LAI-2000<br />

LAI-2000 PCA<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Figure 3-2: LI-CO<br />

The standard method to<br />

<strong>of</strong> <strong>the</strong> above-mentioned Poisson model. In wavelengths where t<br />

radiation, leaves are assumed to be non-reflecting and non-transmittin<br />

fraction and is fur<strong>the</strong>r averaged over azimuth, Equation (3.3) becomes<br />

LAI<br />

G<br />

cos<br />

o e<br />

.<br />

(3.8)<br />

WELLES Welles & Norman NORMAN (1991) <strong>of</strong>fer <strong>of</strong>fer a numerical a numerical solution<br />

<strong>for</strong> Equation 3<br />

who <strong>for</strong> Equation introduced 3.8. <strong>the</strong> It is measurement <strong>based</strong> on Lang <strong>of</strong> et light al. (1985), transmittance who in different su<br />

transmittance introduced <strong>the</strong> data measurement <strong>for</strong> <strong>the</strong> five <strong>of</strong> different light transmittance<br />

rings:<br />

in different sun angles, and includes <strong>the</strong> measured<br />

5<br />

transmittance LAI 2 lndata<br />

i cos <strong>for</strong> <strong>the</strong><br />

Wi<br />

five , different rings:<br />

i1<br />

where i stands <strong>for</strong> rings 1-5 and Wi are normalized weighting factors a<br />

with <strong>the</strong> different rings. Whereas cos values are constant <strong>for</strong> each<br />

viewing angle, weighting factors can vary according to <strong>the</strong> number <strong>of</strong><br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings are employed<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly fo<br />

<br />

,<br />

which can also be written as<br />

G ( ) LAI ln<br />

( ) cos<br />

.<br />

For flat leaves with a random distribution <strong>of</strong> foliage and leaf azimuth a<br />

/ 2<br />

G ( ) sind<br />

0.<br />

5 .<br />

0<br />

Inserting Equation 3.6 in 3.7 leads to<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

<br />

9<br />

where I trans ( ) is <strong>the</strong> intensity <strong>of</strong> <strong>the</strong> transmitted light as<br />

intensity <strong>of</strong> <strong>the</strong> hemispheric light source Table (i.e. <strong>the</strong> 3-2: sun) Nom as<br />

function <strong>of</strong> zenith angle .<br />

rings <strong>of</strong> LAI-2000<br />

LAI-2000 PCA<br />

1<br />

2<br />

3<br />

4<br />

5<br />

Figure 3-2: LI-CO<br />

The standard m<br />

<strong>of</strong> <strong>the</strong> above-mentioned Poisson model. In wavelength<br />

radiation, leaves are assumed to be non-reflecting and non<br />

fraction and is fur<strong>the</strong>r averaged over azimuth, Equation (3.<br />

LAI<br />

G<br />

cos<br />

o e<br />

.<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical solution <strong>for</strong> Equation 3<br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance in different su<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

,<br />

(3.9)<br />

i1<br />

where i i stands <strong>for</strong> rings 1-5 1-5 and and Wi are normalized weighting factors a<br />

with <strong>the</strong> different rings. Whereas cos values are constant <strong>for</strong> each<br />

viewing angle, weighting factors can vary according to <strong>the</strong> number <strong>of</strong><br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings are employed<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly fo<br />

<br />

,<br />

which can also be written as<br />

G ( ) LAI ln<br />

( ) cos<br />

.<br />

For flat leaves with a random distribution <strong>of</strong> foliage and lea<br />

/ 2<br />

G ( ) sind<br />

0.<br />

5 .<br />

0<br />

Inserting Equation 3.6 in 3.7 leads to<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

<br />

Table<br />

rings o<br />

LAI<br />

Figure<br />

.<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical solution <strong>for</strong><br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance in<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

, are normalized<br />

i1<br />

weighting factors associated to <strong>the</strong> relative area<br />

where i stands <strong>for</strong> rings 1-5 and Wi are normalized weightin<br />

with <strong>the</strong> different rings. Whereas cos values are constan<br />

viewing angle, weighting factors can vary according to <strong>the</strong> n<br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings ar<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ


Inserting Equation 3.6 in 3.7 leads to<br />

3 Theoretical background<br />

35<br />

/ 2<br />

LAI 2 lnocos<br />

sin<br />

d<br />

0<br />

covered with <strong>the</strong> different rings. Whereas cos θ<br />

values are constant <strong>for</strong> each ring as <strong>the</strong>y depend on<br />

<strong>the</strong> zenithal viewing angle, weighting factors can<br />

vary according to <strong>the</strong> number <strong>of</strong> rings used <strong>for</strong> LAI<br />

calculation. Table 3-3 shows <strong>the</strong><br />

<br />

LAI 2 ln o cos<br />

sin<br />

d<br />

0<br />

. (3.8)<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical use solution <strong>of</strong> four <strong>for</strong> rings Equation (as used 3.8. in It is this <strong>based</strong> <strong>the</strong>sis, on cf. LANG Chapter et al. (1985),<br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance 5.2.1). The in different sun angles, and includes <strong>the</strong> measured<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

, factors in <strong>the</strong><br />

(3.9)<br />

i1<br />

event that all five rings are employed, as well as <strong>the</strong><br />

where i stands <strong>for</strong> rings 1-5 and Wi are normalized weighting factors associated to <strong>the</strong> relative area covered<br />

with <strong>the</strong> different rings. Whereas cos values are constant <strong>for</strong> each ring as <strong>the</strong>y depend on <strong>the</strong> zenithal<br />

viewing angle, weighting factors can vary according to <strong>the</strong> number <strong>of</strong> rings used <strong>for</strong> LAI calculation. Table<br />

3-3 shows <strong>the</strong> W i factors in <strong>the</strong> event that all five rings are employed, as well as <strong>the</strong> use <strong>of</strong> four rings (as<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly <strong>for</strong> both cases, but <strong>the</strong>y always sum up<br />

.<br />

WELLES & NORMAN (1991) <strong>of</strong>fer a numerical solution <strong>for</strong> Equation 3.8. It is<br />

who introduced <strong>the</strong> measurement <strong>of</strong> light transmittance in different sun angle<br />

transmittance data <strong>for</strong> <strong>the</strong> five different rings:<br />

5<br />

LAI 2 ln<br />

i cos Wi<br />

, factors differ slightly <strong>for</strong> both cases,<br />

i1<br />

but <strong>the</strong>y always sum up to unity. LAI is <strong>the</strong>re<strong>for</strong>e <strong>the</strong><br />

where i stands weighted <strong>for</strong> rings sum 1-5 <strong>of</strong> and measured Wi are normalized canopy transmittance,<br />

weighting factors associate<br />

with <strong>the</strong> different where <strong>the</strong> rings. highest Whereas weight cos is put values on <strong>the</strong> are outmost constant ring <strong>for</strong> as each ring as<br />

viewing angle, it covers weighting <strong>the</strong> largest factors part can <strong>of</strong> vary <strong>the</strong> according canopy (as to illustrated <strong>the</strong> number <strong>of</strong> rings us<br />

3-3 shows <strong>the</strong> in Figure W i factors 3-3). in <strong>the</strong> event that all five rings are employed, as we<br />

used in this <strong>the</strong>sis, cf. Chapter 5.2.1). The W i factors differ slightly <strong>for</strong> both c<br />

Over <strong>the</strong> last 15 years <strong>the</strong> LAI-2000 PCA has gained<br />

wide acceptance. Though several constraints remain<br />

(cf. Chapter 3.3.3), <strong>the</strong> device has been used in a<br />

large number <strong>of</strong> studies in <strong>for</strong>ested environments.<br />

Examples <strong>for</strong> applications in boreal <strong>for</strong>ests can be<br />

found in Chen et al. (1997), Fassnacht et al. (1994),<br />

Stenberg et al. (1994). Cutini et al. (1998), Dufrêne &<br />

Bréda (1995), and Eriksson et al. (2005). Strachan &<br />

McCaughey (1996) used <strong>the</strong> LAI-2000 PCA <strong>for</strong> LAI<br />

determination in deciduous <strong>for</strong>ests and De Wasseige<br />

et al. (2003) and Kalácska et al. (2005) in tropical<br />

rain <strong>for</strong>ests.<br />

Figure 3-2<br />

Table 3-2<br />

LI-COR LAI-2000 PCA (LI-COR 2007).<br />

Nominal angular coverage <strong>of</strong> <strong>the</strong> different rings<br />

<strong>of</strong> LAI-2000 PCA (LI-COR 1992).<br />

LAI-2000 PCA ring Angular coverage [°]<br />

1 0.0°-12.3°<br />

2 16.7°-28.6°<br />

3 32.4°-43.4°<br />

4 47.3°-58.1°<br />

5 62.3°-74.1°<br />

Table 3-3 cos θ and W values <strong>for</strong> LAI-2000 PCA rings 1-5<br />

i<br />

and rings 1-4 respectively (LI-COR 1992).<br />

Ring cos θ W i (rings 1-5) W i (rings 1-4)<br />

1 0.993 0.034 0.034<br />

2 0.921 0.104 0.103<br />

3 0.788 0.160 0.158<br />

4 0.602 0.218 0.706<br />

5 0.375 0.484 n/a<br />

<br />

<br />

Digital hemispherical photography<br />

Gap fraction (and in consequence LAI) can also<br />

be derived from hemispherical photography.<br />

Hemispherical photographs have been used since <strong>the</strong><br />

1960s to study canopy structural in<strong>for</strong>mation, with<br />

Anderson (1964) being <strong>the</strong> first to compute light<br />

penetration through <strong>the</strong> canopy with that technique.<br />

Hemispherical pictures have been analysed with<br />

regard to LAI by various authors (e.g. Anderson<br />

1981, Bonhomme et al. 1974, Jonckheere et al. 2005,<br />

Leblanc et al. 2005, Van Gardingen et al. 1999, Zhang<br />

et al. 2005 among o<strong>the</strong>rs), with studies on <strong>for</strong>est LAI<br />

obtained by Leblanc et al. 2005, Jonckheere et al.<br />

2005 (boreal <strong>for</strong>ests), Neumann et al. 1989, Wang &<br />

Miller 1987 (temperate deciduous <strong>for</strong>ests), and Frazer<br />

et al. 2001 (temperate mixed <strong>for</strong>ests).<br />

Pictures are taken through a fisheye lens (cf. Figure<br />

3-4) from below <strong>the</strong> canopy or, in <strong>the</strong> case <strong>of</strong> low


36<br />

View angle (θ) 7° 23° 38° 53° 68°<br />

Figure 3-3<br />

Illustration <strong>of</strong> <strong>the</strong> different viewing angles <strong>of</strong> <strong>the</strong> LAI-2000 PCA.<br />

canopies or understorey vegetation, placed above<br />

looking downward. The hemispherical view results<br />

in an angular field <strong>of</strong> sight <strong>of</strong> –90 to 90° in zenith and<br />

360° in azimuth directions (cf. Figure 3-5). Levelling<br />

guarantees that <strong>the</strong> camera is oriented towards<br />

<strong>the</strong> zenith, allowing <strong>for</strong> a detailed analysis <strong>of</strong> <strong>the</strong><br />

zenithal and azimuthal variation <strong>of</strong> gap fraction. LAI<br />

is subsequently derived from inversion <strong>of</strong> Equation<br />

3.3, e.g. applying a LUT approach and certain<br />

assumptions concerning leaf angle distribution<br />

(fur<strong>the</strong>r details will be given in Chapter 5.2.2).<br />

In contrast to <strong>the</strong> LAI-2000 PCA, hemispherical<br />

photographs provide not only a single gap fraction<br />

value <strong>for</strong> each image, but also a graphic record <strong>of</strong><br />

site-, species- and age-specific differences in canopy<br />

architecture (Jonckheere et al. 2004).<br />

The development <strong>of</strong> digital cameras eliminated<br />

potential error sources resulting from <strong>the</strong> development<br />

and scanning <strong>of</strong> films needed <strong>for</strong> analogue techniques.<br />

Digital cameras nowadays have a high spatial<br />

resolution (decreasing <strong>the</strong> problem <strong>of</strong> mixed pixels)<br />

and a better radiometric image quality than analogue<br />

cameras. Critical steps in acquisition and analysis <strong>of</strong><br />

digital hemispherical photographs (DHP), however,<br />

remain <strong>the</strong> selection <strong>of</strong> an adequate exposure time and<br />

shutter speed (see Zhang et al. 2005), as well as <strong>the</strong><br />

right threshold to distinguish vegetation from gaps.<br />

As <strong>the</strong> latter is somewhat arbitrary and subjective<br />

it is a potential source <strong>of</strong> error <strong>for</strong> gap fraction and<br />

subsequent LAI retrieval.<br />

Figure 3-4 Digital camera with fish-eye lens.<br />

Figure 3-5 Hemispherical photograph taken in Budongo<br />

Forest, Nature Reserve on 11 October, 2005.


3 Theoretical background<br />

Several s<strong>of</strong>tware packages have been developed <strong>for</strong><br />

<strong>the</strong> processing <strong>of</strong> DHPs, such as HemiView (Delta-T<br />

Devices Ltd., Cambridge, UK), Gap Light Analyzer<br />

(GLA, Frazer et al. 1999), WinSCANOPY (Régent<br />

Instruments Inc., Canada) and CAN-EYE (Baret<br />

2004a).<br />

3.3.3<br />

Limitations <strong>of</strong> indirect<br />

optical methods<br />

The assessment <strong>of</strong> leaf area index using indirect<br />

optical methods <strong>of</strong>fers <strong>the</strong> possibility <strong>of</strong> fast and<br />

non-destructive sampling over large areas compared<br />

to direct or semi-direct methods. However, certain<br />

assumptions and simplifications are made concerning<br />

<strong>the</strong> structural and radiative properties <strong>of</strong> plant<br />

canopies that are sometimes unrealistic. As a result,<br />

errors associated with <strong>the</strong> respective LAI estimation<br />

have been recorded; <strong>the</strong>se have been described and<br />

analysed by various authors (e.g. Cutini et al. 1998,<br />

Hyer & Goetz 2004 and Inoue et al. 2004 <strong>for</strong> LAI-<br />

2000 PCA; Frazer et al. 2001 and Zhang et al. 2005<br />

<strong>for</strong> hemispherical photography, Jonckheere et al.<br />

2005, and Mussche et al. 2001 <strong>for</strong> both).<br />

Usually validation <strong>of</strong> indirect optical methods is<br />

per<strong>for</strong>med by comparison to direct or semi-direct<br />

approaches (e.g. Cutini et al. 1998, Leblanc &<br />

Chen 2001, Mussche et al. 2001). Although a high<br />

correlation usually indicates a good sensitivity <strong>of</strong><br />

optical instruments, a general underestimation <strong>of</strong><br />

LAI with respect to direct and semi-direct methods<br />

cannot be disregarded. This underestimation seems<br />

to be higher <strong>for</strong> coniferous stands than <strong>for</strong> broadleaf<br />

vegetation types (Nilson 1999), but can still result in<br />

an underestimation <strong>of</strong> “true LAI” <strong>of</strong> up to 30% <strong>for</strong><br />

<strong>the</strong> latter (Cutini et al. 1998) under certain conditions.<br />

The main limitations to LAI assessment with indirect<br />

optical methods will <strong>the</strong>re<strong>for</strong>e be described in <strong>the</strong><br />

following section toge<strong>the</strong>r with appropriate correction<br />

methods.<br />

Simplification <strong>of</strong> leaf optical properties<br />

37<br />

When LAI is calculated from transmittance<br />

measurements (e.g. from LAI-2000 PCA data), one<br />

<strong>of</strong> <strong>the</strong> main assumptions used to derive Equation 3.5<br />

is that foliage elements are only absorbing incoming<br />

radiation, i.e. <strong>the</strong>y are non-transmitting and nonreflecting.<br />

Transmission recorded below <strong>the</strong> canopy<br />

is equivalent to gap fraction only if <strong>the</strong>se conditions<br />

are fulfilled. If <strong>the</strong> conditions are not met, radiometric<br />

measurements will be biased and lead to an<br />

overestimation <strong>of</strong> below canopy light transmittance.<br />

In consequence LAI will be underestimated (Hyer &<br />

Goetz 2004). Although <strong>the</strong> LAI-2000 PCA records<br />

radiation in wavelengths, where foliage absorption<br />

reaches 95% and more, a small amount is also<br />

transmitted and reflected. Chen (2001) estimated <strong>the</strong><br />

bias due to <strong>the</strong> ignorance <strong>of</strong> first and second order<br />

scattering effects to be around 8%.<br />

Several studies recommended <strong>the</strong> exclusion <strong>of</strong> certain<br />

rings from calculation in order to retrieve better<br />

results. Leblanc & Chen (2001) e.g. found that LAI<br />

measured with LAI-2000 PCA under direct sunlight<br />

conditions was underestimated by 20% compared<br />

to a canopy radiative transfer model. To correct <strong>for</strong><br />

that, <strong>the</strong>y developed an empirical correction <strong>of</strong> LAI<br />

as a function <strong>of</strong> sun zenith angle θ . They state that<br />

sun<br />

LAI calculated only from ring 4 proved to be more<br />

consistent. Planchais & Pontailler (1999), however,<br />

found in a both <strong>the</strong>oretical and experimental analysis<br />

that <strong>the</strong> underestimation <strong>of</strong> LAI must ra<strong>the</strong>r result<br />

from an inappropriate use <strong>of</strong> <strong>the</strong> Poisson model.<br />

Clumping effects<br />

Foliage in plant canopies – and especially <strong>for</strong>est<br />

canopies – is usually not distributed randomly, ei<strong>the</strong>r<br />

horizontally or vertically as assumed by <strong>the</strong> Poisson<br />

model. In fact, leaves are grouped along branches<br />

and branches are associated with boles resulting in a


38<br />

clumped distribution <strong>of</strong> canopy elements at various LAI is <strong>the</strong>re<strong>for</strong>e <strong>the</strong> true variable measured with<br />

e<br />

scales (Hyer & Goetz 2004). This clumped and thus optical devices if no correction <strong>for</strong> clumping is<br />

non-random spatial distribution has consequences <strong>for</strong><br />

light transmittance through <strong>the</strong> canopy. Clumping<br />

applied.<br />

results in a higher canopy transmittance than Various studies have dealt with <strong>the</strong> correction <strong>for</strong> non-<br />

13 predicted by random models (such as <strong>the</strong> Poisson randomness and <strong>the</strong> THEORETICAL derivation <strong>of</strong> BACKGROUND<br />

λ, especially when<br />

model, which assumes random distribution) and indirect optical methods were applied to coniferous<br />

thus in an underestimation <strong>of</strong> LAI (Black et al. 1991, canopies. According to Walter et al. (2003) <strong>the</strong>y can<br />

Chen & Cihlar 1995a, Fassnacht et al. 1994, among be grouped into optical-spectral methods (e.g. Chen<br />

o<strong>the</strong>rs). Though clumping is higher in conifer <strong>for</strong>ests et al. 1997, Fassnacht et al. 1994, Kuusk et al. 2002,<br />

compared 13 to broadleaf <strong>for</strong>ests as conifer needles are Smith et al. 1993 or THEORETICAL Stenberg et al. 1994) BACKGROUND and methods<br />

additionally grouped closely in shoots (see Figure incorporating stand dimension analyses (Kucharik et<br />

3-6 <strong>for</strong> illustration), it also affects measurements in<br />

broadleaf <strong>for</strong>ests.<br />

al. 1999, Nilson 1999).<br />

Chen & Cihlar (1995a) developed a procedure to<br />

Figure 3-6: Effect <strong>of</strong> foliage clustering on gap fraction according to NILSON (1999). Same amount <strong>of</strong><br />

needles<br />

According<br />

is dispersed<br />

to Nilson<br />

(a)<br />

(1971)<br />

randomly<br />

leaf<br />

and<br />

area<br />

b)<br />

index<br />

clustered.<br />

can be derive a clumping index <strong>based</strong> on gap size and gap<br />

derived from gap fraction in <strong>the</strong> presence <strong>of</strong> clumping fraction analyses. The underlying <strong>the</strong>ory is that within<br />

According through an to adjustment NILSON (1971) <strong>of</strong> Equation leaf area 3.3, index that includes can be derived a random from gap canopy fraction <strong>the</strong> in probability <strong>the</strong> presence <strong>of</strong> <strong>of</strong> <strong>the</strong> clumping occurrence<br />

through <strong>the</strong> clumping an adjustment index λ, <strong>of</strong> Equation 3.3, that includes <strong>the</strong> clumping <strong>of</strong> large gaps index can , be derived from <strong>the</strong> distribution <strong>of</strong><br />

LAI<br />

gap sizes. The iterative removal <strong>of</strong> large gaps from<br />

Figure 3-6: Effect<br />

G,<br />

<br />

cos <strong>of</strong><br />

<br />

foliage P ,<br />

<br />

clustering on gap fraction according to NILSON (1999). Same amount <strong>of</strong><br />

o e<br />

(3.10) <strong>the</strong> total gap fraction is equivalent to <strong>the</strong> (3.10) design <strong>of</strong><br />

needles is dispersed (a) randomly and b) clustered.<br />

an imaginary plant canopy with gaps inserted at<br />

with < 1 in <strong>the</strong> case <strong>of</strong> aggregated canopies, > 1 <strong>for</strong> regular canopies and 1 in <strong>the</strong> case <strong>of</strong> true random<br />

with According λ < 1 in to <strong>the</strong> NILSON case <strong>of</strong> (1971) aggregated leaf area canopies, index can > 1 be <strong>for</strong> derived random from gap and fraction having <strong>the</strong> in <strong>the</strong> same presence total gap <strong>of</strong> clumping fraction as<br />

leaf spatial distribution (BLACK et al. 1991). It is clear that if <strong>the</strong>re is no account <strong>for</strong> clumping in non-<br />

regular through canopies an adjustment and 1 in <strong>of</strong> <strong>the</strong> Equation case <strong>of</strong> 3.3, true that random includes leaf <strong>the</strong> <strong>the</strong> clumping real canopy. index , From this new canopy <strong>the</strong> canopy<br />

random canopies, <strong>the</strong> LAI derived from optical measurements will be under- or overestimated. To<br />

spatial distribution (Black et al. 1991). It is clear that element area index is calculated and compared to <strong>the</strong><br />

LAI<br />

account <strong>for</strong> that,<br />

G,<br />

BLACK et al. (1991) introduced <strong>the</strong> term “effective leaf area index”, which is defined as<br />

if <strong>the</strong>re is<br />

no account cos<br />

P ,<br />

<br />

<strong>for</strong><br />

clumping in non-random<br />

o e<br />

original canopy to derive <strong>the</strong> clumping index. (3.10) Gap<br />

LAIe= canopies, LAI <strong>the</strong> LAI derived from optical measurements size distribution may be estimated with <strong>the</strong> (3.11) TRAC<br />

with < 1 in <strong>the</strong> case <strong>of</strong> aggregated canopies, > 1 <strong>for</strong> regular canopies and 1 in <strong>the</strong> case <strong>of</strong> true random<br />

LAIe<br />

will be is <strong>the</strong>re<strong>for</strong>e under- or <strong>the</strong> overestimated. true variable To measured account with <strong>for</strong> that, optical devices instrument if no (Tracing correction Radiation <strong>for</strong> clumping and Architecture is applied. <strong>of</strong><br />

leaf spatial distribution (BLACK et al. 1991). It is clear that if <strong>the</strong>re is no account <strong>for</strong> clumping in non-<br />

Black et al. (1991) introduced <strong>the</strong> term “effective leaf Canopies, 3rd Wave Engineering).<br />

Various random studies canopies, have <strong>the</strong> dealt LAI with derived <strong>the</strong> correction from optical <strong>for</strong> measurements non-randomness will and be <strong>the</strong> under- derivation or overestimated. <strong>of</strong> , especially To<br />

area index”, which is defined as<br />

when account indirect <strong>for</strong> that, optical BLACK methods et al. (1991) were applied introduced to coniferous <strong>the</strong> term “effective canopies. leaf According area index”, to WALTER which is defined et al. (2003) as<br />

A simpler approach was developed by Lang & Xiang<br />

<strong>the</strong>y can be grouped into optical-spectral methods (e.g. CHEN et al. 1997, FASSNACHT et al. 1994, KUUSK<br />

LAIe= LAI .<br />

(3.11) (1986). They showed that in <strong>the</strong> case <strong>of</strong> large-scale (3.11)<br />

et al. 2002, SMITH et al. 1993 or STENBERG et al. 1994) and methods incorporating stand dimension<br />

analyses LAIe is <strong>the</strong>re<strong>for</strong>e (KUCHARIK <strong>the</strong> et true al. variable 1999, NILSON measured 1999). with optical devices if no correction <strong>for</strong> clumping is applied.<br />

CHEN Various & studies CIHLAR have (1995a) dealt developed with <strong>the</strong> a correction procedure <strong>for</strong> to non-randomness derive a clumping and index <strong>the</strong> <strong>based</strong> derivation on gap <strong>of</strong> size , especially and gap<br />

fraction when indirect analyses. optical The underlying methods were <strong>the</strong>ory applied is that to within coniferous a random canopies. canopy According <strong>the</strong> probability to WALTER <strong>of</strong> <strong>the</strong> et occurrence al. (2003)<br />

<strong>of</strong> <strong>the</strong>y large can gaps be grouped can be derived into optical-spectral from <strong>the</strong> distribution methods (e.g. <strong>of</strong> gap CHEN sizes. et The al. 1997, iterative FASSNACHT removal <strong>of</strong> et al. large 1994, gaps KUUSK from<br />

<strong>the</strong> et al. total 2002, gap SMITH fraction et is al. equivalent 1993 or STENBERG to <strong>the</strong> design et al. <strong>of</strong> 1994) an imaginary and methods plant incorporating canopy with stand gaps inserted dimension at<br />

random analyses and (KUCHARIK having <strong>the</strong> et al. same 1999, total NILSON gap fraction 1999). as <strong>the</strong> real canopy. From this new canopy <strong>the</strong> canopy<br />

element CHEN & area CIHLAR index (1995a) is calculated developed and compared a procedure to <strong>the</strong> to derive original a clumping canopy to index derive <strong>based</strong> <strong>the</strong> clumping on gap size index. and Gap gap<br />

size distribution may be estimated with <strong>the</strong> TRAC instrument (Tracing Radiation and Architecture <strong>of</strong><br />

Figure fraction 3-6analyses.<br />

Effect <strong>of</strong> The foliage underlying clustering on <strong>the</strong>ory gap fraction is that according within a to random Nilson (1999). canopy Same <strong>the</strong> amount probability <strong>of</strong> needles <strong>of</strong> is <strong>the</strong> dispersed occurrence<br />

Canopies, <strong>of</strong> large gaps 3rd (a) can Wave randomly be Engineering). derived and b) clustered. from <strong>the</strong> distribution <strong>of</strong> gap sizes. The iterative removal <strong>of</strong> large gaps from<br />

A <strong>the</strong> simpler total gap approach fraction was is developed equivalent by to LANG <strong>the</strong> design & XIANG <strong>of</strong> (1986). an imaginary They showed plant canopy that in <strong>the</strong> with case gaps <strong>of</strong> inserted large-scale at<br />

clumping random and (i.e. having at plant <strong>the</strong> level) same a total procedure gap fraction <strong>of</strong> spatial as <strong>the</strong> logarithmic real canopy. averaging From <strong>of</strong> this transmittance new canopy can <strong>the</strong> give canopy an<br />

approximation element area index <strong>of</strong> true is calculated LAI. Assuming and compared <strong>the</strong> Poisson to <strong>the</strong> model original – and canopy its condition to derive <strong>the</strong> <strong>of</strong> randomly clumping index. distributed Gap<br />

foliage size distribution – is more applicable may be estimated locally to with smaller <strong>the</strong> canopy TRAC sectors instrument (as e.g. (Tracing derived Radiation from DHPs), and Architecture Po is calculated <strong>of</strong><br />

Canopies, 3rd Wave Engineering).<br />

<strong>for</strong> each sector as well as its logarithm. Instead <strong>of</strong> averaging Po over whole azimuth ranges (as e.g.


3 Theoretical background<br />

clumping (i.e. at plant level) a procedure <strong>of</strong> spatial<br />

logarithmic averaging <strong>of</strong> transmittance can give an<br />

approximation <strong>of</strong> true LAI. Assuming <strong>the</strong> Poisson<br />

model – and its condition <strong>of</strong> randomly distributed<br />

foliage – is more applicable locally to smaller canopy<br />

sectors (as e.g. derived from DHPs), P is calculated<br />

O<br />

<strong>for</strong> each sector as well as its logarithm. Instead <strong>of</strong><br />

averaging P over whole azimuth ranges (as e.g.<br />

O<br />

determined with <strong>the</strong> LAI-2000 PCA) its logarithm<br />

derived <strong>for</strong> small sectors is averaged, approaching<br />

true LAI in <strong>the</strong> presence <strong>of</strong> clumping. This method<br />

can easily be used to derive <strong>the</strong> clumping index from<br />

hemispherical photographs and is implemented e.g.<br />

in <strong>the</strong> CAN-EYE s<strong>of</strong>tware.<br />

Chen & Cihlar (1995) and Law et al. (2001) noticed<br />

that it is more difficult to estimate clumping (and<br />

<strong>the</strong>re<strong>for</strong>e true LAI) <strong>for</strong> high and dense canopies<br />

due to darkness and multiple scattering inside <strong>the</strong><br />

canopy. According to Russell et al. (1989) it has been<br />

estimated that without any degree <strong>of</strong> leaf grouping, a<br />

tree could not sustain an LAI greater than 6.0 because<br />

<strong>of</strong> self-shading.<br />

Radiation interception by non-foliage elements<br />

Ano<strong>the</strong>r issue associated with LAI estimation from<br />

light transmittance measurements is absorption by<br />

non-photosyn<strong>the</strong>tic canopy elements. Stems and<br />

branches prevent light from reaching <strong>the</strong> optical<br />

device (cf. Figure 3-5) and as most instruments<br />

cannot discriminate between <strong>the</strong> effects caused by<br />

foliage or woody elements (cf. Cutini et al. 1998,<br />

Hyer & Goetz 2004, Kucharik et al. 1998, among<br />

o<strong>the</strong>rs) <strong>the</strong>y estimate Plant <strong>Area</strong> <strong>Index</strong> (PAI) instead<br />

<strong>of</strong> LAI. In order to derive LAI from PAI ano<strong>the</strong>r term<br />

is introduced here: <strong>the</strong> Woody <strong>Area</strong> <strong>Index</strong> (WAI),<br />

where<br />

PAI = LAI + WAI.<br />

(3.12)<br />

39<br />

Some studies (e.g. Cutini et al. 1998, Dufrêne &<br />

Bréda 1995, Kalácska et al. 2005) used leafless times<br />

during <strong>the</strong> year to determine WAI with LAI-2000<br />

PCA or DHP. However, care has to be taken when<br />

using this approach, as <strong>the</strong> contribution <strong>of</strong> stems and<br />

branches to LAI when leaf area is at its maximum<br />

is far less than <strong>the</strong> WAI determined during leafless<br />

periods. Alternatively Chapman (2007) developed an<br />

interesting method with hemispherical photography<br />

in <strong>the</strong> near-infrared to better distinguish foliage and<br />

woody elements. However, it is hard to estimate<br />

<strong>the</strong> amount <strong>of</strong> leaves that is obscured by stems and<br />

branches, so if <strong>the</strong> latter are subtracted from PAI, this<br />

leads to an underestimation <strong>of</strong> real LAI.<br />

Kucharik et al. (1998) demonstrated with<br />

measurements in <strong>the</strong> visible and near-infrared<br />

wavelengths <strong>of</strong> a Multiband Vegetation Imager<br />

(MVI) that <strong>for</strong> different boreal <strong>for</strong>est types in Canada<br />

<strong>the</strong> contribution <strong>of</strong> branch area to PAI did not exceed<br />

10% and did <strong>the</strong>re<strong>for</strong>e not significantly bias estimated<br />

LAI. However, <strong>the</strong>y suggest that account should be<br />

taken <strong>of</strong> stem area, as stems comprised 30 to 50%<br />

<strong>of</strong> <strong>the</strong> total woody area in <strong>the</strong>ir study. Stems may in<br />

most cases not, or only partially, be shed by leaves. In<br />

<strong>the</strong> same context Smolander & Stenberg (1996) state<br />

that <strong>the</strong> contribution <strong>of</strong> branches and stems to PAI is<br />

a function <strong>of</strong> foliage amount.<br />

Apart from <strong>the</strong> above-mentioned, fur<strong>the</strong>r errors exist<br />

when LAI is estimated from measurements with<br />

optical devices. These include instrument-inherent<br />

errors, such as instrument sensitivity or accuracy<br />

<strong>of</strong> <strong>the</strong> analogue-to-digital signal converter <strong>for</strong> <strong>the</strong><br />

LAI-2000 PCA. With regard to DHPs, possible lens<br />

distortion has to be taken into account. According<br />

to Jonckheere et al. (2005) even small amounts <strong>of</strong><br />

uncorrected angular distortion might cause substantial<br />

error in <strong>the</strong> measurement <strong>of</strong> gap area and distribution.<br />

As every lens invariably exhibits a small deviation<br />

from <strong>the</strong> <strong>the</strong>oretical projection, a lens correction<br />

should be applied.


40<br />

Apart from that, measurement errors may also<br />

result from <strong>the</strong> inappropriate use <strong>of</strong> <strong>the</strong> devices.<br />

According to Welles & Norman (1991) LAI-<br />

2000 PCA measurements, <strong>for</strong> example, should<br />

be made under diffuse light conditions, i.e. under<br />

homogeneously overcast skies or close to sunrise/<br />

sunset (Chason et al. 1991, Fassnacht et al. 1994). In<br />

this case <strong>the</strong> assumption that leaves are exclusively<br />

absorbing radiation is closest to reality. In spite<br />

<strong>of</strong> <strong>the</strong> optical filter, sunlit leaves in <strong>the</strong> canopy are<br />

expected to transmit and reflect light, which adds<br />

to <strong>the</strong> recorded radiation below <strong>the</strong> canopy. This is<br />

supported by De Wasseige et al. (2003) and Leblanc<br />

& Chen (2001), who found an influence <strong>of</strong> θ on sun<br />

<strong>the</strong> retrieved transmission values and consequently<br />

LAI. Additionally, if measurements are taken under<br />

direct sunlight and <strong>the</strong> optics <strong>of</strong> <strong>the</strong> LAI-2000 PCA<br />

are not carefully shaded, <strong>the</strong> ring containing <strong>the</strong> sun<br />

measures incoming radiation at θ , φ ra<strong>the</strong>r than<br />

sun sun<br />

over <strong>the</strong> entire azimuth range (Hyer & Goetz 2004).<br />

In addition, heterogeneous skies should be avoided<br />

with <strong>the</strong> LAI-2000 PCA instrument (Hyer & Goetz<br />

2004). If above and below canopy measurements are<br />

not taken at <strong>the</strong> same spot (as it is e.g. <strong>the</strong> case in<br />

<strong>for</strong>est canopies, where <strong>the</strong> reference sensor is located<br />

in large gaps or outside <strong>the</strong> <strong>for</strong>est), heterogeneous<br />

skies or moving clouds introduce a bias in <strong>the</strong><br />

transmission measurements, as I is not retrieved<br />

0<br />

from <strong>the</strong> same sky sector.<br />

For hemispherical photographs taken with analogue<br />

cameras, diffuse illumination conditions were also<br />

recommended (Jonckheere et al. 2004). With <strong>the</strong><br />

use <strong>of</strong> digital cameras however, <strong>the</strong> problem <strong>of</strong><br />

distinguishing sunlit leaves from small underexposed<br />

canopy gaps becomes less important. Better<br />

radiometric image quality and a spatial resolution<br />

comparable to that <strong>of</strong> an analogue camera allows<br />

image acquisition even under direct sunlight<br />

conditions. However, care has still to be taken with<br />

respect to exposure. Zhang et al (2005) showed that<br />

automatic exposure is not reliable <strong>for</strong> <strong>the</strong> calculation<br />

<strong>of</strong> gap fraction and resulted in an overestimation <strong>of</strong><br />

gap fraction (i.e. underestimation <strong>of</strong> LAI).<br />

3.3.4<br />

Spatial sampling strategies<br />

The set-up <strong>of</strong> field measurements and <strong>the</strong> spatial<br />

sampling strategy are key issues when per<strong>for</strong>ming<br />

ground-<strong>based</strong> measurements. The main factors in<br />

devising a sampling scheme include <strong>the</strong> size <strong>of</strong> <strong>the</strong><br />

complete test site, <strong>the</strong> size and spacing <strong>of</strong> individual<br />

sampling units, <strong>the</strong> sampling pattern, number<br />

<strong>of</strong> measurements to be made, distance between<br />

measurements and measurement height.<br />

The size <strong>of</strong> <strong>the</strong> complete test site mainly depends on<br />

<strong>the</strong> study objectives. Whereas <strong>for</strong> <strong>the</strong> characterization<br />

<strong>of</strong> different <strong>for</strong>est stands test sites <strong>of</strong> several km<br />

border length might be necessary, <strong>the</strong> description<br />

<strong>of</strong> LAI variability on small-scale agricultural fields<br />

could be accomplished within a much smaller area.<br />

Test sites should be flat to avoid terrain effects in<br />

satellite data and should be relatively homogeneous<br />

at medium resolution scale.<br />

On <strong>the</strong> test sites, smaller sampling units are usually<br />

established where <strong>the</strong> actual ground measurements<br />

occur. According to McCoy (2005) random sampling,<br />

stratified random sampling, systematic sampling and<br />

clustered sampling approaches can be applied to<br />

identify appropriate locations <strong>for</strong> <strong>the</strong> sampling units.<br />

If ground measurements are to be related to satellite<br />

imagery, Justice & Towneshend (1981) suggest a<br />

minimum area <strong>of</strong><br />

a = p (1+2l), (3.13)<br />

with a being <strong>the</strong> minimum area <strong>of</strong> a sampling unit, p<br />

<strong>the</strong> pixel dimension and l <strong>the</strong> location error in pixels.<br />

Especially when geolocation is difficult and l is large,<br />

a homogeneous stand geometry is desirable. This


3 Theoretical background<br />

can be evaluated in advance by using meaningful<br />

reference data and recent satellite imagery.<br />

To cover <strong>the</strong> whole variability <strong>of</strong> <strong>the</strong> measured<br />

variable, it is fur<strong>the</strong>r important to set up a large<br />

enough number <strong>of</strong> sampling units on <strong>the</strong> test site,<br />

which should also be equally spread. This can e.g.<br />

be accomplished by stratifying <strong>the</strong> test site into<br />

landscape units, <strong>based</strong> on vegetation, terrain and soil<br />

features (Aragão et al. 2005). With respect to <strong>the</strong><br />

later validation <strong>of</strong> remote sensing data, care should<br />

be taken that sampling units can be accurately linked<br />

to satellite pixels. Scale effects should be taken into<br />

account (cf. Chapters 3.4.2 and 3.4.4). To avoid <strong>the</strong><br />

effect <strong>of</strong> mixed pixels, sampling units should be<br />

located far enough from landscape boundaries.<br />

With respect to <strong>the</strong> sampling scheme, Garrigues<br />

et al. (2002) compared different patterns with<br />

geostatistical methods (cf. Figure 3-7). According<br />

to <strong>the</strong>m <strong>the</strong>re is no significant difference in <strong>the</strong><br />

spatial representativeness between <strong>the</strong> random<br />

square and cross sampling. Apart from <strong>the</strong>se two,<br />

transect sampling is very common in <strong>for</strong>ests (e.g. De<br />

Wasseige et al. 2003, Erikson et al. 2005, Le Dantec<br />

et al. 2000). O<strong>the</strong>r studies try to get an area-wide<br />

coverage <strong>of</strong> <strong>the</strong> sampling units (Kalácska et al. 2004,<br />

Stenberg et al. 2004).<br />

The number <strong>of</strong> individual measurements per ESU<br />

mainly depends on ESU size and canopy height.<br />

Ideally single measurements represent well LAI<br />

variability on a given ESU but are at <strong>the</strong> same time<br />

spatially independent to meet all conditions <strong>for</strong> later<br />

statistical analysis and to estimate mean LAI according<br />

to a given precision (De Wasseige et al. 2003,<br />

Ferment et al. 2001, and Weiss et al. 2004). There<strong>for</strong>e<br />

an adequate sampling distance must be selected;<br />

this will depend mainly on <strong>the</strong> instrument’s field <strong>of</strong><br />

view, but also on <strong>the</strong> vegetation type. Especially in<br />

dense vegetation, <strong>the</strong> area actually represented by a<br />

single measurement can be much smaller than <strong>the</strong><br />

41<br />

<strong>the</strong>oretical one. For transect measurements varying<br />

sampling distances can be found in <strong>the</strong> literature.<br />

Whereas generally distances <strong>of</strong> 2 to 30 m are<br />

reported (e.g. Dufrêne & Bréda 1995, Eriksson et al.<br />

2005, Mussche et al. 2001), many studies use a 10<br />

m lag (e.g. Aragão et al. 2005, Chen et al. 1997, Le<br />

Dantec et al. 2000). Very few studies however really<br />

tested whe<strong>the</strong>r individual measurements are spatially<br />

autocorrelated.<br />

Figure 3-7<br />

3.4<br />

a) Random square and b) cross sampling<br />

schemes (Garrigues et al. 2002).<br />

Satellite-<strong>based</strong> derivation <strong>of</strong> LAI<br />

The scientific community discovered <strong>the</strong> potential<br />

<strong>of</strong> remote sensing to monitor vegetation properties<br />

in <strong>the</strong> mid-seventies with <strong>the</strong> availability <strong>of</strong> Landsat<br />

data. The earliest attempts to derive LAI were<br />

purely empirical and mainly focused on agricultural<br />

landscapes. Kanemasu (1974) was among <strong>the</strong> first<br />

to correlate in situ measured LAI <strong>of</strong> soybean and<br />

sorghum with band ratios derived from Landsat<br />

MSS data. Comparable studies followed, all <strong>based</strong><br />

on band ratios or spectral vegetation indices (e.g.<br />

Pollock & Kanemasu 1979, Kanemasu et al. 1977).<br />

After promising results were received <strong>for</strong> various<br />

crop types, LAI estimation <strong>of</strong> <strong>for</strong>ested areas started in<br />

<strong>the</strong> mid-eighties. Peterson et al. (1987) were <strong>the</strong> first<br />

to assess LAI <strong>of</strong> coniferous <strong>for</strong>ests in <strong>the</strong> U.S. with<br />

Landsat TM data. O<strong>the</strong>r empirical studies dealing<br />

with <strong>for</strong>est LAI followed (e.g. Curran et al. 1992,<br />

Herwitz et al. 1990, Nemani et al. 1993, and Spanner<br />

et al. 1994, with Landsat TM; Spanner et al. 1990b<br />

with AVHRR).


42<br />

Concurrently <strong>the</strong> development <strong>of</strong> canopy reflectance<br />

models led to physical approaches <strong>of</strong> LAI estimation.<br />

Several studies at <strong>the</strong> beginning <strong>of</strong> <strong>the</strong> eighties<br />

showed that LAI could be estimated with sufficient<br />

accuracy from in situ measured canopy reflectance<br />

data and model inversion if ancillary data on leaf<br />

optical properties as well as soil reflectance are<br />

provided (Goel & Strebel 1983, Goel & Thompson<br />

1984a, Goel & Thompson 1984b). However, due to<br />

various constraints <strong>the</strong>se techniques were transferred<br />

to data acquired by spaceborne sensors only in <strong>the</strong><br />

mid-nineties (Jacquemoud et al. 1995, Kuusk 1995).<br />

Both methods, empirical and physical approaches <strong>of</strong><br />

LAI derivation from remote sensing data, have been<br />

applied and fur<strong>the</strong>r developed in various studies since<br />

<strong>the</strong> 1990s. They will consequently be described in <strong>the</strong><br />

following toge<strong>the</strong>r with <strong>the</strong>ir applicability to different<br />

sensor types. The subchapter closes with a discussion<br />

<strong>of</strong> <strong>the</strong> limitations <strong>of</strong> remotely-sensed LAI derivation<br />

with respect to tropical rain<strong>for</strong>est ecosystems.<br />

3.4.1<br />

Empirical approaches<br />

Empirical approaches <strong>of</strong> LAI derivation rely on<br />

relationships between in situ and satellite reflectance<br />

data or SVIs calculated from <strong>the</strong> latter. A prerequisite<br />

is <strong>the</strong> assumption that variations in surface reflectance<br />

(or SVI) are due to variations in LAI only. Figure<br />

3-8 shows <strong>the</strong> variation <strong>of</strong> surface reflectance with<br />

increasing LAI. SVIs make use <strong>of</strong> <strong>the</strong>se distinct<br />

characteristics in <strong>the</strong> NIR and SWIR wavelengths:<br />

typically NIR and SWIR reflectances increase with<br />

increasing LAI whereas reflectance in <strong>the</strong> visible<br />

wavelengths (especially in <strong>the</strong> red part <strong>of</strong> <strong>the</strong><br />

spectrum) remains low due to chlorophyll absorption.<br />

SVIs ideally maximize <strong>the</strong> sensitivity to biophysical<br />

variables and normalize both external (e.g. sun and<br />

viewing geometry, atmospheric disturbance) and<br />

internal effects (e.g. topography, soil variation) at<br />

<strong>the</strong> same time. The SVIs used most <strong>of</strong>ten include <strong>the</strong><br />

NDVI, <strong>the</strong> SR or SWIR corrected versions <strong>of</strong> both<br />

(cf. Chapter 3.4.3)<br />

Once a relationship between in situ measured LAI and<br />

an SVI is found, an empirical model is established<br />

through parametric or non-parametric regression<br />

analyses or artificial neural networks (ANN). Most<br />

<strong>of</strong>ten, <strong>the</strong> relationship between spectral data and LAI<br />

is modelled by simple or multiple regression analysis<br />

(Cohen et al. 2003). Here, both linear and non-linear<br />

models have been used to predict LAI from SVIs.<br />

Whereas Heiksanen (2006), Schlerf et al. (2005)<br />

and Stenberg et al. (2004) applied linear regression<br />

to estimate LAI from SVIs <strong>for</strong> temperate and boreal<br />

<strong>for</strong>ests, Chen & Cihlar (1996), Turner et al. (1999)<br />

and Kalácska et al. (2004) observed non-linear<br />

regression to be more accurate <strong>for</strong> boreal <strong>for</strong>ests in<br />

Canada, temperate <strong>for</strong>ests in <strong>the</strong> U.S. and a moist<br />

tropical rain<strong>for</strong>est in Costa Rica. Multiple linear<br />

regressions were found to be useful <strong>for</strong> <strong>the</strong> estimation<br />

<strong>of</strong> LAI <strong>of</strong> Wisconsin <strong>for</strong>ests and tropical rain <strong>for</strong>ests<br />

in Brazil by Fassnacht et al. (1997) and Aragão et<br />

Figure 3-8<br />

Reflectance from one to six layers <strong>of</strong><br />

cotton leaves (McCoy 2005).


3 Theoretical background<br />

al. (2005). However, no general conclusion can be<br />

drawn from this as certainly <strong>the</strong> choice <strong>of</strong> regression<br />

model depends on <strong>the</strong> used SVI (e.g. NDVI is very<br />

<strong>of</strong>ten reported to have a non-linear relationship to<br />

LAI, whereas simple ratio (SR) is more <strong>of</strong>ten used<br />

in linear models) and unique site characteristics. This<br />

will be discussed in more detail in Chapter 3.4.4.<br />

As atmospheric effects, viewing geometry and<br />

anisotropic reflectance behaviour <strong>of</strong> vegetation<br />

canopies are usually not taken into fur<strong>the</strong>r account, <strong>the</strong><br />

resulting regressions are mostly site and time specific.<br />

There<strong>for</strong>e operational use <strong>of</strong> empirical models <strong>for</strong><br />

LAI derivation is scarcely possible because in situ<br />

measurements are needed <strong>for</strong> every major land cover<br />

type to establish correct relationships. Consequently<br />

empirical relationships between remotely-sensed<br />

SVIs and in situ LAI are used most <strong>of</strong>ten <strong>for</strong> singledate<br />

and local to regional applications. The method is<br />

easy to implement and provides optimal results when<br />

extensive ground measurements were per<strong>for</strong>med.<br />

An exception is an approach that was developed by<br />

Sellers et al. (1994) <strong>based</strong> on empirically derived<br />

vegetation type properties and NDVI data from <strong>the</strong><br />

National Oceanic and Atmospheric Administration’s<br />

(NOAA) Advanced Very High Resolution<br />

Radiometer (AVHRR). Calculation <strong>of</strong> LAI is here<br />

<strong>based</strong> on <strong>the</strong> linear relationship <strong>of</strong> <strong>the</strong> fraction <strong>of</strong><br />

photosyn<strong>the</strong>tically active radiation absorbed by<br />

vegetation (FPAR) to NDVI after which LAI can<br />

be estimated from FPAR. As potential error sources<br />

exist in <strong>the</strong> original data due to cloud contamination,<br />

atmospheric constituents, and viewing geometry<br />

(Sellers et al. 1994), and NDVI seems to saturate <strong>for</strong><br />

LAI values above 2.0 to 3.0 (Sellers et al. 1996), this<br />

approach must be utilized with care. Never<strong>the</strong>less, it<br />

was <strong>the</strong> first method that allowed <strong>the</strong> derivation <strong>of</strong><br />

LAI data sets from operational processing (see, e.g.,<br />

Dech et al. 1998). A similar approach, but directly<br />

<strong>based</strong> on NDVI values, is used by <strong>the</strong> ECOCLIMAP<br />

project (Masson et al. 2003).<br />

3.4.2<br />

Physical models<br />

43<br />

In contrast to empirical methods, physical models<br />

are <strong>based</strong> on a <strong>the</strong>oretical description <strong>of</strong> canopy<br />

radiative transfer. They characterize <strong>the</strong> interaction<br />

<strong>of</strong> incoming electromagnetic radiation with canopy<br />

elements and its reflection <strong>based</strong> on certain leaf and<br />

canopy structural properties and soil background<br />

(Baret 1995). Ross (1981) gives a detailed description<br />

<strong>of</strong> radiative transfer modelling. Fur<strong>the</strong>r explicit<br />

reviews on this topic were compiled by Myneni &<br />

Ross (1991), and Myneni (1990a and b).<br />

One-dimensional models act on <strong>the</strong> presumption that<br />

<strong>the</strong> canopy is horizontally homogeneous and infinite,<br />

but variable and finite in <strong>the</strong> vertical direction. The<br />

canopy is assumed to contain only small and flat<br />

leaves that are randomly distributed in space. So are<br />

leaf azimuths (cf. Figure 3-9). This turbid medium<br />

approach allows a number <strong>of</strong> approximations and<br />

simplifications with respect to canopy scattering<br />

behaviour (Disney 2001). These models are <strong>the</strong>re<strong>for</strong>e<br />

<strong>based</strong> on a restricted amount <strong>of</strong> input parameters and<br />

are computationally efficient. Based on <strong>the</strong> work <strong>of</strong><br />

Suits (1972), Verhoef (1984) developed a canopy<br />

radiative transfer model called SAIL. This model<br />

was adapted and extended by several authors (e.g.,<br />

Gastellu-Etchegorry et al. 1996a, Kuusk 1995), and<br />

coupled with leaf optical models, e.g., PROSPECT<br />

(Jacquemoud et al. 1995). Comparisons to similar<br />

models gave promising results (Jacquemoud et al.<br />

2000). Bacour et al. (2006) successfully tested <strong>the</strong><br />

generation <strong>of</strong> LAI products <strong>based</strong> on MERIS data<br />

using a coupled SAIL+PROSPECT radiative transfer<br />

model within <strong>the</strong> CYCLOPES programme. According<br />

to Disney (2001) a major drawback <strong>of</strong> <strong>the</strong> onedimensional<br />

turbid-medium approach is that <strong>the</strong> size<br />

<strong>of</strong> scattering objects, i.e. leaves, is not considered. As<br />

certain properties <strong>of</strong> <strong>the</strong> observed canopy reflectance<br />

are directly controlled by <strong>the</strong> size and orientation<br />

<strong>of</strong> scattering objects (e.g. hotspot effects), different<br />

approaches are thus needed. Fur<strong>the</strong>r, as lateral


44<br />

homogeneity at <strong>the</strong> scale <strong>of</strong> interest is presumed <strong>for</strong><br />

1D models, <strong>the</strong>ir application leads to problems if<br />

input data is heterogeneous (e.g. mixed pixels in <strong>the</strong><br />

case <strong>of</strong> coarser resolution satellite data). Never<strong>the</strong>less,<br />

1D models provide a valid basis <strong>for</strong> radiative transfer<br />

modelling in small and homogeneous canopies (e.g.<br />

grasslands), that can be considered as turbid media<br />

(Shabanov et al. 2000).<br />

To account also <strong>for</strong> lateral heterogeneity in vegetation<br />

stands (e.g. single trees in <strong>for</strong>est stands), threedimensional<br />

models were developed. They divide<br />

<strong>the</strong> canopy in a rectangular matrix <strong>of</strong> parallelipipedic<br />

cells (voxels). Optical properties are computed with<br />

optical and structural element characteristics within<br />

each voxel, where also multiple scattering is allowed<br />

(Gastellu-Etchegorry et al. 1996b). Radiation<br />

transport is restricted to propagate only in a finite<br />

number <strong>of</strong> directions (Liang 2004). A widely applied<br />

example <strong>of</strong> a 3D canopy radiative transfer model is<br />

<strong>the</strong> DART model (Gastellu-Etchegorry et al. 1996b).<br />

Disney (2001) gives a good overview <strong>of</strong> more<br />

recent developments in 3D canopy radiative transfer<br />

modelling. Although 3D radiative transfer models<br />

have been developed since <strong>the</strong> 1980s, computational<br />

ef<strong>for</strong>t restricted <strong>the</strong> operational use <strong>of</strong> such algorithms<br />

<strong>for</strong> more than a decade (Jacquemoud et al. 2000). In<br />

principle, numerically solving 1D radiative transfer<br />

equations is carried out <strong>for</strong> one horizontal layer after<br />

<strong>the</strong> o<strong>the</strong>r, but solving 3D equations must be carried<br />

out cell by cell, which makes <strong>the</strong> whole process very<br />

time-consuming.<br />

Apart from <strong>the</strong> above-mentioned physically-<strong>based</strong><br />

models, o<strong>the</strong>r approaches were developed (Disney<br />

2001, Liang 2002). Li and Strahler introduced e.g.<br />

a geometric optical model <strong>for</strong> conifer <strong>for</strong>ests (Li<br />

& Strahler 1985 and 1986) that consists <strong>of</strong> simple<br />

geometric objects (e.g. cones or spheroids). According<br />

to Shabanov et al. (2000) this model suffers from an<br />

inaccurate description <strong>of</strong> multiple scattering and <strong>the</strong><br />

propagation <strong>of</strong> radiation into deeper canopy parts.<br />

Figure 3-9<br />

Simulation <strong>of</strong> vegetation leaf canopy as onedimensional<br />

turbid medium (Pinty et al. 2001).<br />

Computer simulation models, like <strong>the</strong> Monte Carlo<br />

Ray Tracing (cf. Figure 3-10) and <strong>the</strong> radiosity<br />

method (Disney 2001, Liang 2002, Shanbanov et al.<br />

2000), are on <strong>the</strong> o<strong>the</strong>r hand computationally very<br />

extensive.<br />

A critical step in deriving LAI from canopy radiative<br />

transfer models is model inversion. It allows <strong>the</strong><br />

retrieval <strong>of</strong> biophysical variables from remotelysensed<br />

measurements, but is only possible if a unique<br />

solution <strong>of</strong> <strong>the</strong> inverse problem exists (Knyazikhin<br />

et al. 1998). Model uncertainties and uncertainties<br />

in sensor-derived radiation measurements may,<br />

however, lead to a large number <strong>of</strong> possible solutions<br />

(Weiss et al. 2000). Different approaches exist <strong>for</strong><br />

model inversion (see, e.g., Kimes et al. 2000, Liang<br />

2004) with varying computational demand.<br />

A very efficient method <strong>of</strong> canopy radiative transfer<br />

inversion is <strong>the</strong> LUT method, which is also used in<br />

<strong>the</strong> <strong>MODIS</strong> LAI algorithm. As <strong>the</strong> in<strong>for</strong>mation on<br />

remotely-measured mean canopy-leaving radiance<br />

averaged over <strong>the</strong> three-dimensional radiation field<br />

is not sufficient to derive LAI, <strong>the</strong> range <strong>of</strong> variables<br />

determining <strong>the</strong> plant radiative regime has to be<br />

specified. Based on a global biome map, certain<br />

structural parameters <strong>of</strong> individual plants and <strong>the</strong><br />

canopy as a whole, as well as optical properties <strong>of</strong><br />

vegetation and soil, are predefined within certain<br />

ranges <strong>for</strong> each biome (Knyazikhin et al. 1998).<br />

Directional reflectance is computed in advance <strong>based</strong><br />

on a 3D model and <strong>the</strong> results are stored in a LUT.


3 Theoretical background<br />

Spectral reflectances derived from <strong>the</strong> <strong>MODIS</strong> sensor<br />

are <strong>the</strong>n compared to <strong>the</strong>se model entries in order to<br />

find LAI (Myneni et al. 2002).<br />

O<strong>the</strong>r inversion approaches include numerical<br />

optimization methods (e.g. Goel & Thompson 1984a<br />

and b, Jacquemoud et al. 1995, Liang 2002) and ANN<br />

(e.g. Atzberger et al. 2003, Baret et al. 1995, Weiss &<br />

Baret, 1999).<br />

3.4.3<br />

Spatial and spectral scale <strong>of</strong> sensors<br />

LAI and o<strong>the</strong>r biophysical variables have been<br />

derived from various remote-sensing data sources<br />

over <strong>the</strong> past 30 years. Usually <strong>the</strong> choice <strong>of</strong> spatial<br />

and spectral resolution <strong>of</strong> <strong>the</strong> input data depends on<br />

<strong>the</strong> study objectives and on sensor availability.<br />

With respect to <strong>for</strong>est studies, research on localto-regional<br />

scales is mainly aimed at an improved<br />

assessment <strong>of</strong> <strong>for</strong>est structure and contributions<br />

to <strong>for</strong>est management and monitoring (Rautiainen<br />

2005). Due to data availability most studies since <strong>the</strong><br />

1980s have focused on high resolution multispectral<br />

Figure 3-10<br />

Simulation <strong>of</strong> a 3D vegetation canopy with<br />

a Monte Carlo Ray Tracing model (Disney 2001).<br />

45<br />

satellite data, e.g., sensors aboard <strong>the</strong> Landsat, SPOT<br />

and IRS satellites (e.g. Berterreche et al. 2005, Brown<br />

et al. 2000, Chen & Cihlar 1996, Cohen et al. 2003,<br />

Fassnacht et al. 1994, Rautiainen 2005, Spanner et al.<br />

1990a, 1994, Stenberg et al. 1994, Turner et al. 1999,<br />

White et al. 1997, among many o<strong>the</strong>rs).<br />

In recent years airborne hyperspectral and very high<br />

resolution multispectral data has become available.<br />

Lee et al. (2004) compared, <strong>for</strong> instance, <strong>the</strong> power<br />

<strong>of</strong> hyperspectral AVIRIS data with simulated<br />

multispectral ETM+ data <strong>for</strong> predicting LAI <strong>of</strong><br />

needleleaf and broadleaf <strong>for</strong>ests in North America.<br />

According to <strong>the</strong>ir results, <strong>the</strong> use <strong>of</strong> narrow bands<br />

has no clear advantage over <strong>the</strong> use <strong>of</strong> broadband<br />

data. Although regression models using AVIRIS data<br />

per<strong>for</strong>med better, <strong>the</strong>y suspected an overfit <strong>of</strong> <strong>the</strong><br />

applied models. Schlerf et al. (2005) however found<br />

a better correlation between LAI and SVIs derived<br />

from hyperspectral HyMap data than simulated<br />

multispectral data <strong>for</strong> a needleleaf <strong>for</strong>est in Germany.<br />

Whereas <strong>the</strong> studies dealing with high resolution<br />

multispectral data found empirical regressions with<br />

SVIs to be best suited <strong>for</strong> LAI derivation, studies<br />

using very high resolution data as IKONOS or<br />

QuickBird, reported texture measures to be better<br />

related to LAI (Colombo et al. 2003, Leboeuf et<br />

al. 2007, Peddle & Johnston 2000, Seed & King<br />

2003). This is clearly a function <strong>of</strong> spatial scale: <strong>the</strong><br />

IFOV <strong>of</strong> very high spatial resolution sensors does<br />

not integrate over <strong>the</strong> reflectance caused by clusters<br />

<strong>of</strong> ground elements, e.g. a group <strong>of</strong> trees, but only<br />

represents parts <strong>of</strong> <strong>the</strong>se elements (e.g. parts <strong>of</strong> a<br />

shadowed crown). There<strong>for</strong>e very high resolution<br />

pictures have a much higher internal variance caused<br />

by stand structure and shadow fractions. This, in turn,<br />

is related to <strong>the</strong> complexity <strong>of</strong> <strong>for</strong>est stands, which<br />

increases with stand age and thus LAI.<br />

Though <strong>the</strong> major part <strong>of</strong> studies at very high and<br />

high spatial resolution used empirical approaches


46<br />

<strong>based</strong> on in situ data <strong>for</strong> LAI derivation, attempts<br />

to infer LAI from <strong>the</strong> inversion <strong>of</strong> canopy radiative<br />

transfer models were also made. Gascon et al. (2004)<br />

used <strong>for</strong> example a LUT inversion <strong>of</strong> a 3D canopy<br />

radiative transfer model to derive LAI <strong>of</strong> a mixed<br />

<strong>for</strong>est in France from SPOT-HRV and IKONOS data.<br />

Whereas <strong>the</strong> per<strong>for</strong>mance with SPOT data was good<br />

(see also Rautiainen 2005), problems occurred due<br />

to <strong>the</strong> high spatial resolution <strong>of</strong> IKONOS data. Here<br />

<strong>the</strong> inversion procedure had to be applied <strong>based</strong> on<br />

segmentation <strong>of</strong> homogeneous areas ra<strong>the</strong>r than on a<br />

pixel-per-pixel basis.<br />

In contrast to <strong>the</strong> above-mentioned studies that<br />

required a comparatively high spatial resolution <strong>for</strong><br />

<strong>the</strong>ir applications, a lower spatial resolution (300 m<br />

to 7 km) is sufficient <strong>for</strong> research on continental to<br />

global scales. Biophysical variables derived from<br />

<strong>the</strong>se sensors (e.g. <strong>MODIS</strong>, MERIS, POLDER,<br />

VEGETATION, AVHRR) mainly serve as input<br />

<strong>for</strong> climate or ecosystem modelling, <strong>for</strong> example<br />

trying to quantify net primary production and carbon<br />

sequestration under different climate scenarios.<br />

Though empirical approaches do exist at <strong>the</strong>se<br />

relatively coarse spatial scales (e.g. Chen et al.<br />

2002 <strong>for</strong> Canada, Sellers et al. 1996), <strong>the</strong> majority<br />

<strong>of</strong> models is physically <strong>based</strong>. This is mainly due to<br />

<strong>the</strong> fact that at <strong>the</strong> spatial scale <strong>of</strong> medium to coarse<br />

resolution sensors, observed pixels always represent<br />

a mixture <strong>of</strong> different land cover types. Empirical<br />

relationships are on <strong>the</strong> o<strong>the</strong>r hand very sensitive<br />

to <strong>the</strong> land surface and viewing geometry, so that<br />

universal relationships are hard – if not impossible –<br />

to establish.<br />

Physical models do not require ground data or a priori<br />

knowledge about canopies. Weiss & Baret (1999)<br />

consequently suggest <strong>the</strong> estimation <strong>of</strong> biophysical<br />

variables <strong>based</strong> on <strong>the</strong> modelling <strong>of</strong> radiative transfer<br />

process in <strong>the</strong> canopy <strong>for</strong> <strong>the</strong>se sensors. Recent<br />

examples are manifold and comprise operational and<br />

non-operational models presented by e.g. by Myneni<br />

et al. (2002), Bacour et al. (2006), Bicheron & Leroy<br />

(1999), Deng et al. (2006) and Bicheron et al. (1998).<br />

Table 3-4 gives an overview <strong>of</strong> <strong>the</strong> above-mentioned<br />

studies. The summary shows that empirical<br />

approaches are mainly applied to very high and high<br />

spatial resolution multispectral imagery. However,<br />

as empirical models are limited by vegetation<br />

type and related optical properties, sun-sensor<br />

geometry, background reflectance and <strong>the</strong> quality<br />

<strong>of</strong> atmospheric correction, <strong>the</strong>y are not suitable <strong>for</strong><br />

operational processing. Consequently <strong>for</strong> medium to<br />

coarse spatial resolution data mainly physical models<br />

are applied.<br />

With respect to spectral resolution, <strong>the</strong> above<br />

mentioned results Schlerf et al. (2005) <strong>of</strong>fer<br />

an interesting perspective especially <strong>for</strong> future<br />

hyperspectral satellite missions, such as EnMAP.<br />

With a spatial resolution <strong>of</strong> 30 m this sensor seems<br />

to be well suited <strong>for</strong> LAI derivation with empirical<br />

and physically-<strong>based</strong> models. It is acknowledged that<br />

LAI can also be derived from passive microwave<br />

sensors, which have certain advantages especially<br />

over tropical regions (e.g. independency from cloud<br />

cover). Yet as radar data was not available <strong>for</strong> this<br />

<strong>the</strong>sis, this kind <strong>of</strong> remote sensing data is not fur<strong>the</strong>r<br />

taken into account.<br />

3.4.4<br />

Suitability and limitations<br />

<strong>of</strong> empirical methods<br />

As mentioned in <strong>the</strong> previous chapter <strong>the</strong> major<br />

part <strong>of</strong> studies dealing with regional applications<br />

focused on empirical methods with high resolution<br />

multispectral data. As multispectral data will be used<br />

in this <strong>the</strong>sis <strong>for</strong> <strong>the</strong> upscaling <strong>of</strong> in situ measurements<br />

<strong>based</strong> on empirical relationships, a special focus <strong>of</strong><br />

this subchapter will be on SVIs and <strong>the</strong> influence <strong>of</strong><br />

structural characteristics on <strong>the</strong> establishment <strong>of</strong> such<br />

a model. As only few studies have analysed data <strong>for</strong>


3 Theoretical background<br />

tropical rain <strong>for</strong>ests, publications on temperate and<br />

boreal <strong>for</strong>ests will also be taken into account.<br />

Sensitivity <strong>of</strong> SVIs<br />

As mentioned above, empirical models <strong>based</strong> on SVIs<br />

derived from broadband multispectral sensors have<br />

shown to be useful <strong>for</strong> LAI assessment in various<br />

Table 3-4<br />

Data type Spatial<br />

resolution<br />

Multispectral Very high<br />

(0.60 m-4 m)<br />

Multispectral High<br />

(10 m-30 m)<br />

Multispectral Medium to<br />

coarse<br />

(300 m-7 km)<br />

Hyperspectral Very high<br />

to high<br />

(5 m)<br />

Optical remote sensing data types, mainly used LAI derivation methods, applications and recent examples.<br />

LAI derivation method Applications Recent examples<br />

Mainly empirical models,<br />

best regressions models<br />

reported <strong>for</strong> texture<br />

measures<br />

Mainly empirical models<br />

established with in situ<br />

data and SVIs,<br />

some physically-<strong>based</strong><br />

approaches with canopy<br />

radiative transfer models<br />

Mainly physical<br />

modelling, inversion <strong>of</strong><br />

radiative transfer models,<br />

some empirical or semiempirical<br />

approaches<br />

Site and time specific<br />

applications, single<br />

date (<strong>for</strong>est inventories,<br />

<strong>for</strong>est structure, <strong>for</strong>est<br />

monitoring)<br />

Site and time specific<br />

applications, mostly<br />

single date (<strong>for</strong>est<br />

inventories, <strong>for</strong>est<br />

structure, <strong>for</strong>est<br />

monitoring)<br />

Global derivation<br />

<strong>of</strong> biophysical<br />

variables as input<br />

<strong>for</strong> process models<br />

(evapotranspiration,<br />

crop yield, primary<br />

productivity, changes in<br />

vegetation structure),<br />

high temporal resolution<br />

Mainly empirical models Site and time specific<br />

applications, single<br />

date (<strong>for</strong>est inventories,<br />

<strong>for</strong>est structure, <strong>for</strong>est<br />

monitoring)<br />

CASI: Wulder et al. (1998)<br />

IKONOS: Colombo et al. (2003)<br />

QuickBird: Leboeuf et al. (2007)<br />

Aster: Eckert (2006), Heiskanen (2006)<br />

Landsat 5/7: Aragão et al. (2005),<br />

Berterreche et al. (2005),<br />

Brown et al. (2000), Butson &<br />

Fernandes (2004), Cohen et al.<br />

(2003), Kalácska et al. (2004),<br />

among many o<strong>the</strong>rs<br />

SPOT-4/5: Gascon et al. (2004),<br />

Rautiainen (2005)<br />

MERIS: Bacour et al. (2006)<br />

<strong>MODIS</strong>: Myneni et al. (2002)<br />

POLDER: Bicheron et al. (1998)<br />

SPOT-VGT: Chen et al. (2002),<br />

Deng et al. (2006)<br />

AVIRIS: Lee et al. (2004)<br />

HyMap: Schlerf et al. (2005)<br />

47<br />

studies (see Jensen & Bin<strong>for</strong>d 2004, and Liang 2004<br />

<strong>for</strong> general overviews). However, <strong>the</strong> major part<br />

<strong>of</strong> <strong>the</strong>se studies focused on natural and agricultural<br />

vegetation types in semi-arid, temperate and boreal<br />

regions. Especially <strong>for</strong> biomes with little above ground<br />

biomass, SVIs showed a considerable sensitivity to<br />

LAI variation (e.g. Cohen et al. 2003, Lu et al. 2004,<br />

among o<strong>the</strong>rs). Problems occur more frequently when<br />

higher LAI values are to be analysed.


48<br />

Gobron et al. (1997) investigated <strong>the</strong> estimation <strong>of</strong><br />

LAI from visible and near-infrared satellite data on<br />

a <strong>the</strong>oretical basis. According to <strong>the</strong>ir findings <strong>the</strong><br />

main constraint <strong>of</strong> LAI estimation from reflectance<br />

data is that <strong>the</strong> radiation field emerging at <strong>the</strong> top<br />

<strong>of</strong> <strong>the</strong> canopy does not always remain sensitive to<br />

increasing LAI values. Especially when <strong>the</strong> canopy<br />

is optically not thin enough to allow a significant<br />

illumination <strong>of</strong> <strong>the</strong> underlying soil, <strong>the</strong> radiation<br />

field becomes independent from fur<strong>the</strong>r increases in<br />

foliage area. If this happens, LAI can consequently<br />

not directly be estimated from reflectance data at<br />

particular wavelengths without fur<strong>the</strong>r input. This<br />

problem mainly occurs in dense canopies with high<br />

LAI values, as e.g. <strong>for</strong>ests. Consequently saturation<br />

<strong>of</strong> SVIs at higher LAI values is reported <strong>for</strong> various<br />

studies. Especially <strong>the</strong> classic NDVI shows saturation<br />

<strong>for</strong> high LAI values as its dynamic range is stretched<br />

in favour <strong>of</strong> low-biomass conditions.<br />

open and very dense canopies respectively. These<br />

findings are in line with Horler & Ahern (1986), who<br />

reported that SWIR bands contain more in<strong>for</strong>mation<br />

on <strong>for</strong>est structure in Western Canadian conifer and<br />

hardwood <strong>for</strong>ests than o<strong>the</strong>rs.<br />

SWIR data is also included in <strong>the</strong> normalized<br />

difference moisture index (NDMI), which is basically<br />

an NDVI including SWIR in<strong>for</strong>mation instead <strong>of</strong><br />

red reflectance (sometimes also called normalized<br />

difference water index or infrared index). Hardisky et<br />

al. (1983) reported that NDMI was highly correlated<br />

to canopy water content and tracked changes in plant<br />

biomass better than <strong>the</strong> NDVI.<br />

Ano<strong>the</strong>r SVI incorporating SWIR in<strong>for</strong>mation was<br />

introduced by Nemani et al. (1993), who applied a<br />

SWIR correction factor to <strong>the</strong> NDVI in order to<br />

improve <strong>the</strong> relation to LAI. This index is called <strong>the</strong><br />

corrected NDVI (NDVI ). According to <strong>the</strong>m <strong>the</strong><br />

c<br />

In contrary, SR is more sensitive to LAI variation in SWIR correction acts here as a scalar <strong>for</strong> canopy<br />

high biomass vegetation as indicated in Figure 3-11<br />

(assuming that high LAI corresponds to high NDVI<br />

closure.<br />

values) and shows a higher linearity to biophysical In this context, one <strong>of</strong> <strong>the</strong> rare studies in tropical rain<br />

variables (Chen 1996). The same is true <strong>for</strong> <strong>the</strong> <strong>for</strong>est environment was undertaken by Kalácska et<br />

modified simple ratio (MSR), as it is a function <strong>of</strong> SR al. (2004). They tested <strong>the</strong> relationship between LAI<br />

(cf. Table 3-5). Yet MSR increases more slowly with and different SVIs <strong>for</strong> a tropical moist <strong>for</strong>est in Costa<br />

NDVI.<br />

Rica. Although MSR and SR were especially sensitive<br />

24 at higher LAI values (RTHEORETICAL BACKGROUND<br />

2 =0.77 and 0.80 respectively),<br />

Apart from SVIs that are solely THEORETICAL <strong>based</strong> on in<strong>for</strong>mation BACKGROUND<br />

In contrary, SR is more sensitive to LAI variation in high biomass vegetation as indicated in Figure 3-11<br />

in <strong>the</strong> red and NIR wavelengths, many studies<br />

LAI variation in (assuming high biomass that vegetation high LAI corresponds as indicated in to Figure high NDVI 3-11 values) and shows a higher linearity to biophysical<br />

dealing with <strong>for</strong>est environments have also included<br />

ds to high NDVI variables values) (CHEN and shows 1996). a higher The same linearity is true to <strong>for</strong> biophysical <strong>the</strong> modified simple ratio (MSR), as it is a function <strong>of</strong> SR (cf.<br />

SWIR in<strong>for</strong>mation in <strong>the</strong>ir analyses. In order to<br />

true <strong>for</strong> <strong>the</strong> modified Table simple 3-5). Yet ratio MSR (MSR), increases as it is more a function slowly with <strong>of</strong> SR NDVI. (cf.<br />

create nation-wide LAI maps <strong>for</strong> Canada, Chen et al<br />

slowly with NDVI.<br />

(2002) Apart e.g. from applied SVIs a SWIR that are modification solely <strong>based</strong> <strong>of</strong> on <strong>the</strong> in<strong>for</strong>mation SR in <strong>the</strong> red and NIR wavelengths, many studies<br />

sed on in<strong>for</strong>mation called dealing <strong>the</strong> in reduced <strong>the</strong> with red <strong>for</strong>est simple and environments NIR ratio wavelengths, (RSR). have According also many included studies to SWIR in<strong>for</strong>mation in <strong>the</strong>ir analyses. In order to create<br />

ve also included <strong>the</strong>m, SWIR nation-wide RSR in<strong>for</strong>mation improves LAI LAI maps in <strong>the</strong>ir derivation <strong>for</strong> analyses. Canada, <strong>for</strong> In mixed CHEN order cover et to al create (2002) e.g. applied a SWIR modification <strong>of</strong> <strong>the</strong> SR called<br />

CHEN et al (2002) types <strong>the</strong> and e.g. reduced suppresses applied simple a SWIR background ratio modification (RSR). in<strong>for</strong>mation, According <strong>of</strong> <strong>the</strong> as to SR e.g. <strong>the</strong>m, called RSR improves LAI derivation <strong>for</strong> mixed cover types<br />

ording to <strong>the</strong>m, soil, RSR and as <strong>the</strong> suppresses improves SWIR LAI is background most derivation sensitive in<strong>for</strong>mation, <strong>for</strong> to mixed <strong>the</strong> amount cover as e.g. types soil, as <strong>the</strong> SWIR is most sensitive to <strong>the</strong> amount <strong>of</strong><br />

ation, as e.g. <strong>of</strong> soil, vegetation as <strong>the</strong> SWIR containing is most liquid sensitive water. water. to swir <strong>the</strong> max amount and swir <strong>of</strong> min (cf. Table3-5) are directly SR derived from <strong>the</strong> study<br />

swir max and swirarea<br />

min (cf. and Table3-5) represent are SWIR are directly reflectance derived <strong>of</strong> from very <strong>the</strong> open <strong>the</strong> study and Figure very 3-11 dense Relationship canopies <strong>of</strong> respectively. NDVI versus SR These derived findings from are<br />

<strong>of</strong> very open study and in very area line dense and with represent canopies HORLER respectively. SWIR & AHERN reflectance (1986), These <strong>of</strong> findings who very reported are that SWIR ASTER bands data <strong>for</strong> contain <strong>the</strong> study more area <strong>of</strong> in<strong>for</strong>mation Budongo Forest. on<br />

986), who reported <strong>for</strong>est that structure SWIR in bands Western contain Canadian more conifer in<strong>for</strong>mation and hardwood on <strong>for</strong>ests than o<strong>the</strong>rs.<br />

conifer and hardwood <strong>for</strong>ests than o<strong>the</strong>rs.<br />

SWIR data is also included in <strong>the</strong> normalized difference moisture index (NDMI), which is basically an<br />

normalized difference NDVI moisture including index SWIR (NDMI), in<strong>for</strong>mation which is instead basically <strong>of</strong> an red reflectance (sometimes also called normalized<br />

on instead <strong>of</strong> red difference reflectance water (sometimes index or infrared also called index). normalized HARDISKY et al. (1983) reported that NDMI was highly<br />

index). HARDISKY correlated et al. to (1983) canopy reported water content that NDMI and tracked was highly changes in plant biomass better than <strong>the</strong> NDVI.<br />

NDVI


3 Theoretical background<br />

Kalácska et al. (2004) report a saturation <strong>of</strong> <strong>the</strong>se<br />

SVIs above LAI <strong>of</strong> 4.7. In <strong>the</strong>ir case this affects all<br />

late <strong>for</strong>est stages, which have LAI values <strong>of</strong> 4.9-<br />

8.9. For boreal <strong>for</strong>ests however, sensitive relations<br />

between SVIs and LAI were found at much higher<br />

LAI values (see Schlerf et al. 2005 <strong>for</strong> an overview).<br />

This could possibly be attributed to <strong>the</strong> higher<br />

clumping <strong>of</strong> coniferous <strong>for</strong>ests and thus a higher<br />

sensitivity <strong>of</strong> SVIs to high LAI values (Stenberg<br />

et al. 2004). Curran et al. (1992) e.g. showed that<br />

NDVI and LAI were positively correlated <strong>for</strong> North<br />

American pine <strong>for</strong>ests with LAI values between 1.84<br />

and 9.24 (R2 =0.35-0.86 <strong>for</strong> different dates). O<strong>the</strong>r<br />

studies producing similar results were e.g. undertaken<br />

by Spanner (1990a and b) <strong>for</strong> various conifers (LAI<br />

<strong>of</strong> 1-16).<br />

However, <strong>for</strong> broadleaf <strong>for</strong>est stands in Wisconsin<br />

(LAI <strong>of</strong> 4.4-8.4) Fassnacht et al. (1997) found that<br />

after stratification, regressions between LAI and SVIs<br />

Table 3-5<br />

Overview <strong>of</strong> most important SVIs with respect to LAI derivation <strong>for</strong> tropical rain <strong>for</strong>ests.<br />

49<br />

with R2 between 0.60 and 0.98 could be achieved.<br />

A study on mixed <strong>for</strong>est types (aspen, pine and<br />

spruce) by Brown et al. (2000) showed that RSR had<br />

a linear relationship to LAI up to around 8. Here, no<br />

saturation occurred. The respective R2 (0.55) fur<strong>the</strong>r<br />

increased if only data from coniferous stands were<br />

analysed (R2 =0.66). Brown et al. (2000) attribute <strong>the</strong><br />

good per<strong>for</strong>mance <strong>of</strong> RSR to <strong>the</strong> greater sensitivity <strong>of</strong><br />

<strong>the</strong> SWIR band to LAI.<br />

Generally it must be noted that SVIs have only a<br />

limited robustness, especially when transferred<br />

to o<strong>the</strong>r sites. They are sensitive to atmospheric<br />

conditions and viewing geometry (anisotropic<br />

behaviour <strong>of</strong> vegetation). For tropical rain <strong>for</strong>est<br />

environments a fur<strong>the</strong>r problem can be <strong>the</strong> saturation<br />

<strong>of</strong> indices.<br />

25<br />

SVI 25 Definition<br />

THEORETICAL BACKGROUND<br />

THEORETICAL<br />

Reference<br />

BACKGROUND<br />

25<br />

25 SVI<br />

Simple 25 SVI Simple ratio ratio (SR) (SR)<br />

Simple SVI ratio (SR)<br />

SVI<br />

SVI Simple ratio (SR)<br />

Normalized Normalized Simple Normalized ratio difference (SR) difference vegetation vegetation index (NDVI) index<br />

Simple<br />

Normalized (NDVI) ratio (SR)<br />

(NDVI) difference vegetation index<br />

(NDVI) Corrected Normalized<br />

Corrected NDVI<br />

difference (NDVIc)<br />

vegetation index<br />

Corrected NDVI (NDVI )<br />

Corrected (NDVI)<br />

Normalized difference vegetation index<br />

Normalized c<br />

(NDVI) NDVI difference (NDVIc) vegetation index<br />

(NDVI) Corrected NDVI (NDVIc)<br />

Modified Corrected Modified simple NDVI ratio (NDVIc) (MSR)<br />

Corrected Modified simple NDVI ratio (NDVIc) (MSR)<br />

Modified Modified simple simple ratio (MSR) ratio (MSR)<br />

Modified simple ratio (MSR)<br />

Modified simple ratio (MSR)<br />

Reduced simple ratio (RSR)<br />

Reduced simple ratio (RSR)<br />

Reduced Reduced simple ratio (RSR)<br />

Normalized Reduced simple simple ratio (RSR)<br />

difference ratio (RSR)<br />

Reduced Normalized simple difference ratio (RSR) moisture index<br />

Normalized (NDMI) difference moisture index<br />

(NDMI) Normalized difference moisture index<br />

(NDMI)<br />

Normalized difference moisture index<br />

Normalized Normalized (NDMI) difference difference moisture index moisture (NDMI) index<br />

(NDMI)<br />

Definition<br />

THEORETICAL Reference<br />

BACKGROUND<br />

THEORETICAL BACKGROUND<br />

Definition <br />

THEORETICAL Reference BIRTH & MCVEY Birth & McVey (1968)<br />

NIR<br />

BIRTH & MCVEY BACKGROUND (1968)<br />

NIR<br />

Definition <br />

BIRTH<br />

<br />

Reference & MCVEY (1968) (1968)<br />

Definition NIR<br />

red<br />

Reference<br />

Definition <br />

<br />

NIR<br />

Reference BIRTH & MCVEY (1968)<br />

NIR<br />

red <br />

ROUSE et al. (1974)<br />

NIR <br />

BIRTH red<br />

ROUSE & et MCVEY al. (1974)<br />

red<br />

Rouse et (1968)<br />

NIR<br />

al.<br />

<br />

<br />

BIRTH & MCVEY<br />

NIR <br />

<br />

NIR <br />

ROUSE et<br />

red<br />

(1974) al. (1974)<br />

(1968)<br />

NIR red red<br />

red red<br />

<br />

NIR red <br />

<br />

<br />

NIR <br />

red<br />

SWIR NEMANI ROUSE et red<br />

et al. (1993)<br />

SWIR min<br />

<br />

<br />

<br />

1* <br />

<br />

<br />

<br />

<br />

NIR <br />

red<br />

SWIR NEMANI et<br />

al.<br />

al.<br />

(1974)<br />

red<br />

<br />

(1993)<br />

NIR <br />

SWIR min ROUSE et<br />

NIR <br />

red <br />

<br />

1*<br />

<br />

SWIR max <br />

SWIR min <br />

<br />

<br />

Nemani al. (1974)<br />

red<br />

et al.<br />

NIR <br />

NIR <br />

<br />

red SWIR max <br />

<br />

NIR <br />

<br />

NIR red<br />

red<br />

SWIR ROUSE NEMANI et et al. al. (1974)<br />

red<br />

(1993)<br />

SWIR min<br />

NIR <br />

red<br />

SWIR min<br />

<br />

1*<br />

<br />

<br />

<br />

(1993)<br />

NIR <br />

red <br />

<br />

SWIR max <br />

SWIR min <br />

<br />

NIR <br />

red<br />

SWIR NEMANI et al. (1993)<br />

NIR <br />

red<br />

SWIR min CHEN (1996)<br />

NIR<br />

<br />

<br />

1 <br />

<br />

1* <br />

<br />

<br />

NIR <br />

red<br />

SWIR NEMANI et al. (1993)<br />

SWIR min<br />

NIR <br />

red <br />

CHEN (1996)<br />

<br />

1*<br />

SWIR max <br />

SWIR min <br />

<br />

<br />

NIR <br />

red<br />

SWIR NEMANI et al. (1993)<br />

SWIR min<br />

NIR NIR<br />

red <br />

red1<br />

<br />

<br />

1* SWIR max <br />

SWIR min <br />

<br />

NIR CHEN (1996)<br />

NIR<br />

red SWIR max <br />

SWIR min <br />

Chen<br />

<br />

red<br />

NIR 1<br />

CHEN (1996)<br />

NIR<br />

NIR <br />

<br />

11<br />

red 1<br />

CHEN (1996) (1996)<br />

NIR<br />

<br />

NIR<br />

red <br />

1 red<br />

1<br />

red<br />

NIR<br />

red<br />

BROWN<br />

NIR<br />

SWIR et al. (2000)<br />

NIR<br />

SWIR min<br />

<br />

<br />

<br />

1<br />

1* <br />

<br />

<br />

<br />

BROWN<br />

NIR<br />

SWIR et al. (2000)<br />

<br />

SWIR min<br />

red <br />

<br />

red <br />

<br />

<br />

1 NIR<br />

1*<br />

1<br />

SWIR max <br />

<br />

SWIR min <br />

<br />

BROWN<br />

NIR<br />

SWIR <br />

red SWIR max <br />

et al. (2000)<br />

red<br />

SWIR <br />

min<br />

<br />

SWIR min<br />

<br />

1*<br />

<br />

<br />

<br />

red<br />

<br />

<br />

<br />

red NIR<br />

<br />

HARDISKY et al. (1983)<br />

NIR <br />

SWIR max <br />

BROWN<br />

NIR<br />

SWIR et al. (2000)<br />

SWIR<br />

SWIR<br />

min<br />

<br />

min<br />

<br />

<br />

1* <br />

SWIR<br />

HARDISKY et al. (1983)<br />

SWIR<br />

<br />

<br />

BROWN<br />

NIR<br />

SWIR et Brown al. (2000) et al.<br />

SWIR min<br />

<br />

<br />

red<br />

NIR<br />

NIR <br />

<br />

1*<br />

<br />

<br />

NIR <br />

SWIR max <br />

<br />

SWIR min <br />

BROWN (2000)<br />

NIR<br />

SWIR et al. (2000)<br />

SWIR min<br />

<br />

HARDISKY et al. (1983)<br />

red <br />

SWIR <br />

1*<br />

<br />

SWIR<br />

SWIR max <br />

SWIR min <br />

<br />

<br />

<br />

NIR<br />

NIR<br />

red <br />

<br />

HARDISKY <br />

<br />

et al. (1983)<br />

SWIR<br />

SWIR<br />

SWIR max <br />

SWIR min <br />

NIR<br />

<br />

HARDISKY et al. (1983)<br />

SWIR<br />

NIR<br />

NIR <br />

HARDISKY<br />

SWIR<br />

Hardisky et al. (1983) et al.<br />

SWIR<br />

NIR <br />

SWIR<br />

NIR <br />

(1983)<br />

SWIR<br />

Generally it must be noted that SVIs have only a limited robustness, especially when transferred to o<strong>the</strong>r<br />

Generally sites. They it are must sensitive be noted to that atmospheric SVIs have conditions only a limited and robustness, viewing geometry especially (anisotropic when transferred behaviour to o<strong>the</strong>r<br />

sites. They are sensitive to atmospheric conditions and viewing geometry (anisotropic behaviour <strong>of</strong><br />

sites. Generally<br />

vegetation). They it<br />

vegetation). For are must<br />

For tropical sensitive be noted<br />

tropical rain to that<br />

rain <strong>for</strong>est atmospheric SVIs have<br />

<strong>for</strong>est environments conditions only a limited<br />

environments a fur<strong>the</strong>r and robustness,<br />

fur<strong>the</strong>r problem viewing<br />

problem can geometry especially<br />

can be <strong>the</strong> saturation (anisotropic when transferred<br />

saturation <strong>of</strong> indices. behaviour to o<strong>the</strong>r<br />

Generally it must be noted that SVIs have only a limited robustness, especially when transferred <strong>of</strong><br />

indices. to o<strong>the</strong>r<br />

Generally<br />

vegetation). sites. They it must<br />

For are tropical sensitive be noted<br />

rain to that<br />

<strong>for</strong>est atmospheric SVIs have<br />

environments conditions only a limited<br />

a fur<strong>the</strong>r and robustness,<br />

problem viewing can geometry especially<br />

be <strong>the</strong> saturation (anisotropic when transferred<br />

<strong>of</strong> indices. behaviour to o<strong>the</strong>r <strong>of</strong><br />

sites. They are sensitive to atmospheric conditions and viewing geometry (anisotropic behaviour <strong>of</strong><br />

sites. vegetation). They For are tropical sensitive rain to <strong>for</strong>est atmospheric environments conditions a fur<strong>the</strong>r and problem viewing can geometry be <strong>the</strong> saturation (anisotropic <strong>of</strong> indices. behaviour <strong>of</strong><br />

vegetation). For tropical rain <strong>for</strong>est environments a fur<strong>the</strong>r problem can be <strong>the</strong> saturation <strong>of</strong> indices.<br />

vegetation). For tropical rain <strong>for</strong>est environments a fur<strong>the</strong>r problem can be <strong>the</strong> saturation <strong>of</strong> indices.<br />

Influence <strong>of</strong> structural characteristics<br />

Influence <strong>of</strong> structural characteristics


50<br />

Influence <strong>of</strong> structural characteristics<br />

Schlerf et al. (2005) state that in coniferous<br />

<strong>for</strong>est plantations <strong>the</strong> relationship between LAI<br />

and vegetation indices can be disturbed as o<strong>the</strong>r<br />

biophysical stand characteristics may influence <strong>the</strong><br />

reflectance signal. Especially differences in canopy<br />

structure and stand density can affect <strong>the</strong> reflectance<br />

signal. Canopy structure, which is dependent on<br />

canopy depth, leaf area index, leaf orientation and leaf<br />

distribution, becomes more complex as <strong>the</strong> vegetation<br />

amount increases, so that incoming photons are<br />

trapped within <strong>the</strong> canopy causing a decrease in <strong>the</strong><br />

intensity <strong>of</strong> reflected radiation.<br />

The same can be true <strong>for</strong> <strong>the</strong> highly complex<br />

structure <strong>of</strong> tropical primary rain <strong>for</strong>est ecosystems.<br />

Figure 3-12 shows <strong>the</strong> differences in vertical and<br />

horizontal canopy structure between primary <strong>for</strong>est<br />

stands and disturbed secondary <strong>for</strong>ests. Emergent<br />

trees are missing in <strong>the</strong> upper canopy <strong>of</strong> selectively<br />

logged <strong>for</strong>ests causing <strong>the</strong> upper canopy boundary to<br />

be smoo<strong>the</strong>r. In contrast, <strong>the</strong> complexity <strong>of</strong> canopy<br />

structure increases with stand age in undisturbed<br />

<strong>for</strong>ested areas. As illustrated, <strong>the</strong> primary <strong>for</strong>est site<br />

has a very rough upper canopy causing mutual shading<br />

and thus reducing <strong>the</strong> reflectance signal, even though<br />

<strong>the</strong> leaf area might be comparable between both sites<br />

(selectively logged <strong>for</strong>ests tend to have a denser<br />

ground cover, compare Chapter 2.1.2 <strong>for</strong> Budongo<br />

Forest). Especially if satellite data is acquired at low<br />

sun elevation angles, <strong>the</strong> rough canopy surface <strong>of</strong><br />

primary <strong>for</strong>est stands can trap a large proportion <strong>of</strong><br />

solar irradiation. The positive relationship between<br />

SVIs and LAI might thus be disturbed.<br />

Several studies have <strong>the</strong>re<strong>for</strong>e used structural<br />

in<strong>for</strong>mation to improve <strong>the</strong> relation between SVIs and<br />

LAI. According to Seed & King (2003) <strong>the</strong> shadow<br />

fraction remains sensitive to increases in LAI even<br />

after saturation <strong>of</strong> SVIs. Aragão et al. (2005) applied<br />

<strong>for</strong> example spectral unmixing to extract <strong>the</strong> shadow<br />

fraction <strong>for</strong> tropical rain <strong>for</strong>est in Brazil from Landsat<br />

ETM+ imagery. Toge<strong>the</strong>r with SVIs it improved<br />

<strong>the</strong> derivation <strong>of</strong> LAI through <strong>the</strong> application <strong>of</strong> a<br />

multiple linear regression model.<br />

Apart from spectral unmixing, <strong>the</strong> shadow component<br />

<strong>of</strong> vegetation stands can also be derived indirectly<br />

from statistical measures such as variance, standard<br />

deviation, and mean brightness over a predefined<br />

region (see e.g. Leboeuf et al. 2007). Here, image<br />

texture is analysed, i.e. <strong>the</strong> spatial variation in surface<br />

reflectance. As <strong>the</strong> spatial variability <strong>of</strong> pixel values is<br />

related to <strong>the</strong> spatial distribution <strong>of</strong> <strong>for</strong>est vegetation<br />

elements, <strong>for</strong>est structure is accordingly represented<br />

by texture (Wulder et al. 1998). Texture measures<br />

are usually statistics that are calculated <strong>for</strong> <strong>the</strong><br />

central pixel within a moving window <strong>of</strong> fixed size.<br />

The so-called first-order texture measures calculate<br />

a representative statistical value <strong>for</strong> <strong>the</strong> respective<br />

central cell <strong>of</strong> that window and are <strong>based</strong> on <strong>the</strong><br />

original image values. Second-order measures in<br />

turn consider <strong>the</strong> relationship between neighbouring<br />

windows (Hall-Beyer 2007, Wulder et al. 1998).<br />

There<strong>for</strong>e a grey level co-occurrence matrix (GLCM)<br />

has to be calculated, which according to Wulder et al.<br />

(1998), p. 66, and after Haralick et al. (1973) can be<br />

defined as a “matrix <strong>of</strong> relative frequencies <strong>for</strong> which<br />

two neighbouring pixels that are separated by a userdefined<br />

distance and angle can occur in <strong>the</strong> image.”<br />

Statistical measures are <strong>the</strong>n applied to <strong>the</strong> GLCM.<br />

These can include homogeneity, contrast and entropy,<br />

which characterize <strong>the</strong> specific textural characteristics<br />

<strong>of</strong> <strong>the</strong> image, or diversity, mean, variation and<br />

standard deviation that relate to <strong>the</strong> complexity <strong>of</strong><br />

grey level transitions (Wulder et al. 1998).<br />

Texture indices were successfully applied to a study<br />

site in nor<strong>the</strong>rn Italy by Colombo et al. (2003). They<br />

found that <strong>the</strong> inclusion <strong>of</strong> texture indices strongly<br />

increased <strong>the</strong> linear relationship between NDVI<br />

derived from IKONOS data and LAI (increase in<br />

R2 from 0.19 to 0.73, mean LAI values from 3 to 6).


3 Theoretical background<br />

Also Wulder et al. (1998) found that image texture<br />

provides a quantification <strong>of</strong> structural variability <strong>of</strong><br />

mixed <strong>for</strong>est stands. If texture in<strong>for</strong>mation was added,<br />

<strong>the</strong> prediction <strong>of</strong> LAI from NDVI was significantly<br />

improved. The inclusion <strong>of</strong> texture measures,<br />

however, requires that <strong>the</strong> image from which texture<br />

is calculated has an adequate spatial resolution to<br />

represent canopy objects. Often, very high resolution<br />

imagery such as IKONOS or QuickBird is used<br />

(Colombo et al. 2003, Leboeuf et al. 2007). An<br />

alternative are airborne sensors (Seed & King 2003,<br />

Wulder et al. 1998).<br />

3.4.5<br />

Figure 3-12<br />

<strong>Validation</strong> approaches<br />

When focusing on <strong>the</strong> validation <strong>of</strong> biophysical<br />

properties derived from satellite data, in situ<br />

measurements <strong>for</strong>m an independent reference data set.<br />

As described in Chapter 3.3.4 <strong>the</strong> sampling scheme<br />

is <strong>of</strong> major importance. When dealing with satellite<strong>based</strong><br />

LAI at very high to high spatial resolution, in<br />

situ sampling units can be scaled correspondingly<br />

(cf. Equation 3.13) and field data can be compared<br />

directly to satellite-derived LAI.<br />

However, when biophysical variables derived from<br />

medium resolution sensors as <strong>MODIS</strong> are to be<br />

validated, an upscaling <strong>of</strong> in situ measurements to<br />

high spatial resolution satellite data is needed as an<br />

intermediate step (cf. Figure 3-13). Here a direct<br />

comparison between in situ data and <strong>the</strong> <strong>MODIS</strong><br />

output is not feasible due to possible geolocation<br />

Structural differences in primary rain <strong>for</strong>est and selectively logged <strong>for</strong>est (NIES 2007, modified).<br />

51<br />

errors and surface heterogeneity at <strong>MODIS</strong> resolution<br />

(Liang et al. 2002). Upscaling can be achieved by<br />

<strong>the</strong> methods mentioned in Chapters 3.4.1 and 3.4.2,<br />

i.e. ei<strong>the</strong>r empirical approaches or physically-<strong>based</strong><br />

models. The latter is, however, not recommended by<br />

VALERI (Baret et al. submitted), as <strong>the</strong> majority <strong>of</strong><br />

operational algorithms <strong>for</strong> estimating LAI employ<br />

a radiative transfer model. This means that circular<br />

dependencies with <strong>the</strong> LAI product to be validated<br />

might occur.<br />

For <strong>the</strong> validation purpose in this <strong>the</strong>sis, empirical<br />

methods are sufficient as only single time steps are<br />

needed <strong>for</strong> validation. Additionally, in contrast to<br />

physical models that are limited by saturation <strong>of</strong> nadir<br />

spectral measurements unless a priori in<strong>for</strong>mation is<br />

applied (Knyazikhin et al. 1998a), empirical models<br />

are only constrained by <strong>the</strong> sampling distribution.<br />

In consequence, ecophysiological limits on LAI are<br />

incorporated to make predictions past <strong>the</strong> saturation<br />

point <strong>of</strong> reflectance data (Fernandes et al. 2005).<br />

Within <strong>the</strong> framework <strong>of</strong> CEOS-LPV, LAI product<br />

validation has been per<strong>for</strong>med by various academic<br />

groups and research networks, as summarized later<br />

in Table 4-16). Although each <strong>of</strong> <strong>the</strong> groups applied<br />

slightly different methodologies <strong>for</strong> validation, e.g. in<br />

terms <strong>of</strong> instruments chosen <strong>for</strong> in situ LAI estimation<br />

or in terms <strong>of</strong> <strong>the</strong> upscaling procedure used, <strong>the</strong>re are<br />

common guidelines on how to validate medium scale<br />

resolution satellite data <strong>based</strong> on field measurements.<br />

They will be described in <strong>the</strong> following and is <strong>based</strong><br />

mainly on <strong>the</strong> syn<strong>the</strong>sis publications <strong>of</strong> Yang et al.


52<br />

(2006), Morisette et al. (2006a), and Baret et al.<br />

(submitted).<br />

First <strong>of</strong> all, <strong>the</strong> spatial distribution <strong>of</strong> predominant<br />

land cover types on <strong>the</strong> test site is to be assessed. LAI<br />

measurements must be taken in several homogeneous<br />

patches within each land cover type in order to sample<br />

<strong>the</strong> natural range <strong>of</strong> LAI. In situ measurements<br />

should be as accurate as possible and <strong>the</strong> precision<br />

<strong>of</strong> <strong>the</strong> measurements should be known. Usually high<br />

resolution broadband satellite data, e.g. derived from<br />

SPOT, ASTER or ETM+, is used <strong>for</strong> <strong>the</strong> generation<br />

<strong>of</strong> high spatial resolution LAI maps. If available, also<br />

very high resolution optical, hyperspectral, or radar<br />

data could be used. Based on site-specific relationships<br />

between high resolution satellite data and ground<br />

measurements, a so-called transfer function needs to<br />

be established. This transfer function links both data<br />

sets <strong>based</strong> on empirical methods and ideally results<br />

in high resolution LAI maps covering <strong>the</strong> whole test<br />

site.<br />

The next step is <strong>the</strong> aggregation <strong>of</strong> <strong>the</strong> high resolution<br />

LAI map to 1 km <strong>MODIS</strong> resolution. Usually<br />

<strong>the</strong> fine resolution LAI map is averaged and <strong>the</strong>n<br />

Figure 3-13<br />

Schematic illustration <strong>of</strong> validation strategy (Yang et al. 2006, modified).<br />

aggregated. According to Yang et al. (2006) a pixelby-pixel<br />

comparison is not recommended due to e.g.<br />

geolocation uncertainties and pixel-shift errors caused<br />

by <strong>the</strong> point spread function. Fur<strong>the</strong>r <strong>the</strong> <strong>MODIS</strong> LAI<br />

algorithm is not designed to retrieve a deterministic<br />

value, but instead generates a mean LAI value from<br />

all possible solutions within a specified level <strong>of</strong> input<br />

satellite data and model uncertainties. Comparison<br />

should thus be ra<strong>the</strong>r per<strong>for</strong>med at multipixel (patch)<br />

scale, where LAI is statistically stable.<br />

According to Morisette et al. (2006a) global validation<br />

<strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product is <strong>the</strong> final stage <strong>of</strong><br />

validation exercises. The prerequisite is, however, <strong>the</strong><br />

availability <strong>of</strong> data coming from various validation<br />

sites that represent <strong>the</strong> global variability and range<br />

<strong>of</strong> LAI and canopy types. Based on <strong>the</strong>se global<br />

validation ef<strong>for</strong>ts but also on in<strong>for</strong>mation coming<br />

from individual validation initiatives, an algorithm<br />

refinement is also envisaged. Within CEOS-LPV,<br />

<strong>the</strong> VALERI network, especially, has put some<br />

ef<strong>for</strong>t into <strong>the</strong> development <strong>of</strong> adequate sampling<br />

strategies <strong>for</strong> medium resolution satellite biophysical<br />

product validation (Baret et al. submitted). VALERI<br />

recommends test sites <strong>of</strong> at least 3 km x 3 km, with


3 Theoretical background<br />

larger sizes being desirable <strong>for</strong> coarser resolution<br />

sensors as e.g. POLDER (Baret et al. submitted).<br />

Local ground measurements are per<strong>for</strong>med in each<br />

<strong>of</strong> <strong>the</strong> nine 1 km2 subsquares (cf. Figure 3-14) in<br />

so-called elementary sampling units (ESUs). ESUs<br />

ideally correspond approximately to <strong>the</strong> pixel size<br />

<strong>of</strong> a high spatial resolution satellite image, which<br />

can be later used <strong>for</strong> upscaling. Three to five equally<br />

spread ESUs should at least be sampled per 1 km2 .<br />

They should be located far enough from landscape<br />

boundaries and represent all cover types <strong>of</strong> <strong>the</strong> entire<br />

test site.<br />

The sampling strategy proposed by VALERI was<br />

followed in this <strong>the</strong>sis as far as possible. However,<br />

<strong>for</strong> several reasons a suite <strong>of</strong> modifications had to be<br />

applied to <strong>the</strong> VALERI scheme when transferring it<br />

to <strong>the</strong> tropical rain <strong>for</strong>est environment. These will be<br />

discussed in Chapter 5.1.1.<br />

Figure 3-14<br />

VALERI validation scheme <strong>of</strong> medium resolution<br />

satellite data (Baret et al. submitted).<br />

53


4<br />

Remote sensing data<br />

and preprocessing<br />

In <strong>the</strong> tropics <strong>the</strong> main constraint on <strong>the</strong> availability<br />

<strong>of</strong> optical satellite imagery is cloud cover. Although<br />

radar data acquisition is independent <strong>of</strong> clouds,<br />

it was not available <strong>for</strong> <strong>the</strong> study sites. Hence a<br />

suite <strong>of</strong> optical high resolution satellite imagery<br />

was analysed. Recent Landsat ETM+ scenes made<br />

available to <strong>the</strong> BIOTA East Africa project were<br />

utilized <strong>for</strong> visual assessments prior to fieldwork.<br />

However, as <strong>the</strong> sensor’s scan line corrector failed<br />

in 2003, Landsat data was not taken into fur<strong>the</strong>r<br />

account <strong>for</strong> <strong>the</strong> derivation <strong>of</strong> LAI. Instead <strong>the</strong><br />

OASIS (Optimising Access to Spot Infrastructure <strong>for</strong><br />

Science) programme financed by <strong>the</strong> EC provided<br />

two SPOT-4 scenes <strong>for</strong> Budongo and Kakamega<br />

Table 4-1<br />

55<br />

Forests. The cooperation with CEOS-LPV led to <strong>the</strong><br />

fur<strong>the</strong>r possibility <strong>of</strong> analysing level 2 ASTER data<br />

<strong>for</strong> Budongo Forest (courtesy <strong>of</strong> Jaime Nickeson and<br />

Jeff Morisette from Goddard Space Flight Center,<br />

GSFC). In <strong>the</strong> following <strong>the</strong> available data sets <strong>for</strong><br />

this study will be briefly described with respect to <strong>the</strong><br />

sensor characteristics and preprocessing steps.<br />

4.1<br />

4.1.1<br />

Available high resolution<br />

satellite data<br />

ASTER<br />

The Advanced Spaceborne Thermal Emission and<br />

Reflection radiometer (ASTER) was built in a<br />

joint ef<strong>for</strong>t between NASA and Japan’s Ministry <strong>of</strong><br />

Economy, Trade and Industry and is orbiting aboard<br />

<strong>the</strong> Earth Observing System (EOS) plat<strong>for</strong>m Terra.<br />

Launched in December 1999, Terra is in a sunsynchronous<br />

near-polar orbit with a local equator<br />

crossing time <strong>of</strong> 10:30 am. The ASTER instrument<br />

consists <strong>of</strong> three different subsystems (VNIR, SWIR<br />

and TIR) and has a swath width <strong>of</strong> 60 km. Table 4-1<br />

shows ASTER’s spatial, spectral and radiometric<br />

characteristics <strong>for</strong> <strong>the</strong> two subsystems used in this<br />

<strong>the</strong>sis. VNIR consists <strong>of</strong> two telescopes, one with a<br />

nadir-viewing three-spectral band detector (visible to<br />

Spectral and spatial properties <strong>of</strong> <strong>the</strong> ASTER VNIR and SWIR instruments (Abrams & Hook 2002).<br />

Band/System Wavelength [µm] Spatial resolution at nadir [m] Quantization [bit]<br />

(1) VNIR (green) 0.52 - 0.60 15 8<br />

(2) VNIR (red) 0.63 - 0.69 15 8<br />

(3a) VNIR nadir looking 0.76 - 0.86 15 8<br />

(3b) VNIR backward 0.76 - 0.86 15 8<br />

(4) SWIR 1.60 - 1.70 30 8<br />

(5) SWIR 2.145 - 2.185 30 8<br />

(6) SWIR 2.185 - 2.225 30 8<br />

(7) SWIR 2.235 - 2.285 30 8<br />

(8) SWIR 2.295 - 2.365 30 8<br />

(9) SWIR 2.360 - 2.430 30 8


56<br />

near-infrared wavelengths) and <strong>the</strong> o<strong>the</strong>r backward<br />

looking in <strong>the</strong> near-infrared range. All VNIR<br />

telescopes can rotate 24° across track. The SWIR<br />

instrument consists <strong>of</strong> a nadir-looking telescope that<br />

operates in six spectral bands. With <strong>the</strong> help <strong>of</strong> a<br />

pointing mirror, imagery can be acquired 8.55° across<br />

track. ASTER data is only acquired on demand. For<br />

more detailed in<strong>for</strong>mation please refer to Abrams &<br />

Hook (2002).<br />

Compared to <strong>the</strong> o<strong>the</strong>r high resolution satellite data<br />

used in this study, ASTER has <strong>the</strong> highest spectral<br />

resolution with nine bands in VIS to SWIR and <strong>the</strong><br />

highest spatial resolution in <strong>the</strong> VIS/NIR with 15 m.<br />

Un<strong>for</strong>tunately ASTER data was only available <strong>for</strong><br />

Budongo Forest exactly one year after fieldwork.<br />

However, <strong>the</strong> fact that imagery was also acquired<br />

at <strong>the</strong> end <strong>of</strong> <strong>the</strong> rainy season justifies <strong>the</strong> analysis<br />

<strong>of</strong> this data set. Acquisition details <strong>for</strong> <strong>the</strong> respective<br />

scenes are shown in Table 4-4.<br />

4.1.2<br />

SPOT-HRVIR<br />

Launched in March 1998, <strong>the</strong> fourth member <strong>of</strong><br />

<strong>the</strong> Satellite Pour l’Observation de la Terre (SPOT)<br />

family is one <strong>of</strong> <strong>the</strong> three SPOT satellites that are still<br />

in operation. It is in a sun-synchronous near-polar<br />

orbit at an altitude <strong>of</strong> 822 km with an inclination <strong>of</strong><br />

98° and a repetition time <strong>of</strong> 26 days <strong>for</strong> <strong>the</strong> same<br />

ground track. The local equator crossing time on<br />

a descending node is around 10:30 am. Designed<br />

and developed by CNES, SPOT 4 carries a linear<br />

pushbroom radiometer, <strong>the</strong> HRVIR instrument.<br />

Table 4-2<br />

It features a pointable optics and <strong>of</strong>fers <strong>of</strong>f nadir<br />

viewing up to 27° (Henry & Meygret 2001). The<br />

spatial, spectral and radiometric properties <strong>of</strong> SPOT-<br />

4 are introduced in Table 4-2. Table 4-4 provides <strong>the</strong><br />

acquisition details <strong>for</strong> <strong>the</strong> study sites.<br />

Preprocessing <strong>of</strong> high resolution<br />

satellite data<br />

Band Wavelength [µm] Spatial resolution at nadir [m] Quantization [bit]<br />

(1) green 0.50 - 0.59 20 8<br />

(2) red 0.61 - 0.68 20 8<br />

(3) NIR 0.78 - 0.89 20 8<br />

(4) SWIR 1.58 - 1.75 20 8<br />

4.2<br />

The preprocessing <strong>of</strong> satellite data aims to take into<br />

account <strong>the</strong> spatial and spectral comparability <strong>of</strong><br />

different data sources. Geometric correction is <strong>the</strong><br />

prerequisite <strong>for</strong> a precise co-registration <strong>of</strong> multiple<br />

satellite scenes. Although atmospheric correction is<br />

not necessarily needed to apply empirical models,<br />

Turner et al. (1999) found a significant improvement<br />

in <strong>the</strong> relationship between in situ measured LAI and<br />

SVIs after conversion <strong>of</strong> DNs to surface reflectance<br />

<strong>for</strong> three sites in temperate zones including a broadleaf<br />

<strong>for</strong>est. Effects <strong>of</strong> topographic corrections were in<br />

turn less evident. Consequently in this <strong>the</strong>sis <strong>the</strong><br />

preprocessing steps <strong>of</strong> high resolution satellite scenes<br />

comprised geometric and atmospheric correction.<br />

4.2.1<br />

Spectral and spatial properties <strong>of</strong> SPOT-4 HRVIR (Henry & Meygret 2001).<br />

Geometric correction<br />

According to Richards & Jia (2006) <strong>the</strong>re are many<br />

sources <strong>for</strong> geometric distortion, both systematic<br />

and non-systematic. While systematic distortions are<br />

mainly caused by <strong>the</strong> sensor’s <strong>for</strong>ward movement<br />

(scan skew), its scanning system (mirror velocity) or<br />

<strong>the</strong> earth curvature (panoramic effect), nonsystematic


4.2.1 Geometric correction<br />

According to RICHARDS & JIA (2006) <strong>the</strong>re are many sources <strong>for</strong> geometric distortion, both systematic<br />

4 Remote sensing data and preprocessing<br />

and non-systematic. While systematic distortions are mainly caused by <strong>the</strong> sensor’s <strong>for</strong>ward movement<br />

(scan skew), its scanning system (mirror velocity) or <strong>the</strong> earth curvature (panoramic effect), nonsystematic<br />

distortions can be related to variations in altitude, attitude and velocity <strong>of</strong> <strong>the</strong> satellite plat<strong>for</strong>m and terrain-<br />

distortions can be related to variations in altitude, 4.2.2 Atmospheric correction<br />

induced effects. Systematic distortion is usually corrected <strong>for</strong> level 1 imagery by <strong>the</strong> data provider, but<br />

attitude and velocity <strong>of</strong> <strong>the</strong> satellite plat<strong>for</strong>m and<br />

non-systematic distortions still need to be rectified by <strong>the</strong> user. In this <strong>the</strong>sis topographic map sheets<br />

terrain-induced effects. Systematic distortion is Radiance reaching <strong>the</strong> sensor is usually not only<br />

1:50,000 <strong>of</strong> both study sites (made available by <strong>the</strong> BIOTA East project) were used as reference data <strong>for</strong><br />

usually corrected <strong>for</strong> level 1 imagery by <strong>the</strong> data influenced by <strong>the</strong> properties <strong>of</strong> <strong>the</strong> reflecting ground<br />

geometric correction (cf. Table 4-3).<br />

provider, but non-systematic distortions still need to surface in <strong>the</strong> sensor’s instantaneous field <strong>of</strong> view<br />

be rectified by <strong>the</strong> user. In this <strong>the</strong>sis topographic map<br />

Table 4-3: Reference data <strong>for</strong> geometric correction<br />

(IFOV), but also by relief induced topographic<br />

sheets 1:50,000 <strong>of</strong> both study sites (made available illumination effects as well as atmospheric scattering<br />

by <strong>the</strong> BIOTA REFERENCE East project) were used COVERAGE as reference SCALE and absorption. YEAR Especially SOURCE <strong>for</strong> multitemporal<br />

data <strong>for</strong><br />

Topographic<br />

geometric correction<br />

Map Sheets<br />

(cf. Table 4-3).<br />

Budongo<br />

30/III, 30/IV, 39/I, 39/II, 38/IV, 39/III Forest<br />

applications and<br />

East<br />

<strong>the</strong><br />

Africa<br />

derivation<br />

1:50 000<br />

<strong>of</strong><br />

(Uganda),<br />

quantities<br />

Series<br />

<strong>based</strong> on<br />

1:50,000 1966<br />

physical models, image Y732 calibration (D.O.S. 426), and Edition <strong>the</strong> I correction<br />

Each scene Topographic was georeferenced Map Sheets with 15-20 Kakamega evenly 1:50,000 <strong>of</strong> atmospheric East<br />

1970 effects Africa is 1:50 crucial. 000 (Kenya) Many Series different<br />

102/2 and 102/4<br />

Forest<br />

Y731 (D.O.S. 423), Edition 6<br />

distributed ground control points (GCPs) and radiative transfer codes <strong>for</strong> atmospheric correction<br />

Each assigned scene to was UTM georeferenced projection (Zone with 15-20 36N), evenly WGS 84 distributed exist, ground including control <strong>the</strong> points widely (GCPs) applied and 6S assigned (Second<br />

to datum. UTM Table projection 4-5 indicates (Zone 36N), <strong>the</strong> respective WGS 84 root datum. mean Table Simulation 4-5 indicates <strong>of</strong> <strong>the</strong> respective Satellite root Signal mean in square <strong>the</strong> Solar<br />

error square (RMSE), error (RMSE), which is which calculated is calculated as as<br />

Spectrum; Vermote et al. 1997) code or MODTRAN<br />

(Berk et al. 1998). In this <strong>the</strong>sis an atmospheric<br />

RMSE <br />

n 1 2 2<br />

XRi<br />

YRi<br />

n i1<br />

(4.1) correction with <strong>the</strong> IDL code <strong>of</strong> ATCOR-3 (4.1) was<br />

per<strong>for</strong>med (Richter 1998, 2007).<br />

with n n being <strong>the</strong> <strong>the</strong> number <strong>of</strong> GCPs, <strong>of</strong> GCPs, and and XR and XR YR and <strong>the</strong> residual at GCP i in X and Y direction. An RMSE<br />

<strong>of</strong> YR less <strong>the</strong> than residual one pixel at GCP was i in achieved X and <strong>for</strong> Y direction. all scenes. An It should ATCOR-3 be noted, uses however, an atmospheric that GCPs database were mainly compiled<br />

located RMSE at <strong>of</strong> roads less than and intersections one pixel was outside achieved <strong>the</strong> <strong>for</strong>est <strong>for</strong> all as distinct from features <strong>the</strong> MODTRAN-4 are hard to identify code to in represent <strong>the</strong> <strong>for</strong>ested a wide<br />

areas. scenes. The It should clearing be <strong>of</strong> noted, <strong>the</strong> Sonso however, Camp that Site GCPs in were Budongo range Forest <strong>of</strong> is atmospheric <strong>the</strong> only exception conditions because <strong>based</strong> buildings on various<br />

were mainly already located indicated at roads on and <strong>the</strong> intersections topographic outside maps <strong>of</strong> <strong>the</strong> 1966. combinations It should be mentioned <strong>of</strong> aerosol that types, <strong>the</strong> water RMSE vapour indicates content,<br />

<strong>the</strong> <strong>for</strong>est error as close distinct to features <strong>the</strong> selected are hard GCPs to and identify does in not <strong>the</strong> necessarily ground serve elevation, as an absolute solar zenith quality angles measure and inside visibility<br />

<strong>the</strong> <strong>for</strong>ested <strong>for</strong>ests. areas. Never<strong>the</strong>less The clearing detailed <strong>of</strong> <strong>the</strong> comparisons Sonso Camp <strong>of</strong> Site <strong>the</strong> different (Richter images 2007). and Based <strong>the</strong>ir on good an atmosphere agreement in type overlay selected<br />

suggest in Budongo that <strong>the</strong> Forest remaining is <strong>the</strong> internal only exception variation <strong>of</strong> because <strong>the</strong> scenes by is low. <strong>the</strong> user, an iterative process modifies properties<br />

buildings were already indicated on <strong>the</strong> topographic <strong>of</strong> <strong>the</strong> selected standard aerosol type to match scene<br />

maps <strong>of</strong> 1966. It should be mentioned that <strong>the</strong> RMSE conditions. The so-called dark dense vegetation/dark<br />

4.2.2 indicates Atmospheric <strong>the</strong> error close to correction <strong>the</strong> selected GCPs and does dense object approach uses dark scene areas <strong>of</strong> low<br />

not necessarily serve as an absolute quality measure reflectance to derive <strong>the</strong> scene specific optical depth.<br />

Radiance reaching <strong>the</strong> sensor is usually not only influenced by <strong>the</strong> properties <strong>of</strong> <strong>the</strong> reflecting ground<br />

inside <strong>the</strong> <strong>for</strong>ests. Never<strong>the</strong>less detailed comparisons For details <strong>of</strong> this approach see Song et al. (2001).<br />

surface in <strong>the</strong> sensor’s instantaneous field <strong>of</strong> view (IFOV), but also by relief induced topographic<br />

<strong>of</strong> <strong>the</strong> different images and <strong>the</strong>ir good agreement in<br />

illumination effects as well as atmospheric scattering and absorption. Especially <strong>for</strong> multitemporal<br />

applications<br />

overlay suggest<br />

and<br />

that<br />

<strong>the</strong><br />

<strong>the</strong><br />

derivation<br />

remaining<br />

<strong>of</strong><br />

internal<br />

quantities<br />

variation<br />

<strong>based</strong> on<br />

For<br />

physical<br />

<strong>the</strong> calculation<br />

models,<br />

<strong>of</strong><br />

image<br />

surface<br />

calibration<br />

reflectance<br />

and<br />

ρ,<br />

<strong>the</strong><br />

sensor<br />

correction<br />

<strong>of</strong> <strong>the</strong> scenes<br />

<strong>of</strong><br />

is<br />

atmospheric<br />

low.<br />

effects is crucial. Many different<br />

specific<br />

radiative<br />

calibration<br />

transfer<br />

files<br />

codes<br />

and<br />

<strong>for</strong><br />

a digital<br />

atmospheric<br />

elevation<br />

correction exist, including <strong>the</strong> widely applied 6S (Second Simulation <strong>of</strong> <strong>the</strong> Satellite Signal in <strong>the</strong> Solar<br />

Table 4-3<br />

Reference data <strong>for</strong> geometric correction.<br />

Reference Coverage Scale Year Source<br />

Topographic Map Sheets<br />

30/III, 30/IV, 39/I, 39/II, 38/IV, 39/III<br />

Topographic Map Sheets<br />

102/2 and 102/4<br />

Budongo Forest 1:50,000 1966 East Africa 1:50 000 (Uganda),<br />

Series Y732 (D.O.S. 426), Edition I<br />

Kakamega Forest 1:50,000 1970 East Africa 1:50 000 (Kenya)<br />

Series Y731 (D.O.S. 423), Edition 6<br />

57


58<br />

model (DEM) that allows <strong>the</strong> correction <strong>of</strong> radiation<br />

scattered into <strong>the</strong> satellite’s IFOV from adjacent<br />

terrain (in this case a combined SRTM/ERS DEM<br />

from <strong>the</strong> W42 database at DLR-DFD) are needed<br />

as additional input. ρ can <strong>the</strong>n be approximated<br />

iteratively according to Equation A.1 in <strong>the</strong> Annex.<br />

For fur<strong>the</strong>r in<strong>for</strong>mation on ATCOR and atmospheric<br />

correction <strong>the</strong>ory please refer to Jensen (2007) and<br />

Richter (2007).<br />

All SPOT-4 scenes were corrected with ATCOR-3.<br />

Since ASTER data had only been available as ondemand<br />

level 2 surface reflectance product (USGS<br />

2006) and already contained atmospherically<br />

corrected data <strong>for</strong> both <strong>the</strong> VNIR and SWIR bands,<br />

ATCOR was not applied. The ASTER SWIR data<br />

had been fur<strong>the</strong>r corrected automatically <strong>for</strong> <strong>the</strong><br />

crosstalk phenomenon <strong>based</strong> on statistically derived<br />

radiometric sensitivity coefficients (USGS 2006).<br />

Acquisition and preprocessing details <strong>of</strong> <strong>the</strong> high<br />

resolution satellite scenes that were used in this <strong>the</strong>sis<br />

are summarized in Table 4-4.<br />

Table 4-4<br />

4.3<br />

Acquisition and preprocessing in<strong>for</strong>mation <strong>for</strong> <strong>the</strong> high resolution satellite data.<br />

Study site Sensor Acqu.<br />

date<br />

Budongo Forest ASTER 15 Oct.<br />

2006<br />

Budongo Forest SPOT-4<br />

HRVIR<br />

Budongo Forest Landsat<br />

ETM+<br />

Kakamega Forest SPOT-4<br />

HRVIR<br />

Kakamega Forest Landsat<br />

ETM+<br />

01 Dec.<br />

2005<br />

23 May<br />

2000<br />

14 Dec.<br />

2004<br />

10 Jan.<br />

2003<br />

Solar<br />

azim. [°]<br />

Solar<br />

elev. [°]<br />

The <strong>MODIS</strong> LAI product<br />

In <strong>the</strong> following <strong>the</strong> <strong>MODIS</strong> LAI product<br />

(MOD15A2) will be described briefly with respect to<br />

instrument and data set characteristics, <strong>the</strong> algorithm<br />

used to derive LAI and <strong>the</strong> current validation status.<br />

It should be noted that both <strong>MODIS</strong> collection 4 (C4)<br />

and collection 5 (C5) were available <strong>for</strong> this <strong>the</strong>sis, as<br />

<strong>the</strong> <strong>for</strong>mer were processed within <strong>the</strong> BIOTA EAST<br />

AFRICA project. However, as C5 is <strong>the</strong> current<br />

version <strong>of</strong> <strong>the</strong> algorithm, validation will focus on<br />

<strong>the</strong>se products. Because <strong>the</strong> following description<br />

only covers in<strong>for</strong>mation relevant <strong>for</strong> validation, <strong>the</strong><br />

interested reader is referred to Knyazikhin et al.<br />

(1998a, b) and Myneni et al. (1997) <strong>for</strong> fur<strong>the</strong>r details<br />

concerning <strong>the</strong> <strong>the</strong>oretical basis <strong>of</strong> <strong>the</strong> algorithm.<br />

Tian et al. (2000) and Zhang et al. (2000) describe<br />

<strong>the</strong> algorithm prototyping, Panferov et al. (2001),<br />

Shabanov et al. (2003), Tian et al. (2002a), Wang<br />

et al. (2003a), and Shabanov et al. (2005) deal with<br />

<strong>the</strong> evaluation <strong>of</strong> <strong>the</strong> algorithm physics. Myneni et<br />

al. (2002), Tan et al. (2006), Wang et al. (2001), and<br />

Yang et al. (2006) provide product diagnostics.<br />

Tilt<br />

ang. [°]<br />

RMSE<br />

x [m]<br />

RMSE<br />

y [m]<br />

RMSE MODTRAN<br />

total [m] atmosphere<br />

118.20 68.51 n/a 6.95 9.23 11.55 Atm. correction<br />

already applied<br />

142.50 60.20 4.2 (L) 6.52 6.39 9.14 Tropical rural<br />

52.81 56.69 n/a 7.86 8.03 11.24 Tropical rural<br />

136.30 56.80 29.0 (R) 10.52 7.72 13.05 Tropical rural<br />

53.77 55.61 n/a 11.53 10.38 15.52 Tropical rural


4 Remote sensing data and preprocessing<br />

4.3.1<br />

The <strong>MODIS</strong> instrument<br />

The <strong>MODIS</strong> sensor is a key instrument aboard<br />

<strong>the</strong> Terra and Aqua plat<strong>for</strong>ms (<strong>for</strong>mer EOS-AM<br />

and EOS-PM). The satellites were launched on<br />

18 December, 1999 and 4 May, 2002 respectively<br />

in a sun-synchronous, near-polar orbit at an altitude<br />

<strong>of</strong> 705 km. While Terra is on a descending node with<br />

a mean equator crossing time <strong>of</strong> 10:30 a.m. (see also<br />

Chapter 4.1.1 on ASTER data), Aqua operates on<br />

an ascending node with an equatorial overpass at<br />

1:30 p.m. Because afternoon cloud convection in<br />

tropical regions hinders ground observations with<br />

optical data, only data acquired by <strong>MODIS</strong>-Terra<br />

were used <strong>for</strong> this <strong>the</strong>sis.<br />

<strong>MODIS</strong> is a cross-track scanning spectroradiometer.<br />

It acquires data in 36 spectral bands with bands in<br />

<strong>the</strong> visible, near-, shortwave- and <strong>the</strong>rmal infrared<br />

portions <strong>of</strong> <strong>the</strong> electromagnetic spectrum ranging<br />

from 0.4 to 14.54 µm (cf. Table 4-5 <strong>for</strong> bands relevant<br />

<strong>for</strong> MOD15A2, see Guen<strong>the</strong>r et al. (2002) <strong>for</strong> all<br />

bands) in different spatial resolutions (250 m to 1000<br />

m). With a cross-track swath dimension <strong>of</strong> 2,330 km<br />

and 10 km along-track at nadir, global coverage can<br />

be achieved daily at latitudes higher than 30° and<br />

every two days <strong>for</strong> lower latitudes. Because pixel<br />

size increases with swath angle, scan lines overlap<br />

Table 4-5<br />

59<br />

towards <strong>the</strong> scene edges (known as bow-tie effect,<br />

see Masouka et al. [1998] <strong>for</strong> illustration).<br />

According to Wolfe et al. (2002) <strong>the</strong> mean geolocation<br />

error <strong>of</strong> <strong>MODIS</strong> data is quantified as 27 m alongtrack<br />

and 25 m along-scan. For fur<strong>the</strong>r in<strong>for</strong>mation<br />

on <strong>the</strong> <strong>MODIS</strong> instrument please refer to Barnes et<br />

al. (1998) and Guen<strong>the</strong>r et al. (2002), as well as <strong>the</strong><br />

special issues in Remote Sensing <strong>of</strong> <strong>the</strong> Environment<br />

(2002, Volume 83, Issues 1-2) and IEEE Transactions<br />

on Geoscience and Remote Sensing (1998, Vol. 36,<br />

Issue 4).<br />

Following reception, data are sent to <strong>the</strong> EOS Data<br />

and Operations System (EDOS) at <strong>the</strong> GSFC, where<br />

level 1A, level 1B, geolocation and cloud mask<br />

products as well as higher-level <strong>MODIS</strong> land and<br />

atmosphere products are generated by <strong>the</strong> <strong>MODIS</strong><br />

Adaptive Processing System (MODAPS). The<br />

products are supplied to <strong>the</strong> user by three Distributed<br />

Active Archive Centers (DAACs) at GSFC, EROS<br />

Data Center (EDC) and <strong>the</strong> National Snow and Ice<br />

Data Center (NSIDC). Land <strong>Product</strong>s are available<br />

through <strong>the</strong> Land Processes DAAC at EDC, also<br />

including data relevant to this research. Figure 4-1<br />

shows <strong>the</strong> global distribution <strong>of</strong> <strong>MODIS</strong> tiles with <strong>the</strong><br />

tile h08v21, marked in red, covering <strong>the</strong> study sites.<br />

Spectral and spatial properties <strong>of</strong> <strong>the</strong> first seven bands <strong>of</strong> <strong>MODIS</strong> (relevant <strong>for</strong> MOD15A2).<br />

For fur<strong>the</strong>r in<strong>for</strong>mation on bands 8-36 (1000 m resolution) please refer to NASA (2007a).<br />

Band Wavelength [µm] IFOV at nadir [M] Quantization [bit] Key Applications<br />

1 0.620 – 0.670 250 12 Chlorophyll, land cover trans<strong>for</strong>mations<br />

2 0.841 – 0.876 250 12 Cloud amount, vegetation land cover trans<strong>for</strong>mation<br />

3 0.459 – 0.479 500 12 Soil/vegetation differences<br />

4 0.545 – 0.565 500 12 Green vegetation<br />

5 1.230 – 1.250 500 12 <strong>Leaf</strong>/canopy differences<br />

6 1.628 – 1.652 500 12 Snow/cloud differences<br />

7 2.105 – 2.155 500 12 Cloud properties, land properties


60<br />

Figure 4-1<br />

4.3.2<br />

<strong>MODIS</strong> tiles <strong>for</strong> processing levels 2G, 3 and 4. Study area marked in red (Wolfe 2002, modified).<br />

Data set characteristics<br />

The <strong>MODIS</strong> LAI/FPAR product (MOD15A2) is<br />

processed at 1 km spatial resolution and composited<br />

over 8 days <strong>based</strong> on <strong>the</strong> maximum FPAR value. Daily<br />

surface reflectance and cloud states (MOD09) as well<br />

as a biome map derived from <strong>the</strong> MOD12 land cover<br />

product and ancillary in<strong>for</strong>mation serve as input to<br />

<strong>the</strong> <strong>MODIS</strong> LAI algorithm (cf. Figure 4-2). As LAI is<br />

a modelled output, MOD15A2 is regarded as level 4<br />

product. Fur<strong>the</strong>r descriptions <strong>of</strong> <strong>the</strong> algorithm will be<br />

provided in Chapter 4.3.3.<br />

Similar to most o<strong>the</strong>r <strong>MODIS</strong> data products,<br />

MOD15A2 is stored in <strong>the</strong> EOS-HDF <strong>for</strong>mat<br />

and provided in <strong>the</strong> Sinusoidal Projection (cf.<br />

Figure 4-1). Table 4-6 shows <strong>the</strong> scientific data<br />

sets (SDS) and data specifications. Relevant<br />

<strong>for</strong> this <strong>the</strong>sis are only <strong>the</strong> SDS LAI_1km and<br />

LAIStDev_1km as well as <strong>the</strong> quality control data<br />

set FPAR_LAI_QC, <strong>for</strong> which <strong>the</strong> specifications<br />

are given in <strong>the</strong> Appendix in Table A-2. The key<br />

indicator <strong>for</strong> <strong>the</strong> retrieval quality is <strong>the</strong> SCF_QC<br />

bitfield. In order to provide a complete overview, <strong>the</strong><br />

second quality control SDS, FparExtra_QC can be<br />

found in Table A-3 as well as <strong>the</strong> values assigned to<br />

non-vegetated surfaces <strong>of</strong> <strong>the</strong> SDS FPAR_1km and<br />

LAI_1km (cf. Table A-4).<br />

All available C4 MOD15A2 data sets from 2000<br />

to 2006 were processed within <strong>the</strong> BIOTA project<br />

to derive phonological characteristics <strong>for</strong> <strong>the</strong> study<br />

sites. For C5, at <strong>the</strong> time <strong>of</strong> writing only data sets<br />

covering <strong>the</strong> years 2000 through 2005 were available<br />

to analyse temporal consistency. Preprocessing was<br />

accomplished with TiSeG (Colditz et al. 2008). C5<br />

data sets <strong>of</strong> Julian Days 281 <strong>of</strong> 2004 (Kakamega<br />

Forest) and 329 <strong>of</strong> 2005 (Budongo Forest) served<br />

as input <strong>for</strong> validation with in situ data (coinciding<br />

with <strong>the</strong> field campaigns at Budongo and Kakamega<br />

Forest, respectively).


4 Remote sensing data and preprocessing<br />

Figure 4-2<br />

Table 4-6<br />

Schematic processing chain <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product (Knyazikhin et al. 1998, modified).<br />

MOD15A2 data set characteristics (USGS 2006b).<br />

SDS Units Data type (bit) Fill values Valid range Scale factors<br />

FPAR_1km % 8-bit unsigned integer 249-255 0 - 100 0.01<br />

LAI_1km m 2 /m 2 8-bit unsigned integer 249-255 0 - 100 0.10<br />

FPARLAI_QC Class flag 8-bit unsigned integer 255 0 - 254 n/a<br />

FPARExtra_QC Class flag 8-bit unsigned integer 255 0 - 254 n/a<br />

FPARStdDev_1km Percent 8-bit unsigned integer 248-255 0 - 100 0.01<br />

LAIStdDev_1km m 2 plant/m 2 ground 8-bit unsigned integer 248-255 0 - 100 0.10<br />

4.3.3<br />

Algorithm description<br />

As briefly addressed in Chapter 3.4.2, <strong>the</strong> physical<br />

modelling <strong>of</strong> LAI given surface reflectance is a<br />

complex 3D inversion problem. Radiation leaving<br />

<strong>the</strong> canopy <strong>of</strong> a vegetation stand depends not only<br />

on <strong>the</strong> LAI, sun and viewing geometries, but also<br />

on o<strong>the</strong>r canopy parameters, such as canopy closure<br />

and understorey vegetation. There<strong>for</strong>e <strong>the</strong> socalled<br />

“space <strong>of</strong> canopy realization” is introduced<br />

(Knyazikhin et al. 1998b). In C5 it is <strong>based</strong> on an<br />

eight biome map, which describes canopy structural<br />

types <strong>of</strong> global vegetated land cover. However, as this<br />

61<br />

map is not available online so far, <strong>the</strong> classes <strong>of</strong> <strong>the</strong><br />

old six biome map are summarized in Table 4-7 (land<br />

cover type 3 in <strong>the</strong> MOD12Q1 product). According<br />

to NASA (2007b) <strong>the</strong> only difference is that <strong>for</strong> C5<br />

<strong>the</strong> classes “broadleaf <strong>for</strong>est” and “needleleaf <strong>for</strong>est”<br />

are split up into deciduous and evergreen types.<br />

Each biome represents structural patterns <strong>of</strong> trees<br />

and/or o<strong>the</strong>r canopy types as well as o<strong>the</strong>r structural<br />

and spectral attributes relevant to radiative transport<br />

<strong>the</strong>ory (cf. Table 4-8). For each variable <strong>the</strong> range<br />

<strong>of</strong> variation is limited within <strong>the</strong> respective biome<br />

(Knyazikhin et al. 1998b).


vegetated land cover. However, as this map is not available online so far, <strong>the</strong> classes <strong>of</strong> <strong>the</strong> old six biome<br />

map are summarized in Table 4-7 (land cover type 3 in <strong>the</strong> MOD12Q1 product). According to NASA<br />

62<br />

Table 4-7: Biome classes in (2007b) <strong>the</strong> only difference is that <strong>for</strong> C5 <strong>the</strong> classes “broadleaf <strong>for</strong>est”<br />

MOD12 land cover product as and “needleleaf <strong>for</strong>est” are split up into deciduous and evergreen types.<br />

underlying <strong>the</strong> MOD15A2<br />

The algorithm relation (USGS between 2006b). measured Also<br />

Each biome represents structural patterns <strong>of</strong> trees and/or o<strong>the</strong>r canopy<br />

surface reflectance<br />

refer to Table 4-10.<br />

types as well as o<strong>the</strong>r structural and spectral attributes relevant to<br />

and canopy structure is <strong>based</strong> on a 3D radiative<br />

radiative transport <strong>the</strong>ory (cf. Table 4-8). For each variable <strong>the</strong> range <strong>of</strong><br />

transfer model (compare Chapter 3.4.2). Bi-directional<br />

variation is limited within <strong>the</strong> respective biome (KNYAZIKHIN et al.<br />

reflectance factors (BRF) are modelled <strong>for</strong> <strong>the</strong> above-<br />

1998b).<br />

mentioned suite <strong>of</strong> canopy structures and soil patterns<br />

covering a range <strong>of</strong> expected natural The conditions relation between (Tian measured surface reflectance and canopy structure<br />

et al. 2000). The transfer model fur<strong>the</strong>r is <strong>based</strong> fulfils on <strong>the</strong> a 3D law radiative transfer model (compare Chapter 3.4.2). Bi-<br />

<strong>of</strong> energy conservation, i.e. incident directional radiation equals reflectance factors (BRF) are modelled <strong>for</strong> <strong>the</strong> aboveabsorbed,<br />

transmitted and reflected mentioned radiation suite by <strong>the</strong> <strong>of</strong> canopy structures and soil patterns covering a range<br />

canopy (Knyazikhin et al. 1999). <strong>of</strong> The expected results natural are conditions (TIAN et al. 2000). The transfer model<br />

stored in a LUT, which is <strong>the</strong> interface fur<strong>the</strong>r to <strong>the</strong> fulfils <strong>MODIS</strong> <strong>the</strong> law <strong>of</strong> energy conservation, i.e. incident radiation<br />

LAI algorithm. The algorithm compares equals observed absorbed, and transmitted and reflected radiation by <strong>the</strong> canopy<br />

modelled BRFs, taking into account (KNYAZIKHIN <strong>the</strong> uncertainties et al. 1999). The results are stored in a LUT, which is <strong>the</strong><br />

in both (Myneni et al. 2002). Only interface those solutions to <strong>the</strong> <strong>MODIS</strong> LAI algorithm. The algorithm compares<br />

are accepted, where <strong>the</strong> magnitude observed <strong>of</strong> residuals and in modelled <strong>the</strong> BRFs, taking into account <strong>the</strong> uncertainties in<br />

comparison both (MYNENI does et not al. exceed 2002). uncertainties Only those in solutions modelled are accepted, where <strong>the</strong> magnitude <strong>of</strong> residuals in <strong>the</strong><br />

and comparison observed does BRFs, not i.e. exceed uncertainties in modelled and observed BRFs, i.e.<br />

<br />

<br />

2<br />

obs<br />

mdl<br />

1 BRF ( , L,<br />

, ,<br />

) ( , , , ,<br />

) <br />

<br />

LC BRF L LC<br />

1 , (4.2)<br />

N <br />

<br />

<br />

<br />

<br />

obs<br />

mdl<br />

where BRF denotes observed and BRF modelled BRFs <strong>for</strong> N spectral bands with <strong>the</strong>ir<br />

uncertainties . <br />

and 1998b). In C5 it is <strong>based</strong> on an eight biome map, which describes canopy structural types <strong>of</strong> global<br />

vegetated land cover. However, as this map is not available online so far, <strong>the</strong> classes <strong>of</strong> <strong>the</strong> old six biome<br />

map are summarized in Table 4-7 (land cover type 3 in <strong>the</strong> MOD12Q1 product). According to NASA<br />

Table 4-7: Biome classes in (2007b) <strong>the</strong> only difference is that <strong>for</strong> C5 <strong>the</strong> classes “broadleaf <strong>for</strong>est”<br />

MOD12 land cover product as and “needleleaf <strong>for</strong>est” are split up into deciduous and evergreen types.<br />

underlying <strong>the</strong> MOD15A2<br />

algorithm (USGS 2006b). Also<br />

Each biome represents structural patterns <strong>of</strong> trees and/or o<strong>the</strong>r canopy<br />

refer to Table 4-10.<br />

types as well as o<strong>the</strong>r structural and spectral attributes relevant to<br />

CLASS LAI/FPAR radiative transport <strong>the</strong>ory (cf. Table 4-8). For each variable <strong>the</strong> range <strong>of</strong><br />

variation is limited within <strong>the</strong> respective biome (KNYAZIKHIN et al.<br />

0 water<br />

1998b).<br />

1 grasses/cereal crops<br />

2 shrubs<br />

The relation between measured surface reflectance and canopy structure<br />

is <strong>based</strong> on a 3D radiative transfer model (compare Chapter 3.4.2). Bi-<br />

3 broadleaf crops<br />

directional reflectance factors (BRF) are modelled <strong>for</strong> <strong>the</strong> above-<br />

4 savanna<br />

mentioned suite <strong>of</strong> canopy structures and soil patterns covering a range<br />

5 broadleaf <strong>for</strong>est<br />

<strong>of</strong> expected natural conditions (TIAN et al. 2000). The transfer model<br />

6 needleleaf <strong>for</strong>est fur<strong>the</strong>r fulfils <strong>the</strong> law <strong>of</strong> energy conservation, i.e. incident radiation<br />

7 unvegetated<br />

equals absorbed, transmitted and reflected radiation by <strong>the</strong> canopy<br />

8 urban<br />

(KNYAZIKHIN et al. 1999). The results are stored in a LUT, which is <strong>the</strong><br />

interface to <strong>the</strong> <strong>MODIS</strong> LAI algorithm. The algorithm compares<br />

254 unclassified<br />

observed and modelled BRFs, taking into account <strong>the</strong> uncertainties in<br />

both (MYNENI et al. 2002). Only those solutions are accepted, where <strong>the</strong> magnitude <strong>of</strong> residuals in <strong>the</strong><br />

comparison does not exceed uncertainties in modelled and observed BRFs, i.e.<br />

<br />

<br />

2<br />

obs<br />

mdl<br />

1 BRF ( , L,<br />

, , ) ( , , , , ) <br />

<br />

LC BRF L <br />

LC<br />

1 , (4.2)<br />

N <br />

<br />

<br />

<br />

<br />

obs<br />

mdl<br />

where<br />

BRF denotes observed and BRF modelled BRFs <strong>for</strong> N spectral bands with <strong>the</strong>ir<br />

indicate <strong>the</strong> sun and view directions and LC stands <strong>for</strong> <strong>the</strong> land cover type.<br />

uncertainties . Although <strong>the</strong> use <strong>of</strong> biome dependent canopy attributes reduces <strong>the</strong> possible number <strong>of</strong> solutions <strong>of</strong> <strong>the</strong><br />

inverse problem/<strong>for</strong> <strong>the</strong> inversion, <strong>the</strong>re are still multiple possible results because observed BRFs can<br />

relate to several combinations <strong>of</strong> different LAI values and soil background. There<strong>for</strong>e, mean LAI values<br />

averaged over all acceptable solutions are computed as output <strong>of</strong> <strong>the</strong> algorithm. Whereas only average LAI<br />

values <strong>for</strong> each pixel were provided by <strong>the</strong> C4 algorithm, a new SDS in C5 specifies also <strong>the</strong> standard<br />

deviation corresponding to <strong>the</strong> acceptable solution indicated (cf. Table 4-6).<br />

If <strong>the</strong> main algorithm fails to produce a valid solution <strong>for</strong> a certain pixel, a backup algorithm is applied. It<br />

estimates LAI <strong>based</strong> on biome dependent regressions as a non-linear function <strong>of</strong> NDVI. For low and<br />

medium NDVI values SHABANOV et al. (2005) report consistency <strong>of</strong> <strong>the</strong> LAI-NDVI curve with main<br />

<br />

and 1998b). In C5 it is <strong>based</strong> on an eight biome map, which describes canopy structural types <strong>of</strong> global<br />

vegetated land cover. However, as this map is not available online so far, <strong>the</strong> classes <strong>of</strong> <strong>the</strong> old six biome<br />

map are summarized in Table 4-7 (land cover type 3 in <strong>the</strong> MOD12Q1 product). According to NASA<br />

Table 4-7: Biome classes in (2007b) <strong>the</strong> only difference is that <strong>for</strong> C5 <strong>the</strong> classes “broadleaf <strong>for</strong>est”<br />

MOD12 land cover product as and “needleleaf <strong>for</strong>est” are split up into deciduous and evergreen types.<br />

underlying <strong>the</strong> MOD15A2<br />

algorithm (USGS 2006b). Also<br />

Each biome represents structural patterns <strong>of</strong> trees and/or o<strong>the</strong>r canopy<br />

refer to Table 4-10.<br />

types as well as o<strong>the</strong>r structural and spectral attributes relevant to<br />

radiative transport <strong>the</strong>ory (cf. Table 4-8). For each variable <strong>the</strong> range <strong>of</strong><br />

CLASS LAI/FPAR<br />

variation is limited within <strong>the</strong> respective biome (KNYAZIKHIN et al.<br />

0 water<br />

1998b).<br />

1 grasses/cereal crops<br />

The relation between measured surface reflectance and canopy structure<br />

2 shrubs<br />

is <strong>based</strong> on a 3D radiative transfer model (compare Chapter 3.4.2). Bi-<br />

3 broadleaf crops directional reflectance factors (BRF) are modelled <strong>for</strong> <strong>the</strong> above-<br />

4 savanna<br />

mentioned suite <strong>of</strong> canopy structures and soil patterns covering a range<br />

5 broadleaf <strong>for</strong>est <strong>of</strong> expected natural conditions (TIAN et al. 2000). The transfer model<br />

fur<strong>the</strong>r fulfils <strong>the</strong> law <strong>of</strong> energy conservation, i.e. incident radiation<br />

6 needleleaf <strong>for</strong>est<br />

equals absorbed, transmitted and reflected radiation by <strong>the</strong> canopy<br />

7 unvegetated<br />

(KNYAZIKHIN et al. 1999). The results are stored in a LUT, which is <strong>the</strong><br />

8 urban<br />

interface to <strong>the</strong> <strong>MODIS</strong> LAI algorithm. The algorithm compares<br />

254 unclassified observed and modelled BRFs, taking into account <strong>the</strong> uncertainties in<br />

both (MYNENI et al. 2002). Only those solutions are accepted, where <strong>the</strong> magnitude <strong>of</strong> residuals in <strong>the</strong><br />

comparison does not exceed uncertainties in modelled and observed BRFs, i.e.<br />

<br />

<br />

2<br />

obs<br />

mdl<br />

1 BRF ( , L,<br />

, , ) ( , , , , ) <br />

<br />

LC BRF L <br />

LC<br />

1 , (4.2)<br />

N <br />

<br />

<br />

<br />

<br />

obs<br />

mdl<br />

where BRF denotes denotes observed and and BRF modelled BRFs <strong>for</strong> N spectral bands with <strong>the</strong>ir<br />

uncertainties . indicate <strong>the</strong> sun and view directions and LC stands <strong>for</strong> <strong>the</strong> land cover type.<br />

Although <strong>the</strong> use <strong>of</strong> biome dependent canopy attributes reduces <strong>the</strong> possible number <strong>of</strong> solutions <strong>of</strong> <strong>the</strong><br />

inverse problem/<strong>for</strong> <strong>the</strong> inversion, <strong>the</strong>re are still multiple possible results because observed BRFs can<br />

relate to several combinations <strong>of</strong> different LAI values and soil background. There<strong>for</strong>e, mean LAI values<br />

averaged over all acceptable solutions are computed as output <strong>of</strong> <strong>the</strong> algorithm. Whereas only average LAI<br />

values <strong>for</strong> each pixel were provided by <strong>the</strong> C4 algorithm, a new SDS in C5 specifies also <strong>the</strong> standard<br />

deviation corresponding to <strong>the</strong> acceptable solution indicated (cf. Table 4-6).<br />

If <strong>the</strong> main algorithm fails to produce a valid solution <strong>for</strong> a certain pixel, a backup algorithm is applied. It<br />

estimates LAI <strong>based</strong> on biome dependent regressions as a non-linear function <strong>of</strong> NDVI. For low and<br />

medium NDVI values SHABANOV et al. (2005) report consistency <strong>of</strong> <strong>the</strong> LAI-NDVI curve with main<br />

<br />

and vegetation. There<strong>for</strong>e <strong>the</strong> so-called “space <strong>of</strong> canopy realization” is introduced (KNYAZIKHIN et al.<br />

1998b). In C5 it is <strong>based</strong> on an eight biome map, which describes canopy structural types <strong>of</strong> global<br />

vegetated land cover. However, as this map is not available online so far, <strong>the</strong> classes <strong>of</strong> <strong>the</strong> old six biome<br />

map are summarized in Table 4-7 (land cover type 3 in <strong>the</strong> MOD12Q1 product). According to NASA<br />

Table 4-7: Biome classes in (2007b) <strong>the</strong> only difference is that <strong>for</strong> C5 <strong>the</strong> classes “broadleaf <strong>for</strong>est”<br />

MOD12 land cover product as and “needleleaf <strong>for</strong>est” are split up into deciduous and evergreen types.<br />

underlying <strong>the</strong> MOD15A2<br />

algorithm (USGS 2006b). Also<br />

Each biome represents structural patterns <strong>of</strong> trees and/or o<strong>the</strong>r canopy<br />

refer to Table 4-10.<br />

types as well as o<strong>the</strong>r structural and spectral attributes relevant to<br />

CLASS LAI/FPAR radiative transport <strong>the</strong>ory (cf. Table 4-8). For each variable <strong>the</strong> range <strong>of</strong><br />

0 water<br />

variation is limited within <strong>the</strong> respective biome (KNYAZIKHIN et al.<br />

1 grasses/cereal crops 1998b).<br />

2 shrubs<br />

The relation between measured surface reflectance and canopy structure<br />

3 broadleaf crops is <strong>based</strong> on a 3D radiative transfer model (compare Chapter 3.4.2). Bi-<br />

4 savanna<br />

directional reflectance factors (BRF) are modelled <strong>for</strong> <strong>the</strong> abovementioned<br />

suite <strong>of</strong> canopy structures and soil patterns covering a range<br />

5 broadleaf <strong>for</strong>est<br />

<strong>of</strong> expected natural conditions (TIAN et al. 2000). The transfer model<br />

6 needleleaf <strong>for</strong>est<br />

fur<strong>the</strong>r fulfils <strong>the</strong> law <strong>of</strong> energy conservation, i.e. incident radiation<br />

7 unvegetated<br />

equals absorbed, transmitted and reflected radiation by <strong>the</strong> canopy<br />

8 urban<br />

(KNYAZIKHIN et al. 1999). The results are stored in a LUT, which is <strong>the</strong><br />

254 unclassified interface to <strong>the</strong> <strong>MODIS</strong> LAI algorithm. The algorithm compares<br />

observed and modelled BRFs, taking into account <strong>the</strong> uncertainties in<br />

both (MYNENI et al. 2002). Only those solutions are accepted, where <strong>the</strong> magnitude <strong>of</strong> residuals in <strong>the</strong><br />

comparison does not exceed uncertainties in modelled and observed BRFs, i.e.<br />

<br />

<br />

2<br />

obs<br />

mdl<br />

1 BRF ( , L,<br />

, , ) ( , , , , ) <br />

<br />

LC BRF L <br />

LC<br />

1 , (4.2)<br />

N <br />

<br />

<br />

<br />

<br />

modelled<br />

obs<br />

mdl<br />

where BRF BRFs denotes <strong>for</strong> N spectral observed<br />

indicate bands and with <strong>the</strong><br />

BRF<br />

sun <strong>the</strong>ir and<br />

modelled uncertainties view directions<br />

BRFs <strong>for</strong><br />

and<br />

N<br />

LC<br />

spectral<br />

stands <strong>for</strong><br />

bands<br />

<strong>the</strong> land<br />

with<br />

cover<br />

<strong>the</strong>ir<br />

type.<br />

uncertainties<br />

Although <strong>the</strong><br />

. <br />

use <strong>of</strong> biome dependent canopy attributes reduces <strong>the</strong> possible number <strong>of</strong> solutions <strong>of</strong> <strong>the</strong><br />

inverse problem/<strong>for</strong> <strong>the</strong> inversion, <strong>the</strong>re are still multiple possible results because observed BRFs can<br />

relate to several combinations <strong>of</strong> different LAI values and soil background. There<strong>for</strong>e, mean LAI values<br />

averaged over all acceptable solutions are computed as output <strong>of</strong> <strong>the</strong> algorithm. Whereas only average LAI<br />

values <strong>for</strong> each pixel were provided by <strong>the</strong> C4 algorithm, a new SDS in C5 specifies also <strong>the</strong> standard<br />

deviation corresponding to <strong>the</strong> acceptable solution indicated (cf. Table 4-6).<br />

If <strong>the</strong> main algorithm fails to produce a valid solution <strong>for</strong> a certain pixel, a backup algorithm is applied. It<br />

estimates LAI <strong>based</strong> on biome dependent regressions as a non-linear function <strong>of</strong> NDVI. For low and<br />

medium NDVI values SHABANOV et al. (2005) report consistency <strong>of</strong> <strong>the</strong> LAI-NDVI curve with main<br />

<br />

and egetation. There<strong>for</strong>e <strong>the</strong> so-called “space <strong>of</strong> canopy realization” is introduced (KNYAZIKHIN et al.<br />

998b). In C5 it is <strong>based</strong> on an eight biome map, which describes canopy structural types <strong>of</strong> global<br />

egetated land cover. However, as this map is not available online so far, <strong>the</strong> classes <strong>of</strong> <strong>the</strong> old six biome<br />

ap are summarized in Table 4-7 (land cover type 3 in <strong>the</strong> MOD12Q1 product). According to NASA<br />

able 4-7: Biome classes in (2007b) <strong>the</strong> only difference is that <strong>for</strong> C5 <strong>the</strong> classes “broadleaf <strong>for</strong>est”<br />

OD12 land cover product as and “needleleaf <strong>for</strong>est” are split up into deciduous and evergreen types.<br />

nderlying <strong>the</strong> MOD15A2<br />

lgorithm (USGS 2006b). Also<br />

Each biome represents structural patterns <strong>of</strong> trees and/or o<strong>the</strong>r canopy<br />

efer to Table 4-10.<br />

types as well as o<strong>the</strong>r structural and spectral attributes relevant to<br />

radiative transport <strong>the</strong>ory (cf. Table 4-8). For each variable <strong>the</strong> range <strong>of</strong><br />

CLASS LAI/FPAR<br />

variation is limited within <strong>the</strong> respective biome (KNYAZIKHIN et al.<br />

0 water<br />

1998b).<br />

1 grasses/cereal crops<br />

The relation between measured surface reflectance and canopy structure<br />

2 shrubs<br />

is <strong>based</strong> on a 3D radiative transfer model (compare Chapter 3.4.2). Bi-<br />

3 broadleaf crops<br />

directional reflectance factors (BRF) are modelled <strong>for</strong> <strong>the</strong> above-<br />

4 savanna<br />

mentioned suite <strong>of</strong> canopy structures and soil patterns covering a range<br />

5 broadleaf <strong>for</strong>est<br />

<strong>of</strong> expected natural conditions (TIAN et al. 2000). The transfer model<br />

fur<strong>the</strong>r fulfils <strong>the</strong> law <strong>of</strong> energy conservation, i.e. incident radiation<br />

6 needleleaf <strong>for</strong>est<br />

equals absorbed, transmitted and reflected radiation by <strong>the</strong> canopy<br />

7 unvegetated<br />

(KNYAZIKHIN et al. 1999). The results are stored in a LUT, which is <strong>the</strong><br />

8 urban<br />

interface to <strong>the</strong> <strong>MODIS</strong> LAI algorithm. The algorithm compares<br />

254 unclassified<br />

observed and modelled BRFs, taking into account <strong>the</strong> uncertainties in<br />

oth (MYNENI et al. 2002). Only those solutions are accepted, where <strong>the</strong> magnitude <strong>of</strong> residuals in <strong>the</strong><br />

omparison does not exceed uncertainties in modelled and observed BRFs, i.e.<br />

<br />

<br />

2<br />

obs<br />

mdl<br />

1 BRF ( , L,<br />

, , ) ( , , , , ) <br />

<br />

LC BRF L <br />

LC<br />

1 , (4.2)<br />

N <br />

<br />

<br />

<br />

<br />

obs<br />

mdl<br />

where BRF denotes observed and BRF modelled BRFs <strong>for</strong> N spectral bands with <strong>the</strong>ir<br />

ncertainties . indicate <strong>the</strong> sun and view directions and LC stands <strong>for</strong> <strong>the</strong> land cover type.<br />

Although <strong>the</strong> use <strong>of</strong> biome dependent canopy attributes reduces <strong>the</strong> possible number <strong>of</strong> solutions <strong>of</strong> <strong>the</strong><br />

inverse problem/<strong>for</strong> <strong>the</strong> inversion, <strong>the</strong>re are still multiple possible results because observed BRFs can<br />

relate to several combinations <strong>of</strong> different LAI values and soil background. There<strong>for</strong>e, mean LAI values<br />

averaged over all acceptable solutions are computed as output <strong>of</strong> <strong>the</strong> algorithm. Whereas only average LAI<br />

values <strong>for</strong> each pixel were provided by <strong>the</strong> C4 algorithm, a new SDS in C5 specifies also <strong>the</strong> standard<br />

deviation corresponding to <strong>the</strong> acceptable solution indicated (cf. Table 4-6).<br />

If <strong>the</strong> main algorithm fails to produce a valid solution <strong>for</strong> a certain pixel, a backup algorithm is applied. It<br />

estimates LAI <strong>based</strong> on biome dependent regressions as a non-linear function <strong>of</strong> NDVI. For low and<br />

medium NDVI values SHABANOV et al. (2005) report consistency <strong>of</strong> <strong>the</strong> LAI-NDVI curve with main<br />

<br />

and vegetation. There<strong>for</strong>e <strong>the</strong> so-called “space <strong>of</strong> canopy realization” is introduced (KNYAZIKHIN et al.<br />

1998b). In C5 it is <strong>based</strong> on an eight biome map, which describes canopy structural types <strong>of</strong> global<br />

vegetated land cover. However, as this map is not available online so far, <strong>the</strong> classes <strong>of</strong> <strong>the</strong> old six biome<br />

map are summarized in Table 4-7 (land cover type 3 in <strong>the</strong> MOD12Q1 product). According to NASA<br />

able 4-7: Biome classes in (2007b) <strong>the</strong> only difference is that <strong>for</strong> C5 <strong>the</strong> classes “broadleaf <strong>for</strong>est”<br />

OD12 land cover product as and “needleleaf <strong>for</strong>est” are split up into deciduous and evergreen types.<br />

nderlying <strong>the</strong> MOD15A2<br />

lgorithm (USGS 2006b). Also<br />

Each biome represents structural patterns <strong>of</strong> trees and/or o<strong>the</strong>r canopy<br />

efer to Table 4-10.<br />

types as well as o<strong>the</strong>r structural and spectral attributes relevant to<br />

radiative transport <strong>the</strong>ory (cf. Table 4-8). For each variable <strong>the</strong> range <strong>of</strong><br />

LASS LAI/FPAR<br />

variation is limited within <strong>the</strong> respective biome (KNYAZIKHIN et al.<br />

water<br />

1998b).<br />

grasses/cereal crops<br />

The relation between measured surface reflectance and canopy structure<br />

shrubs<br />

is <strong>based</strong> on a 3D radiative transfer model (compare Chapter 3.4.2). Bi-<br />

broadleaf crops<br />

directional reflectance factors (BRF) are modelled <strong>for</strong> <strong>the</strong> abovesavanna<br />

mentioned suite <strong>of</strong> canopy structures and soil patterns covering a range<br />

broadleaf <strong>for</strong>est<br />

<strong>of</strong> expected natural conditions (TIAN et al. 2000). The transfer model<br />

fur<strong>the</strong>r fulfils <strong>the</strong> law <strong>of</strong> energy conservation, i.e. incident radiation<br />

needleleaf <strong>for</strong>est<br />

equals absorbed, transmitted and reflected radiation by <strong>the</strong> canopy<br />

unvegetated<br />

(KNYAZIKHIN et al. 1999). The results are stored in a LUT, which is <strong>the</strong><br />

urban<br />

interface to <strong>the</strong> <strong>MODIS</strong> LAI algorithm. The algorithm compares<br />

54 unclassified<br />

observed and modelled BRFs, taking into account <strong>the</strong> uncertainties in<br />

both (MYNENI et al. 2002). Only those solutions are accepted, where <strong>the</strong> magnitude <strong>of</strong> residuals in <strong>the</strong><br />

comparison does not exceed uncertainties in modelled and observed BRFs, i.e.<br />

<br />

<br />

2<br />

obs<br />

mdl<br />

1 BRF ( , L,<br />

, , ) ( , , , , ) <br />

<br />

LC BRF L <br />

LC<br />

1 , (4.2)<br />

N <br />

<br />

<br />

<br />

<br />

obs<br />

mdl<br />

where BRF denotes observed and BRF modelled BRFs <strong>for</strong> N spectral bands with <strong>the</strong>ir<br />

uncertainties . and indicate <strong>the</strong> sun and view directions and LC stands <strong>for</strong> <strong>the</strong> land cover type.<br />

lthough <strong>the</strong> use <strong>of</strong> biome dependent canopy attributes reduces <strong>the</strong> possible number <strong>of</strong> solutions <strong>of</strong> <strong>the</strong><br />

nverse problem/<strong>for</strong> <strong>the</strong> inversion, <strong>the</strong>re are still multiple possible results because observed BRFs can<br />

elate to several combinations <strong>of</strong> different LAI values and soil background. There<strong>for</strong>e, mean LAI values<br />

veraged over all acceptable solutions are computed as output <strong>of</strong> <strong>the</strong> algorithm. Whereas only average LAI<br />

alues <strong>for</strong> each pixel were provided by <strong>the</strong> C4 algorithm, a new SDS in C5 specifies also <strong>the</strong> standard<br />

eviation corresponding to <strong>the</strong> acceptable solution indicated (cf. Table 4-6).<br />

f <strong>the</strong> main algorithm fails to produce a valid solution <strong>for</strong> a certain pixel, a backup algorithm is applied. It<br />

stimates LAI <strong>based</strong> on biome dependent regressions as a non-linear function <strong>of</strong> NDVI. For low and<br />

edium NDVI values SHABANOV et al. (2005) report consistency <strong>of</strong> <strong>the</strong> LAI-NDVI curve with main<br />

<br />

and Table 4-7 Biome classes in MOD12 land cover product as<br />

underlying <strong>the</strong> MOD15A2 algorithm<br />

(USGS 2006b). Also refer to Table 4-10.<br />

DN Biome class<br />

CLASS LAI/FPAR<br />

0 Water<br />

0 water<br />

1 Grasses/cereal crops<br />

2 Shrubs<br />

1 grasses/cereal crops<br />

3 Broadleaf crops<br />

2 shrubs<br />

4 Savanna<br />

3 broadleaf crops<br />

5 Broadleaf <strong>for</strong>est<br />

4 savanna<br />

6 Needleleaf <strong>for</strong>est<br />

5 broadleaf <strong>for</strong>est<br />

7 Unvegetated<br />

6 needleleaf <strong>for</strong>est<br />

8 Urban<br />

7 unvegetated<br />

254 Unclassified<br />

8 urban<br />

254 unclassified<br />

(4.2)<br />

low and medium NDVI values Shabanov et al.<br />

(2005) report consistency <strong>of</strong> <strong>the</strong> LAI-NDVI curve<br />

indicate <strong>the</strong> sun and view directions directions and with LC stands main algorithm <strong>for</strong> <strong>the</strong> land retrievals. cover type. However <strong>for</strong> NDVI<br />

Although <strong>the</strong> and use LC <strong>of</strong> biome stands dependent <strong>for</strong> <strong>the</strong> land canopy cover type. attributes Although reduces <strong>the</strong> <strong>the</strong> possible values greater number than <strong>of</strong> 0.82 solutions a constant <strong>of</strong> <strong>the</strong> LAI value <strong>of</strong> 6.1<br />

inverse problem/<strong>for</strong> use <strong>of</strong> biome <strong>the</strong> inversion, dependent <strong>the</strong>re canopy are still attributes multiple reduces possible results is computed, because leading observed to retrieval BRFs anomalies can especially<br />

relate to several <strong>the</strong> combinations possible number <strong>of</strong> different <strong>of</strong> solutions LAI values <strong>of</strong> <strong>the</strong> and inverse soil background. <strong>for</strong> broadleaf There<strong>for</strong>e, <strong>for</strong>ests. mean LAI values<br />

averaged over problem/<strong>for</strong> all acceptable <strong>the</strong> solutions inversion, are computed <strong>the</strong>re are as still output multiple <strong>of</strong> <strong>the</strong> algorithm. Whereas only average LAI<br />

values <strong>for</strong> each possible pixel results were provided because observed by <strong>the</strong> C4 BRFs algorithm, can relate a new to SDS Even in C5 though specifies <strong>the</strong> radiative also <strong>the</strong> transfer standard model is designed<br />

deviation corresponding several combinations to <strong>the</strong> acceptable <strong>of</strong> different solution LAI values indicated and (cf. soil Table to 4-6). incorporate up to seven <strong>MODIS</strong> bands, currently<br />

background. There<strong>for</strong>e, mean LAI values averaged only surface reflectances <strong>of</strong> bands 1 (red) and 2<br />

If <strong>the</strong> main algorithm fails to produce a valid solution <strong>for</strong> a certain pixel, a backup algorithm is applied. It<br />

over all acceptable solutions are computed as output (NIR) are incorporated (Shabanov 2001). In previous<br />

estimates LAI <strong>based</strong> on biome dependent regressions as a non-linear function <strong>of</strong> NDVI. For low and<br />

<strong>of</strong> <strong>the</strong> algorithm. Whereas only average LAI values versions <strong>of</strong> <strong>the</strong> MOD15A2 product (C1 to C3)<br />

medium NDVI values SHABANOV et al. (2005) report consistency <strong>of</strong> <strong>the</strong> LAI-NDVI curve with main<br />

<strong>for</strong> each pixel were provided by <strong>the</strong> C4 algorithm, a AVHRR land cover maps and modelled BRF values<br />

new SDS in C5 specifies also <strong>the</strong> standard deviation <strong>based</strong> on SeaWiFS data were processed (Cohen et al.<br />

corresponding to <strong>the</strong> acceptable solution indicated 2006).<br />

(cf. Table 4-6).<br />

If <strong>the</strong> main algorithm fails to produce a valid<br />

solution <strong>for</strong> a certain pixel, a backup algorithm is<br />

applied. It estimates LAI <strong>based</strong> on biome dependent<br />

regressions as a non-linear function <strong>of</strong> NDVI. For<br />

Although some product anomalies have been resolved,<br />

o<strong>the</strong>r problems did still exist <strong>for</strong> C4 <strong>MODIS</strong> LAI.<br />

An analysis by Shabanov et al. (2005) indicated, <strong>for</strong><br />

instance, a retrieval anomaly <strong>for</strong> broadleaf <strong>for</strong>ests, <strong>for</strong><br />

which a low portion <strong>of</strong> best quality main algorithm


4 Remote sensing data and preprocessing<br />

retrievals are observed over broadleaf <strong>for</strong>ests in<br />

nor<strong>the</strong>rn America and Amazonia. Apparently <strong>the</strong>re<br />

were still some inconsistencies between modelled and<br />

observed surface reflectances in <strong>the</strong> C4 algorithm. At<br />

<strong>the</strong> time <strong>of</strong> writing <strong>the</strong>re are no available publications<br />

which analyse or address possible improvements <strong>of</strong><br />

<strong>the</strong>se issues with <strong>the</strong> C5 algorithm (cf. Chapter 4.3.4).<br />

Fur<strong>the</strong>r details on <strong>the</strong> <strong>the</strong>oretical background <strong>of</strong> <strong>the</strong><br />

<strong>MODIS</strong> LAI algorithm are provided by Knyazikhin<br />

et al. (1998b), Knyazikhin et al. (1999), Myneni et al.<br />

(1997), Myneni et al. (2002), Tian et al. (2000), and<br />

Wang et al. (2001).<br />

4.3.4<br />

Algorithm refinements in C5<br />

Beginning March 2005 <strong>the</strong> C5 version <strong>of</strong> <strong>MODIS</strong><br />

level 1 science algorithms was released. Reprocessing<br />

<strong>of</strong> <strong>MODIS</strong>/Terra land products started in July 2006<br />

and was almost completed at <strong>the</strong> time <strong>of</strong> writing.<br />

Table 4-8<br />

63<br />

Apart from an improved atmospheric correction <strong>of</strong><br />

surface reflectance products that serve as input to <strong>the</strong><br />

<strong>MODIS</strong> LAI algorithm, major refinements include<br />

<strong>the</strong> replacement <strong>of</strong> six biome map with <strong>the</strong> new eight<br />

biome map (NASA 2007b, cf. Chapter 4.3.3). LUT<br />

values were fur<strong>the</strong>r defined and refined <strong>for</strong> all eight<br />

biomes. A new stochastic radiative transfer model<br />

<strong>based</strong> on <strong>the</strong> work <strong>of</strong> Shabanov et al. (2007) was<br />

incorporated, which improved <strong>the</strong> representation <strong>of</strong><br />

canopy structure and spatial heterogeneity intrinsic<br />

to woody biomes. Fur<strong>the</strong>r, biome-dependent<br />

uncertainties were introduced (cf. Equation 4.2)<br />

as thresholds <strong>of</strong> allowable discrepancies between<br />

modelled and observed BRFs. For woody vegetation<br />

<strong>the</strong> uncertainties equal 30% <strong>for</strong> red and 15% <strong>for</strong> NIR<br />

reflectances.<br />

As a result <strong>of</strong> <strong>the</strong> above-mentioned changes, higher<br />

retrieval rates and improved consistencies with field<br />

measurements were observed <strong>for</strong> savannas, broadleaf<br />

Canopy structural attributes <strong>of</strong> <strong>the</strong> six biomes used in <strong>the</strong> C4 LAI algorithm (Knyazikhin et al. 1998).<br />

Grasses and<br />

cereal crops<br />

Shrubs Broadleaf<br />

crops<br />

Savannas Broadleaf<br />

<strong>for</strong>ests<br />

Horizontal heterogeneity No Yes Variable Yes Yes Yes<br />

Needleleaf<br />

<strong>for</strong>ests<br />

<strong>Ground</strong> cover 100% 20-60% 10-100% 20-40% > 70% > 70%<br />

Vertical heterogeneity<br />

(leaf optics and LAD)<br />

No No No Yes Yes Yes<br />

Stems/trunks No No Green stems Yes Yes Yes<br />

Understorey No No No Grasses Yes Yes<br />

Foliage dispersion Minimal<br />

clumping<br />

Random Regular Minimal<br />

clumping<br />

Clumped Severe<br />

clumping<br />

Crown shadowing No Not mutual No No Yes, mutual Yes, mutual<br />

Brightness <strong>of</strong> canopy<br />

ground<br />

Medium Bright Dark Medium Dark Dark


64<br />

and needleleaf <strong>for</strong>ests (NASA 2007b). Improvements<br />

are also obvious <strong>for</strong> <strong>the</strong> study sites <strong>of</strong> this <strong>the</strong>sis and<br />

will be discussed in Chapter 7.<br />

4.3.5<br />

<strong>Validation</strong> status <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI<br />

product<br />

In order to evaluate <strong>the</strong> per<strong>for</strong>mance <strong>of</strong> <strong>the</strong> <strong>MODIS</strong><br />

LAI algorithm with respect to various circumstances,<br />

i.e. <strong>for</strong> different biomes, seasons and atmospheric<br />

conditions, a comprehensive validation is needed.<br />

Generally and depending on <strong>the</strong> ef<strong>for</strong>ts already made<br />

<strong>for</strong> <strong>the</strong> validation <strong>of</strong> a certain satellite product, CEOS<br />

distinguishes three validation stages (Morisette et al.<br />

2006b):<br />

Stage 1 <strong>Product</strong> accuracy has been estimated using a<br />

small number <strong>of</strong> independent measurements<br />

obtained from selected locations and time<br />

periods and ground-truth/field program<br />

ef<strong>for</strong>t.<br />

Stage 2 <strong>Product</strong> accuracy has been assessed over<br />

a widely distributed set <strong>of</strong> locations and<br />

time periods via several ground-truth and<br />

validation ef<strong>for</strong>ts.<br />

Stage 3 <strong>Product</strong> accuracy has been assessed, and<br />

uncertainties in <strong>the</strong> product are wellestablished<br />

via independent measurements<br />

made in a systematic and statistically robust<br />

way that represent global conditions.<br />

The <strong>MODIS</strong> LAI product is currently classified as<br />

stage 1, indicating that more ground truth data and<br />

validation ef<strong>for</strong>ts are needed. As already mentioned<br />

in Chapter 1 validation activities have mainly been<br />

completed within <strong>the</strong> CEOS-LPV context (including<br />

<strong>the</strong> VALERI and BigFoot networks). Table 4-9 lists<br />

published studies concerned with <strong>the</strong> analysis <strong>of</strong> <strong>the</strong><br />

C4 <strong>MODIS</strong> LAI product, <strong>the</strong> applied in situ methods,<br />

upscaling approaches and validation results. Fur<strong>the</strong>r<br />

validation work, e.g. by Privette et al. (2002), Cohen<br />

et al. (2003), Kang et al. (2003), Scholes et al. (2004)<br />

or Tian et al. (2002a and b) was not taken into<br />

account as those studies employed previous versions<br />

<strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product (C1 and C3). Studies on<br />

C5 data have not been published so far.<br />

Generally, sites with low LAI values are reported to<br />

be mapped with satisfying accuracy by <strong>the</strong> <strong>MODIS</strong><br />

LAI product (Cohen et al. 2006, Hill et al. 2006,<br />

Pandya et al. 2006, Tan et al. 2005). For woody<br />

biomes a weak correlation between in situ LAI and<br />

<strong>MODIS</strong> LAI is reported (Abuelgasim et al. 2006,<br />

Aragão et al. 2005) as well as an overestimation <strong>of</strong><br />

<strong>MODIS</strong> LAI (Abuelgasim et al. 2006, Aragão et al.<br />

2005, Cohen et al. 2006, Hill et al. 2006). The studies<br />

mentioned in Table 4-9 also confirm a dominance<br />

<strong>of</strong> backup algorithm retrievals in C4 <strong>for</strong> <strong>for</strong>est sites<br />

(Aragão et al. 2005, Cohen et al. 2006).<br />

With respect to validation <strong>of</strong> C4 <strong>MODIS</strong> LAI <strong>for</strong><br />

tropical rain <strong>for</strong>ests three studies have been published<br />

so far (Aragão et al. 2005, Cohen et al. 2006 and Hill<br />

et al. 2006). However <strong>the</strong> methodology applied in Hill<br />

et al. (2006) to a tropical rain <strong>for</strong>est site in Australia<br />

involved <strong>the</strong> direct comparison <strong>of</strong> field measurements<br />

and <strong>MODIS</strong> LAI without intermediate upscaling<br />

steps. This methodology is seen as problematic<br />

by Tan et al. (2005) and Yang et al. (2006) among<br />

o<strong>the</strong>rs, because <strong>of</strong> scale mismatch. Hill et al. (2006)<br />

also mention that <strong>the</strong> observed variations between<br />

<strong>MODIS</strong> LAI and in situ measurements are likely to<br />

be due to <strong>the</strong> exclusion <strong>of</strong> understorey vegetation,<br />

which was not included in field measurements.<br />

Cohen et al. (2006) took LAI-2000 PCA measurements<br />

as reference <strong>for</strong> validation <strong>of</strong> <strong>the</strong> C4 <strong>MODIS</strong> LAI<br />

product <strong>for</strong> a BigFoot primary rain <strong>for</strong>est site in<br />

<strong>the</strong> Tapajós National Forest in Brazil. Understorey<br />

vegetation was also not included in <strong>the</strong>ir study.<br />

Similar to Hill et al. (2006), this can partly explain


4 Remote sensing data and preprocessing<br />

<strong>the</strong> reported overestimation <strong>of</strong> <strong>MODIS</strong> LAI. Cohen<br />

et al. (2006) fur<strong>the</strong>r acknowledge problems with<br />

geolocation, which restricted <strong>the</strong> available test sites<br />

to only 10 <strong>of</strong> 100 plots <strong>for</strong> analysis. This limitation<br />

has been partly attributed to <strong>the</strong> small plot size <strong>of</strong><br />

25 x 25 m, which had to be registered accurately to<br />

<strong>the</strong> respective Landsat ETM+ pixel to avoid errors in<br />

regression analysis.<br />

Aragão et al. (2005) also validated <strong>the</strong> C4 <strong>MODIS</strong><br />

LAI product <strong>for</strong> <strong>the</strong> Tapajós region, but used larger<br />

plot sizes with a nested sampling design <strong>of</strong> three<br />

50 x 50 m subplots within an area <strong>of</strong> 270 x 270 m.<br />

Field measurements were again conducted with<br />

LAI-2000 PCA, thus excluding understorey<br />

vegetation. Aragão et al. (2005) also report an<br />

overestimation <strong>of</strong> <strong>MODIS</strong> LAI compared to a<br />

fine resolution LAI map with an RMSE <strong>of</strong> 1.54.<br />

Apparently no difference was made between <strong>MODIS</strong><br />

LAI data derived from main and backup algorithms.<br />

To <strong>the</strong> author’s knowledge, no studies have been<br />

published on <strong>the</strong> validation <strong>of</strong> <strong>the</strong> C5 <strong>MODIS</strong> LAI<br />

product <strong>based</strong> on in situ data.<br />

Although <strong>the</strong>re are some common approaches in<br />

<strong>the</strong> scientific community (cf. Table 4-16) especially<br />

with respect to <strong>the</strong> applied upscaling approaches, <strong>the</strong><br />

description <strong>of</strong> validation results is fairly different. It<br />

ranged from purely descriptive quality assessments<br />

(Hill et al. 2006) to quantitative estimates <strong>of</strong> accuracy,<br />

precision and uncertainty (Tan et al. 2005).<br />

65


66<br />

comparison between aggregated fine resolution map and<br />

<strong>MODIS</strong> data<br />

Table 4-9<br />

Wang et al.<br />

(2004)<br />

Needleleaf <strong>for</strong>est<br />

(Finland)<br />

LAI-2000 PCA Empirical transfer function, univariate regression (RSR<br />

derived from Landsat ETM+ and in situ data), patch<br />

C4 <strong>MODIS</strong> LAI is accurate within a range <strong>of</strong> 0.5,<br />

precision: 48%<br />

average LAI values derived from fine resolution maps<br />

factor <strong>of</strong> 2<br />

measurement errors), comparison: <strong>MODIS</strong> LAI and<br />

Tan et al.<br />

(2005)<br />

Grasses & cereal crops<br />

(France)<br />

LAI-2000 PCA Empirical transfer function (SR derived from Landsat<br />

ETM+ and in situ data, including observation &<br />

C4 <strong>MODIS</strong> LAI is accurate within a range <strong>of</strong> 0.3<br />

with a precision <strong>of</strong> 20% and an uncertainty <strong>of</strong> 25%,<br />

biome misclassification decreases <strong>the</strong> accuracy by<br />

<strong>MODIS</strong> LAI and average LAI values derived from fine<br />

resolution maps<br />

Pandya et al.<br />

(2006)<br />

Grasses & cereal crops<br />

(India)<br />

LAI-2000 PCA Empirical transfer functions, non-linear regressions (NDVI<br />

derived from LISS-III and in situ data), comparison <strong>of</strong><br />

<strong>MODIS</strong> C4 data were positively correlated to LISS-<br />

III derived LAI maps (r = 0.62-0.73), RMSE better<br />

than with C3 data (0.55-0.82)<br />

in tropical rain <strong>for</strong>est. Misclassifications in biome<br />

map strongly affect quality <strong>of</strong> <strong>MODIS</strong> LAI.<br />

Hill et al.<br />

(2006)<br />

Broadleaf <strong>for</strong>est,<br />

savanna (Australia)<br />

LAI-2000 PCA,<br />

hemispherical<br />

photography<br />

Direct comparison between in situ and <strong>MODIS</strong> LAI C4 <strong>MODIS</strong> LAI shows good correspondence with<br />

field measurements in savanna, overestimates LAI<br />

Fensholt et al.<br />

(2004)<br />

Savanna<br />

(Senegal)<br />

LAI-2000 PCA Direct comparison between in situ and <strong>MODIS</strong> LAI<br />

(homogeneity <strong>of</strong> sites evaluated with Landsat ETM+)<br />

Seasonal dynamics <strong>of</strong> in situ LAI was captured well<br />

by <strong>MODIS</strong> LAI (C4), but overestimation <strong>of</strong> <strong>MODIS</strong><br />

LAI by approximately 2-15%<br />

and South America)<br />

> 50% <strong>for</strong> evergreen broadleaf <strong>for</strong>ests<br />

Overview on validation studies <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product (C4).<br />

Cohen et al.<br />

(2006)<br />

Grasses & cereal crops,<br />

needle-leaf <strong>for</strong>est,<br />

broadleaf <strong>for</strong>est (North<br />

LAI-2000 PCA,<br />

destructive sampling,<br />

allometry<br />

Empirical transfer functions (Landsat ETM+ and in situ<br />

data), comparison <strong>of</strong> <strong>MODIS</strong> LAI and average LAI values<br />

derived from fine resolution maps<br />

Overprediction <strong>of</strong> C4 <strong>MODIS</strong> LAI in <strong>for</strong>ested biome<br />

sites with low LAIs correctly mapped (especially<br />

arid and semiarid regions), backup algorithm used<br />

<strong>MODIS</strong> LAI with respect to field LAI <strong>of</strong> 1-2<br />

Aragão et al.<br />

(2005)<br />

Broadleaf <strong>for</strong>est (Brazil) LAI-2000 PCA Empirical transfer functions (multivariate regression with<br />

terrain and spectral variables derived from Landsat ETM+)<br />

Weak correlation between C4 <strong>MODIS</strong> LAI and fine<br />

resolution LAI maps (predominance <strong>of</strong> backup<br />

algorithm over region), general overprediction <strong>of</strong><br />

resolution maps, both aggregated to 3 km resolution<br />

<strong>for</strong>ests (bias = -2.2 to 2.4)<br />

Abuelgasim<br />

et al. (2006)<br />

Needleleaf <strong>for</strong>est, and<br />

broadleaf <strong>for</strong>est,<br />

(Canada)<br />

LAI-2000 PCA,<br />

TRAC, hemispherical<br />

photography<br />

Empirical transfer functions, linear regression (RSR<br />

derived from Landsat ETM+ and in situ data), comparison<br />

<strong>of</strong> <strong>MODIS</strong> LAI and average LAI values derived from fine<br />

C4 <strong>MODIS</strong> LAI showed weak correlation to mixed<br />

<strong>for</strong>est stands (R2 =-0.06 to 0.34), overestimation<br />

<strong>of</strong> <strong>MODIS</strong> LAI <strong>for</strong> mixed and broadleaf dominated<br />

Reference Biome Field instrumentation Upscaling approach Results


5 <strong>Ground</strong>-<strong>based</strong><br />

LAI measurements<br />

As indicated earlier, methods <strong>of</strong> in situ LAI sampling<br />

need an adaptation to <strong>the</strong> tropical rain <strong>for</strong>est<br />

environment. Based on <strong>the</strong> <strong>the</strong>oretical background<br />

given in Chapter 3, <strong>the</strong> establishment <strong>of</strong> an appropriate<br />

spatial sampling strategy will be described in <strong>the</strong><br />

following. Subsequently, processing and analysis <strong>of</strong><br />

field data will be presented toge<strong>the</strong>r with a correction<br />

method that was developed <strong>for</strong> LAI-2000 PCA data<br />

acquired under non-diffuse illumination conditions.<br />

Finally <strong>the</strong> results <strong>of</strong> <strong>the</strong> LAI field sampling will<br />

be described, <strong>for</strong>ming an independent basis <strong>for</strong> <strong>the</strong><br />

validation <strong>of</strong> <strong>MODIS</strong> LAI data.<br />

Lessons learnt during <strong>the</strong> first field campaign in<br />

Kakamega Forest in October/November 2004 led<br />

to improvements in <strong>the</strong> sampling strategy and an<br />

elaborate correction <strong>of</strong> in situ data. Thus data <strong>of</strong><br />

higher quality could be derived from <strong>the</strong> second<br />

field campaign that took place in Budongo Forest in<br />

October 2005. Data presented in Chapters 5.1 and<br />

5.2 originate from this study site only. Results from<br />

Kakamega Forest will, however, also be presented<br />

<strong>for</strong> comparison purposes in Chapter 5.3.2.<br />

5.1<br />

5.1.1<br />

Spatial sampling strategy<br />

Modifications made to <strong>the</strong> CEOS-<br />

LPV/VALERI methodology<br />

67<br />

The validation strategy <strong>of</strong> CEOS-LPV/VALERI was<br />

introduced in Chapter 3.4.5. These recommendations<br />

had to be modified in this <strong>the</strong>sis with respect to <strong>the</strong><br />

complex vegetation structure in tropical rain <strong>for</strong>ests,<br />

<strong>the</strong> area sampled with individual in situ measurements,<br />

as well as considerations <strong>of</strong> geolocation accuracy and<br />

upscaling. The main modifications related to <strong>the</strong> size<br />

<strong>of</strong> <strong>the</strong> whole test site, <strong>the</strong> size <strong>of</strong> individual sampling<br />

units and <strong>the</strong> spatial sampling scheme in terms <strong>of</strong><br />

pattern and distance.<br />

Table 5-1 gives an overview <strong>of</strong> <strong>the</strong> recommendations<br />

made by VALERI on <strong>the</strong> above-mentioned issues and<br />

necessary modifications. First <strong>of</strong> all, <strong>the</strong> recommended<br />

size <strong>of</strong> <strong>the</strong> test site – 3 x 3 km – is too small to cover<br />

all <strong>for</strong>est stages and thus <strong>the</strong> variability in structure<br />

and LAI (compare logging compartments in Budongo<br />

Forest in Figure 2-9). The size <strong>of</strong> <strong>the</strong> test sites was<br />

<strong>the</strong>re<strong>for</strong>e increased to cover <strong>the</strong> whole <strong>for</strong>est.<br />

Second, <strong>the</strong> VALERI scheme is <strong>based</strong> on sampling<br />

<strong>of</strong> ESUs. These correspond to <strong>the</strong> size <strong>of</strong> one pixel <strong>of</strong><br />

<strong>the</strong> high resolution satellite image that is later used<br />

<strong>for</strong> upscaling <strong>of</strong> field measurements. In tropical rain<br />

<strong>for</strong>ests this leads to difficulties <strong>for</strong> two reasons:<br />

a) Geolocation with GPS is difficult as<br />

<strong>the</strong> satellite signal is obstructed by <strong>the</strong><br />

dense canopy. Often <strong>the</strong> reception <strong>of</strong> a<br />

sufficient number <strong>of</strong> satellites to locate <strong>the</strong><br />

respective position with an acceptable error<br />

(i.e. 5-10 m) cannot be accomplished.<br />

Thus it can not be certain that <strong>the</strong> mean<br />

field LAI <strong>of</strong> an ESU can be assigned to <strong>the</strong><br />

correct reflectance values retrieved from<br />

<strong>the</strong> corresponding pixel <strong>of</strong> high resolution<br />

satellite data (cf. Equation 3.13).


68<br />

b) To provide a statistically sound basis <strong>for</strong><br />

later analysis, single LAI measurements<br />

should be spatially independent. At <strong>the</strong> same<br />

time a sufficient number <strong>of</strong> measurements<br />

per ESU should be taken (cf. Chapter<br />

3.3.4). This prerequisite cannot be fulfilled<br />

in tropical rain <strong>for</strong>ests if <strong>the</strong> ESU size<br />

corresponds to pixel sizes <strong>of</strong> high resolution<br />

satellite imagery (i.e. 15 to 30 m) as <strong>the</strong><br />

<strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> <strong>the</strong> instruments<br />

used is more than 30 m <strong>based</strong> on <strong>the</strong> settings<br />

applied in this <strong>the</strong>sis (cf. Figure A-1 and<br />

Equations A.2 and A.3 in <strong>the</strong> Appendix). On<br />

<strong>the</strong> o<strong>the</strong>r hand, in reality canopy elements<br />

(especially in dense canopies) block <strong>the</strong><br />

sensor’s field <strong>of</strong> view at a smaller distance<br />

than <strong>the</strong> <strong>the</strong>oretical one. For example, in<br />

a seasonal moist tropical <strong>for</strong>est in French<br />

Guiana, Ferment et al. (2001) found LAI<br />

measurements derived from LAI-2000 PCA<br />

and DHP to be independent <strong>of</strong> each o<strong>the</strong>r at<br />

a distance <strong>of</strong> 10 m (at a significance level <strong>of</strong><br />

5%). According to <strong>the</strong>ir results, a minimum<br />

distance between two measurements <strong>of</strong><br />

10 m was defined, decreasing to 5 m in<br />

lower canopy types <strong>of</strong> young <strong>for</strong>est stages.<br />

In consequence <strong>of</strong> a) and b), and taking account<br />

<strong>of</strong> Equation 3.13, ESU size was increased to 200<br />

x 200 m. Also <strong>the</strong> spatial sampling pattern had to<br />

be modified. Instead <strong>of</strong> following <strong>the</strong> sampling in<br />

square or cross patterns recommended by Garrigues<br />

et al. (2002), a transect approach was followed as<br />

in De Wasseige et al. (2003), Leblanc et al. (2002),<br />

Le Dantec et al. (2000), Mussche et al. (2001),<br />

and Strachan & McCaughey (1996). This had <strong>the</strong><br />

advantage that in both study sites existing transects<br />

established <strong>for</strong> scientific purposes by ei<strong>the</strong>r <strong>the</strong><br />

BIOTA EAST AFRICA project or <strong>the</strong> Budongo<br />

Conservation Field Station could be used as <strong>the</strong> test<br />

sites were simply not o<strong>the</strong>rwise accessible due to<br />

dense understorey vegetation on many ESUs. As <strong>the</strong><br />

different sampling patterns tested by Garrigues et<br />

al. (2002) gave similar results with respect to <strong>the</strong>ir<br />

representativeness <strong>of</strong> ESUs, <strong>the</strong> impact <strong>of</strong> <strong>the</strong> transect<br />

sampling is considered to be marginal.<br />

Table 5-1 Methodology proposed by VALERI <strong>for</strong> in situ sampling and modifications made with respect to <strong>the</strong> rain <strong>for</strong>est<br />

environment.<br />

Parameter VALERI recommendation Modification Reason <strong>for</strong> modification<br />

Size <strong>of</strong> test 3 x 3 km Size <strong>of</strong> <strong>for</strong>est Forests not easy to access (larger area<br />

site<br />

needed), focal point on sampling <strong>of</strong><br />

representative amount <strong>of</strong> ESUs <strong>for</strong> each<br />

<strong>for</strong>est type<br />

Size <strong>of</strong> ESU Pixel size <strong>of</strong> high resolution 200 x 200 m GPS location difficult (larger ESU size<br />

satellite data (e.g. 15 x 15 m <strong>for</strong><br />

minimizes error), individual measurement<br />

ASTER data)<br />

covers area larger in high canopies<br />

Sampling<br />

pattern<br />

Sampling<br />

distance<br />

Square or cross pattern 3 transects á 200 m Established transects could be used in<br />

Budongo and Kakamega Forest<br />

Not specified, 12 measurements<br />

per ESU recommended<br />

(cf. Figure 3-7)<br />

Ei<strong>the</strong>r 20 or 40 per transect<br />

(distance 10 m/5 m inside/<br />

outside <strong>for</strong>est respectively)<br />

Distance between individual measurements<br />

chosen according to canopy height to<br />

guarantee spatial independence


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

5.1.2<br />

Field campaigns<br />

The field measurements took place from 25 October to<br />

29 November, 2004 in Kakamega Forest (main <strong>for</strong>est<br />

block and Kisere Fragment) and from 5 October to<br />

3 November, 2005 in Budongo Forest (Budongo and<br />

Kanyo-Pabidi <strong>for</strong>est blocks). In situ measurements <strong>of</strong><br />

LAI were conducted in both study sites with DHP and<br />

LAI-2000 PCA. 30 ESUs were selected in each <strong>of</strong> <strong>the</strong><br />

<strong>for</strong>ests <strong>based</strong> on a stratified random approach. Land<br />

cover classifications (Lung 2004, cf. Figure A-2)<br />

and in<strong>for</strong>mation on <strong>for</strong>est management and logging<br />

regimes in different compartments (cf. Chapter 2.1.3,<br />

available <strong>for</strong> Budongo only) were used to establish<br />

Figure 5-1<br />

69<br />

ESUs in different <strong>for</strong>est stages. Here <strong>the</strong> concept <strong>of</strong><br />

Kalácska et al. 2004 was adopted, who instead <strong>of</strong> only<br />

distinguishing between primary and secondary <strong>for</strong>est<br />

areas, focused in <strong>the</strong>ir analyses on early, intermediate<br />

(disturbed) and late (undisturbed) <strong>for</strong>est stages.<br />

Early <strong>for</strong>est stages are thus referring to very young<br />

secondary (i.e. colonizing) <strong>for</strong>ests, intermediate<br />

<strong>for</strong>est stages are selectively logged secondary <strong>for</strong>ests<br />

and late <strong>for</strong>est stages signify natural (or near natural<br />

in case <strong>of</strong> Kakamega) primary <strong>for</strong>est areas.<br />

Whereas <strong>the</strong> first two correspond to early succession<br />

stages and disturbed secondary <strong>for</strong>est, <strong>the</strong> latter<br />

resembles nearly undisturbed primary <strong>for</strong>est. ESUs<br />

Location <strong>of</strong> ESUs in Kakamega Forest. a) Buyango Nature Reserve, Kambiri, Shanderema, Lugusi,<br />

b) Kisere fragment, c) Isecheno and Yala and d) Ikuywa (background: Landsat-ETM+ scene, 10 January, 2003).


70<br />

were always located more than 100 m away from<br />

<strong>the</strong> <strong>for</strong>est border to avoid mixed pixels in <strong>the</strong> high<br />

resolution satellite imagery. Yet <strong>the</strong> main restriction<br />

<strong>for</strong> ESU distribution in both <strong>for</strong>ests was <strong>the</strong><br />

accessibility to test sites.<br />

Figure 5-1 illustrates <strong>the</strong> spatial distribution <strong>of</strong> <strong>the</strong><br />

30 ESUs on <strong>the</strong> test site in Kenya. The major part <strong>of</strong><br />

ESUs was situated in <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> Kakamega<br />

Forest (21 ESUs in Buyangu Nature Reserve and <strong>the</strong><br />

areas <strong>of</strong> Kambiri, Shanderema, Lugusi and <strong>the</strong> Kisere<br />

fragment, cf. Figure 5-1a and b). In addition, ef<strong>for</strong>ts<br />

were made to sample Isecheno and Yala (6 ESUs,<br />

Figure 5-1c), and Ikuywa (1 ESU, Figure 5-1d).<br />

Figure 5-2<br />

In Budongo Forest, all 30 ESUs were located in <strong>the</strong><br />

<strong>for</strong>est blocks <strong>of</strong> Budongo and Kanyo-Pabidi (cf.<br />

Figure 5-2a-d). Whereas <strong>the</strong> major part was situated<br />

within a 3 km radius around Sonso camp site (Figure<br />

5-2a, 17 ESUs), some remote areas in Kanyo-Pabidi<br />

(Figure 5-2b, 4 ESUs), Busingiro (Figure 5-2c,<br />

3 ESUs) and Kidwera (Figure 5-2d, 5 ESUs) were<br />

sampled as well.<br />

Recent satellite imagery (Landsat 7 ETM+ satellite<br />

scenes acquired on 17 February 2000 <strong>for</strong> Budongo<br />

Forest and 10 January 2003 <strong>for</strong> Kakamega Forest,<br />

available within <strong>the</strong> BIOTA project) was analysed<br />

to determine relative homogeneity or heterogeneity<br />

Location <strong>of</strong> ESUs in Budongo Forest, a) around Sonso Camp site, b) in Kanyo-Pabidi, c) Busingiro and d) Kidwera<br />

(background: Landsat ETM+scene, 17 February, 2000).


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

<strong>of</strong> surface reflectance in <strong>the</strong> proposed area. Here<br />

especially brightness values <strong>of</strong> ETM bands 3, 4 and<br />

5 were taken into account. As larger changes in <strong>the</strong><br />

<strong>for</strong>est canopy could have occurred in between <strong>the</strong><br />

acquisition <strong>of</strong> <strong>the</strong> older satellite data and <strong>the</strong> time<br />

<strong>of</strong> fieldwork, visual assessments were accomplished<br />

in <strong>the</strong> field be<strong>for</strong>e sampling <strong>the</strong> respective ESU.<br />

Proposed ESUs were only accepted if <strong>the</strong> site and its<br />

surroundings proved to be homogeneous in structure<br />

(at ESU scale) and <strong>the</strong> topography was flat in order to<br />

minimize errors due to terrain effects.<br />

Within each ESU measurements were taken along<br />

three transects (Figure 5-3). In Budongo Forest<br />

transects established <strong>for</strong> chimpanzee research could<br />

be used in all <strong>for</strong>est stages except <strong>the</strong> early ones. In<br />

Kakamega Forest existing transects established by<br />

BIOTA scientists could be used on 14 <strong>of</strong> <strong>the</strong> 30 ESUs.<br />

On <strong>the</strong> o<strong>the</strong>rs new transects had to be established.<br />

Examples from both test sites are displayed in<br />

Figure 5-4.<br />

In order to determine <strong>the</strong> correct distance between <strong>the</strong><br />

individual measurements, two tapes <strong>of</strong> 100 m length<br />

each were laid out on <strong>the</strong> transects. GPS readings<br />

were taken with a non-differential GPS (GARMIN<br />

GPS III) at <strong>the</strong> beginning and end <strong>of</strong> <strong>the</strong> transects.<br />

Care was taken to only include readings with dilution<br />

<strong>of</strong> precision < 5 (i.e. satellites well distributed over<br />

<strong>the</strong> sky) and error <strong>of</strong> precision < 10 m, so that most<br />

measurements should be accurate to within 20 m. If<br />

valid readings could not be retrieved <strong>for</strong> a certain<br />

point, coordinates were calculated <strong>based</strong> on readings<br />

taken at o<strong>the</strong>r locations along <strong>the</strong> respective transect.<br />

5.1.3<br />

Measurement set-up<br />

LAI-2000 Plant Canopy Analyzer<br />

To calculate transmission (cf. Equation 3.4), reference<br />

readings were collected by a so-called “A sensor”<br />

Figure 5-3<br />

Figure 5-4<br />

71<br />

Modified sampling scheme (crosses<br />

represent single measurements; note that <strong>the</strong><br />

distances in between varied according to canopy<br />

height, cf. chapter 5.1.1).<br />

Transects on ESUs in a) Kakamega Forest<br />

(ESU 2) and b) Budongo Forest (ESU 9).<br />

that was placed outside <strong>the</strong> <strong>for</strong>est or in <strong>for</strong>est glades<br />

as close as possible to <strong>the</strong> respective ESU (Figure<br />

5-5). The sensor was levelled and <strong>the</strong> unit set to<br />

remote logging mode with a sampling frequency <strong>of</strong><br />

60 seconds. A shading construction behind <strong>the</strong> sensor<br />

(outside its field <strong>of</strong> view) assured that no direct<br />

radiation could hit <strong>the</strong> lens and cause errors due to<br />

reflections. As ideal illumination conditions <strong>for</strong><br />

LAI-2000 PCA measurements, i.e. diffuse radiation<br />

only, are seldom present in <strong>the</strong> tropics and sunset<br />

and sunrise are too short <strong>for</strong> sampling a whole ESU,<br />

measurements were also made during <strong>the</strong> day (cf. De


72<br />

Wasseige et al. 2003). The influence <strong>of</strong> changing sky<br />

conditions on <strong>the</strong> measurements was recorded with<br />

a stationary LAI-2000 PCA unit set up at <strong>the</strong> ESU<br />

center (B1 sensor, cf. Figure 5-6a). The device took<br />

continuous measurements with <strong>the</strong> same frequency as<br />

<strong>the</strong> A sensor. This later helped to analyse <strong>the</strong> influence<br />

<strong>of</strong> θ and errors introduced by <strong>the</strong> spatial distance<br />

sun<br />

between <strong>the</strong> ESU and <strong>the</strong> A device due to moving<br />

cloud cover. Following <strong>the</strong> recommendations <strong>of</strong><br />

Leblanc & Chen (2001) <strong>for</strong> <strong>the</strong> use <strong>of</strong> LAI-2000 PCA<br />

under sunny conditions, <strong>the</strong> sensors were always<br />

oriented towards north, i.e. away from <strong>the</strong> sun. The<br />

third device was used <strong>for</strong> readings along transects<br />

(B2 sensor). Prior to fieldwork, <strong>the</strong> instruments<br />

were intercalibrated under diffuse light conditions<br />

(LI-COR 1992, Weiss 2002). Intercalibration was<br />

repeated several times during <strong>the</strong> field campaigns.<br />

As LAI is highly non-linear to gap fraction (cf.<br />

Equation 3.3), but each reading represents a linear<br />

average <strong>of</strong> radiation recorded in <strong>the</strong> azimuthal view<br />

range, gaps in <strong>the</strong> canopy will be overweighted if<br />

<strong>the</strong> sensor sees large gaps and dense foliage at <strong>the</strong><br />

same time. To minimize this error and <strong>the</strong> subsequent<br />

underestimation <strong>of</strong> LAI, all three sensors were<br />

equipped with a view cap with a 45° azimuth opening.<br />

Measurement <strong>of</strong> ground cover with <strong>the</strong> LAI-2000 PCA<br />

was avoided, as <strong>the</strong> instrument also responds non-<br />

Figure 5-5<br />

linearly to light interception by vegetation close to<br />

<strong>the</strong> sensor (Hyer & Goetz 2004).<br />

With <strong>the</strong> LAI-2000 PCA a minimum distance to<br />

foliage has to be assured that varies according to<br />

stand architecture, leaf size and to <strong>the</strong> view cap<br />

employed (LI-COR 1992). Consequently <strong>for</strong> many<br />

studies in <strong>for</strong>est environments LAI was sampled<br />

above <strong>the</strong> understorey vegetation (Mussche et al.<br />

2001, Stenberg et al. 2004). De Wasseige et al.<br />

(2003) fur<strong>the</strong>r recommends a minimum distance<br />

to understorey foliage <strong>of</strong> 2 m when measuring<br />

in tropical rain <strong>for</strong>ests. Consequently <strong>the</strong> optical<br />

sensor was levelled at 80 cm above ground inside<br />

<strong>the</strong> <strong>for</strong>est, thus disregarding <strong>the</strong> possible influence<br />

<strong>of</strong> understorey vegetation below that height (cf.<br />

Figure 5-6b). In <strong>the</strong> wooded grasslands outside <strong>the</strong><br />

<strong>for</strong>est <strong>the</strong> sensor was held as close to ground level<br />

as possible. Measurements were repeated 4 times at<br />

each location.<br />

Digital hemispherical photography<br />

DHPs were acquired on <strong>the</strong> same spots and at <strong>the</strong><br />

same heights as LAI-2000 PCA measurements in<br />

order to be able to compare both methods. They<br />

were obtained with a Nikon Coolpix 4300 camera<br />

Schematic illustration <strong>of</strong> measurement set-up with three LAI-2000 PCA devices (shaded area represents <strong>for</strong>est).


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

in Budongo Forest and a Nikon Coolpix 4500 (both<br />

Nikon Corporation, Tokyo, Japan) in Kakamega<br />

Forest (courtesy <strong>of</strong> Frédéric Baret, INRA, France and<br />

Jörg Szyarzynski, <strong>for</strong>merly ZEF, Bonn). Both have a<br />

charge-coupled-device (CCD) array with 4.1 million<br />

pixels (2272 x 1704). The cameras were equipped<br />

with a Nikon FC-E8 fisheye converter (183° field <strong>of</strong><br />

view) with equidistant projection. Images were stored<br />

in JPEG <strong>for</strong>mat in fine resolution mode <strong>for</strong> storage<br />

reasons (compression factor 1:4). Frazer et al. (2001)<br />

found no significant influence <strong>of</strong> this compression<br />

on mean stand estimates <strong>of</strong> LAI compared to<br />

uncompressed TIFF images.<br />

The camera was mounted on a tripod, levelled at<br />

each location and oriented (cf. Figure 3-4) so that<br />

magnetic north was always at <strong>the</strong> top <strong>of</strong> <strong>the</strong> picture.<br />

Images were taken with 1 f-stop underexposure as<br />

underexposure was found to give better contrasts<br />

between foliage and sky (Frazer et al. 2001, Hale &<br />

Edwards 2002, Leblanc et al. 2005, Mussche et al.<br />

2001). Quality and contrast were controlled according<br />

to Leblanc et al. (2005) using <strong>the</strong> instant viewing<br />

mode. To capture LAI <strong>of</strong> <strong>the</strong> understorey vegetation,<br />

additional DHPs were taken at <strong>the</strong> same height with<br />

<strong>the</strong> camera looking downwards.<br />

5.2<br />

5.2.1<br />

Processing, correction and<br />

error analysis<br />

LAI-2000 PCA<br />

Data processing and LAI calculation<br />

Files stored in <strong>the</strong> LAI-2070 control unit were<br />

downloaded and processed with <strong>the</strong> FV 2000 s<strong>of</strong>tware<br />

(LI-COR 2005). Prior to LAI calculation, false<br />

readings (e.g. double readings) were removed. Then<br />

B1/B2 readings were combined with <strong>the</strong> temporally<br />

closest A record to calculate transmission (Equation<br />

3.4). LAI was calculated according to Equation 3.9<br />

Figure 5-6<br />

Measurement set-up <strong>of</strong> LAI-2000 PCA.<br />

a) Stationary set-up <strong>of</strong> B1 instrument on ESU 5<br />

in Budongo Forest, b) measurements with B2<br />

instrument along transect 2 on <strong>the</strong> same ESU.<br />

73<br />

<strong>based</strong> on rings 1-4 only as <strong>for</strong>est gaps, in which <strong>the</strong> A<br />

sensor was placed, were not always large enough in<br />

diameter to assure that that ring 5 was not influenced<br />

by vegetation. No corrections <strong>for</strong> clumping were<br />

made, so that <strong>the</strong> output refers to effective LAI, from<br />

now on called LAI . As Leblanc & Chen (2001)<br />

e (LAI2000)<br />

recommended using solely <strong>the</strong> fourth ring when<br />

measurements are made under sunlit conditions, this<br />

calculation was also per<strong>for</strong>med.<br />

Figure 5-7 shows LAI e (LAI2000) calculated from rings<br />

1-4 plotted against LAI e (LAI2000) calculated from<br />

ring 4 only. The results indicate that despite a high<br />

correlation between both data sets (adjusted R2 =0.97),<br />

mean LAI values retrieved from ring 1-4<br />

e (LAI2000)<br />

are with 4.51 (±2.17) slightly higher than those<br />

calculated from ring 4 only (4.18±2.36). In order to<br />

test whe<strong>the</strong>r both samples differ significantly from<br />

each o<strong>the</strong>r, a non-parametric Wilcoxon signed-rank<br />

test was per<strong>for</strong>med. Although measurements from<br />

ring 4 are attributed <strong>the</strong> highest weight <strong>of</strong> all rings<br />

(cf. Table 3-4) and <strong>the</strong>re<strong>for</strong>e contribute significantly<br />

to <strong>the</strong> calculation <strong>of</strong> LAI from rings 1-4, <strong>the</strong><br />

e (LAI2000)<br />

results show a significant difference (p < 0.001%).<br />

Two arguments can, however, be made to justify <strong>the</strong><br />

usage <strong>of</strong> LAI <strong>based</strong> on rings 1-4: first, one <strong>of</strong><br />

e (LAI2000)


74<br />

<strong>the</strong> underlying assumptions <strong>of</strong> Equation 3.9 is that<br />

gap fraction is integrated over <strong>the</strong> whole range <strong>of</strong><br />

zenith angles from 0 to π/2. This range is never fully<br />

met by <strong>the</strong> LAI-2000 PCA (cf. Chapter 3.3.2), with<br />

<strong>the</strong> outmost ring being always overweighted, yet it is<br />

better represented by four rings instead <strong>of</strong> only one.<br />

Second, LAI is likely to underestimate true<br />

e (LAI2000)<br />

LAI (cf. Fassnacht et al. 1994). As LAI <strong>based</strong><br />

e (LAI2000)<br />

on rings 1-4 has a higher mean, it is probably closer<br />

to true LAI.<br />

Figure 5-7 Scatter plot showing LAI calculated from<br />

e (LAI2000)<br />

rings 1-4 versus LAI calculated from<br />

e (LAI2000)<br />

Table 5-2<br />

ring 4 only (N=12,659).<br />

Quantification <strong>of</strong> instrument inherent error<br />

In order to quantify <strong>the</strong> instrument inherent error <strong>of</strong> <strong>the</strong><br />

LAI-2000 PCA, an experiment was conducted under<br />

stable light conditions in <strong>the</strong> spectroscopy laboratory<br />

<strong>of</strong> <strong>the</strong> Remote Sensing Technology Institute (Institut<br />

highe high<br />

LAI.<br />

Figu<br />

LAIe LAIe LAI<br />

versu vers<br />

only<br />

für Methodik der Fernerkundung, IMF) at <strong>the</strong> German<br />

only<br />

Quantification Aerospace Quantification Center <strong>of</strong> (Deutsches <strong>of</strong> instrument Zentrum inherent für Luft- error und<br />

In Raumfahrt In order to e. quantify V., DLR). <strong>the</strong> instrument Measurements inherent were error made <strong>of</strong> <strong>the</strong> LAI-2000<br />

under under stable a stable calibrated light conditions lamp <strong>for</strong> six in in hours <strong>the</strong> spectroscopy (no vegetation laboratory <strong>of</strong> <strong>the</strong> R<br />

(Institut present). (Institut für The für Methodik analysis <strong>of</strong> der <strong>the</strong> Fernerkundung, results indicated IMF) that <strong>the</strong> at at <strong>the</strong> German Aer Ae<br />

für instrument für Luft- und noise Raumfahrt is negligible e. e. V., DLR). with DLR). a Measurements coefficient <strong>of</strong> were made unde und<br />

vegetation variation vegetation (c present). ) <strong>of</strong> 1% or below <strong>for</strong> each single ring (cf.<br />

v present). The analysis <strong>of</strong> <strong>the</strong> results indicated that <strong>the</strong> in ini<br />

coefficient Table coefficient 5-2), <strong>of</strong> with <strong>of</strong> variation c being defined as<br />

v (cv) <strong>of</strong> 1% or below <strong>for</strong> each single ring (cf. Tab Ta<br />

<br />

<br />

c v <br />

. v . .<br />

<br />

<br />

(5.1)<br />

and are standard deviation and and mean mean <strong>of</strong> <strong>the</strong> <strong>of</strong> measurements.<br />

<strong>the</strong><br />

measurements.<br />

Table 5-2: Statistical measures <strong>for</strong> light intensity recorded by rings 1-4<br />

conditions (N= 397).<br />

Influence Ring <strong>of</strong> illumination N conditions Minimum Maximum <br />

1 1 397 0.506 0.518 0.512 0.51<br />

In order 2 2 to analyse 397 <strong>the</strong> effect <strong>of</strong> 0.760 non-ideal ideal 0.776 0.767 0.76<br />

illumination 3 3 conditions 397 on LAI-2000 0.636 PCA 0.660 0.651 0.65<br />

4 4 397 0.185 0.198<br />

measurements, test measurements were conducted 0.198<br />

on 30 October, 2005 in Budongo Forest over <strong>the</strong><br />

whole day. The three devices were put up around<br />

Sonso Camp. The A sensor was placed at <strong>the</strong> camp<br />

side <strong>for</strong> reference readings, <strong>the</strong> B1 and B2 sensors in<br />

two different locations in secondary <strong>for</strong>est close to<br />

<strong>the</strong> camp. The canopy on both sites had a relatively<br />

dense understorey with trees <strong>of</strong> up to 3 m height. The<br />

remote logging mode <strong>of</strong> all devices was set to one<br />

minute. Figure 5-8 indicates <strong>the</strong> problems related to<br />

changing sun zenith angle and moving clouds. Figure<br />

5-8a shows LAI values plotted against θ .<br />

e (LAI2000) sun<br />

0.193 0.19<br />

Statistical measures <strong>for</strong> light intensity recorded by rings 1-4 <strong>of</strong> LAI-2000 PCA under laboratory conditions (N= 397).<br />

Ring N Minimum Maximum μ σ c v<br />

1 397 0.506 0.518 0.512 0.002 > 0.01<br />

2 397 0.760 0.776 0.767 0.003 > 0.01<br />

3 397 0.636 0.660 0.651 0.005 0.01<br />

4 397 0.185 0.198 0.193 0.002 0.01


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

It is obvious that LAI e (LAI2000) shows a certain variation<br />

throughout <strong>the</strong> day, probably due to moving clouds<br />

and scattering effects. What is remarkable, however,<br />

is <strong>the</strong> strong decrease in LAI with higher<br />

e (LAI2000)<br />

θ , i.e. with <strong>the</strong> sun being close to <strong>the</strong> horizon. If<br />

sun<br />

LAI values at θ > 70° are ignored, mean<br />

e (LAI2000) sun<br />

LAI over <strong>the</strong> day is 7.19 (cf. Table 5-5). At<br />

e (LAI2000)<br />

θ higher than 70° LAI strongly decreases,<br />

sun e (LAI2000)<br />

with a minimum <strong>of</strong> 2.54 at 90°. Transmission values<br />

plotted against θ in turn show a strong increase<br />

sun<br />

at higher θ (Figure 5-8b). As transmission is<br />

sun<br />

calculated as quotient <strong>of</strong> I (θ) and I (θ) (see<br />

trans o<br />

Equation 3.4) this indicates that <strong>the</strong> LAI-2000 PCA<br />

is probably not sensitive enough <strong>for</strong> below canopy<br />

measurements under a dense vegetation canopy<br />

and faint light conditions (see also De Wasseige et<br />

al. 2003). Light variations at high zenith angles are<br />

obviously not captured by <strong>the</strong> B1 sensor inside <strong>the</strong><br />

<strong>for</strong>est, whereas <strong>the</strong> A sensor outside is still sensitive<br />

(cf. Figures 5-8c and d). The restriction to a 45° field<br />

<strong>of</strong> view additionally contributes to this error as less<br />

radiation can be recorded by <strong>the</strong> device. Figure 5-9, a<br />

hemispherical picture taken besides <strong>the</strong> B1 sensor at<br />

6:45 p.m. illustrates <strong>the</strong> dense canopy and <strong>the</strong> small<br />

amount <strong>of</strong> light reaching <strong>the</strong> <strong>for</strong>est floor at that time.<br />

In order to test <strong>the</strong> dependence <strong>of</strong> LAI from<br />

e (LAI2000)<br />

θ , erroneous measurements were excluded <strong>for</strong> all<br />

sun<br />

three devices (A, B1 and B2), i.e. values recorded<br />

at θ >70°. Table 5-3 displays <strong>the</strong> descriptive<br />

sun<br />

statistics <strong>of</strong> <strong>the</strong> resulting data sets. Mean LAIe (LAI2000)<br />

values calculated from A/B1 and A/B2 sensors are<br />

very similar (7.19 and 7.27) as well as <strong>the</strong> standard<br />

deviations (0.29 and 0.27 respectively). The relatively<br />

large range <strong>of</strong> values (2.36 and 2.71) however<br />

indicates a strong variation over <strong>the</strong> day, probably<br />

due to errors introduced by changes in blue light<br />

scattering (changing θ ) and moving clouds.<br />

sun<br />

A Kolmogorov-Smirnov test was applied to test<br />

whe<strong>the</strong>r LAI is normally distributed (Sachs<br />

e (LAI2000)<br />

2004). As <strong>the</strong> resulting probability <strong>of</strong> error p was<br />

Figure 5-8<br />

Figure 5-9<br />

Test measurements with LAI-2000 PCA made<br />

on 30 October, 2005. a) LAI calculated<br />

e (LAI2000)<br />

from continuous A/B1 readings, b) transmission<br />

calculated from continuous A/B2 readings. The<br />

respective solar zenith angle is shown on <strong>the</strong> x<br />

axis.<br />

75<br />

Hemispherical photograph taken alongside to <strong>the</strong><br />

above-mentioned B1 sensor at 6:45 p.m.


76<br />

in both cases > 0.05, <strong>the</strong> null hypo<strong>the</strong>sis (normal<br />

distribution) can be accepted. In consequence <strong>the</strong><br />

relation between LAI and θ was tested with<br />

e (LAI2000) sun<br />

Pearson’s correlation coefficient r. For both test sites a<br />

correlation significant at 0.01 level could be retrieved<br />

between LAI and θ , yet with <strong>the</strong> relation<br />

e (LAI2000) sun<br />

between <strong>the</strong> variables being relatively weak (r = 0.39<br />

and 0.35 respectively). As A, B1 and B2 readings<br />

are presumably influenced by both θ effects and<br />

sun<br />

errors introduced by moving clouds (i.e. changing<br />

illumination conditions), <strong>the</strong>se error sources will be<br />

analysed in more detail. Additionally a correction<br />

scheme <strong>for</strong> both effects is proposed in <strong>the</strong> following.<br />

Table 5-3<br />

Correction <strong>of</strong> illumination effects<br />

Apart from <strong>the</strong> instrument inherent error, in situ LAI<br />

calculated with LAI-2000 PCA is mainly biased by<br />

non-diffuse illumination conditions. The following<br />

two error components influence <strong>the</strong> readings:<br />

a) errors introduced by changing sun zenith<br />

angle (and with it <strong>the</strong> amount <strong>of</strong> blue light<br />

scattered within <strong>the</strong> canopy), e , and<br />

θsun<br />

b) errors introduced by differences in incident<br />

radiation due to cloud movement and <strong>the</strong><br />

spatial distance between A sensor and B<br />

sensor, e . cloud<br />

Descriptive statistics <strong>of</strong> continuous LAI-2000 PCA measurements (A/B1 and A/B2) recorded on 30 October, 2005.<br />

A/B1 devices A/B2 devices<br />

LAI recorded at starting time (8:04 a.m.) 6.90 7.48<br />

LAI recorded at ending time (5:16 p.m.) 6.49 7.66<br />

LAI recorded at minimum θ sun (14.86° at 12:40 am) 6.83 6.83<br />

Maximum LAI 8.04 (at 12:45 a.m.) 7.80 (at 8:52 a.m.)<br />

Minimum LAI 5.68 (at 2:21 p.m.) 5.09 (at 5:10 p.m.)<br />

Mean LAI (θ sun < 70°) 7.19 7.27<br />

Standard deviation (θ sun < 70°) 0.29 0.27<br />

N 553 539<br />

Table 5-4<br />

Allocation <strong>of</strong> ESUs to <strong>the</strong> respective radiation regimes and necessary processing steps.<br />

Radiation regime ESU number Necessary processing steps<br />

Diffuse radiation 1, 5, 14, 15, 16, 20, 24, 28 No correction necessary<br />

Direct radiation 2, 3, 7, 11, 18 Correction <strong>for</strong> e θsun<br />

Changing radiation regime 4, 6, 8, 9, 10, 12, 13, 17, 19, 21, 22,<br />

23, 25, 26, 27, 28, 30<br />

Filter <strong>for</strong> valid measurements


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

Figure 5-10 LAI calculated from continuous B1 readings made on ESU 7 on 10 October, 2005, toge<strong>the</strong>r with hemispherical<br />

e (LAI2000)<br />

pictures, taken at θ =38° (10:14 a.m.) and 61° (8:41 a.m.) respectively. Note cloud presence in right, cloud absence<br />

sun<br />

in right picture.<br />

Under perfectly diffuse light conditions, both e θsun<br />

and e cloud are 0. In case <strong>of</strong> clear skies e cloud is 0, so<br />

that LAI e (LAI2000) is influenced by e θsun only. In case<br />

<strong>of</strong> changing illumination conditions, LAI e (LAI2000)<br />

derivation is biased by both e θsun and e cloud .<br />

Figure 5-10 illustrates <strong>the</strong> influence <strong>of</strong> e θsun and e cloud<br />

on LAI e (LAI2000) retrieved from B1 readings taken<br />

on 10 October, 2005 on ESU 7. The A sensor had<br />

been situated in <strong>the</strong> camp. This was <strong>the</strong> closest large<br />

enough clearing and approximately 1.6 km away.<br />

Measurements were made between 8:12 a.m. and<br />

11:20 a.m. (corresponding to θ <strong>of</strong> 67° and 22°). The<br />

sun<br />

morning hours up until 9:47 a.m. (θ = 44°) were<br />

sun<br />

influenced by passing clouds, thus both, e and e θsun cloud<br />

77<br />

biased <strong>the</strong> measurements over both A and B1 sensors.<br />

After 10 a.m. <strong>the</strong> skies remained clear, so that only<br />

e influenced <strong>the</strong> reading.<br />

θsun<br />

As <strong>the</strong> location <strong>of</strong> <strong>the</strong> B1 sensors remained stable<br />

during <strong>the</strong> time <strong>of</strong> measurement, <strong>the</strong> derived<br />

LAI is not influenced by local differences in<br />

e (LAI2000)<br />

canopy structure, unlike <strong>the</strong> B2 device. Consequently<br />

B1 measurements derived <strong>for</strong> each ESU were used <strong>for</strong><br />

analysing and partly correcting <strong>the</strong> above-mentioned<br />

errors. There<strong>for</strong>e ESUs were separated into three<br />

groups <strong>based</strong> on notes made during measurements<br />

and evidence coming from <strong>the</strong> digital hemispherical<br />

pictures (cf. Table 5-4):


78<br />

1) ESUs on which <strong>the</strong> measurements were made<br />

under diffuse radiation conditions,<br />

2) ESUs that were visited under completely clear<br />

skies, and<br />

3) ESUs where <strong>the</strong> conditions changed during <strong>the</strong><br />

time <strong>of</strong> measurement.<br />

As expected, <strong>the</strong> majority <strong>of</strong> ESUs (17 out <strong>of</strong> 30) visited<br />

had experienced a changing radiation regime over<br />

time. As overcast skies represent ideal conditions <strong>for</strong><br />

LAI-2000 PCA measurements, no fur<strong>the</strong>r correction<br />

is applied to group 1. In this case, <strong>the</strong> mean LAIe (LAI2000)<br />

<strong>for</strong> <strong>the</strong> respective ESU was calculated directly from<br />

A and B2 measurements. For <strong>the</strong> o<strong>the</strong>r two cases<br />

correction schemes or filters had to be applied.<br />

Figure 5-11 LAI derived from continuous B1<br />

e (LAI2000)<br />

measurements on ESUs 16 (blue), 18 (red)<br />

and 30 (green).<br />

Figure 5-11 shows LAI calculated from B1<br />

e (LAI2000)<br />

readings on three ESUs <strong>for</strong> different θ . Each ESU<br />

sun<br />

represents one <strong>of</strong> <strong>the</strong> above-mentioned radiation<br />

regimes. Whereas LAI values on ESU 16<br />

e (LAI2000)<br />

(blue) show only little variation, <strong>the</strong> increasing eθsun with decreasing θ is obvious <strong>for</strong> ESU 18 (red). ESU<br />

sun<br />

30 (green) is a characteristic example <strong>of</strong> <strong>the</strong> strong<br />

influence <strong>of</strong> e on derived LAI .<br />

cloud e (LAI2000)<br />

Table 5-5 indicates that, as expected, ESU 16 has<br />

<strong>the</strong> smallest c v <strong>of</strong> LAI e (LAI2000) values with 1%. This<br />

corresponds to <strong>the</strong> instrument inherent error. Although<br />

c <strong>for</strong> ESU 18 also seems to be relatively small (3%),<br />

v<br />

it could be much higher at ESUs measured at smaller<br />

θ sun since under direct radiation conditions e θsun is<br />

assumed to be 0 at dawn (only diffuse radiation<br />

present) and increasing with decreasing θ (amount<br />

sun<br />

<strong>of</strong> direct radiation penetrating <strong>the</strong> canopy increases).<br />

The largest c v was retrieved <strong>for</strong> ESU 30. Here <strong>the</strong><br />

variation accounts <strong>for</strong> 10% <strong>of</strong> <strong>the</strong> arithmetic mean and<br />

individual LAI values show a range <strong>of</strong> almost<br />

e (LAI2000)<br />

2. Certainly <strong>the</strong> large variation is also a function <strong>of</strong><br />

distance between <strong>the</strong> A and B1/B2 sensors.<br />

To develop a correction <strong>for</strong> e , all five ESUs that<br />

θsun<br />

were measured at least partly under cloud-free<br />

conditions were analysed. B1 measurements taken<br />

under changing radiation conditions (as e.g. at<br />

θ > 44° on ESU 7, cf. Figure 5-10) were excluded<br />

sun<br />

from analysis. To compare LAI values from different<br />

ESUs LAI was centred <strong>for</strong> θ = 50° (as LAI sun e (LAI2000)<br />

values measured at a sun zenith angle <strong>of</strong> 50° were<br />

present on all ESUs) with<br />

Table 5-5 Descriptive statistics <strong>for</strong> LAI derived from continuous B1 measurements on <strong>the</strong> ESUs shown in Figure 5-10.<br />

e (LAI2000)<br />

ESU N Minimum Maximum μ σ c v<br />

16 117 8.23 8.60 8.38 0.070 0.01<br />

18 118 6.09 6.77 6.45 0.166 0.03<br />

30 130 3.08 5.07 3.98 0.394 0.10


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

cent. LAI e (LAI2000) = LAI e (LAI2000) -LAI e (LAI2000, 50°) ,<br />

(5.2)<br />

where LAI e (LAI2000, 50°) is LAI e (LAI2000) measured at<br />

θ sun = 50° on <strong>the</strong> respective ESU. Consequently<br />

centered LAI e (LAI2000) is 0 at θ sun = 50° with negative<br />

values at smaller and positive values at higher sun<br />

zenith angles on average. Figure 5-12a displays<br />

centered LAI values <strong>for</strong> <strong>the</strong> five ESUs<br />

e (LAI2000)<br />

measured under direct radiation plotted against θ . sun<br />

Linear regression was per<strong>for</strong>med subsequently with<br />

θ as an independent variable. The relation between<br />

sun<br />

centered LAI and θ can thus be modelled as<br />

e (LAI2000) sun<br />

cent. LAI e (LAI2000) = 0.012* θ sun - 0.532.<br />

(5.3)<br />

Presumably LAI values measured at very<br />

e (LAI2000)<br />

high sun zenith angles (i.e. close to dawn or dusk<br />

with mostly diffuse radiation) are not influenced by<br />

e θsun , so that <strong>the</strong>y represent <strong>the</strong> actual LAI e (LAI2000) . In<br />

order to avoid sensitivity problems with <strong>the</strong> device<br />

(cf. Figure 5-8), centered LAI at θ = 80°<br />

e (LAI2000) sun<br />

(and not 90°) was considered to be unaffected by e . θsun<br />

Consequently a correction factor θ /LAI sun e (LAI2000, 80°)<br />

could be derived <strong>for</strong> each ESU, with LAIe (LAI2000, 80°)<br />

being centered LAI calculated with Equation<br />

e (LAI2000)<br />

R²=0.84<br />

Y=-0.532+0.012 X<br />

Figure 5-12 a) Scatter plot <strong>of</strong> LAIe<br />

(LAI2000, cen) and θsun , solid line represents linear regression and b) same data corrected <strong>for</strong> eθsun .<br />

79<br />

5.3 <strong>for</strong> θ = 80°. Figure 5-12b displays this correction<br />

sun<br />

factor applied to centered LAI , which is thus<br />

e (LAI2000)<br />

corrected <strong>for</strong> e θsun . To derive corrected LAI e (LAI2000) <strong>for</strong><br />

each ESU, <strong>the</strong> corrected centered LAI must<br />

e (LAI2000)<br />

be summed up with <strong>the</strong> ESU specific LAIe (LAI2000, 50°)<br />

(cf. Equation 5.2). B2 measurements were corrected<br />

correspondingly, assuming that e had a similar<br />

θsun<br />

effect on B1 and B2 measurements taken at <strong>the</strong> same<br />

θ . Whereas e can thus be eliminated, variability<br />

sun θsun<br />

due to random noise (e.g. caused by sunflecks) still<br />

remains (cf. Figure 5-12b).<br />

For several reasons it is difficult to establish a<br />

correction <strong>for</strong> e . First <strong>of</strong> all, <strong>the</strong> changing radiation<br />

cloud<br />

regimes and proportions <strong>of</strong> direct and diffuse<br />

radiation influencing A and B1 sensors can probably<br />

not be reconstructed. Modelled curves <strong>of</strong> light<br />

intensity would be needed each day <strong>for</strong> <strong>the</strong> reference<br />

site and <strong>the</strong> ESU. Whereas this could probably be<br />

accomplished <strong>for</strong> <strong>the</strong> A sensor (as it was usually<br />

taking readings over <strong>the</strong> whole day), problems occur<br />

with <strong>the</strong> B1 sensor, as measurements rarely cover<br />

more than three hours per ESU. Effects introduced by<br />

moving clouds can thus not sufficiently be modelled<br />

and corrected <strong>for</strong>.


80<br />

Consequently instead <strong>of</strong> developing a correction<br />

<strong>for</strong> e , a filter was applied to LAI transect<br />

cloud e (LAI2000)<br />

measurements that were taken under changing sky<br />

conditions. B1 measurements served here as “quality<br />

control”. Mean LAI and standard deviation<br />

e (LAI2000)<br />

were calculated first <strong>for</strong> B1 derived LAI <strong>for</strong> each<br />

ESU. If <strong>the</strong> corresponding value was within μ±σ, <strong>the</strong><br />

B1 value was declared as valid. LAI calculated<br />

e (LAI2000)<br />

from transect measurements were <strong>the</strong>n compared to<br />

<strong>the</strong> time-corresponding B1 measurements. If <strong>the</strong>se<br />

were valid, LAI calculated from B2 were<br />

e (LAI2000)<br />

included in fur<strong>the</strong>r analysis.<br />

5.2.2<br />

DHP<br />

Data processing and LAI calculation<br />

All DHPs were processed using <strong>the</strong> CAN-<br />

EYE s<strong>of</strong>tware (Version 4.2, cf. Baret 2004a).<br />

Compared to o<strong>the</strong>r available s<strong>of</strong>tware packages<br />

(e.g. GapLightAnalyzer by Frazer et al. 1999, or<br />

HemiView Canopy Analysis S<strong>of</strong>tware by Delta-T<br />

Devices Ltd., 1999) <strong>the</strong> programme has several<br />

advantages, such as, <strong>for</strong> instance, a simultaneous<br />

analysis <strong>of</strong> up to 20 pictures acquired under <strong>the</strong><br />

same canopy and illumination conditions (with<br />

most o<strong>the</strong>r s<strong>of</strong>tware packages, pictures have to be<br />

analysed individually). In addition, <strong>the</strong> s<strong>of</strong>tware is<br />

<strong>based</strong> on colour photography, which simplifies <strong>the</strong><br />

classification process and is also more accurate than<br />

single- or two-value thresholds <strong>based</strong> on black and<br />

white pictures. Be<strong>for</strong>e processing, <strong>the</strong> optical centre<br />

<strong>of</strong> <strong>the</strong> camera system was determined. Azimuthal and<br />

zenithal computations are <strong>based</strong> on this, but usually<br />

<strong>the</strong>re is slight deviation from <strong>the</strong> centre point <strong>of</strong> <strong>the</strong><br />

images due to lens distortions. Not correcting this<br />

factor would introduce an error to LAI calculation.<br />

The determination <strong>of</strong> <strong>the</strong> optical centre was done<br />

according to <strong>the</strong> method proposed by Baret (2004b).<br />

The black cross in Figure 5-13 (see below) marks <strong>the</strong><br />

centre point <strong>of</strong> an image (pixel coordinates 1136, 852)<br />

taken with <strong>the</strong> Nikon Coolpix 4300 camera. Blue dots<br />

mark <strong>the</strong> position <strong>of</strong> holes in an objective lid that had<br />

been rotated after image acquisition. The procedure<br />

was repeated 12 times. As <strong>the</strong> holes rotate around<br />

<strong>the</strong> optical centre <strong>of</strong> <strong>the</strong> system, its coordinates can<br />

be calculated. For <strong>the</strong> Nikon Coolpix 4300 system,<br />

<strong>the</strong> optical centre is slightly shifted towards to<br />

<strong>the</strong> upper left compared to <strong>the</strong> image centre (from<br />

pixel coordinates 1180, 862 to 1163, 852). Be<strong>for</strong>e<br />

processing, all images were fur<strong>the</strong>r checked <strong>for</strong><br />

quality, i.e. same illumination conditions <strong>for</strong> pictures<br />

processed toge<strong>the</strong>r.<br />

The CAN-EYE s<strong>of</strong>tware starts with <strong>the</strong> definition <strong>of</strong><br />

a calibration parameter file that identifies <strong>the</strong> image<br />

dimensions, <strong>the</strong> optical centre, as well as <strong>the</strong> circle <strong>of</strong><br />

interest (COI) in °, i.e. <strong>the</strong> zenithal area to be analysed<br />

(cf. Table A-5 <strong>for</strong> <strong>the</strong> camera system used in Budongo<br />

Forest). According to Frédéric Baret (personal<br />

communication), <strong>the</strong> COI should be restricted to<br />

60° since, depending on <strong>the</strong> camera system, <strong>the</strong><br />

image sharpness may decrease significantly at larger<br />

zenith angles and mixed pixels (i.e. containing gaps<br />

and foliage) occur. Also Frazer et al. (2001) noticed<br />

strong blurring at zenith angles higher than 45° with<br />

1,500<br />

1,000<br />

500<br />

0<br />

0 500 1,000 1,500 2,000<br />

Figure 5-13<br />

Optical centre (red cross) <strong>of</strong> Nikon Coolpix 4300<br />

camera used in Budongo Forest. X and Y units<br />

are pixels.


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

a Nikon Coolpix 950 camera and FC-E8 fisheye.<br />

However, <strong>the</strong> loss <strong>of</strong> fine canopy details in <strong>the</strong>ir<br />

study may partly be attributed to <strong>the</strong> comparably low<br />

resolution <strong>of</strong> <strong>the</strong>ir camera system (1.92 Megapixel).<br />

For processing, all images are divided into small<br />

sectors <strong>of</strong> 2.5° zenith and 5° azimuth angular<br />

resolution (cf. Figure 5-14). The <strong>for</strong>mer are needed to<br />

compare derived gap fraction with <strong>the</strong> modelled one<br />

Figure 5-14<br />

Figure 5-15<br />

Illustration <strong>of</strong> DHP area used <strong>for</strong> LAI derivation:<br />

COI (60°) split up into 2.5° zenith and 5° azimuth<br />

sectors.<br />

Subset <strong>of</strong> hemispherical photograph as a) original and b) classified image. Red circles illustrate <strong>the</strong> problem <strong>of</strong> mixed<br />

pixels and <strong>the</strong>ir classification.<br />

81<br />

(cf. Equation 5.6) and estimate LAI. Azimuth sectors<br />

help to derive <strong>the</strong> clumping index.<br />

As <strong>the</strong> s<strong>of</strong>tware cannot cope with more than 20<br />

pictures per ESU, processing was done on transect<br />

basis and compiled later with an extension <strong>of</strong> CAN-<br />

EYE developed by Weiss (2006). The images were<br />

masked if necessary (appearance <strong>of</strong> solar disk,<br />

illuminated stems and branches, feet <strong>of</strong> <strong>the</strong> user<br />

in case <strong>of</strong> downward-looking pictures, etc.). After<br />

a reduction to 324 colours, gaps are classified<br />

interactively (cf. Figure A-3 <strong>for</strong> illustration).<br />

A critical step in determining gap fraction from<br />

hemispherical photographs is <strong>the</strong> classification<br />

process, as illustrated in Figure 5-15. Transitions<br />

between sky and foliage are not always sharp (Figure<br />

5-15a), so that mixed pixels occur. This makes <strong>the</strong><br />

selection <strong>of</strong> pixels belonging to gaps in <strong>the</strong> canopy<br />

somewhat arbitrary and subjective (cf. classified<br />

image in Figure 5-15b). The consequences <strong>of</strong> this<br />

operator effect will be analysed later.<br />

Based on <strong>the</strong> classification, gap fraction is computed<br />

<strong>for</strong> each sector as <strong>the</strong> relation <strong>of</strong> pixels classified as<br />

foliage and non-foliage. In principle and analogue<br />

to <strong>the</strong> LAI retrieval with LAI-2000 PCA, Miller’s<br />

derivation (see Equation 3.7) could be used to solve


82<br />

inclination angle (WEISS et al. 2004)<br />

(2) The second method is <strong>based</strong> on model inversion. Assuming an<br />

Fig. 5-16: Variation <strong>of</strong> <strong>the</strong> G-function (mean projection <strong>of</strong> unit foliage area) with <strong>the</strong> average leaf<br />

inclination angle (WEISS et al. 2004) a large range <strong>of</strong> random LAI and ALA combinations and th<br />

(2) The second method is <strong>based</strong> on model inversion. defined zenith Assuming angles an are ellipsoidal stored in leaf a LUT. inclination Possible distribution, LAI values<br />

Equation 3.3 and compute LAI from gap fraction weighted cost-function, where <strong>the</strong> distance <strong>of</strong><br />

derived from hemispherical photographs. As already <strong>the</strong> k<br />

mentioned, <strong>the</strong> basic assumption with Miller’s<br />

<strong>for</strong>mula however is that gap fraction is integrated<br />

from 0 to π/2. Gap fraction retrieved from DHP at<br />

larger zenith angles might, however, include large<br />

uncertainties due to blurring. There<strong>for</strong>e an approach,<br />

that allows LAI retrieval from gap fraction at lower<br />

zenith angles only, is preferred.<br />

th a large range <strong>of</strong> random LAI and ALA and combinations ALA 10° and 80° <strong>the</strong> (in corresponding steps <strong>of</strong> 2°). The P LUT element close<br />

o values <strong>for</strong> userdefined<br />

zenith angles are stored in a LUT.<br />

chosen<br />

Possible element <strong>based</strong><br />

LAI<br />

on<br />

values <strong>of</strong> a <strong>the</strong> weighted<br />

lie LUT between to cost-function, measured 0 and 10 gap (in<br />

where<br />

steps <strong>of</strong><br />

<strong>the</strong><br />

0.01)<br />

distanc<br />

and ALA 10° and 80° (in steps <strong>of</strong> 2°). The<br />

measured fraction LUT is element<br />

gap computed fraction<br />

closest as is computed<br />

to <strong>the</strong> derived<br />

as<br />

gap fraction is <strong>the</strong>n<br />

chosen <strong>based</strong> on a weighted cost-function, where no.<br />

zenith <strong>the</strong> angles distance <strong>of</strong> <strong>the</strong> k<br />

LUT ( k ) MES 2<br />

wi<br />

P0( i<br />

) P0<br />

( i<br />

) <br />

LUT ( k)<br />

i1<br />

ALA 60<br />

ck<br />

<br />

<br />

MES<br />

( ( ))<br />

30<br />

MOD P0<br />

i<br />

th element <strong>of</strong> <strong>the</strong> LUT to<br />

measured gap fraction is computed as<br />

no.<br />

zenith angles<br />

LUT ( k ) MES 2<br />

wi<br />

P0( i<br />

) P0<br />

( i<br />

) <br />

LUT ( k)<br />

i1<br />

ALA 60<br />

c <br />

<br />

(5.6)<br />

k<br />

(5.6)<br />

MES<br />

( ( ))<br />

30<br />

MOD P0<br />

i<br />

IN SITU LAI MEASUREMENTS<br />

17<br />

16<br />

LUT(k) MES The IN SITU CAN-EYE LAI MEASUREMENTS s<strong>of</strong>tware <strong>of</strong>fers two solutions <strong>for</strong> with P (θi ) and P (θi ) being modelled<br />

0<br />

0<br />

16<br />

Equation derived 3.3 from (Weiss hemispherical et al. 2006): photographs. As already mentioned, and measured <strong>the</strong> basic gap assumption fractions <strong>for</strong> with azimuthally Miller’s<br />

derived from hemispherical photographs. As already mentioned, 17<br />

<strong>for</strong>mula however is that gap fraction is integrated from 0 to averaged <strong>the</strong> basic assumption with Miller’s<br />

/2. Gap LUTzenith<br />

( kfraction<br />

) sector retrieved MES i. w from is a DHP weighting<br />

i with P at<br />

0 ( i ) and P0 ( i ) being modelled and measured<br />

(1) <strong>for</strong>mula<br />

larger According however<br />

zenith angles to is that<br />

might, Warren-Wilson gap fraction is<br />

however, include (1963) integrated<br />

large <strong>the</strong> from 0 to<br />

uncertainties function /2. Gap fraction retrieved from DHP at<br />

zenith due sector to blurring. i. wi is a There<strong>for</strong>e weighting an function approach,<br />

that<br />

larger<br />

allows G-function zenith<br />

LAI<br />

angles<br />

retrieval can might, be considered from<br />

however,<br />

gap fraction as include independent at lower<br />

large uncertainties<br />

zenith angles only,<br />

due LUT to<br />

is ( blurring.<br />

preferred.<br />

k ) There<strong>for</strong>e MES an approach,<br />

with P<br />

that allows <strong>of</strong> leaf LAI inclination retrieval at from a zenith gap angle fraction <strong>of</strong> 57.5°, at lower i.e. zenith angles only,<br />

Npix 0<br />

is preferred.<br />

i (<br />

Nmask i ) and P<br />

i 0 ( i ) being modelled and measured<br />

The CAN-EYE s<strong>of</strong>tware <strong>of</strong>fers two solutions <strong>for</strong> Equation 3.3<br />

zenith<br />

w i <br />

sector<br />

(WEISS et Npix al.<br />

i. (5.7)<br />

wi 2006):<br />

is a weighting ,<br />

function<br />

i<br />

The CAN-EYE here G=0.5 (cf. s<strong>of</strong>tware Figure <strong>of</strong>fers 5-16). two Based solutions on Warren <strong>for</strong> Equation 3.3 (WEISS Npixet<br />

al. 2006):<br />

i Nmaski<br />

(1) According Wilson’s work, to WARREN-WILSON Bonhomme et (1963) al. (1974) <strong>the</strong> G-function where wcan<br />

i be Npix considered iis<br />

is <strong>the</strong> number as independent <strong>of</strong> <strong>of</strong> pixels in <strong>of</strong> in zenith leaf sector i and Nm<br />

i<br />

(1) According<br />

inclination derived LAI at<br />

to WARREN-WILSON<br />

<strong>for</strong> a zenith young angle crops and <strong>of</strong> 57.5°,<br />

(1963)<br />

found i.e.<br />

<strong>the</strong><br />

a good here<br />

G-function<br />

G=0.5 (cf.<br />

can<br />

Figure<br />

be considered Npix i<br />

5-16). Based<br />

as<br />

on<br />

independent<br />

Warren Wilson’s<br />

<strong>of</strong> leaf<br />

sector i. The function i and Nmask takes is into <strong>the</strong> account number that <strong>of</strong> masked<br />

i some zenith directions<br />

work,<br />

inclination at a zenith<br />

agreement BONHOMME to direct et measurements. al.<br />

angle<br />

(1974)<br />

<strong>of</strong> 57.5°,<br />

derived<br />

i.e.<br />

There<strong>for</strong>e LAI<br />

here<br />

<strong>for</strong><br />

G=0.5<br />

young<br />

(cf.<br />

crops where Figure<br />

and found<br />

5-16). Based<br />

a good<br />

on<br />

agreement<br />

Warren Wilson’s<br />

pixels in and Npix i. is The inormalized<br />

is function <strong>the</strong> number takes by <strong>of</strong> into pixels account to in direct zenith that sector i and Nm<br />

measurements.<br />

work, BONHOMME<br />

LAI can be calculated There<strong>for</strong>e<br />

et al. (1974)<br />

according LAI can<br />

derived<br />

tobe<br />

calculated<br />

LAI <strong>for</strong><br />

according<br />

young crops<br />

to<br />

and found a good agreement to direct<br />

i. some The function zenith directions takes into may account contain that a some large zenith directions m<br />

no.<br />

zenith angles<br />

measurements. There<strong>for</strong>e LAI can be calculated according to<br />

LAI<br />

pixels amount and w<strong>of</strong><br />

is<br />

masked i 1. normalized pixels by and is normalized by<br />

0.<br />

5<br />

P 57. 5<br />

cos 57.<br />

7<br />

o e LAI<br />

i 1<br />

0.<br />

5 , (5.4)<br />

(5.4)<br />

no.<br />

zenith angles<br />

P 57. 5<br />

cos 57.<br />

7<br />

o e , The root (5.4)<br />

mean w square error between modelled and measured gap<br />

i 1.<br />

(5.8)<br />

which is equivalent to<br />

i 1<br />

which is is equivalent to<br />

sectors (cf. Equation 5.6) is divided by a modelled standard deviatio<br />

lnPo57. 5<br />

The root mean square error between modelled and measured gap<br />

LAI <br />

lnP . This modelled standard deviation is derived (5.5) by fitting a second o<br />

0o.<br />

93 57. 5<br />

LAI <br />

. . (5.5) The sectors (5.5)<br />

0.<br />

93<br />

empirical<br />

root (cf. mean Equation<br />

standard<br />

square 5.6)<br />

deviation<br />

error is divided<br />

<strong>of</strong><br />

between by a modelled<br />

measured<br />

modelled standard deviatio<br />

gap fraction in a certai<br />

Yet as this method was verified only <strong>for</strong> agricultural and vegetation This images<br />

measured modelled processed types gap standard and toge<strong>the</strong>r<br />

fractions <strong>the</strong> portion deviation and<br />

over<br />

belonging <strong>of</strong> all is mixed derived user-defined<br />

to pixels <strong>the</strong> by same fitting transect. a second Tho<br />

Yet<br />

increases Yet as as this this with<br />

method method zenith<br />

was<br />

angle, was verified<br />

this verified calculation<br />

only <strong>for</strong> only agricultural<br />

was <strong>for</strong> not zenith per<strong>for</strong>med.<br />

vegetation types and <strong>the</strong> portion <strong>of</strong> mixed pixels<br />

empirical called<br />

sectors<br />

regularization standard (cf. Equation deviation term that<br />

5.6) <strong>of</strong> puts measured is<br />

constraints<br />

divided gap by<br />

on fraction a<br />

<strong>the</strong> retrieved in a certai ALA<br />

increases agricultural with vegetation zenith angle, types this and calculation <strong>the</strong> portion was <strong>of</strong> not modelled per<strong>for</strong>med.<br />

images processed standard toge<strong>the</strong>r deviation and <strong>of</strong> belonging <strong>the</strong> measured to <strong>the</strong> gap same transect. The<br />

mixed pixels increases with zenith angle, this fraction The LUT gap fraction providing <strong>the</strong> minimum value <strong>of</strong> c k and its c<br />

called regularization values. This modelled term that standard puts constraints deviation on is <strong>the</strong> retrieved ALA<br />

calculation was not per<strong>for</strong>med.<br />

derived <strong>the</strong>n considered by fitting as a solution. second This order output polynomial will be fur<strong>the</strong>r to called LAIe<br />

The MES LUT gap fraction providing <strong>the</strong> minimum value <strong>of</strong> c k and its c<br />

σ (P (θi Quantification )), <strong>the</strong> empirical <strong>of</strong> system standard inherent deviation errors <strong>of</strong><br />

0<br />

(2) The second method is <strong>based</strong> on model inversion. measured <strong>the</strong>n considered gap fraction as solution. in a certain This zenithal output will direction be fur<strong>the</strong>r called LAIe (<br />

Assuming an ellipsoidal leaf inclination resulting<br />

Apart from <strong>the</strong> above-described deviation <strong>of</strong> <strong>the</strong> optical centre f<br />

Quantification from all images <strong>of</strong> system processed inherent toge<strong>the</strong>r errors and<br />

distribution, a large range <strong>of</strong> random LAI and belonging<br />

distortions<br />

to <strong>the</strong><br />

can<br />

same<br />

also introduce<br />

transect. The<br />

errors<br />

second<br />

in <strong>the</strong><br />

term<br />

calculation<br />

<strong>of</strong><br />

<strong>of</strong> LAI. Fi<br />

Apart<br />

ALA combinations and <strong>the</strong> corresponding P<br />

between from <strong>the</strong> <strong>the</strong> image above-described radius r in pixels deviation and <strong>the</strong> <strong>of</strong> corresponding <strong>the</strong> optical centre zenit f<br />

Equation 5.6 is a so-called regularization term that<br />

0<br />

distortions equidistant can projection, also introduce which is errors in <strong>the</strong> calculation <strong>of</strong> LAI. Fig<br />

values <strong>for</strong> user-defined zenith angles are stored puts constraints on <strong>the</strong> retrieved ALA values (Weiss<br />

between <strong>the</strong> image radius r in pixels and <strong>the</strong> corresponding zenith<br />

in a LUT. Possible LAI values lie between 0 2006). The d LUT gap fraction providing <strong>the</strong> minimum<br />

e<br />

equidistant . projection, which is<br />

and 10 (in steps <strong>of</strong> 0.01) and ALA 10° and 80° value 90<strong>of</strong><br />

cr and its corresponding LAI and ALA values<br />

k<br />

(in steps <strong>of</strong> 2°). The LUT element closest to <strong>the</strong> are <strong>the</strong>n<br />

dconsidered In order e as solution. This output will be<br />

to . test <strong>the</strong> projection quality <strong>for</strong> <strong>the</strong> camera system in us<br />

derived gap fraction is <strong>the</strong>n chosen <strong>based</strong> on a fur<strong>the</strong>r 90<br />

called r LAI .<br />

Fig. 5-16: Variation <strong>of</strong> <strong>the</strong> G-function (mean projection<br />

(2004b)<br />

<strong>of</strong> unit<br />

was built e (DHP)<br />

foliage<br />

(cf.<br />

area)<br />

Figure<br />

with<br />

A-4).<br />

<strong>the</strong><br />

Based<br />

average<br />

on this<br />

leaf<br />

<strong>the</strong> projection fu<br />

inclination Fig. 5-16: Variation angle (WEISS <strong>of</strong> <strong>the</strong> et G-function al. 2004) (mean projection In The order results <strong>of</strong> unit to test in foliage Figure <strong>the</strong> projection area) 5-18 with show quality <strong>the</strong> that average <strong>for</strong> <strong>the</strong> <strong>the</strong> projection leaf camera system functions in use<br />

inclination angle (WEISS et al. 2004)<br />

(2) The second method is <strong>based</strong> on model inversion. Assuming<br />

(2004b) equidistant an<br />

was<br />

ellipsoidal projection. built (cf.<br />

leaf<br />

Figure It inclination can A-4). thus be Based<br />

distribution, assumed on this that <strong>the</strong> no projection fur<strong>the</strong>r error fu<br />

(2) The second method is <strong>based</strong> on model inversion. Assuming The lens system. results an ellipsoidal in Figure leaf 5-18 inclination show that distribution, <strong>the</strong> projection functions<br />

a large range <strong>of</strong> random LAI and ALA combinations and <strong>the</strong> corresponding P o values <strong>for</strong> user-<br />

a large range <strong>of</strong> random LAI and ALA combinations equidistant and <strong>the</strong> projection. corresponding It can P thus<br />

o values be assumed <strong>for</strong> user- that no fur<strong>the</strong>r error<br />

defined zenith angles are stored in a LUT. Possible LAI<br />

lens<br />

values<br />

system.<br />

lie between 0 and 10 (in steps <strong>of</strong> 0.01)<br />

and<br />

defined<br />

ALA<br />

zenith<br />

10° and<br />

angles<br />

80°<br />

are<br />

(in<br />

stored<br />

steps<br />

in<br />

<strong>of</strong><br />

a<br />

2°).<br />

LUT.<br />

The<br />

Possible<br />

LUT element<br />

LAI values<br />

closest<br />

lie between<br />

to <strong>the</strong> derived<br />

0 and 10<br />

gap<br />

(in<br />

fraction<br />

steps <strong>of</strong><br />

is<br />

0.01)<br />

<strong>the</strong>n


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

17 IN SITU LAI MEASUREMENTS<br />

LUT ( k )<br />

MES<br />

with P0 ( i ) and P0 ( i ) being modelled and measured gap fractions <strong>for</strong> azimuthally averaged<br />

zenith sector i. wi is a weighting function<br />

w<br />

Npix<br />

Nmask<br />

i<br />

i<br />

i (5.7)<br />

Npix i<br />

where Npix i is <strong>the</strong> number <strong>of</strong> pixels in zenith sector i and Nmaski is <strong>the</strong> number <strong>of</strong> masked pixels in<br />

i. The function takes into account that some zenith directions may contain a large amount <strong>of</strong> masked<br />

pixels and is normalized by<br />

no.<br />

zenith angles<br />

w i<br />

i 1<br />

1. (5.8)<br />

The root mean square error between modelled and measured gap fractions over all user-defined zenith<br />

sectors (cf. Equation 5.6) is divided by a modelled standard deviation <strong>of</strong> <strong>the</strong> measured gap fraction values.<br />

MES<br />

Variation <strong>of</strong> <strong>the</strong> G-function (mean projection <strong>of</strong> unit foliage area) with <strong>the</strong> average leaf inclination ( P0angle( θ i<br />

(Weiss et al. 2004).<br />

This Figure modelled 5-16 standard deviation is derived by fitting a second order polynomial to )) , <strong>the</strong><br />

empirical standard deviation <strong>of</strong> measured gap fraction in a certain zenithal direction resulting from all<br />

images processed toge<strong>the</strong>r and belonging to <strong>the</strong> same transect. The second term <strong>of</strong> Equation 5.6 is a socalled<br />

Quantification regularization <strong>of</strong> system term that inherent puts constraints errors on <strong>the</strong> retrieved ALA values (WEISS 2006).<br />

The LUT gap fraction providing <strong>the</strong> minimum value <strong>of</strong> c k and its corresponding LAI and ALA values are<br />

Apart from <strong>the</strong> above-described deviation <strong>of</strong> <strong>the</strong><br />

<strong>the</strong>n considered as solution. This output will be fur<strong>the</strong>r called LAIe (DHP).<br />

optical centre from <strong>the</strong> image centre, geometric<br />

lens Quantification distortions can <strong>of</strong> also system introduce inherent errors errors in <strong>the</strong><br />

Apart<br />

calculation<br />

from<br />

<strong>of</strong><br />

<strong>the</strong><br />

LAI.<br />

above-described<br />

Figure 5-17 illustrates<br />

deviation<br />

<strong>the</strong><br />

<strong>of</strong><br />

linear<br />

<strong>the</strong> optical centre from <strong>the</strong> image centre, geometric lens<br />

distortions<br />

relation between<br />

can also<br />

<strong>the</strong><br />

introduce<br />

image radius<br />

errors<br />

r in<br />

in<br />

pixels<br />

<strong>the</strong> calculation<br />

and <strong>the</strong><br />

<strong>of</strong> LAI. Figure 5-17 illustrates <strong>the</strong> linear relation<br />

between<br />

corresponding<br />

<strong>the</strong> image<br />

zenith<br />

radius<br />

angles<br />

r in<br />

<strong>for</strong><br />

pixels<br />

a hemispherical<br />

and <strong>the</strong> corresponding<br />

lens<br />

zenith angles <strong>for</strong> a hemispherical lens with<br />

equidistant<br />

with equidistant<br />

projection,<br />

projection,<br />

which<br />

which<br />

is<br />

is<br />

de <br />

90<br />

r<br />

. (5.9)<br />

(5.9)<br />

In order order to to test <strong>the</strong> projection quality <strong>for</strong> <strong>the</strong> camera<br />

system in use, an experimental design after BARET<br />

(2004b) system in was use, built an (cf. experimental Figure A-4). design Based after on this Baret <strong>the</strong> projection function was calculated (BARET 2004b).<br />

The (2004b) results was in built Figure (cf. Figure 5-18 show A-4). Based that <strong>the</strong> on this projection <strong>the</strong> functions correspond well with <strong>the</strong> <strong>the</strong>oretical<br />

equidistant projection projection. function was It can calculated thus be assumed (Baret 2004b). that no fur<strong>the</strong>r error is introduced through <strong>the</strong> camera and<br />

lens The system. results in Figure 5-18 show that <strong>the</strong> projection<br />

functions correspond well with <strong>the</strong> <strong>the</strong>oretical<br />

equidistant projection. It can thus be assumed that<br />

no fur<strong>the</strong>r error is introduced through <strong>the</strong> camera and Figure 5-17 Equidistant projection<br />

lens system.<br />

(Yamashita et al. 2004, modified).<br />

83


84<br />

Influence <strong>of</strong> operator and processing<br />

In order to assess <strong>the</strong> operator effect in <strong>the</strong><br />

classification process, DHPs <strong>of</strong> all ESUs were<br />

processed twice by two different individuals. Figure<br />

5-19 shows <strong>the</strong> resulting LAI values per ESU<br />

e (DHP)<br />

plotted against each o<strong>the</strong>r. Mean LAI and<br />

e (DHP)<br />

standard deviation are very close to each o<strong>the</strong>r (see<br />

Table 5-6), with LAI calculated by operator 2<br />

e (DHP)<br />

being slightly higher (R2 =0.99).<br />

A more detailed analysis <strong>of</strong> outliers revealed that<br />

<strong>the</strong>re were some misclassifications on <strong>the</strong> part <strong>of</strong> <strong>the</strong><br />

less experienced operator 2. It is thus acknowledged<br />

that <strong>the</strong> operator has a certain influence on <strong>the</strong><br />

resulting variable LAI . But in order to avoid <strong>the</strong><br />

e (DHP)<br />

introduction <strong>of</strong> new errors, only LAI processed<br />

e (DHP)<br />

by <strong>the</strong> first operator was included in <strong>the</strong> study instead<br />

<strong>of</strong> taking <strong>the</strong> mean <strong>of</strong> <strong>the</strong> two results.<br />

angle (°)<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

NIKON Coolpix 4300, INRA, 10/01/2006<br />

R² = 0.9987<br />

angle= 0.09581 * radius<br />

0<br />

0 100 200 300 400<br />

radius (pixels)<br />

500 600 700 800<br />

Figure 5-18<br />

Influence <strong>of</strong> illumination conditions<br />

In order to assess <strong>the</strong> influence <strong>of</strong> θ sun variation on<br />

LAI derived from DHP, a test was carried out in <strong>the</strong><br />

centre <strong>of</strong> ESU 5 (late <strong>for</strong>est stage in Budongo Forest).<br />

Pictures were taken every 15 minutes from 7:45 a.m.<br />

to 6:00 p.m. The resulting LAI values are<br />

e (DHP)<br />

shown in Figure 5-20. LAI varied between 3.50<br />

e (DHP)<br />

(θ =55°/4.15 p.m.) and 4.35 (θ = 36°/10.30 a.m.),<br />

sun sun<br />

with a mean <strong>of</strong> 3.93 (±0.19).<br />

Table 5-6 Mean and standard deviation <strong>of</strong> LAIe<br />

(DHP)<br />

per ESU, processed by two different operators.<br />

Illustration <strong>of</strong> calculated projection function <strong>for</strong> NIKON coolpix 4300 and FC-E8 fisheye.<br />

μ σ<br />

LAI e (DHP) processed by operator 1 5.11 1.41<br />

LAI e (DHP) processed by operator 2 5.15 1.37


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

was taken per hour so that <strong>the</strong> result is less<br />

19 IN SITU LAI MEASUREMENTS<br />

19 IN SITU LAI MEASUREMENTS<br />

19<br />

representative and<br />

IN SITU LAI<br />

<strong>the</strong>re<strong>for</strong>e not shown here.<br />

MEASUREMENTS<br />

Figure<br />

A more<br />

more<br />

detailed<br />

detailed<br />

analysis<br />

analysis<br />

<strong>of</strong><br />

<strong>of</strong><br />

outliers<br />

outliers<br />

revealed<br />

revealed<br />

that<br />

that<br />

<strong>the</strong>re<br />

<strong>the</strong>re<br />

were<br />

were<br />

some<br />

some<br />

misclassifications<br />

misclassifications<br />

on<br />

on<br />

<strong>the</strong><br />

<strong>the</strong><br />

part<br />

part<br />

<strong>of</strong><br />

<strong>of</strong><br />

<strong>the</strong><br />

<strong>the</strong><br />

less<br />

less<br />

A more detailed analysis <strong>of</strong> outliers revealed that <strong>the</strong>re were experienced some misclassifications Figure 5-20: LAIe operator 2. It on is thus <strong>the</strong> (DHP) part plotted<br />

acknowledged <strong>of</strong> <strong>the</strong> less against sun. Measu<br />

that<br />

Fig. 5-19: LAIe (DHP) derived from classification experienced operator 2. It is thus acknowledged that<br />

Fig. 5-19: LAIe (DHP) derived from classification<br />

<strong>of</strong> gap fraction by two different operators with experienced <strong>the</strong> operator operator<br />

Measurements<br />

has 2. a certain It is thus<br />

made<br />

influence acknowledged<br />

on 1<br />

<strong>the</strong> operator has a certain influence on<br />

on<br />

<strong>the</strong><br />

<strong>the</strong> that<br />

Fig. resulting<br />

resulting<br />

<strong>of</strong> 5-19: gap LAIe fraction (DHP) by derived two different from classification operators with<br />

<strong>of</strong> gap CAN-EYE.<br />

CAN-EYE. fraction by two different operators with <strong>the</strong> operator variable has LAIe a certain<br />

variable (DHP). influence But in order on <strong>the</strong> to resulting avoid <strong>the</strong><br />

LAIe (DHP). But in order to avoid <strong>the</strong><br />

CAN-EYE.<br />

variable Correction<br />

introduction LAIe (DHP). <strong>of</strong><br />

<strong>of</strong> new But clumping<br />

errors, in order only LAIe to avoid (DHP) introduction <strong>of</strong> new errors, only<br />

processed <strong>the</strong><br />

LAIe (DHP) processed<br />

by<br />

by<br />

<strong>the</strong><br />

<strong>the</strong><br />

first<br />

first<br />

operator<br />

operator<br />

was<br />

was<br />

included<br />

included<br />

in<br />

in<br />

<strong>the</strong><br />

<strong>the</strong><br />

study<br />

study<br />

instead<br />

instead introduction<br />

<strong>of</strong><br />

<strong>of</strong> In<br />

taking<br />

taking order <strong>of</strong> new<br />

<strong>the</strong><br />

<strong>the</strong> to<br />

mean<br />

mean correct errors,<br />

<strong>of</strong><br />

<strong>of</strong> only<br />

<strong>the</strong><br />

<strong>for</strong> <strong>the</strong><br />

two<br />

foliage two LAIe<br />

results.<br />

results. (DHP) clumping, processed <strong>the</strong> logarithm gap fraction<br />

by <strong>the</strong> first operator was included in <strong>the</strong> study instead <strong>of</strong> taking (1986) <strong>the</strong> mean was <strong>of</strong> applied <strong>the</strong> two (cf. results. Chapter 3.3.3). Hence gap fraction cal<br />

azimuth sectors (as specified in <strong>the</strong> CAN-EYE parameter file) is t<br />

Influence<br />

Influence<br />

<strong>of</strong><br />

<strong>of</strong><br />

illumination<br />

illumination<br />

conditions<br />

conditions<br />

Influence <strong>of</strong> illumination conditions<br />

st Figu<br />

November, 2005<br />

Mea<br />

Correction <strong>of</strong> clumping<br />

Correction <strong>of</strong> clumping<br />

R²=0.96<br />

In order to correct <strong>for</strong> foliage clumping, <strong>the</strong> logarithm<br />

In<br />

gap<br />

order<br />

fraction<br />

to correct<br />

averaging<br />

<strong>for</strong> foliage<br />

method<br />

clumping,<br />

by LANG<br />

<strong>the</strong><br />

&<br />

logarithm<br />

XIANG<br />

gap frac<br />

Y=-0.009+0.982 X<br />

(1986) was applied (cf. Chapter 3.3.3). Hence gap<br />

(1986)<br />

fraction<br />

was<br />

calculated<br />

applied<br />

<strong>for</strong><br />

(cf.<br />

<strong>the</strong><br />

Chapter<br />

predefined<br />

3.3.3).<br />

zenith<br />

Hence<br />

and<br />

gap fraction<br />

azimuth sectors (as specified in <strong>the</strong> CAN-EYE parameter<br />

azimuth<br />

file)<br />

sectors<br />

is taken<br />

(as<br />

into<br />

specified<br />

account.<br />

in<br />

The<br />

<strong>the</strong> CAN-EYE<br />

clumping factor<br />

parameter file) i<br />

Figure 5-19 LAI derived from classification <strong>of</strong> gap fraction Figure 5-20 LAI<br />

e (DHP) e (DHP)<br />

by two different operators with CAN-EYE. 1:1 line<br />

shown <strong>for</strong> comparison.<br />

plotted against θ . Measurements<br />

sun<br />

made on 1st o <strong>for</strong> each zenithal ring can <strong>the</strong>n be derived according to WEISS et<br />

In order to assess <strong>the</strong> influence <strong>of</strong> sun In order<br />

o <strong>for</strong><br />

to<br />

each<br />

assess<br />

zenithal<br />

<strong>the</strong> influence<br />

ring can<br />

<strong>of</strong><br />

<strong>the</strong>n be variation on LAI derived from DHP, a test was carried out in <strong>the</strong><br />

sun variation<br />

derived<br />

on<br />

according<br />

LAI derived<br />

to o <strong>for</strong><br />

WEISS<br />

each<br />

from<br />

et<br />

zenithal<br />

DHP,<br />

al. (2006)<br />

ring<br />

a test<br />

as<br />

can <strong>the</strong>n be derived according to WEISS<br />

sec<br />

was carried out in <strong>the</strong><br />

centre <strong>of</strong> ESU 5 (late <strong>for</strong>est stage in Budongo Forest). Pictures ln( Po<br />

( were ))<br />

centre <strong>of</strong> ESU 5 (late <strong>for</strong>est stage in Budongo Forest). <br />

taken every 15 minutes from<br />

o Pictures sec<br />

sec<br />

were taken November, every 15 2005 minutes every 15 minutes. from<br />

7:45 a.m. to ln( 6:00 P ( p.m. ))<br />

ln( sec Po<br />

( ))<br />

o The resulting LAIe (DHP) values are shown in Figure 5-20. LAIe (DHP) 7:45 a.m. <br />

ln Po( ) <br />

to 6:00 p.m. The resulting<br />

varied between<br />

o <br />

o<br />

<br />

LAIe (DHP) values are shown in Figure sec 5-20. LAIe (DHP) varied between (5.10)<br />

sec<br />

3.50 (sun=55°/4.15 ln P( ) p.m.) <br />

ln P and 4.35 (sun 3.50 (sun=55°/4.15 p.m.) and 4.35<br />

= 36°/10.30 a.m.), with a mean o ( ) <br />

o<br />

<strong>of</strong> 3.93 (± 0.19).<br />

(sun = 36°/10.30 a.m.), with a mean <strong>of</strong> 3.93 (± 0.19).<br />

sec<br />

where ln( P o ( )) is <strong>the</strong> logarithm <strong>of</strong> gap fraction derived <strong>for</strong> e<br />

Pearson’s correlation coefficient reveals that <strong>the</strong>re is is Pearson’s <strong>the</strong><br />

Pearson’s<br />

logarithm correlation<br />

correlation sec<br />

<strong>of</strong> <strong>the</strong> coefficient<br />

coefficient<br />

previously reveals<br />

reveals<br />

averaged that<br />

that<br />

sec<br />

where ln( P gap<br />

o ( )) is <strong>the</strong> logarithm <strong>of</strong> gap fraction derived<br />

where ln(<br />

<strong>for</strong><br />

P oeach<br />

( ))<br />

sector<br />

is <strong>the</strong><br />

on<br />

logarithm<br />

<strong>the</strong> DHPs,<br />

<strong>of</strong><br />

subsequently<br />

gap fraction derived fo<br />

Pearson’s<br />

no significant correlation between <strong>the</strong> two variables<br />

averaged<br />

<strong>the</strong>re<br />

<strong>the</strong>re<br />

is<br />

is correlation<br />

no<br />

no<br />

significant<br />

significant coefficient<br />

correlation<br />

correlation reveals<br />

between<br />

between that sec<br />

fraction over<br />

over<br />

azimuth<br />

azimuth.<br />

(comparable<br />

In contrary to<br />

to<br />

that,<br />

gap fraction<br />

ln<br />

<strong>the</strong><br />

<strong>the</strong> P o ( ) is <strong>the</strong> log<br />

sec<br />

averaged over azimuth. In contrary to that, ln <strong>the</strong>re sec<br />

P ( averaged ) two is<br />

no variables significant at 0.05 correlation level (r = between -0.31). Pictures <strong>the</strong><br />

o is <strong>the</strong><br />

over<br />

logarithm<br />

azimuth.<br />

<strong>of</strong> <strong>the</strong><br />

In contrary<br />

previously<br />

to<br />

averaged<br />

that, ln P gap o ( ) is <strong>the</strong><br />

In order to assess <strong>the</strong> influence <strong>of</strong> sun variation on LAI derived from DHP, a test was carried out in <strong>the</strong><br />

centre <strong>of</strong> ESU 5 (late <strong>for</strong>est stage in Budongo Forest). Pictures were taken every 15 minutes from<br />

7:45 a.m. to 6:00 p.m. The resulting LAIe (DHP) values are shown in Figure 5-20. LAIe (DHP) varied between<br />

3.50 (sun=55°/4.15 p.m.) and 4.35 (sun = 36°/10.30 a.m.), with a mean <strong>of</strong> 3.93 (± 0.19).<br />

at 0.05 level (r=-0.31). Pictures taken on 30 October,<br />

2005 in secondary <strong>for</strong>est close to <strong>the</strong> Sonso campsite<br />

show a comparable result. However, only one<br />

picture was taken per hour so that <strong>the</strong> result is less<br />

representative and <strong>the</strong>re<strong>for</strong>e not shown here.<br />

Correction <strong>of</strong> clumping<br />

Figure 5-20: LAIe (DHP) plotted against sun.<br />

Measurements made on 1<br />

In order to correct <strong>for</strong> foliage clumping, <strong>the</strong> logarithm<br />

gap fraction averaging method by Lang & Xiang<br />

(1986) was applied (cf. Chapter 3.3.3). Hence gap<br />

fraction calculated <strong>for</strong> <strong>the</strong> predefined zenith and<br />

azimuth sectors (as specified in <strong>the</strong> CAN-EYE<br />

parameter file) is taken into account. The clumping<br />

factor λ <strong>for</strong> each zenithal ring can <strong>the</strong>n be derived<br />

0<br />

according to Weiss et al. (2006) as<br />

st taken on 30 October, 2005 in secondary<br />

<strong>for</strong>est close to <strong>the</strong> Sonso campsite show a<br />

comparable result. However, only one picture<br />

was taken per hour so that <strong>the</strong> result is less<br />

representative and <strong>the</strong>re<strong>for</strong>e not shown here.<br />

Figure 5-20: LAIe (DHP) plotted against sun.<br />

November, 2005<br />

Measurements made on 1<br />

Correction <strong>of</strong> clumping<br />

In order to correct <strong>for</strong> foliage clumping, <strong>the</strong> logarithm gap fraction averaging method by LANG & XIANG<br />

(1986) was applied (cf. Chapter 3.3.3). Hence gap fraction calculated <strong>for</strong> <strong>the</strong> predefined zenith and<br />

azimuth sectors (as specified in <strong>the</strong> CAN-EYE parameter file) is taken into account. The clumping factor<br />

o <strong>for</strong> each zenithal ring can <strong>the</strong>n be derived according to WEISS et al. (2006) as<br />

sec<br />

ln( Po<br />

( ))<br />

o<br />

(5.10)<br />

(5.10)<br />

sec<br />

ln Po( ) <br />

sec<br />

where ln( P o ( ))<br />

where<br />

is <strong>the</strong> logarithm <strong>of</strong> gap fraction derived <strong>for</strong> each sector on <strong>the</strong> DHPs, subsequently<br />

sec<br />

averaged over azimuth. In contrary to that, ln P o ( ) is <strong>the</strong> logarithm <strong>of</strong> <strong>the</strong> previously averaged gap<br />

fraction over azimuth (comparable to gap fraction values given by LAI-2000 PCA). In <strong>the</strong> case that<br />

sec<br />

sec<br />

P o , <br />

= 0, i.e. only foliage occurring in <strong>the</strong> respective sector, Po , <br />

is assumed to be equal to<br />

sat<br />

P o , which stands <strong>for</strong> <strong>the</strong> gap fraction derived from a simple Poisson law using a prescribed LAI<br />

st two variables at 0.05 level (r = -0.31). Pictures<br />

taken on 30 October, 2005 in secondary<br />

<strong>for</strong>est close to <strong>the</strong> Sonso campsite show a<br />

comparable result. However, only one picture<br />

was taken per hour so that <strong>the</strong> result is less<br />

representative and <strong>the</strong>re<strong>for</strong>e not shown here.<br />

Figure 5-20: LAIe (DHP) plotted against sun.<br />

November, 2005<br />

Measurements made on 1<br />

Correction <strong>of</strong> clumping<br />

In order to correct <strong>for</strong> foliage clumping, <strong>the</strong> logarithm gap fraction averaging method by LANG & XIANG<br />

(1986) was applied (cf. Chapter 3.3.3). Hence gap fraction calculated <strong>for</strong> <strong>the</strong> predefined zenith and<br />

azimuth sectors (as specified in <strong>the</strong> CAN-EYE parameter file) is taken into account. The clumping factor<br />

o <strong>for</strong> each zenithal ring can <strong>the</strong>n be derived according to WEISS et al. (2006) as<br />

sec<br />

ln( Po<br />

( ))<br />

o<br />

(5.10)<br />

sec<br />

ln Po( ) <br />

sec<br />

where ln( P o ( )) is <strong>the</strong> logarithm logarithm <strong>of</strong> gap fraction derived <strong>for</strong> each sector on <strong>the</strong> DHPs, subsequently<br />

derived <strong>for</strong> each sector on <strong>the</strong> DHPs, subsequently sec<br />

averaged over azimuth. In contrary to that, ln P o ( ) is <strong>the</strong> logarithm <strong>of</strong> <strong>the</strong> previously averaged gap<br />

averaged over azimuth. In contrary to that,<br />

fraction over azimuth (comparable to gap fraction values given by LAI-2000 PCA). In <strong>the</strong> case that<br />

sec<br />

sec<br />

P o , <br />

= 0, i.e. only foliage occurring in <strong>the</strong> respective sector, Po , <br />

is assumed to be equal to<br />

sat<br />

P o , which stands <strong>for</strong> <strong>the</strong> gap fraction derived from a simple Poisson law using a prescribed LAI<br />

st fraction values given over by azimuth LAI-2000 (comparable PCA). In to <strong>the</strong> gap case fraction that values given<br />

fraction over azimuth (comparable to gap fraction two sec fraction variables at 0.05 level (r = -0.31). Pictures<br />

P values , given<br />

over<br />

o = 0, i.e. by<br />

azimuth<br />

i.e. LAI-2000<br />

(comparable<br />

only foliage PCA). occurring In<br />

to<br />

<strong>the</strong><br />

gap<br />

in case<br />

fraction<br />

in <strong>the</strong> <strong>the</strong> respective that<br />

values giv<br />

secto<br />

taken sec<br />

sec<br />

on 30 October, 2005 in secondary<br />

P o , <br />

= 0, i.e. only foliage occurring in <strong>the</strong> respective P<br />

sat o , <strong>for</strong>est<br />

P o ,<br />

close<br />

which<br />

sector,<br />

<br />

sector, = 0, sec<br />

to<br />

stands<br />

P<br />

i.e.<br />

<strong>the</strong> Sonso<br />

<strong>for</strong> only o ,<br />

<strong>the</strong><br />

<br />

is<br />

foliage<br />

is<br />

campsite<br />

gap<br />

assumed<br />

occurring<br />

fraction<br />

to to be<br />

show<br />

derived<br />

be<br />

in<br />

equal<br />

<strong>the</strong><br />

a<br />

from<br />

to<br />

respective se<br />

a simple<br />

sat<br />

sat<br />

P o , which stands <strong>for</strong> <strong>the</strong> gap fraction derived<br />

to<br />

comparable<br />

from<br />

P o a<br />

, which which<br />

simple<br />

stands stands<br />

result.<br />

Poisson<br />

<strong>for</strong> However,<br />

law<br />

<strong>the</strong> only<br />

using<br />

gap fraction<br />

one<br />

a<br />

fraction<br />

picture<br />

prescribed<br />

derived<br />

LAI<br />

from a sim<br />

from a simple Poisson law using a prescribed LAI<br />

was taken per hour so that <strong>the</strong> result is less<br />

saturation value (LAI=10 <strong>for</strong> this study). After that<br />

representative and <strong>the</strong>re<strong>for</strong>e not shown here.<br />

<strong>the</strong> same LUT approach as described <strong>for</strong> LAIe (DHP)<br />

is applied to <strong>the</strong> modified Poisson model (Equation<br />

5.5). The output will be called LAI . (DHP)<br />

November, 2005<br />

5.2.3 Comparison between LAI-2000 PCA<br />

Correction <strong>of</strong> clumping<br />

and DHPs<br />

In order to correct <strong>for</strong> foliage clumping, <strong>the</strong> logarithm gap fraction averaging method by LANG & XIANG<br />

Analysis at gap fraction level is rarely accomplished,<br />

(1986) was applied (cf. Chapter 3.3.3). Hence gap fraction calculated <strong>for</strong> <strong>the</strong> predefined zenith and<br />

but necessary to understand differences between<br />

azimuth sectors (as specified in <strong>the</strong> CAN-EYE parameter file) is taken into account. The clumping factor<br />

instruments (Leblanc et al. 2005). Accordingly this<br />

o <strong>for</strong> each zenithal ring can <strong>the</strong>n be derived according to WEISS et al. (2006) as<br />

comparison was per<strong>for</strong>med first, i.e. be<strong>for</strong>e logarithmic<br />

sec<br />

trans<strong>for</strong>mation <strong>of</strong> gap fraction and subsequent LAI<br />

ln( Po<br />

( ))<br />

o<br />

calculation. It should be noted that despite (5.10) an exact<br />

sec<br />

ln Po( ) <br />

demarcation <strong>of</strong> <strong>the</strong> respective sampling points along<br />

<strong>the</strong> transects, <strong>the</strong> two instruments were handled by<br />

sec<br />

where ln( P o ( )) is <strong>the</strong> logarithm <strong>of</strong> gap fraction derived <strong>for</strong> each sector on <strong>the</strong> DHPs, subsequently<br />

different operators, so that small displacement errors<br />

sec<br />

averaged over azimuth. In contrary to that, ln P o ( ) is <strong>the</strong> were logarithm still present. <strong>of</strong> <strong>the</strong> previously averaged gap<br />

fraction over azimuth (comparable to gap fraction values given by LAI-2000 PCA). In <strong>the</strong> case that<br />

sec<br />

sec<br />

P o , <br />

= 0, i.e. only foliage occurring in <strong>the</strong> respective sector, Po , <br />

is assumed to be equal to<br />

P , which stands <strong>for</strong> <strong>the</strong> gap fraction derived from a simple Poisson law using a prescribed LAI<br />

sat<br />

o<br />

two variables at 0.05 level (r = -0.31). Pictures<br />

taken on 30 October, 2005 in secondary<br />

85<br />

<strong>for</strong>est close to <strong>the</strong> Sonso campsite show a<br />

comparable result. However, only one picture<br />

taken<br />

two<br />

<strong>for</strong>est<br />

take<br />

compa<br />

<strong>for</strong>e<br />

was<br />

com<br />

ta<br />

repres<br />

was<br />

repr


86<br />

As <strong>the</strong> two instruments have different fields <strong>of</strong> view<br />

(cf. Figure 5-21), gap fraction was extracted from<br />

DHP <strong>for</strong> those sectors that correspond to <strong>the</strong> field <strong>of</strong><br />

view <strong>of</strong> <strong>the</strong> LAI-2000 PCA. Analogue to <strong>the</strong> latter,<br />

gap fraction was averaged over 45° azimuth and over<br />

those zenith angles that correspond to <strong>the</strong> different<br />

rings <strong>of</strong> <strong>the</strong> LAI-2000 PCA. As <strong>the</strong> variables differed<br />

significantly from normal distribution, <strong>the</strong> Spearman-<br />

Rho correlation coefficient r was calculated (Sachs<br />

s<br />

2004). As shown in Table 5-7, it revealed relatively<br />

strong relationships between transmission from<br />

LAI-2000 PCA data and gap fraction derived from<br />

DHP <strong>for</strong> <strong>the</strong> corresponding azimuth sectors (between<br />

0.63 and 0.75, all significant at p


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

sectors, LAI e values calculated from <strong>the</strong> above shown<br />

light variables were also compared.<br />

As <strong>the</strong> s<strong>of</strong>tware code <strong>of</strong> CAN-EYE was not available,<br />

LAI could not be calculated <strong>for</strong> <strong>the</strong> field <strong>of</strong><br />

e (DHP)<br />

view restricted to 45° in azimuth. Consequently<br />

a comparison <strong>for</strong> each sampling point would not<br />

be correct. Instead, mean LAI per ESU derived<br />

e<br />

from both methods was analysed with a Wilcoxon<br />

signed-rank test <strong>for</strong> paired data. On ESU level,<br />

<strong>the</strong> test indicated that LAI and LAI do<br />

e (LAI2000) e (DHP)<br />

not differ significantly from each o<strong>the</strong>r (Z=1.265,<br />

p=0.21, <strong>the</strong>re<strong>for</strong>e <strong>the</strong> null hypo<strong>the</strong>sis that<br />

LAI = LAI cannot be rejected).<br />

e (LAI2000) e (DHP)<br />

A more detailed analysis reveals that <strong>for</strong> lower<br />

values LAI is overestimated with respect<br />

e (DHP)<br />

to LAI . For higher values LAI is<br />

e (LAI2000) e (DHP)<br />

underestimated (cf. Figure 5-23, note that only<br />

LAI derived from upward-looking photographs is<br />

e (DHP)<br />

shown). This corresponds to <strong>the</strong> findings <strong>of</strong> Mussche<br />

et al. (2001) <strong>for</strong> a deciduous <strong>for</strong>est in Belgium where<br />

LAI and LAI were compared with LAI<br />

e (DHP) e (LAI2000)<br />

estimated from leaf litter fall. Here LAI was<br />

e (LAI2000)<br />

not significantly different from reference LAI <strong>for</strong> high<br />

Figure 5-22 Boxplots showing light variables ( P and τ ) <strong>for</strong><br />

0 0<br />

<strong>the</strong> respective rings calculated from DHP and<br />

LAI2000. The box represents 50% <strong>of</strong> <strong>the</strong> values,<br />

whiskers include minimum and maximum values,<br />

solid line in box represents median.<br />

87<br />

LAI values. Lower LAI values were overestimated<br />

by <strong>the</strong> LAI-2000 PCA, probably because <strong>of</strong> <strong>the</strong><br />

contribution <strong>of</strong> woody area to LAI . As low<br />

e (LAI2000)<br />

LAI values (i.e. < 3) in this <strong>the</strong>sis correspond to<br />

ESUs in very early <strong>for</strong>est stages without significant<br />

contribution <strong>of</strong> woody material, it is assumed that<br />

LAI is not overestimated in <strong>the</strong>se cases.<br />

e (LAI2000)<br />

5.3 Results <strong>of</strong> in situ measurements<br />

The previously described preprocessing lead to<br />

three different LAI measures, namely LAI ,<br />

e (LAI2000)<br />

LAI , and LAI . Whereas LAI excludes<br />

e (DHP) (DHP) e (LAI2000)<br />

understorey vegetation under 0.8 m and represents<br />

a LAI measure as used in o<strong>the</strong>r recent studies <strong>of</strong><br />

tropical rain <strong>for</strong>ests (De Wasseige et al. 2003,<br />

Kalácska et al. 2004, Aragao et al. 2005), LAIe (DHP)<br />

and LAI both include understorey vegetation.<br />

(DHP)<br />

This inclusion has been recommended by various<br />

authors (e.g. Wang et al. 2005, Yang et al. 2006) as<br />

<strong>the</strong> understorey can contribute significantly to canopy<br />

LAI. As a postprocessing step <strong>of</strong> DHP, a correction <strong>for</strong><br />

clumping was applied to account <strong>for</strong> <strong>the</strong> non-random<br />

distribution <strong>of</strong> foliage elements, resulting in LAI . (DHP)<br />

R²=0.83<br />

Y=-0.224+1.079 X<br />

Figure 5-23 LAIe<br />

(LAI2000) plotted against LAI . 1:1 line<br />

e (DHP)<br />

shown <strong>for</strong> comparison.


88<br />

Results <strong>of</strong> <strong>the</strong> in situ measurements <strong>of</strong> all three LAI<br />

measures will be presented in <strong>the</strong> following.<br />

5.3.1<br />

Budongo Forest<br />

Representativeness <strong>of</strong> sampling design<br />

In order to analyse <strong>the</strong> representation <strong>of</strong> different<br />

<strong>for</strong>est stages by <strong>the</strong> sampled ESUs, Budongo Forest<br />

area was classified <strong>based</strong> on NFA data on logging<br />

compartments (cf Figure 2-9). All areas without<br />

closed tree cover were assigned to <strong>the</strong> „early <strong>for</strong>est<br />

stage” class. The decision whe<strong>the</strong>r an area was<br />

assigned to <strong>the</strong> category intermediate (disturbed)<br />

or late (undisturbed), was mainly <strong>based</strong> on NFA<br />

in<strong>for</strong>mation on zone classification and logging status.<br />

If compartments were classified as unlogged (such<br />

as <strong>the</strong> nature reserves) or selectively logged in 1945<br />

(as <strong>the</strong> buffer zones around <strong>the</strong> nature reserves) <strong>the</strong>y<br />

were considered as late and mostly undisturbed <strong>for</strong>est<br />

stages. All o<strong>the</strong>r compartments were considered as<br />

intermediate and disturbed (cf. Table 5-12).<br />

Accordingly <strong>the</strong> three <strong>for</strong>est stages were assigned to<br />

<strong>the</strong> 30 ESUs, with one exception. Although ESU 28<br />

was situated in <strong>the</strong> nature reserve N15 which is under<br />

protection and thus classified as late (undisturbed),<br />

Figure 5-24<br />

ESU 28 showed clear signs <strong>of</strong> recent illegal logging<br />

activities (cf. Figure 5-24). Consequently it was<br />

assigned to <strong>the</strong> intermediate (disturbed) <strong>for</strong>est stage<br />

class.<br />

A comparison between NFA data on logging<br />

compartments and ESUs showed that early <strong>for</strong>est<br />

stages cover 21.36% <strong>of</strong> <strong>the</strong> total <strong>for</strong>est area in<br />

Budongo Forest and are represented by 26.67% <strong>of</strong><br />

<strong>the</strong> ESUs (cf. Figure 5-25). 51.39% (ESU: 46.67%)<br />

comprise intermediate (disturbed) <strong>for</strong>est stages and<br />

27.25% (ESU: 26.67%) are undisturbed as <strong>the</strong>y are put<br />

under protection in <strong>the</strong> nature reserves. The sampled<br />

ESUs are <strong>the</strong>re<strong>for</strong>e regarded as representative <strong>for</strong> <strong>the</strong><br />

distribution <strong>of</strong> <strong>for</strong>est stages <strong>of</strong> <strong>the</strong> whole study site.<br />

Spatial variability <strong>of</strong> in situ LAI<br />

At sample level, individual LAI e (LAI2000) ranged<br />

from 0.25 (on ESU 25) to 8.92 (on ESU 8), with a<br />

mean <strong>of</strong> 4.76 (±2.09) <strong>for</strong> a total number <strong>of</strong> 1,652<br />

individual measurements. The frequency distribution<br />

(cf. Figure 5-26) shows that <strong>the</strong>re are local<br />

maxima between 2 to 3 and 6 to 7, indicating that<br />

LAI values <strong>of</strong> 4 to 5 are underrepresented.<br />

e (LAI2000)<br />

Figure 5-27 shows a comparison between an<br />

ESU located in an early <strong>for</strong>est stage and an ESU<br />

a) Thinned upper canopy and b) illegal logging activities on ESU 28 in <strong>the</strong> nature reserve N15.


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

representing an intermediate (disturbed) <strong>for</strong>est area.<br />

Whereas <strong>the</strong> first is characterized by a dense and high<br />

grass layer with some dispersed trees <strong>of</strong> max. 14 m<br />

height, <strong>the</strong> intermediate <strong>for</strong>est stage is a degraded<br />

primary <strong>for</strong>est site that had been selectively logged<br />

during <strong>the</strong> past 50 years. Succession stages in between<br />

<strong>the</strong>se two ESUs were hardly present in Budongo<br />

Forest and if so, only occurred on <strong>for</strong>est edges, where<br />

ESUs with an adequate size could not be established.<br />

The frequency distribution <strong>of</strong> LAI and LAI e (DHP) (DHP)<br />

cannot be shown <strong>for</strong> individual measurements, as<br />

both are only calculated on ESU level (cf. Chapter<br />

Figure 5-25<br />

Percentage <strong>of</strong> <strong>for</strong>est stages a) <strong>of</strong> <strong>the</strong> whole study area <strong>of</strong> Budongo Forest and b) <strong>of</strong> sampled ESUs.<br />

Figure 5-26 Frequency distribution <strong>of</strong> LAI on sample<br />

e (LAI2000)<br />

level.<br />

Figure 5-27 a) ESU 21 (mean LAI <strong>of</strong> 3.1) and<br />

e (LAI2000)<br />

b) ESU 15 (mean LAI <strong>of</strong> 5.53).<br />

e (LAI2000)<br />

89<br />

5.2.2). Yet <strong>the</strong> frequency distribution on ESU level<br />

reflects <strong>the</strong> lack <strong>of</strong> intermediate LAI values <strong>for</strong><br />

all three measures (cf. Figure 5-28). In situ LAI in<br />

Budongo Forest is thus not normally distributed.<br />

Whereas ESUs with LAI e (LAI2000) between 3.5 and 5<br />

are not present at all, LAI e (DHP) and LAI (DHP) show<br />

minima between 4.0 to 5.0 (LAI e (DHP) ) and 6.5 to<br />

8.0 (LAI (DHP) ). As expected – because this does not<br />

include understorey vegetation in intermediate and<br />

late <strong>for</strong>est stages – LAI is lowest with a mean<br />

e (LAI2000)<br />

<strong>of</strong> 5.29 (minimum <strong>of</strong> 1.56 on ESU 22 and maximum<br />

<strong>of</strong> 7.38 on ESU 11). LAI is slightly higher with<br />

e (DHP)


90<br />

a mean <strong>of</strong> 5.46 on ESU level (minimum <strong>of</strong> 2.53 on<br />

ESU 22 and maximum <strong>of</strong> 7.69 on ESU 7). This seems<br />

reasonable as understorey LAI ranged between<br />

e (DHP)<br />

0.18 and 0.68 with a mean <strong>of</strong> 0.36 (cf. Figure 5-29).<br />

LAI ranges between 5.19 and 10.47 (ESUs 22<br />

(DHP)<br />

and 6 respectively) with a mean <strong>of</strong> 8.38. In particular,<br />

<strong>the</strong> minimum value <strong>of</strong> 5.19 seems to be unreasonably<br />

high, as ESU 22 belongs to <strong>the</strong> early <strong>for</strong>est stages, i.e.<br />

covered mainly by grass species.<br />

At <strong>the</strong> <strong>for</strong>est stage level (cf. Figure 5-30), a Mann-<br />

Whitney U test revealed that mean LAI (all measures)<br />

is significantly different <strong>for</strong> early and intermediate,<br />

as well as early and late <strong>for</strong>est stages (p


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

Figure 5-29 DHP <strong>of</strong> understorey on a) ESU 4 (mean understorey LAIe<br />

(DHP) <strong>of</strong> 0.18) and b) ESU 28 (mean LAI <strong>of</strong> 0.53).<br />

e (DHP)<br />

Figure 5-30<br />

Mean LAI per <strong>for</strong>est stage in Budongo Forest.<br />

Upper and lower box boundaries indicate<br />

25th and 75th percentiles, whiskers indicate<br />

minimum and maximum values except outliers<br />

(defined as more than 3 box lengths from 25th or 75th percentile, displayed as point). Mean is<br />

shown as solid line.<br />

91


92<br />

30 Late (undisturbed) Logged in 1945 Protection (buffer) zone N1 6.06 5.43 8.10<br />

29 Late (undisturbed) Unlogged Nature reserve N15 6.13 6.02 8.88<br />

28 Intermediate (disturbed) Unlogged Nature reserve N15 5.39 5.68 9.82<br />

27 Late (undisturbed) Unlogged Nature reserve N15 6.16 5.49 8.31<br />

26 Early Logged in 1970-72 Low impact harvesting KP1 2.15 3.40 6.30<br />

25 Early Logged in 1970-72 Low impact harvesting KP1 2.31 2.98 5.50<br />

24 Early Logged in 1970-72 Low impact harvesting KP1 2.93 4.44 8.34<br />

23 Early Logged in 1970-72 Low impact harvesting KP1 3.48 3.17 5.63<br />

22 Early Logged in 1970-72 Low impact harvesting KP2 1.56 2.53 5.19<br />

21 Early n/a Low impact harvesting Grassland 3.16 3.69 6.70<br />

20 Late (undisturbed) Unlogged Recreation zone KP11 6.07 4.83 7.78<br />

19 Early n/a Protection (buffer) zone KP10 2.69 3.56 6.05<br />

18 Late (undisturbed) Unlogged Recreation zone KP11 6.00 5.36 8.16<br />

17 Early n/a Protection (buffer) zone KP9 3.12 3.74 7.25<br />

Forest stage, logging status, NFA zone, compartment number and LAI measures (LAI<br />

Table 5-8 e (LAI2000) , LAI e (DHP) , and<br />

LAI (DHP) ) <strong>of</strong> <strong>the</strong> 30 ESUs in Budongo Forest.<br />

16 Intermediate (disturbed) Logged in 1947-52 Site <strong>of</strong> scientific interference N3 5.96 6.08 8.47<br />

15 Intermediate (disturbed) Logged in 1947-52 Site <strong>of</strong> scientific interference N3 5.53 6.19 9.59<br />

14 Intermediate (disturbed) Logged in 1947-52 Site <strong>of</strong> scientific interference N3 5.95 5.75 9.22<br />

13 Intermediate (disturbed) Logged in 1943-44 Sawmill harvesting B5 6.32 6.69 9.98<br />

12 Intermediate (disturbed) Logged in 1943-44 Sawmill harvesting B5 6.08 6.35 9.13<br />

11 Intermediate (disturbed) Logged in 1947-52 Site <strong>of</strong> scientific interference N3 7.38 6.08 8.73<br />

10 Late (undisturbed) Unlogged Nature reserve N15 6.47 6.70 9.28<br />

9 Late (undisturbed) Unlogged Nature reserve N15 6.51 6.13 9.82<br />

8 Intermediate (disturbed) Logged in 1945-47 Sawmill harvesting N2 6.06 7.07 10.14<br />

7 Late (undisturbed) Logged in 1945 Protection (buffer) zone N1 6.92 7.69 9.57<br />

6 Late (undisturbed) Unlogged Nature reserve N15 6.42 7.36 10.47<br />

5 Late (undisturbed) Unlogged Nature reserve N15 6.78 6.40 9.29<br />

4 Intermediate (disturbed) Logged in 1947-52 Site <strong>of</strong> scientific interference N3 6.09 6.89 9.80<br />

3 Intermediate (disturbed) Logged in 1947-52 Site <strong>of</strong> scientific interference N3 6.17 5.74 8.04<br />

2 Intermediate (disturbed) Logged in 1947-52 Site <strong>of</strong> scientific interference N3 6.35 6.41 8.90<br />

1 Intermediate (disturbed) Logged in 1947-52 Site <strong>of</strong> scientific interference N3 6.39 6.08 9.10<br />

ESU Forest stage Logging status NFA zone No. <strong>of</strong> compartment LAI e (LAI2000) LAI e (DHP) LAI (DHP)


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

5.3.2<br />

Kakamega Forest<br />

As previously explained, in situ measurements in<br />

Kakamega Forest were technically not as mature<br />

as <strong>the</strong> ones accomplished in Budongo Forest.<br />

Never<strong>the</strong>less <strong>the</strong>y are presented <strong>for</strong> comparison. First<br />

<strong>of</strong> all, differences with respect to <strong>the</strong> Budongo field<br />

campaign will briefly be discussed.<br />

The field campaign in Kakamega Forest was planned<br />

in 2004 according to state <strong>of</strong> <strong>the</strong> art recommendations<br />

from o<strong>the</strong>r studies. First <strong>of</strong> all, two LAI-2000 PCA<br />

instruments are usually used <strong>for</strong> <strong>the</strong> derivation <strong>of</strong> LAI<br />

in <strong>for</strong>ests, as explained in Chapter 5.1.3. The analysis<br />

<strong>of</strong> LAI-2000 PCA data sampled in Kakamega<br />

Forest, however, revealed that non-ideal illumination<br />

conditions could have a large impact on calculated LAI<br />

that might not be quantified or corrected without <strong>the</strong><br />

help <strong>of</strong> a third device. Second, <strong>the</strong> distance between<br />

individual measurements was defined according to<br />

<strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> <strong>the</strong> instruments (cf.<br />

Figure A-1 and Equations A.2 and A.3). Yet tests later<br />

showed that autocorrelation did not occur at smaller<br />

distances due to canopy density. As a higher density<br />

<strong>of</strong> sampling points results in better coverage <strong>of</strong> <strong>the</strong><br />

sampling unit and <strong>the</strong>re<strong>for</strong>e a higher precision in LAI,<br />

<strong>the</strong> distance between individual measurements was<br />

decreased to 10 m in mature and 5 m in early <strong>for</strong>est<br />

stages in Budongo Forest. Third, discussions within<br />

<strong>the</strong> VALERI network and <strong>the</strong> availability <strong>of</strong> s<strong>of</strong>tware<br />

that made possible <strong>the</strong> analysis <strong>of</strong> downward-looking<br />

Table 5-9<br />

Issue Kakamega Forest Budongo Forest Reason <strong>for</strong> modification<br />

Number <strong>of</strong> used LAI-2000<br />

PCA devices<br />

Distance between<br />

individual measurements<br />

Differences in LAI sampling between <strong>the</strong> first and <strong>the</strong> second field campaign.<br />

2 3 Analyse illumination effects/<br />

quantify measurement precision<br />

35/17.5 m<br />

(late/early <strong>for</strong>est stages)<br />

Figure 5-31<br />

10/5 m<br />

(late/early <strong>for</strong>est stages)<br />

Picture <strong>of</strong> ESU 19 in Kakamega Forest.<br />

93<br />

hemispherical pictures led to <strong>the</strong> acquisition <strong>of</strong> <strong>the</strong><br />

latter in Budongo Forest. Here a Nikon Coolpix 4500<br />

with FC-E8 fisheye converter was used (cf. Table A-5<br />

<strong>for</strong> image parameters). Table 5-9 presents a summary<br />

<strong>of</strong> modifications. Results <strong>of</strong> <strong>the</strong> Kakamega field<br />

campaign are given in <strong>the</strong> following.<br />

Representativeness <strong>of</strong> sampling design<br />

In Kakamega Forest a detailed and spatially explicit<br />

logging history was not available <strong>for</strong> <strong>the</strong> classification<br />

<strong>of</strong> different <strong>for</strong>est stages. Instead a recent land cover<br />

classification generated within <strong>the</strong> BIOTA EAST<br />

AFRICA project by T. Lung was used (cf. Figure<br />

A-2, <strong>for</strong> methods refer to Lung [2004]). The land<br />

cover classes bushland/shrubs, secondary bushland/<br />

Psidium guajava, grassland with scattered trees<br />

No autocorrelation at smaller distance/better<br />

coverage <strong>of</strong> sampling unit<br />

Understorey LAI Excluded Included S<strong>of</strong>tware <strong>for</strong> downward DHP not available<br />

until second field campaign


94<br />

and grassland were assigned to early <strong>for</strong>est stages,<br />

secondary <strong>for</strong>est as intermediate (disturbed) <strong>for</strong>est<br />

stage and near natural/old secondary <strong>for</strong>est was<br />

assigned to <strong>the</strong> late <strong>for</strong>est stage class. All o<strong>the</strong>r classes<br />

(plantations and water) were not considered.<br />

As mentioned be<strong>for</strong>e, different <strong>for</strong>est stages are not<br />

systematically distributed in logging compartments in<br />

Kakamega Forest. Forest stages are ra<strong>the</strong>r dispersed<br />

over <strong>the</strong> <strong>for</strong>est. Consequently ESUs were assigned to<br />

<strong>the</strong> <strong>for</strong>ests stage covering <strong>the</strong> major part <strong>of</strong> <strong>the</strong> ESU<br />

with <strong>the</strong> exception <strong>of</strong> ESU 19. Though its area is<br />

mainly deemed to be grassland with scattered trees in<br />

<strong>the</strong> classification by Lung (2004) and would thus be<br />

affiliated to <strong>the</strong> early <strong>for</strong>est stage, it was reassigned<br />

to <strong>the</strong> intermediate (disturbed) <strong>for</strong>est stage. As Figure<br />

5-31 shows it is characterized by a closed canopy <strong>of</strong><br />

up to 15 m high individuals <strong>of</strong> Psidium guajava and<br />

Bisch<strong>of</strong>fia javanica, which justified this reassignment<br />

decision.<br />

A comparison between <strong>the</strong> proportions <strong>of</strong> <strong>the</strong> three<br />

<strong>for</strong>est stages on <strong>the</strong> whole test site and <strong>the</strong> <strong>for</strong>est<br />

stages represented by <strong>the</strong> sampled ESUs indicated<br />

that in Kakamega Forest late <strong>for</strong>est stages are well<br />

covered (cf. Figure 5-32). By contrast, early <strong>for</strong>est<br />

Figure 5-32<br />

stages were underrepresented (ESUs: 9%, whole test<br />

site: 36.88%), whereas intermediate <strong>for</strong>est stages<br />

were overrepresented with 45.03% <strong>of</strong> ESUs (but<br />

only 19.61% <strong>of</strong> whole test site). This might partly<br />

be attributed to misclassifications in <strong>the</strong> land cover<br />

classification provided by Lung (2004), but cannot be<br />

verified as an accuracy assessment is not available.<br />

Spatial variability <strong>of</strong> in situ LAI<br />

Due to <strong>the</strong> lack <strong>of</strong> a third LAI-2000 PCA device, no<br />

correction <strong>of</strong> e could be applied to LAI θsun e (LAI2000)<br />

in Kakamega Forest. In order to filter presumably<br />

bad values <strong>the</strong> method developed by Broich (2005)<br />

was used. It is <strong>based</strong> on a) <strong>the</strong> range between<br />

<strong>the</strong> maximum and minimum ring values <strong>of</strong> <strong>the</strong><br />

LAI-2000 PCA at a certain time step and b) <strong>the</strong><br />

slope between two time steps. Large ranges and<br />

slopes indicate quick changes in illumination<br />

through moving clouds and are thus excluded<br />

from analysis. Yet in contrast to Broich (2005),<br />

only measurements made in a north direction were<br />

included. As <strong>the</strong> impact <strong>of</strong> <strong>the</strong> filter is quite strong,<br />

no valid LAI values could be retrieved <strong>for</strong><br />

e (LAI2000)<br />

ESUs 16 and 20.<br />

Percentage <strong>of</strong> <strong>for</strong>est stages <strong>for</strong> a) <strong>the</strong> whole study area <strong>of</strong> Kakamega Forest and b) sampled ESUs.


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

Figure 5-33 Frequency distribution <strong>of</strong> LAIe<br />

(LAI2000)<br />

on sample level in Kakamega Forest.<br />

After <strong>the</strong> elimination <strong>of</strong> presumably influenced<br />

LAI values individual measurements ranged<br />

e (LAI2000) ,<br />

from 0.00 (ESU 6) to 9.24 (ESU 14), <strong>based</strong> on a<br />

total number <strong>of</strong> 367 measurements. The frequency<br />

distribution <strong>of</strong> LAI on sample level reflects<br />

e (LAI2000)<br />

again <strong>the</strong> underrepresentation <strong>of</strong> early <strong>for</strong>est stages<br />

(cf. Figure 5-34) and low to medium LAI values. As<br />

in Budongo Forest a local maximum occurs around<br />

LAI <strong>of</strong> 6.<br />

e (LAI2000)<br />

The frequency distribution on ESU level is displayed<br />

in Figure 5-35a) to c). Data gaps occur between<br />

LAI 1.5 and 3.5, LAI 3.5 and 4 and<br />

e (LAI2000) e (DHP)<br />

LAI indicating ei<strong>the</strong>r that those values are actually<br />

(DHP)<br />

missing in Kakamega Forest or that LAI variability<br />

was not sufficiently sampled. Mean values on ESU<br />

level were lower than in Budongo Forest with mean<br />

LAI <strong>of</strong> 5.09, mean LAI <strong>of</strong> 3.93 and mean<br />

e (LAI2000) e (DHP)<br />

LAI <strong>of</strong> 6.30. Especially <strong>the</strong> last two measured<br />

(DHP)<br />

were remarkably lower that in Budongo Forest even<br />

though <strong>the</strong>re were more ESUs sampled in early<br />

<strong>for</strong>est stages with low LAI. Once more, it has to be<br />

Figure 5-34 a) ESU 25 with mean LAI <strong>of</strong> 1.49 and<br />

e (LAI2000)<br />

b) ESU 27 with mean LAI 3.96.<br />

e (LAI2000)<br />

95<br />

emphasized that understorey vegetation was not<br />

taken into account in Kakamega Forest.<br />

A comparison on <strong>for</strong>est stage level also revealed<br />

a much higher variance <strong>of</strong> LAI measures than in<br />

Budongo Forest. A Mann-Whitney test indicated that<br />

only LAI differs significantly (p


96<br />

Frequency distribution <strong>of</strong> a) LAI<br />

Figure 5-35 e (LAI2000) , b) LAI e (DHP) and c) LAI (DHP) on ESU level and d) mean LAI on <strong>for</strong>est<br />

stage level in Kakamega Forest. For explanations <strong>of</strong> d) see Figure 5-29.


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

ESU Forest stage Land cover (Lung & Schaab 2004) LAI e (LAI2000) LAI e (DHP) (excl. understorey) LAI (DHP) (excl. understorey)<br />

1 Intermediate (disturbed) Secondary <strong>for</strong>est 3.95 3.54 5.64<br />

2 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 5.51 3.75 6.22<br />

3 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 5.06 4.41 6.91<br />

4 Intermediate (disturbed) Secondary <strong>for</strong>est 5.26 4.56 7.09<br />

5 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 5.93 4.07 6.13<br />

6 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 5.48 4.01 6.59<br />

7 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 5.56 3.77 5.71<br />

8 Early Grassland with scattered trees 1.49 2.80 6.82<br />

9 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 5.72 3.81 5.78<br />

10 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 6.36 4.25 6.37<br />

11 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 6.23 4.06 5.85<br />

12 Intermediate (disturbed) Secondary <strong>for</strong>est 6.11 3.77 5.69<br />

13 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 5.69 3.96 6.20<br />

14 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 7.35 4.59 7.78<br />

15 Intermediate (disturbed) Secondary <strong>for</strong>est 4.71 3.71 5.90<br />

Forest stage, logging status, NFA zone, compartment number and LAI measures (LAI<br />

Table 5-10 e (LAI2000) , LAI e (DHP) , and LAI (DHP) ) <strong>of</strong><br />

<strong>the</strong> 30 ESUs in Kakamega Forest.<br />

16 Late (undisturbed) Near natural and old secondary <strong>for</strong>est n/a 5.10 8.47<br />

17 Intermediate (disturbed) Secondary <strong>for</strong>est 5.53 4.36 6.45<br />

18 Intermediate (disturbed) Secondary <strong>for</strong>est 5.75 4.94 7.60<br />

19 Early Bushland/shrubs 4.31 2.82 4.65<br />

20 Late (undisturbed) Near natural and old secondary <strong>for</strong>est n/a 4.24 6.42<br />

21 Intermediate (disturbed) Secondary <strong>for</strong>est 5.51 4.10 6.91<br />

22 Intermediate (disturbed) Secondary <strong>for</strong>est 4.83 3.69 6.30<br />

23 Intermediate (disturbed) Secondary <strong>for</strong>est 4.39 3.71 5.79<br />

24 Intermediate (disturbed) Secondary <strong>for</strong>est 5.23 4.22 6.21<br />

25 Early Psidium guajava bushland 1.41 2.35 4.63<br />

26 Intermediate (disturbed) Secondary <strong>for</strong>est 6.29 4.06 6.42<br />

27 Intermediate (disturbed) Secondary <strong>for</strong>est 4.01 4.42 7.51<br />

28 Early Grassland with scattered trees 3.96 2.65 4.86<br />

29 Intermediate (disturbed) Secondary <strong>for</strong>est 5.13 4.05 6.13<br />

97<br />

30 Late (undisturbed) Near natural and old secondary <strong>for</strong>est 5.79 4.23 6.04


98<br />

5.4<br />

Conclusion<br />

The preceding chapter presented <strong>the</strong> design <strong>of</strong> a valid<br />

sampling scheme <strong>for</strong> in situ LAI measurements in<br />

tropical rain <strong>for</strong>ests, <strong>the</strong> analysis <strong>of</strong> field data and<br />

<strong>the</strong> development <strong>of</strong> correction methods as well as<br />

<strong>the</strong> results <strong>of</strong> <strong>the</strong>se field measurements <strong>for</strong> <strong>the</strong> test<br />

sites Budongo and Kakamega Forest. Although<br />

sampling schemes recommended by CEOS-LPV<br />

and <strong>the</strong> VALERI network served as a basis, several<br />

modifications had to be made to account <strong>for</strong> problems<br />

associated with <strong>the</strong> tropical rain <strong>for</strong>est environment<br />

(cf. Table 5-1).<br />

One <strong>of</strong> <strong>the</strong> main challenges in validating medium<br />

to coarse spatial resolution satellite data such as<br />

<strong>MODIS</strong> is to find a correct method to compare<br />

satellite data and in situ measurements. Whereas<br />

<strong>the</strong> approach employed by <strong>the</strong> CEOS-LPV and<br />

<strong>the</strong> VALERI networks – namely <strong>the</strong> usage <strong>of</strong> high<br />

resolution satellite data as an intermediate step in<br />

upscaling – is basically correct and applicable to<br />

<strong>the</strong> tropical rain <strong>for</strong>est environment, adjustments<br />

have to be made especially with respect to ESU<br />

size. Whereas in o<strong>the</strong>r studies ESUs equivalent to<br />

<strong>the</strong> size <strong>of</strong> a pixel <strong>of</strong> high resolution satellite data<br />

were set up, <strong>for</strong> several reasons <strong>the</strong> sampling units<br />

had to be increased to 200 x 200 m <strong>for</strong> this <strong>the</strong>sis.<br />

First <strong>of</strong> all, an accurate geolocation <strong>of</strong> <strong>the</strong> sites with<br />

a handheld GPS is critical. Although errors were<br />

minimized as far as possible, a location shift <strong>of</strong> up<br />

to 30 m is still <strong>the</strong>oretically possible. Second, <strong>the</strong><br />

geometric correction <strong>of</strong> high resolution satellite<br />

data is associated with a certain error, as map sheets<br />

applied <strong>for</strong> geometric adjustments were more than<br />

35-40 years old and ground control points could<br />

not be set inside <strong>the</strong> <strong>for</strong>est area. Consequently<br />

<strong>the</strong> relationship between in situ LAI values and<br />

<strong>the</strong> corresponding surface reflectance data might<br />

still contain inaccuracies. This will be subject to<br />

discussion in <strong>the</strong> following chapter. The applied<br />

sampling pattern and sampling distance were<br />

<strong>based</strong> on thorough <strong>the</strong>oretical considerations and<br />

recommendations found in <strong>the</strong> relevant literature.<br />

Indirect optical instruments were used to<br />

derive in situ LAI. Here, <strong>for</strong> <strong>the</strong> first time three<br />

LAI-2000 PCA devices were employed. The<br />

stationary device on each ESU allowed <strong>the</strong> analysis<br />

and quantification <strong>of</strong> θ effects and <strong>the</strong> influence <strong>of</strong><br />

sun<br />

changing illumination conditions (i.e. moving clouds)<br />

during <strong>the</strong> time <strong>of</strong> measurement. Results suggest that<br />

<strong>the</strong> influence <strong>of</strong> direct radiation (which is increasing<br />

with decreasing θ ) must not be disregarded and<br />

sun<br />

is possibly even higher in lower canopies with less<br />

foliage. Yet <strong>the</strong> development <strong>of</strong> a correction scheme<br />

led to LAI-2000 PCA results that are unbiased by<br />

θ . The influence <strong>of</strong> moving clouds on <strong>the</strong> data<br />

sun<br />

was minimized by a filter application. The proposed<br />

methodology can also be transferred to o<strong>the</strong>r<br />

vegetation types and gives a detailed insight to field<br />

data quality. It far exceeds <strong>the</strong> sun zenith angle<br />

correction scheme developed by Leblanc<br />

& Chen (2001), which is <strong>based</strong> on a single<br />

device and mean plot values only. Fur<strong>the</strong>r it<br />

contradicts <strong>the</strong> recommendations made by<br />

De Wasseige et al. (2003), who found θ <strong>of</strong> less<br />

sun<br />

than 50° to be best suited <strong>for</strong> LAI-2000 PCA<br />

measurements in tropical rain <strong>for</strong>ests. This might<br />

occur because <strong>the</strong>ir analysis was <strong>based</strong> on only two<br />

devices and three groups <strong>of</strong> data taken on <strong>the</strong> same<br />

transects at different times <strong>of</strong> <strong>the</strong> day. The results <strong>of</strong><br />

De Wasseige et al. (2003) suggest that measurements<br />

taken on <strong>the</strong> same transect at θ < 50° differ least,<br />

sun<br />

which is also true <strong>for</strong> <strong>the</strong> data presented in this chapter.<br />

Yet if measurements are taken at low sun zenith<br />

angles, e is also higher (cf. Chapter 5.2.1), resulting<br />

θsun<br />

in a necessary correction <strong>for</strong> e . As shown in this<br />

θsun<br />

chapter, this correction can be achieved through <strong>the</strong><br />

application <strong>of</strong> three devices. This, however, is very<br />

expensive and also requires accurate intercalibration.<br />

Fur<strong>the</strong>r, <strong>the</strong> selection <strong>of</strong> <strong>the</strong> measurement location <strong>for</strong><br />

<strong>the</strong> reference sensor is crucial, yet <strong>of</strong>ten hindered in<br />

tropical rain <strong>for</strong>ests as clearings or natural gaps are not


5 <strong>Ground</strong>-<strong>based</strong> LAI measurements<br />

large enough. Consequently a large distance between<br />

reference sensor and ESU had to be accepted as a<br />

trade-<strong>of</strong>f, leading to a stronger influence <strong>of</strong> changing<br />

illumination conditions on <strong>the</strong> results and subsequent<br />

reduction <strong>of</strong> valid readings by filtering.<br />

Digital hemispherical photographs were also analysed<br />

<strong>for</strong> understorey LAI. Yet care has to be taken with<br />

respect to general interpretation, as measurement<br />

heights differed with <strong>for</strong>est stages and understorey<br />

LAI results are thus not comparable between ESUs.<br />

If only intermediate and late <strong>for</strong>est stages are<br />

considered, understorey contributed up to 14% to<br />

total LAI and should thus not be disregarded.<br />

e (DHP)<br />

Though <strong>the</strong> correction <strong>for</strong> non-random foliage<br />

distribution applied to LAI is an attempt to<br />

e (DHP)<br />

derive true LAI values that are corrected <strong>for</strong> foliage<br />

clumping, it must be viewed as a critical step.<br />

Especially LAI values <strong>of</strong> <strong>the</strong> early <strong>for</strong>est stages with<br />

values <strong>of</strong> up to 7.25 (plot 17 in Budongo Forest)<br />

indicate an overestimation <strong>of</strong> λ, as LAI seems to be<br />

much too high <strong>for</strong> wooded grassland. The clumping<br />

factors are derived from DHP processing, and lie<br />

around 0.50 on average <strong>for</strong> ESUs <strong>of</strong> early <strong>for</strong>est<br />

stages.<br />

As a summary, Table 5-11 lists <strong>the</strong> different<br />

properties <strong>of</strong> both devices. With respect to tropical<br />

rain <strong>for</strong>ests, DHP does not only <strong>of</strong>fer a cheaper<br />

solution, but – provided that <strong>the</strong> results are accurate<br />

– also <strong>the</strong> possibility <strong>of</strong> correction <strong>for</strong> clumping.<br />

Table 5-11<br />

Comparison between <strong>the</strong> properties <strong>of</strong> LAI-2000 PCA and DHP.<br />

System Illumination<br />

conditions<br />

LAI-2000 PCA Preferably diffuse 0-74°<br />

in five rings<br />

Digital camera with<br />

fisheye lens<br />

Zenith coverage Azimuth coverage Computer resources Cost in €<br />

Up to 360°<br />

(selectable by view caps)<br />

99<br />

DHP is also less sensitive to illumination conditions<br />

(direct or changing) and has a similar zenith and<br />

azimuth coverage as LAI-2000 PCA (<strong>the</strong>oretically<br />

it has a higher zenithal coverage, however zenith<br />

angles between 60 and 90° are usually not taken<br />

into consideration). Although DHP demands greater<br />

computer resources, <strong>the</strong>re is compensation <strong>for</strong> this in<br />

<strong>the</strong> <strong>for</strong>m <strong>of</strong> a permanent in<strong>for</strong>mation archive <strong>of</strong> each<br />

study site.<br />

Intercomparison at <strong>the</strong> ESU level reveals that <strong>the</strong>re<br />

are differences in <strong>the</strong> results <strong>of</strong> LAI-2000 and DHP.<br />

As LAI includes understorey, it should be higher<br />

e (DHP)<br />

than LAI <strong>for</strong> all ESUs. This is, however, not<br />

e (LAI2000)<br />

always <strong>the</strong> case (11 out <strong>of</strong> 30 ESUs in Budongo<br />

Forest, cf. Table 5-8). Classification problems<br />

associated with mixed pixels might be <strong>the</strong> reason <strong>for</strong><br />

this, especially when DHP are acquired under high<br />

canopies. The distance to <strong>the</strong> highest foliage elements<br />

does not allow single leaves to be represented by<br />

one or more pixels. Pixels ra<strong>the</strong>r represent several<br />

foliage units so that misclassifications can cause<br />

inaccuracies in <strong>the</strong> retrieved LAI. This is underlined<br />

by <strong>the</strong> fact that in Budongo mean LAI and<br />

e (LAI2000)<br />

mean LAI differ <strong>for</strong> <strong>the</strong> early <strong>for</strong>est stage, but<br />

e (DHP)<br />

do not differ <strong>for</strong> <strong>the</strong> intermediate and late <strong>for</strong>est<br />

stages. In Kakamega Forest, where hemispherical<br />

pictures <strong>of</strong> <strong>the</strong> understorey were not taken, LAIe (DHP)<br />

is much lower than LAI <strong>for</strong> <strong>the</strong> last two stages,<br />

e (LAI2000)<br />

indicating an underestimation <strong>of</strong> LAI through DHP<br />

<strong>for</strong> this test site.<br />

Low Ca. 8.900,-<br />

Diffuse/direct 0-90° 360° High Ca. 800,-


100<br />

In order to carry out an absolute calibration <strong>of</strong> <strong>the</strong><br />

applied indirect methods, supplementary direct<br />

or semi-direct methods (assumed to give “true”<br />

LAI) would be necessary. Due to <strong>the</strong> complexity <strong>of</strong><br />

vegetation stands, <strong>the</strong> high species numbers and <strong>the</strong><br />

lack <strong>of</strong> seasonal leaf litter fall in tropical rain <strong>for</strong>ests,<br />

this would have exceeded <strong>the</strong> possible fieldwork<br />

within <strong>the</strong> scope <strong>of</strong> this <strong>the</strong>sis. Only one study<br />

published to date encompasses direct landscapescale<br />

measurements <strong>for</strong> a tropical rain <strong>for</strong>est LAI<br />

(Clark et al. 2008). As part <strong>of</strong> a larger project, in this<br />

study a modular tower was used to harvest leaves<br />

and branches in 55 vertical transects in old-growth<br />

tropical rain <strong>for</strong>est around La Selva Biological<br />

Station in Costa Rica. A mean landscape LAI <strong>of</strong><br />

6.00 was retrieved (with individual measurements<br />

ranging from 1.20 to 12.94), which corresponds<br />

approximately to mean LAI and LAI e (LAI2000) e (DHP)<br />

values retrieved <strong>for</strong> this <strong>the</strong>sis from Budongo Forest.<br />

Fur<strong>the</strong>r comparison to o<strong>the</strong>r studies using indirect<br />

optical measurements in tropical rain <strong>for</strong>ests<br />

revealed that mostly higher values <strong>for</strong> effective LAI<br />

were retrieved <strong>for</strong> Budongo (cf. Table 3-2). Results<br />

reported by Aragão et al. (2005) <strong>for</strong> an evergreen<br />

rain <strong>for</strong>est in Brazil, De Wasseige et al. (2003) <strong>for</strong> a<br />

semi-deciduous <strong>for</strong>est in <strong>the</strong> Car and Laumonier et<br />

al. (1994) <strong>for</strong> an evergreen rain <strong>for</strong>est in Cameroon<br />

are slightly lower on average <strong>for</strong> late <strong>for</strong>est stages,<br />

yet in all three sites uncorrected LAI-2000 PCA<br />

measurements were used. In contrast, <strong>the</strong> LAI<br />

retrieved <strong>for</strong> <strong>the</strong> VALERI site in Counami (evergreen<br />

rain <strong>for</strong>est in French Guiana) and published by<br />

Rossello et al. (2007) seems to be unreasonably low<br />

with a mean LAI <strong>of</strong> 3.38 and a mean LAI <strong>of</strong><br />

e (DHP) (DHP)<br />

4.92. As in <strong>the</strong> above-mentioned studies, understorey<br />

vegetation was not considered. Interestingly, o<strong>the</strong>r<br />

studies found a significant difference between LAI<br />

<strong>of</strong> late and intermediate <strong>for</strong>est stages (primary and<br />

secondary <strong>for</strong>ests respectively), as shown in Table<br />

3-4. Probably <strong>the</strong> inclusion <strong>of</strong> early <strong>for</strong>est stages<br />

explains <strong>the</strong> comparably low LAI values <strong>of</strong> <strong>the</strong> latter.<br />

Summarizing, it remains hard – if not impossible<br />

– to judge which <strong>of</strong> <strong>the</strong> two methods applied leads<br />

to better results. In any case, LAI seems to<br />

(DHP)<br />

give unreasonably high results. With respect to<br />

LAI and LAI measurement precision is<br />

e (LAI2000) e (DHP)<br />

an important factor <strong>for</strong> fur<strong>the</strong>r upscaling that should<br />

not be disregarded and will be discussed thoroughly<br />

in <strong>the</strong> next chapter.


6 Derivation<br />

<strong>of</strong> high<br />

resolution LAI maps<br />

The following chapter describes <strong>the</strong> up scaling<br />

<strong>of</strong> in situ LAI measurements with <strong>the</strong> help <strong>of</strong> high<br />

resolution satellite data introduced in Chapter 4.<br />

Following <strong>the</strong> recommendations <strong>of</strong> Baret el al.<br />

(submitted), empirical relationships between field<br />

data and SVIs or texture measures are established. To<br />

provide a solid validation basis, special focus is here<br />

put on <strong>the</strong> correct establishment <strong>of</strong> transfer functions<br />

as well as <strong>the</strong> consideration <strong>of</strong> measurement errors in<br />

ground and satellite data (cf. Chapter 6.2). For this<br />

purpose a robust regression method, namely Theil-<br />

Sen regression, is introduced and compared to an<br />

approach that accounts <strong>for</strong> measurement precision in<br />

input data (Tan et al. 2005). The analyses <strong>of</strong> in situ<br />

and satellite data revealed that different regression<br />

models had to be applied to early/intermediate and<br />

late <strong>for</strong>est stages. The resulting LAI maps <strong>for</strong> both<br />

study sites represent independent data sets needed <strong>for</strong><br />

<strong>the</strong> validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product (cf. Chapter<br />

7). At least <strong>for</strong> Budongo Forest <strong>the</strong> upscaling could<br />

be per<strong>for</strong>med without any degradation in accuracy<br />

compared to field measurements. Both maps are<br />

presented toge<strong>the</strong>r with <strong>the</strong>ir accuracies at <strong>the</strong> end <strong>of</strong><br />

this chapter.<br />

(submitted), empirical relationships between field data and SVIs o<br />

provide a solid validation basis, special focus is here put on <strong>the</strong> cor<br />

101<br />

as well as <strong>the</strong> consideration <strong>of</strong> measurement errors in ground and<br />

purpose a robust regression method, namely Theil-Sen regressio<br />

approach that accounts <strong>for</strong> measurement precision in input data (<br />

and<br />

6.1<br />

satellite<br />

Calculation<br />

data revealed<br />

<strong>of</strong><br />

that<br />

spectral<br />

different<br />

vegetation<br />

regression models had to b<br />

<strong>for</strong>est stages.<br />

indices<br />

The resulting<br />

and texture<br />

LAI maps<br />

measures<br />

<strong>for</strong> both study sites represen<br />

validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product (cf. Chapter 7). At least <strong>for</strong><br />

per<strong>for</strong>med<br />

For both Kakamega<br />

without<br />

and<br />

any<br />

Budongo<br />

degradation<br />

Forest<br />

in<br />

<strong>the</strong><br />

accuracy<br />

available<br />

compared to<br />

presented<br />

pre-processed<br />

toge<strong>the</strong>r<br />

SPOT-4<br />

with<br />

and<br />

<strong>the</strong>ir<br />

ASTER<br />

accuracies<br />

data<br />

at<br />

were<br />

<strong>the</strong><br />

used<br />

end <strong>of</strong><br />

to<br />

this chapter.<br />

calculate <strong>the</strong> SVIs listed in Table 3-5. SWIR and min<br />

SWIR values as needed in NDVI and RSR were<br />

max c<br />

derived directly from <strong>the</strong> <strong>for</strong>est area <strong>of</strong> <strong>the</strong> respective<br />

For<br />

test<br />

both<br />

sites (<strong>for</strong><br />

Kakamega<br />

ASTER<br />

and<br />

band<br />

Budongo<br />

4 was used).<br />

Forest <strong>the</strong> available pre-process<br />

to calculate <strong>the</strong> SVIs listed in Table 3-5. SWIRmin and SWIRmax val<br />

derived<br />

With respect<br />

directly<br />

to<br />

from<br />

image<br />

<strong>the</strong><br />

texture,<br />

<strong>for</strong>est<br />

second-order<br />

area <strong>of</strong> <strong>the</strong> respective<br />

statistics<br />

test sites (<strong>for</strong><br />

With were calculated respect to <strong>based</strong> image on texture, a GLCM second-order according to statistics Table were calcu<br />

Table 6-1 with 6-1 with<br />

6.1 Calculation <strong>of</strong> spectral vegetation indices an<br />

P<br />

V<br />

<br />

i , j <br />

i , j<br />

N 1<br />

Vi<br />

, j<br />

i , j 0<br />

(6.1)<br />

being a normalization equation within <strong>the</strong> respective<br />

kernel, that<br />

column kernel, that j (starting divides with <strong>the</strong> 0,0) pixel by value <strong>the</strong> sum V at <strong>of</strong> row values. i and According to<br />

(2007) column a j horizontal (starting with <strong>of</strong>fset 0,0) distance by <strong>the</strong> <strong>of</strong> sum one pixel <strong>of</strong> values. was used and diffe<br />

7x7, According 9x9 and to 11x11 <strong>the</strong> recommendations pixels were compared. by Hall-Beyer<br />

(2007) a horizontal <strong>of</strong>fset distance <strong>of</strong> one pixel was<br />

used and different moving window sizes <strong>of</strong> 3x3, 5x5,<br />

7x7, 9x9 and 11x11 pixels were compared.<br />

For all ESUs, mean SVIs and texture measures were<br />

extracted as well as <strong>the</strong>ir standard deviation. All<br />

pixels whose centre lay within <strong>the</strong> 200 m x 200 m<br />

area <strong>of</strong> <strong>the</strong> respective ESU were taken into account.<br />

Additionally mean and standard deviation per ESU<br />

<strong>of</strong> all reflective bands were stored.<br />

6.2<br />

Establishment <strong>of</strong> transfer<br />

functions<br />

In situ data, reflectance, SVIs and texture measures<br />

served as input <strong>for</strong> regression analysis. The goal<br />

is <strong>the</strong> establishment <strong>of</strong> an empirical relationship<br />

that can be used to invert LAI <strong>for</strong> <strong>the</strong> whole test<br />

site from reflectance data (or SVIs and texture<br />

respectively). However, both field LAI values and


accuracy observations <strong>of</strong> In situ <strong>the</strong> data, resulting reflectance, include high measurement resolution SVIs and LAI texture errors observations maps. measures that product Knowledge influence served include is crucial about <strong>the</strong> as measurem input <strong>for</strong> derived <strong>the</strong> a mfo<br />

q<br />

product accuracy is establishment crucial <strong>of</strong> <strong>for</strong> <strong>the</strong> a meaningful <strong>of</strong> resulting an empirical high and accurate resolution relationship accuracy validation LAI in<strong>for</strong>mation that maps. <strong>of</strong> can <strong>of</strong> <strong>the</strong> <strong>the</strong> be Knowledge resulting on used <strong>MODIS</strong> model to high inve LAI abo an<br />

102<br />

in<strong>for</strong>mation product reflectance on model is crucial data and <strong>for</strong> (or data a SVIs meaningful uncertainties and texture and into product accurate respectively). account. errors is validation crucial and As <strong>the</strong>ir a <strong>for</strong> However, result, <strong>of</strong> a impact <strong>the</strong> meaning MO quan bo<br />

errors and in<strong>for</strong>mation observations <strong>the</strong>ir impact on include on model <strong>the</strong> measurement and establishment data uncertainties errors <strong>of</strong> in<strong>for</strong>mation an errors empirical that into influence in on account. both transfer model data <strong>the</strong> As and functio sets, derive a res dat fi<br />

errors in errors both accuracy data and <strong>of</strong> sets, <strong>the</strong>ir <strong>the</strong> field impact resulting LAI on and high <strong>the</strong> remote establishment resolution sensing errors LAI to data, and <strong>of</strong> that, maps. an <strong>the</strong>ir have empirical regression Knowledge impact to be assessed transfer on analys <strong>the</strong> ab<br />

satellite observations include measurement errors to that that, errors regression product 6.2.1 in both is Quantification analysis crucial data sets, <strong>for</strong> is a per<strong>for</strong>med. field meaningful <strong>of</strong> LAI measurement and At and remote errors this accurate point applicable in sensing both validation differing data data, in <strong>the</strong> sets, have approache <strong>of</strong> event field <strong>the</strong> to tha MO be LA<br />

influence <strong>the</strong> derived relationships and thus lower applicable <strong>the</strong> to in<strong>for</strong>mation in that, <strong>the</strong> event regression errors on that model input analysis data and is is data per<strong>for</strong>med. not uncertainties error-free. to that, At this regression into point account. differing analysis As a is re ap<br />

accuracy <strong>of</strong> <strong>the</strong> resulting high resolution LAI maps. applicable errors and in <strong>the</strong>ir <strong>the</strong> event impact that on input <strong>the</strong> data establishment is applicable not error-free. 6.2.1 <strong>of</strong> in an <strong>the</strong> Quantification<br />

empirical event that transfe inpu<br />

Knowledge about <strong>the</strong> quality <strong>of</strong> this intermediate 6.2.1 Quantification errors To pursue in both <strong>the</strong> data quantification <strong>of</strong> sets, measurement field <strong>of</strong> LAI measurement and remote errors To errors, sensing pursue data, <strong>the</strong> quantifica have to b<br />

product is crucial <strong>for</strong> a meaningful and accurate To pursue 6.2.1 to <strong>the</strong> <strong>the</strong> that, terms quantification Quantification regression error, uncertainty, analysis <strong>of</strong> measurement <strong>of</strong> is measurement per<strong>for</strong>med. bias, 6.2.1 precision errors, accuracy At Quantification <strong>the</strong> errors this and terms will point be error, differing defined uncer <strong>of</strong> fi<br />

validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product, which accuracy also To will applicable accuracy pursue be defined will <strong>the</strong> in <strong>the</strong> be quantification first event defined as <strong>the</strong>y that first are input <strong>of</strong> as frequently measurement data <strong>the</strong>y To is are not misused. pursue frequently published error-free. errors, <strong>the</strong> For quantification by <strong>the</strong> this FOODY terms <strong>the</strong>sis <strong>the</strong> erro & o<br />

takes in<strong>for</strong>mation on model and data uncertainties published into accuracy misused. by FOODY will For be & defined this ATKINSON <strong>the</strong>sis first as (2002) <strong>the</strong> <strong>the</strong>y widely are will accuracy frequently be accepted<br />

xfollowed. 0 may<br />

will misused. be<br />

be<br />

defined<br />

defined Accordingly, For first<br />

as<br />

this<br />

<strong>the</strong><br />

as th<br />

d<br />

account. As a result, quantification <strong>of</strong> measurement definitions published by Foody & Atkinson (2002)<br />

x 0 may be published 6.2.1 Quantification<br />

defined as by <strong>the</strong> FOODY difference & ATKINSON <strong>of</strong> measurement<br />

between <strong>the</strong> (2002) published true value will by errors be ( FOODY followed.<br />

0xz ) and <strong>the</strong> & Accor ATKI meas<br />

errors and <strong>the</strong>ir impact on <strong>the</strong> establishment <strong>of</strong> an will be followed. Accordingly, an error ˆ)( 0 at 0 x. 0 )<br />

x<br />

To<br />

0 may<br />

pursue<br />

be defined<br />

<strong>the</strong> quantification<br />

as <strong>the</strong> difference<br />

<strong>of</strong> measurement<br />

between x 0 may <strong>the</strong> be true<br />

errors,<br />

defined value<br />

<strong>the</strong><br />

as ( <strong>the</strong><br />

terms<br />

0xz ) differe and<br />

er<br />

empirical transfer function is needed. Consequently ˆ)( 0 accuracy location 0 . xzxzx will 0 )()(<br />

may be defined be defined first as as <strong>the</strong> <strong>the</strong>y difference are frequently between e misused. For this t<br />

An error is thus data base<br />

errors in both data sets, field LAI and remote sensing <strong>the</strong> true ˆ)( 0 value 0 . xzxzx 0 )()(<br />

and <strong>the</strong> measured value ˆ)( 0<br />

0 , . xz<br />

e<br />

0 )()<br />

An error is<br />

published<br />

thus data<br />

by FOODY<br />

<strong>based</strong> and refers<br />

& ATKINSON<br />

to a single<br />

(2002)<br />

measurement. error<br />

will<br />

is<br />

be<br />

well<br />

followed.<br />

In defined. this case,<br />

Acc<br />

Uncn<br />

data, have to be assessed as a first step. Subsequent so that<br />

error is well An x 0error<br />

may defined. be is thus defined Uncertainty data as <strong>based</strong> <strong>the</strong> in difference and turn refers is related between An to a error to single from <strong>the</strong> prediction is true regression measurement. thus value data <strong>based</strong> analysis, <strong>based</strong> ( on 0xz In ) stat and<br />

thi a<br />

to that, regression analysis is per<strong>for</strong>med. At this point<br />

from regression error is analysis, well defined. and is Uncertainty associated to in three turn error is o<strong>the</strong>r related is well terms, to defined. prediction namely Uncertain bias, <strong>based</strong> pre<br />

differing approaches are introduced that are applicable ˆ)( 0<br />

0 . xzxzx 0 )()( (6.2)<br />

e<br />

from regression analysis, and is associated from to three regression o<strong>the</strong>r analysis, terms, namely and is<br />

in <strong>the</strong> event that input data is not error-free.<br />

An error is thus data <strong>based</strong> and refers to a single measurement. In t<br />

error is well defined. Uncertainty in turn is related to prediction bas<br />

from regression analysis, and is associated to three o<strong>the</strong>r terms, namel<br />

Table 6-1<br />

Second-order texture measures used in this <strong>the</strong>sis (definition according to Hall-Beyer 2007).<br />

Texture measure Equation<br />

Mean (μ i )<br />

2 Variance (σ ) i<br />

Homogeneity<br />

Contrast<br />

Entropy<br />

Correlation<br />

2 DERIVATION OF HIGH RESOLUTION LAI M<br />

Table 6-1: Second-order texture measures used in this <strong>the</strong>sis (definition according to HALL-BEYER 2007)<br />

Texture measure Equation<br />

Mean ( i )<br />

1<br />

,<br />

0,<br />

)(<br />

N<br />

Pi ji<br />

ji<br />

2<br />

Variance ( i )<br />

1<br />

2<br />

, <br />

0,<br />

)(<br />

N<br />

iP ij i<br />

ji<br />

Homogeneity<br />

N 1 P , ji<br />

2<br />

ji 0,<br />

ji )(1<br />

Contrast<br />

1<br />

2<br />

,<br />

0,<br />

)(<br />

jiP<br />

N<br />

ji<br />

ji<br />

<br />

<br />

Entropy<br />

N 1<br />

, ji , ji )l n(<br />

ji 0,<br />

PP<br />

Correlation<br />

<br />

<br />

<br />

<br />

<br />

<br />

1 <br />

<br />

<br />

<br />

0, <br />

<br />

<br />

22<br />

2 DERIVATION OF HIGH RESOLUTION LAI M<br />

Table 6-1: Second-order texture measures used in this <strong>the</strong>sis (definition according to HALL-BEYER 2007)<br />

Texture measure Equation<br />

Mean ( i )<br />

<br />

N<br />

ii<br />

<br />

ji<br />

P , ji<br />

ji ji<br />

For all ESUs, mean SVIs and texture measures were extracted as well as <strong>the</strong>ir standard deviation.<br />

pixels whose centre lay within <strong>the</strong> 200 m x 200 m area <strong>of</strong> <strong>the</strong> respective ESU were taken into accou<br />

Additionally mean and standard deviation per ESU <strong>of</strong> all reflective bands were stored.<br />

6.2 Establishment <strong>of</strong> transfer functions<br />

In situ data, reflectance, SVIs and texture measures served as input <strong>for</strong> regression analysis. The goal is<br />

establishment <strong>of</strong> an empirical relationship that can be used to invert LAI <strong>for</strong> <strong>the</strong> whole test site fr<br />

reflectance data (or SVIs and texture respectively). However, both field LAI values and sate<br />

observations include measurement errors that influence <strong>the</strong> derived relationships and thus lower<br />

accuracy <strong>of</strong> <strong>the</strong> resulting high resolution LAI maps. Knowledge about <strong>the</strong> quality <strong>of</strong> this intermed<br />

1<br />

,<br />

0,<br />

)(<br />

N<br />

Pi ji<br />

ji<br />

2<br />

Variance ( i )<br />

1<br />

2<br />

, <br />

0,<br />

)(<br />

N<br />

iP ij i<br />

ji<br />

Homogeneity<br />

N 1 P , ji<br />

2<br />

ji 0,<br />

ji )(1<br />

Contrast<br />

1<br />

2<br />

,<br />

0,<br />

)(<br />

jiP<br />

N<br />

ji<br />

ji<br />

<br />

<br />

Entropy<br />

N 1<br />

, ji , ji )l n(<br />

ji 0,<br />

PP<br />

Correlation<br />

<br />

<br />

<br />

<br />

<br />

<br />

1 <br />

<br />

<br />

<br />

0, <br />

<br />

<br />

22<br />

2 DERIVATION OF HIGH RESOLUTION LAI M<br />

Table 6-1: Second-order texture measures used in this <strong>the</strong>sis (definition according to HALL-BEYER 2007)<br />

Texture measure Equation<br />

Mean ( i )<br />

<br />

N<br />

ii<br />

<br />

ji<br />

P , ji<br />

ji ji<br />

For all ESUs, mean SVIs and texture measures were extracted as well as <strong>the</strong>ir standard deviation.<br />

pixels whose centre lay within <strong>the</strong> 200 m x 200 m area <strong>of</strong> <strong>the</strong> respective ESU were taken into acco<br />

Additionally mean and standard deviation per ESU <strong>of</strong> all reflective bands were stored.<br />

6.2 Establishment <strong>of</strong> transfer functions<br />

In situ data, reflectance, SVIs and texture measures served as input <strong>for</strong> regression analysis. The goal is<br />

establishment <strong>of</strong> an empirical relationship that can be used to invert LAI <strong>for</strong> <strong>the</strong> whole test site f<br />

reflectance data (or SVIs and texture respectively). However, both field LAI values and sate<br />

observations include measurement errors that influence <strong>the</strong> derived relationships and thus lower<br />

1<br />

,<br />

0,<br />

)(<br />

N<br />

Pi ji<br />

ji<br />

2<br />

Variance ( i )<br />

1<br />

2<br />

, <br />

0,<br />

)(<br />

N<br />

iP ij i<br />

ji<br />

Homogeneity<br />

N 1 P , ji<br />

2<br />

ji 0,<br />

ji )(1<br />

Contrast<br />

1<br />

2<br />

,<br />

0,<br />

)(<br />

jiP<br />

N<br />

ji<br />

ji<br />

<br />

<br />

Entropy<br />

N 1<br />

, ji , ji )l n(<br />

ji 0,<br />

PP<br />

Correlation<br />

<br />

<br />

<br />

<br />

<br />

<br />

1 <br />

<br />

<br />

<br />

0, <br />

<br />

<br />

22<br />

2 DERIVATION OF HIGH RESOLUTION LAI M<br />

Table 6-1: Second-order texture measures used in this <strong>the</strong>sis (definition according to HALL-BEYER 2007)<br />

Texture measure Equation<br />

Mean ( i )<br />

<br />

N<br />

ii<br />

<br />

ji<br />

P , ji<br />

ji ji<br />

For all ESUs, mean SVIs and texture measures were extracted as well as <strong>the</strong>ir standard deviation.<br />

pixels whose centre lay within <strong>the</strong> 200 m x 200 m area <strong>of</strong> <strong>the</strong> respective ESU were taken into acco<br />

Additionally mean and standard deviation per ESU <strong>of</strong> all reflective bands were stored.<br />

6.2 Establishment <strong>of</strong> transfer functions<br />

In situ data, reflectance, SVIs and texture measures served as input <strong>for</strong> regression analysis. The goal is<br />

establishment <strong>of</strong> an empirical relationship that can be used to invert LAI <strong>for</strong> <strong>the</strong> whole test site f<br />

reflectance data (or SVIs and texture respectively). However, both field LAI values and sate<br />

1<br />

,<br />

0,<br />

)(<br />

N<br />

Pi ji<br />

ji<br />

2<br />

Variance ( i )<br />

1<br />

2<br />

, <br />

0,<br />

)(<br />

N<br />

iP ij i<br />

ji<br />

Homogeneity<br />

N 1 P , ji<br />

2<br />

ji 0,<br />

ji )(1<br />

Contrast<br />

1<br />

2<br />

,<br />

0,<br />

)(<br />

jiP<br />

N<br />

ji<br />

ji<br />

<br />

<br />

Entropy<br />

N 1<br />

, ji , ji )l n(<br />

ji 0,<br />

PP<br />

Correlation<br />

<br />

<br />

<br />

<br />

<br />

<br />

1 <br />

<br />

<br />

<br />

0, <br />

<br />

<br />

22<br />

2 DERIVATION OF HIGH RESOLUTION LAI M<br />

Table 6-1: Second-order texture measures used in this <strong>the</strong>sis (definition according to HALL-BEYER 2007)<br />

Texture measure Equation<br />

Mean ( i )<br />

<br />

N<br />

ii<br />

<br />

ji<br />

P , ji<br />

ji ji<br />

For all ESUs, mean SVIs and texture measures were extracted as well as <strong>the</strong>ir standard deviation.<br />

pixels whose centre lay within <strong>the</strong> 200 m x 200 m area <strong>of</strong> <strong>the</strong> respective ESU were taken into acco<br />

Additionally mean and standard deviation per ESU <strong>of</strong> all reflective bands were stored.<br />

6.2 Establishment <strong>of</strong> transfer functions<br />

In situ data, reflectance, SVIs and texture measures served as input <strong>for</strong> regression analysis. The goal is<br />

establishment <strong>of</strong> an empirical relationship that can be used to invert LAI <strong>for</strong> <strong>the</strong> whole test site f<br />

1<br />

,<br />

0,<br />

)(<br />

N<br />

Pi ji<br />

ji<br />

2<br />

Variance ( i )<br />

1<br />

2<br />

, <br />

0,<br />

)(<br />

N<br />

iP ij i<br />

ji<br />

Homogeneity<br />

N 1 P , ji<br />

2<br />

ji 0,<br />

ji )(1<br />

Contrast<br />

1<br />

2<br />

,<br />

0,<br />

)(<br />

jiP<br />

N<br />

ji<br />

ji<br />

<br />

<br />

Entropy<br />

N 1<br />

, ji , ji )l n(<br />

ji 0,<br />

PP<br />

Correlation<br />

<br />

<br />

<br />

<br />

<br />

<br />

1 <br />

<br />

<br />

<br />

0, <br />

<br />

<br />

22<br />

2 DERIVATION OF HIGH RESOLUTION LAI MA<br />

Table 6-1: Second-order texture measures used in this <strong>the</strong>sis (definition according to HALL-BEYER 2007)<br />

Texture measure Equation<br />

Mean ( i )<br />

<br />

N<br />

ii<br />

<br />

ji<br />

P , ji<br />

ji ji<br />

For all ESUs, mean SVIs and texture measures were extracted as well as <strong>the</strong>ir standard deviation.<br />

pixels whose centre lay within <strong>the</strong> 200 m x 200 m area <strong>of</strong> <strong>the</strong> respective ESU were taken into acco<br />

Additionally mean and standard deviation per ESU <strong>of</strong> all reflective bands were stored.<br />

6.2 Establishment <strong>of</strong> transfer functions<br />

In situ data, reflectance, SVIs and texture measures served as input <strong>for</strong> regression analysis. The goal is<br />

1<br />

,<br />

0,<br />

)(<br />

N<br />

Pi ji<br />

ji<br />

2<br />

Variance ( i )<br />

1<br />

2<br />

, <br />

0,<br />

)(<br />

N<br />

iP ij i<br />

ji<br />

Homogeneity<br />

N 1 P , ji<br />

2<br />

ji 0,<br />

ji )(1<br />

Contrast<br />

1<br />

2<br />

,<br />

0,<br />

)(<br />

jiP<br />

N<br />

ji<br />

ji<br />

<br />

<br />

Entropy<br />

N 1<br />

, ji , ji )l n(<br />

ji 0,<br />

PP<br />

Correlation<br />

<br />

<br />

<br />

<br />

<br />

<br />

1 <br />

<br />

<br />

<br />

0, <br />

<br />

<br />

22<br />

N<br />

ii<br />

<br />

ji<br />

P , ji<br />

ji ji<br />

For all ESUs, mean SVIs and texture measures were extracted as well as <strong>the</strong>ir standard deviation. A<br />

pixels whose centre lay within <strong>the</strong> 200 m x 200 m area <strong>of</strong> <strong>the</strong> respective ESU were taken into accoun<br />

Additionally mean and standard deviation per ESU <strong>of</strong> all reflective bands were stored.<br />

6.2 Establishment <strong>of</strong> transfer functions<br />

In situ data, reflectance, SVIs and texture measures served as input <strong>for</strong> regression analysis. The goal is th


6 Derivation <strong>of</strong> high resolution LAI maps<br />

An error is thus data <strong>based</strong> and refers to a single<br />

measurement. In this case, no uncertainty exists as<br />

<strong>the</strong> error is well defined. Uncertainty in turn is related<br />

to prediction <strong>based</strong> on statistical models, e.g. derived<br />

from regression analysis, and is associated to three<br />

o<strong>the</strong>r terms, namely bias, precision and accuracy.<br />

Bias is a measure <strong>of</strong> agreement between a data set bias) and precision, i.e. accuracy = unbias + precision<br />

and values predicted by a statistical model. It is (cf. Figure 6-1 <strong>for</strong> illustration). As displayed in<br />

DERIVATION OF HIGH RESOLUTION LAI MAPS 3<br />

an expectation <strong>of</strong> over- or under-prediction, thus Figure 6-1b, high accuracy can only be achieved if<br />

Bias revealing, is a measure <strong>for</strong> instance, <strong>of</strong> agreement systematic between errors a data in set <strong>the</strong> and both values precision predicted and by unbias a statistical are high. model. It is an<br />

expectation measurement <strong>of</strong> process. over- It or is <strong>of</strong>ten under-prediction, characterized by thus <strong>the</strong><br />

DERIVATION OF OF HIGH HIGH RESOLUTION LAI LAI MAPS MAPS revealing, <strong>for</strong> instance, systematic errors in 3 <strong>the</strong> 3<br />

measurement mean error ME, process. with It is <strong>of</strong>ten characterized by <strong>the</strong> mean If error an independent ME, with data set is used to assess accuracy<br />

Figure 6-1: Illustrations <strong>of</strong> data sets with a) large bias and high preci<br />

Bias Bias is a is measure a measure <strong>of</strong> <strong>of</strong> agreement between a data a data set set and and high values (as values precision in predicted <strong>the</strong> predicted case (high by with by a accuracy), statistical a in statistical situ measurements c) model. low model. bias It is It and an is used an low an <strong>for</strong> precision (low<br />

n 1<br />

ME zˆ( x i ) z(<br />

x i ) .<br />

(6.3)<br />

precision<br />

expectation <strong>of</strong> <strong>of</strong> over- over- or or under-prediction, thus thus revealing, validation<br />

(low<br />

<strong>for</strong> <strong>for</strong> instance, <strong>of</strong><br />

accuracy).<br />

instance, <strong>the</strong> <strong>MODIS</strong> systematic product), errors errors <strong>the</strong>n in in (6.3) <strong>the</strong> accuracy <strong>the</strong><br />

n i 1<br />

measurement process. It is It <strong>of</strong>ten is <strong>of</strong>ten characterized by by <strong>the</strong> by <strong>the</strong> mean mean error If may error an error ME, be independent ME, predicted with with<br />

data directly set is from used <strong>the</strong> to RMSE, assess accuracy which (as in <strong>the</strong> c<br />

Yet ME does not take into account <strong>the</strong> <strong>the</strong> spread <strong>of</strong> <strong>of</strong> errors. Instead this is considered by precision, which<br />

n n<br />

validation is basically <strong>of</strong> <strong>the</strong> <strong>the</strong> same <strong>MODIS</strong> as in product), Equation <strong>the</strong>n 4.1 (RMSE accuracy <strong>of</strong> may be predic<br />

1 1<br />

MEalso<br />

ME errors. ME depends<br />

Instead<br />

zˆ( xz z ˆ<br />

(<br />

on x<br />

this a statistical is considered<br />

model fitted by precision, to a certain data set. Precision is usually predicted using a<br />

i ) i<br />

i)<br />

z)<br />

(<br />

xz<br />

z i(<br />

( ) x<br />

i.<br />

i)<br />

) .<br />

. basically GCPs). Yet <strong>the</strong> variables same as in are Equation different, 4.1 so (RMSE that (6.3) in (6.3)<br />

this <strong>of</strong> GCPs). case Yet variab<br />

n i n<br />

n 1 ii<br />

1<br />

measure which also <strong>of</strong> error depends spread on a around statistical <strong>the</strong> mean model error, fitted e.g. to a <strong>the</strong> standard written it is written as deviation as <strong>of</strong> <strong>the</strong> error e with<br />

Yet Yet certain ME ME does data does not set. not take Precision take into into account is account usually <strong>the</strong> <strong>the</strong> predicted <strong>the</strong> spread spread <strong>of</strong> using <strong>of</strong> errors. errors. Instead this this is considered is considered by by precision, which which<br />

n<br />

n<br />

also also a depends measure depends 2<br />

e<strong>of</strong> on<br />

i error<br />

on eˆ<br />

a<br />

i statistical<br />

a spread statistical around model model <strong>the</strong> fitted mean fitted to error, to a certain a e.g. certain data data set. set. Precision is usually is 2usually<br />

usually predicted using using a a<br />

zizˆi measure i1<br />

<strong>the</strong> e standard<br />

<strong>of</strong> <strong>of</strong> error <strong>of</strong> error deviation spread spread , around <strong>of</strong> around <strong>the</strong> error <strong>the</strong> <strong>the</strong> mean σmean mean with error, error, e.g. e.g. <strong>the</strong> <strong>the</strong> standard RMSE deviation i 1<br />

<strong>of</strong> <strong>of</strong> <strong>the</strong> <strong>of</strong> <strong>the</strong> error error . error (6.4)<br />

e e with e<br />

with<br />

(6.5)<br />

n 1<br />

n<br />

n<br />

n<br />

<br />

i1<br />

e <br />

,<br />

n 1<br />

2 2<br />

where e e i is e<br />

e ie<strong>the</strong><br />

ˆ i <br />

e<br />

ˆ<br />

mean error and êi is <strong>the</strong> predicted error. With<br />

i i<br />

With If <strong>the</strong> respect errors to <strong>the</strong>mselves measurement are not errors known, several o<strong>the</strong>r issues have to be tak<br />

i<br />

i1<br />

i<br />

<br />

1<br />

emeasures<br />

e<br />

<br />

<strong>of</strong> precision , , can be used, such as, <strong>for</strong> (6.4) instance, important have <strong>the</strong> to standard be to taken acknowledge deviation into consideration. that <strong>of</strong> <strong>the</strong> both measurements in First (6.4) situ (6.4)<br />

and <strong>of</strong> all, satellite it reflectance<br />

n n<br />

n 1<br />

1<br />

(TAN et al. 2005) or c v (ARAGÃO et al. 2005, cf. Equation error is important 5.1). sources Note to in that acknowledge field precision LAI estimation is that an expectation both were in situ discussed <strong>of</strong> and in Chapter<br />

where where where ei is e ii<strong>the</strong><br />

is <strong>the</strong> mean error and and êi is ê ii<strong>the</strong><br />

<strong>the</strong> is <strong>the</strong> predicted error.<br />

error. If influenced satellite <strong>the</strong> If <strong>the</strong> errors errors reflectance by <strong>the</strong>mselves erroneous <strong>the</strong>mselves data are sensor are are not not influenced calibration, known, o<strong>the</strong>r by georegistration o<strong>the</strong>r<br />

errors. and atm<br />

<strong>the</strong> spread <strong>of</strong> errors, with a large number corresponding to small errors and vice-versa, e.g. <strong>the</strong> higher e ,<br />

measures If <strong>the</strong> errors <strong>of</strong> <strong>of</strong> precision <strong>the</strong>mselves precision can can are can be not be used, known, used, such such o<strong>the</strong>r as, as, measures <strong>for</strong> <strong>for</strong> instance, Possible <strong>the</strong> <strong>the</strong> standard error deviation sources deviation in <strong>of</strong> field <strong>of</strong> <strong>the</strong> <strong>the</strong> measurements<br />

LAI measurements<br />

estimation were<br />

<strong>the</strong> lower <strong>the</strong> precision!<br />

(TAN (TAN <strong>of</strong> et precision al. et al. 2005) al. 2005) can or be or c used,<br />

v c (ARAGÃO v<br />

(ARAGÃO such et as, al. et <strong>for</strong> al. 2005, al. instance, 2005, cf. cf. Equation <strong>the</strong> Equation discussed 5.1). 5.1). Note Note in that Chapter that precision 3.3.3. is an is Remote is an expectation sensing <strong>of</strong> data <strong>of</strong><br />

is<br />

The standard last term, deviation accuracy <strong>of</strong> <strong>the</strong> (sometimes measurements also referred (Tan et to al. as uncertainty), mainly influenced is defined by as erroneous unbias (i.e. sensor <strong>the</strong> opposite calibration,<br />

<strong>the</strong> <strong>the</strong> spread spread <strong>of</strong> <strong>of</strong> errors, <strong>of</strong> errors, with with a large a large number corresponding to small to small errors errors and and vice-versa, e.g. e.g. <strong>the</strong> <strong>the</strong> higher higher e , e<br />

,<br />

<strong>of</strong> 2005) bias) or and c precision, (Aragão et i.e. al. accuracy 2005, cf. = Equation unbias + 5.1). precision georegistration (cf. Figure 6-1 and <strong>for</strong> atmospheric illustration). correction. As displayed in<br />

v<br />

<strong>the</strong> <strong>the</strong> lower lower <strong>the</strong> <strong>the</strong> precision!<br />

Figure 6-1b, high accuracy can only be achieved if both precision and unbias are high.<br />

The The last last term, term, accuracy (sometimes also also referred to to as to as uncertainty), as uncertainty), is defined is defined as as unbias as unbias (i.e. (i.e. <strong>the</strong> <strong>the</strong> opposite<br />

<strong>of</strong> <strong>of</strong> bias) <strong>of</strong> bias) and and precision, i.e. i.e. accuracy = unbias = unbias + precision + precision (cf. (cf. Figure Figure 6-1 6-1 <strong>for</strong> <strong>for</strong> illustration). As As displayed in in<br />

Figure Figure 6-1b, 6-1b, high high accuracy can can only only be be achieved if both if both precision and and unbias unbias are are high. high.<br />

Figure 6-1: Illustrations <strong>of</strong> data sets with a) large bias and high precision (low accuracy), b) low bias and<br />

high precision (high accuracy), c) low bias and low precision (low accuracy) and d) large bias and low<br />

precision (low accuracy).<br />

Figure 6-1<br />

Figure Figure If 6-1: an 6-1: Illustrations independent Illustrations (high accuracy), <strong>of</strong> data data <strong>of</strong> data set c) sets low is sets used with bias with and a) to large a) low assess large precision bias accuracy bias and (low and high accuracy) (as high in precision <strong>the</strong> precision and case d) large (low with (low bias accuracy), in accuracy), and situ low measurements precision b) b) low b) low bias (low bias accuracy).<br />

and used and<br />

<strong>for</strong><br />

high high precision validation precision (high <strong>of</strong> (high <strong>the</strong> accuracy), <strong>MODIS</strong> accuracy), c) product), c) low c) low bias bias <strong>the</strong>n and and accuracy low low precision may be (low predicted (low accuracy) directly and and d) from d) large large <strong>the</strong> bias RMSE, bias and and low which low<br />

is<br />

precision (low (low accuracy).<br />

basically <strong>the</strong> same as in Equation 4.1 (RMSE <strong>of</strong> GCPs). Yet variables are different, so that in this case it is<br />

If an If written If an independent as data data set set is used is used to to assess to assess accuracy (as (as in in <strong>the</strong> in <strong>the</strong> case case with with in situ in situ measurements used used <strong>for</strong> <strong>for</strong><br />

validation <strong>of</strong> <strong>of</strong> <strong>the</strong> <strong>the</strong> <strong>MODIS</strong> product), <strong>the</strong>n <strong>the</strong>n accuracy may may be be predicted directly from from <strong>the</strong> <strong>the</strong> RMSE, which which is is<br />

n<br />

2<br />

basically <strong>the</strong> <strong>the</strong> same same same zas i in<br />

as zˆ<br />

Equation in i Equation 4.1 4.1 (RMSE <strong>of</strong> <strong>of</strong> GCPs). <strong>of</strong> GCPs). Yet Yet variables are are different, so so that so that in this in this case case it is it is<br />

i 1<br />

written RMSE as as<br />

<br />

. (6.5)<br />

103<br />

where ei is <strong>the</strong> mean error and êi is <strong>the</strong> predicted error. If <strong>the</strong> e<br />

measures <strong>of</strong> precision can be used, such as, <strong>for</strong> instance, <strong>the</strong> st<br />

(TAN et al. 2005) or c v (ARAGÃO et al. 2005, cf. Equation 5.1). N<br />

Note that precision is an expectation <strong>of</strong> <strong>the</strong> spread <strong>of</strong><br />

<strong>the</strong> errors, spread with <strong>of</strong> a errors, large with number a large corresponding number corresponding to small to small e<br />

<strong>the</strong> errors lower and <strong>the</strong> vice-versa, precision! e.g. <strong>the</strong> higher σ , <strong>the</strong> lower <strong>the</strong><br />

e<br />

The<br />

precision!<br />

last term, accuracy (sometimes also referred to as uncertainty<br />

<strong>of</strong> bias) and precision, i.e. accuracy = unbias + precision (cf. Figu<br />

Figure<br />

The last<br />

6-1b,<br />

term,<br />

high<br />

accuracy<br />

accuracy<br />

(sometimes<br />

can only<br />

also<br />

be achieved<br />

referred<br />

if<br />

to<br />

both<br />

as<br />

precision a<br />

uncertainty), is defined as unbias (i.e. <strong>the</strong> opposite <strong>of</strong><br />

Illustrations <strong>of</strong> data sets with a) large bias and high precision (low accuracy), b) low bias and high precision


104<br />

When finally both in situ and satellite data are<br />

combined, a third error source has to be taken into<br />

account: misregistration <strong>of</strong> both measurements.<br />

Spatial misregistration includes spatial location<br />

errors, but also errors introduced through <strong>the</strong> sensor’s<br />

point spread function, velocity smear and atmospheric<br />

scatter. Yet as pixels in a remotely sensed scene<br />

are spatially autocorrelated, this becomes a greater<br />

problem when landscapes are highly diverse in<br />

relation to <strong>the</strong> sensor’s spatial resolution and can be<br />

disregarded <strong>for</strong> <strong>the</strong> study sites in this <strong>the</strong>sis. Temporal<br />

misregistration occurs when satellite reflectance<br />

data is taken as representative <strong>for</strong> in situ conditions<br />

that do not match those <strong>of</strong> individual measurements<br />

(Curran & Hay 1986). Consequently, as in situ data<br />

were acquired over a period <strong>of</strong> four weeks at both<br />

Descriptive statistics <strong>of</strong> LAI<br />

Table 6-2 e (LAI2000) per ESU derived from B1 measurements after correction <strong>of</strong> e θsun and applied<br />

cloud filter.<br />

ESU N Minimum Maximum μ σ c v<br />

1 140 6.03 6.48 6.32 0.119 0.019<br />

2 164 8.09 8.55 8.34 0.090 0.011<br />

3 92 7.38 7.78 7.58 0.098 0.013<br />

4 80 7.09 7.86 7.48 0.225 0.030<br />

5 185 7.88 8.26 8.10 0.099 0.012<br />

6 99 7.66 8.21 7.89 0.146 0.019<br />

7 143 4.31 4.77 4.60 0.088 0.019<br />

8 119 5.58 6.16 5.87 0.151 0.026<br />

9 122 5.62 6.11 5.88 0.122 0.021<br />

10 92 6.16 6.81 6.49 0.192 0.030<br />

11 233 6.30 6.56 6.42 0.065 0.010<br />

12 114 5.42 6.20 5.92 0.172 0.029<br />

13 138 7.75 8.49 8.09 0.193 0.024<br />

15 58 7.89 8.58 8.29 0.230 0.028<br />

16 117 8.23 8.60 8.38 0.069 0.008<br />

17 181 2.25 3.09 2.66 0.246 0.092<br />

18 117 6.47 6.81 6.62 0.077 0.012<br />

19 174 3.45 3.99 3.77 0.117 0.031<br />

20 94 6.67 7.06 6.84 0.092 0.013<br />

21 141 1.51 2.25 1.91 0.190 0.099<br />

22 174 1.44 1.99 1.60 0.136 0.085<br />

23 125 4.51 5.19 4.87 0.197 0.040<br />

24 114 3.27 3.63 3.44 0.077 0.022<br />

25 184 2.79 3.67 3.17 0.259 0.082<br />

26 151 2.63 3.18 2.92 0.148 0.051<br />

27 103 7.94 9.17 8.57 0.353 0.041<br />

28 151 5.85 6.27 6.10 0.100 0.016<br />

29 89 7.08 8.11 7.52 0.295 0.039<br />

30 91 3.60 4.34 3.94 0.216 0.055


6 Derivation <strong>of</strong> high resolution LAI maps<br />

study sites and satellite data was acquired at <strong>the</strong> end<br />

<strong>of</strong> this period (respectively exactly one year later in<br />

<strong>the</strong> case <strong>of</strong> ASTER data), changing LAI conditions<br />

can have an effect on <strong>the</strong> resulting transfer functions.<br />

There was an attempt to minimize this error by taking<br />

measurements at <strong>the</strong> end <strong>of</strong> <strong>the</strong> rainy season when<br />

phenological changes are less pronounced and LAI<br />

is assumed to be relatively stable. In <strong>the</strong> following,<br />

<strong>the</strong> precision <strong>of</strong> in situ measurements and satellite<br />

observations will be assessed as this is a prerequisite<br />

<strong>for</strong> <strong>the</strong> establishment <strong>of</strong> adequate transfer functions.<br />

Precision <strong>of</strong> LAI e (LAI2000)<br />

As previously described, <strong>the</strong> precision <strong>of</strong> in situ<br />

data can be estimated by different measures. For<br />

LAI-2000 PCA data, Tan et al. (2005) used <strong>the</strong><br />

standard deviation <strong>of</strong> measured LAI per ESU. In <strong>the</strong>ir<br />

study field sampling <strong>of</strong> an agricultural area with very<br />

homogeneous vegetation cover and low LAI values<br />

was undertaken, so that differences in individual<br />

measurements were related to noise from instrument<br />

set-up ra<strong>the</strong>r than to real heterogeneities in foliage<br />

amount.<br />

In this <strong>the</strong>sis however, LAI values are generally<br />

much higher and vegetation is less homogeneous,<br />

so that <strong>the</strong> standard deviation <strong>of</strong> LAI on each ESU<br />

is not only influenced by measurement precision,<br />

but also by local changes in <strong>for</strong>est structure and thus<br />

<strong>the</strong> natural variability <strong>of</strong> <strong>for</strong>est cover and foliage.<br />

Accordingly <strong>the</strong> standard deviation <strong>of</strong> LAIe (LAI2000)<br />

calculated from transect measurements (B2) contains<br />

both measurement errors and natural LAI variability.<br />

This is not <strong>the</strong> case <strong>for</strong> LAI derived from<br />

e (LAI2000)<br />

continuous B1 data. In this case, <strong>the</strong> measurement<br />

locality is stable and after correction <strong>of</strong> e and<br />

θsun<br />

filtering only <strong>the</strong> remaining e and instrument<br />

cloud<br />

inherent noise affect <strong>the</strong> results. Accordingly B1<br />

data is used to determine ESU specific LAIe (LAI2000)<br />

precision. Yet in lieu <strong>of</strong> <strong>the</strong> standard deviation, <strong>the</strong><br />

Figure 6-2 Relation between c and LAI <strong>for</strong> all<br />

v e (2000)<br />

ESUs in Budongo Forest.<br />

105<br />

coefficient <strong>of</strong> variation was used to determine <strong>the</strong><br />

measurement precision as <strong>the</strong> latter is independent <strong>of</strong><br />

mean LAI per ESU (cf. Equation 5.1). Table<br />

e (LAI2000)<br />

6-2 shows <strong>the</strong> descriptive statistics <strong>of</strong> LAIe (LAI2000)<br />

derived from corrected B1 measurements with<br />

measurement precision being defined as c <strong>of</strong> <strong>the</strong><br />

v<br />

respective ESU.<br />

As expected, measurement precision is lower (i.e. c v is<br />

higher nota bene) on ESUs measured under changing<br />

illumination conditions (e.g. ESU 4, 8, 10, 12).<br />

Noticeably it is especially low on ESUs belonging to<br />

<strong>the</strong> early <strong>for</strong>est stage, i.e. ESUs 17, 21, 22, 23, 25,<br />

and 26. Logically changing illumination conditions<br />

here have a higher impact on measurement precision<br />

due to <strong>the</strong> open – or non-existent – tree cover. Figure<br />

6-2 illustrates that c decreases with mean LAI (and<br />

v<br />

thus precision increases) with a mean c <strong>of</strong> 0.04 (4%).<br />

v<br />

It should be noted however that this is only true <strong>for</strong><br />

LAI-2000 PCA measurements under at least partly<br />

changing illumination conditions. Precision is always<br />

highest on ESUs measured under diffuse and direct<br />

illumination conditions (with correction <strong>of</strong> eθsun applied to <strong>the</strong> latter).


106<br />

N Min Max N Mi<br />

Test 2 (30/10/2005) 12<br />

Test 1 (01/11/2005)<br />

2.90<br />

31<br />

3.39<br />

3.50 Test<br />

3.21<br />

4.35 1 (01/11/2005) 3.93<br />

0.140<br />

31 3.50<br />

Test 2 (30/10/2005) 12 2.90 Test 3.39 2 (30/10/2005) 3.21 12 2.90<br />

Precision <strong>of</strong> surface reflectance<br />

Precision <strong>of</strong> LAIe (DHP)<br />

In contrast <strong>the</strong> to proximity <strong>the</strong> precision <strong>of</strong> observed <strong>of</strong> field to LAI, true surface <strong>the</strong> precision reflectance. <strong>of</strong> surface reflectance<br />

Ideally field According spectrometer to this measurements m , m , …, m 1 2 N would are true serve values as ground <strong>of</strong> truth <strong>for</strong> s<br />

In contrast to LAI-2000 PCA measurements, difficult only to atmospherically acquire over a corrected tropical rain<strong>for</strong>est surface reflectance canopy, <strong>the</strong> derived approach <strong>of</strong> WANG<br />

transect data was available <strong>for</strong> DHP (equivalent this to <strong>the</strong>sis. from In this high case, spatial precision resolution values satellite derived data from at N ei<strong>the</strong>r spectral satellite data or lite<br />

B2 measurements <strong>of</strong> LAI-2000 PCA). As mean <strong>the</strong> and proximity bands. <strong>of</strong> Various observed error to true sources surface can reflectance. cause deviations According to this m1, m<br />

standard deviation <strong>of</strong> <strong>the</strong> latter are also influenced atmospherically from <strong>the</strong>se corrected values, surface leading reflectance to observations derived d , d 1 from , …, 2 high spatial res<br />

by vegetation heterogeneity, <strong>the</strong>y could not be spectral used bands. d . The N Various latter error are treated sources as can independent cause deviations random from <strong>the</strong>se true valu<br />

2 as a measure <strong>of</strong> precision. Instead <strong>the</strong> two tests d1, on d2, …, variables dN. The latter with finite are treated variances as independent σ , k=1, 2,…, k random N and it variables with fini<br />

<strong>the</strong> influence <strong>of</strong> illumination conditions described<br />

k k <br />

in Chapter 5.2.2 were analysed. As here <strong>the</strong> locality N and it is is supposed that <strong>the</strong> deviations k= 2<br />

k<br />

was not changed during photograph acquisition, it<br />

can be assumed that <strong>the</strong> resulting standard deviation<br />

is influenced by errors due to instrument set-up,<br />

exposure, illumination conditions and processing<br />

only. Again <strong>the</strong> cv <strong>of</strong> both time series was calculated.<br />

The results showed that both test sites had very similar<br />

cv <strong>of</strong> around 0.05 (cf. Table 6-3), which was taken<br />

as measurement precision <strong>for</strong> LAI on all ESUs.<br />

e (DHP)<br />

Although test data <strong>for</strong> early <strong>for</strong>est stages was not<br />

available, it is assumed that measurement precision<br />

<strong>for</strong> this is even higher as <strong>the</strong> smaller distance between<br />

foliage and lens minimizes <strong>the</strong> problem <strong>of</strong> mixed<br />

pixels and thus increases classification accuracy.<br />

Precision <strong>of</strong> surface reflectance<br />

In contrast to <strong>the</strong> precision <strong>of</strong> field LAI, <strong>the</strong> precision<br />

<strong>of</strong> surface reflectance is much harder to assess.<br />

Ideally field spectrometer measurements would Table 6-4: Relative precision <strong>for</strong> red, NIR and SWIR spectral bands <strong>for</strong> <strong>the</strong> high<br />

serve as ground truth <strong>for</strong> satellite data. As <strong>the</strong>se are<br />

difficult to acquire over a tropical rain<strong>for</strong>est canopy,<br />

<strong>the</strong> approach <strong>of</strong> Wang et al. (2001) is followed in this<br />

<strong>the</strong>sis. In this case, precision values derived from<br />

ei<strong>the</strong>r satellite data or literature are used to describe<br />

m d<br />

Precision <strong>of</strong> surface reflectance<br />

In contrast to <strong>the</strong> precision <strong>of</strong> field LAI, <strong>the</strong> precision <strong>of</strong> surface re<br />

Ideally field spectrometer measurements would serve as ground tr<br />

difficult to acquire over a tropical rain<strong>for</strong>est canopy, <strong>the</strong> approach o<br />

this <strong>the</strong>sis. In this case, precision values derived from ei<strong>the</strong>r satellite d<br />

<strong>the</strong> proximity <strong>of</strong> observed to true surface reflectance. According to t<br />

atmospherically corrected surface reflectance derived from high sp<br />

spectral bands. Various error sources can cause deviations from <strong>the</strong>se<br />

d1, d2, …, dN. The latter are treated as independent random variables<br />

follow Gaussian distribu<br />

k k <br />

N and it is supposed that <strong>the</strong> deviations k= 2<br />

Gaussian N distribution. N<br />

2<br />

2<br />

<br />

dkmThe k random variable k<br />

d m <br />

k <br />

2<br />

k1<br />

k1 k<br />

describes <strong>the</strong> proximity <strong>of</strong> observed to true values and has a chi-s<br />

2<br />

N indicates good observation quality. Dispersions 1, 2, …, <br />

atmospherically corrected surface reflectance and parameterized in terms <strong>of</strong> <strong>the</strong><br />

(TAN et al. 2005). k is thus sensor and band specific.<br />

Un<strong>for</strong>tunately k cannot be assessed easily as it is influenced by <strong>the</strong> sensor-sp<br />

<strong>the</strong> applied atmospheric correction algorithm. According to HUANG et al. (200<br />

<strong>of</strong> satellite reflectance data can be derived from invariant targets, i.e. surfaces<br />

temporal processes such as, <strong>for</strong> example, vegetation phenology. Successive an<br />

<strong>of</strong> such surfaces can reveal mean, variance and thus precision, if <strong>the</strong> tar<br />

correction are assumed to be stable. As multi-temporal scenes were not avai<br />

specific relative precision had to be derived from literature in <strong>the</strong> case <strong>of</strong> SPO<br />

delivered as atmospherically corrected surface reflectance product, relative pre<br />

<strong>the</strong> data product description (cf. Table 6-4 <strong>for</strong> <strong>the</strong> bands used in SVI calculation<br />

Table 6-4: Relative precision <strong>for</strong> red, NIR and SWIR spectral bands fo<br />

m d<br />

Precision <strong>of</strong> surface reflectance<br />

In contrast to <strong>the</strong> precision <strong>of</strong> fie<br />

Ideally field spectrometer measur<br />

difficult to acquire over a tropical<br />

this <strong>the</strong>sis. In this case, precision v<br />

<strong>the</strong> proximity <strong>of</strong> observed to true<br />

atmospherically corrected surface<br />

spectral bands. Various error sourc<br />

d1, d2, …, dN. The latter are treated<br />

N and it is supposed<br />

<br />

follow that <strong>the</strong> Gaussia devia<br />

<br />

N N<br />

N N<br />

2<br />

2<br />

<br />

dkmk 2<br />

2<br />

d m <br />

k <br />

<br />

dkmk d m (6.6)<br />

<br />

k <br />

2<br />

2<br />

k1<br />

k1 <br />

k1<br />

k1 k<br />

k<br />

describes <strong>the</strong> proximity <strong>of</strong> observed <strong>of</strong> observed describes to true values to <strong>the</strong> true and proximity values and <strong>of</strong> has ob<br />

2<br />

has 2<br />

a<br />

chi-square N indicates distribution, good observation where quality.<br />

N indicates Dispersions good observa 1, <br />

good observation quality. Dispersions<br />

atmospherically corrected surface reflectance atmospherically σ , σ , …, σ are<br />

1 2 N and parameterized corrected surface in terr<br />

precisions in <strong>the</strong> atmospherically corrected surface<br />

(TAN et al. 2005). k is thus sensor (TAN and band et al. specific. 2005). k is thus sensor<br />

reflectance and parameterized in terms <strong>of</strong> <strong>the</strong>ir<br />

Un<strong>for</strong>tunately relative values k α =σ cannot /d (Tan be assessed et al. Un<strong>for</strong>tunately 2005). easily αas is it thus is k influenced cannot be by assess <strong>the</strong><br />

k k k k<br />

<strong>the</strong> sensor applied and band atmospheric specific. correction <strong>the</strong> algorithm. applied According atmospheric to correction HUANG<br />

<strong>of</strong> satellite reflectance data can be <strong>of</strong> derived satellite from reflectance invariant data targets, can i.e. be<br />

temporal Un<strong>for</strong>tunately processes α cannot such as, be <strong>for</strong> assessed example, temporal easily vegetation processes as it is phenology. such as, <strong>for</strong> Succ ex<br />

k<br />

<strong>of</strong> influenced such surfaces by <strong>the</strong> can sensor-specific reveal mean, calibration <strong>of</strong> variance such surfaces as and well thus can precision, reveal mea if<br />

correction as <strong>the</strong> applied are assumed atmospheric to be stable. correction correction As multi-temporal algorithm. are assumed scenes to be were stab<br />

specific According relative to Huang precision et had al. to (2006) be specific derived <strong>the</strong> relative from relative literature precision in had <strong>the</strong> to cab<br />

delivered precision as αatmospherically <strong>of</strong> satellite reflectance corrected delivered surface data as can reflectance atmospherically be product, correc<br />

k re<br />

<strong>the</strong> derived data product from invariant description targets, (cf. i.e. Table <strong>the</strong> surfaces 6-4 data <strong>for</strong> product that <strong>the</strong> are bands description used in (cf. SVI Tc<br />

not influenced by temporal processes such as, <strong>for</strong><br />

example, vegetation phenology. Table Successive 6-4: Relative and precision <strong>for</strong> re<br />

repetitive measurements <strong>of</strong> such surfaces can reveal<br />

mean, variance and thus precision, if <strong>the</strong> target<br />

itself and atmospheric correction are assumed to be<br />

stable. As multi-temporal scenes were not available<br />

Table 6-3 Descriptive statistics <strong>for</strong> LAI derived from DHP time series taken at two test sites, each over a whole day.<br />

e (DHP)<br />

N Minimum Maximum μ σ c v<br />

Test 1 (01/11/2005) 31 3.50 4.35 3.93 0.182 0.046<br />

Test 2 (30/10/2005) 12 2.90 3.39 3.21 0.140 0.044


6 Derivation <strong>of</strong> high resolution LAI maps<br />

<strong>for</strong> this <strong>the</strong>sis, sensor specific relative precision had is define <strong>the</strong> slope LAI as <strong>of</strong> independent <strong>the</strong> relationship from between satellite X reflectance and Y, and e is <strong>the</strong> err<br />

to be derived from literature in <strong>the</strong> case <strong>of</strong> SPOT-4. between (and not an <strong>the</strong> easy o<strong>the</strong>r to way measure around), variable much (such <strong>of</strong> <strong>the</strong> as remote SVIs) and a costly<br />

As ASTER data was delivered as atmospherically easy sensing to literature measure reports variable <strong>the</strong> is LAI <strong>the</strong>n being used modelled to predict as a <strong>the</strong> costly to<br />

corrected surface reflectance product, relative independent dependent variable variable, (Cohen and it et would al. 2003). be logical Alternatively to define LAI as ind<br />

precisions were available from <strong>the</strong> data product not <strong>the</strong> <strong>the</strong> coefficients o<strong>the</strong>r way <strong>for</strong> around), <strong>the</strong> regression much <strong>of</strong> model <strong>the</strong> remote could sensing be literatur<br />

description (cf. Table 6-4 <strong>for</strong> <strong>the</strong> bands used in SVI dependent derived with variable Y = LAI (COHEN and X et al. can 2003). subsequently Alternatively be <strong>the</strong> coeffic<br />

calculation).<br />

DERIVATION OF HIGH RESOLUTION LAI MAPS derived with by inversion Y = LAI <strong>of</strong> and Equation X can subsequently 6.7 so that be 7 derived by invers<br />

Sensor red<br />

6.2.2 Regression analysis<br />

NIR<br />

Y 0<br />

X e ,<br />

SWIR 1<br />

Source<br />

(6.8)<br />

1% (0.15)<br />

Linear SPOT-4 regression HRVIR is a common 5.8% approach in 6.4% remote each 5.6% observation.<br />

One <strong>of</strong> <strong>the</strong> most HENRY widely & used MEYGRET regression (2001) approaches in remote<br />

sensing to estimate biophysical variables from remote<br />

squares (OLS) regression. Yet very <strong>of</strong>ten <strong>the</strong> underlying conditions<br />

sensing data, ei<strong>the</strong>r in univariate (e.g. Tan et al. One <strong>of</strong> <strong>the</strong> most widely used regression approaches in<br />

one assumption is that X is measured free <strong>of</strong> error (COHEN et al<br />

2005) or multivariate <strong>for</strong>m (e.g. Cohen et al. 2003). remote sensing research is <strong>the</strong> ordinary least squares<br />

6.2.2 Regression analysis<br />

realized as discussed earlier. The inclusion <strong>of</strong> error-related outlier<br />

Bivariate plots are usually used to determine whe<strong>the</strong>r (OLS) regression. Yet very <strong>of</strong>ten <strong>the</strong> underlying<br />

Linear regression is a common approach in remote sensing and to hence estimate <strong>the</strong> biophysical predictions variables <strong>based</strong> on from this remote model. Second, homos<br />

<strong>the</strong> relationship between <strong>the</strong> variables is truly linear conditions <strong>of</strong> this approach are not fulfilled. First,<br />

sensing data, ei<strong>the</strong>r in univariate (e.g. TAN et al. 2005) variances or multivariate (FERNANDES <strong>for</strong>m (e.g. et al. COHEN 2005). If et heteroscedasticity al. 2003). is <strong>the</strong> ca<br />

or if trans<strong>for</strong>mations have to be applied a priori. one assumption is that X is measured free <strong>of</strong> error<br />

Bivariate plots are usually used to determine whe<strong>the</strong>r <strong>the</strong> reduced relationship leading between to over <strong>the</strong> or variables under prediction is truly linear (COHEN et al. 2003).<br />

In <strong>the</strong> simple linear case <strong>the</strong> <strong>for</strong>m <strong>of</strong> <strong>the</strong> regression (Cohen et al. 2003). In practice this can scarcely be<br />

or if trans<strong>for</strong>mations have to be applied a priori. In <strong>the</strong> simple degree linear <strong>of</strong> variance case <strong>the</strong> <strong>for</strong>m attenuation <strong>of</strong> <strong>the</strong> regression is a linear model function <strong>of</strong> <strong>the</strong> co<br />

model is<br />

realized as discussed earlier. The inclusion <strong>of</strong> error-<br />

is<br />

investigation. The lower <strong>the</strong> correlation, <strong>the</strong> higher is <strong>the</strong> compress<br />

related outliers biases <strong>the</strong> retrieved model parameters<br />

Y 0 1X<br />

e<br />

(6.7) As and an hence alternative <strong>the</strong> to predictions OLS, FERNANDES <strong>based</strong> on et this al. (6.7) (2005) model. introduced <strong>the</strong><br />

<strong>the</strong> Second, remote homoscedasticity sensing community. is assumed, It is <strong>based</strong> i.e. equal on <strong>the</strong> error median <strong>of</strong> ranke<br />

where Y is <strong>the</strong> variable to be predicted, X is <strong>the</strong> variable from which Y is predicted, 0 is <strong>the</strong> intercept, 1<br />

where Y is <strong>the</strong> variable to be predicted, X is <strong>the</strong> robust variances in <strong>the</strong> (Fernandes presence et <strong>of</strong> al. measurement 2005). If heteroscedasticity<br />

errors. A comparison to se<br />

is <strong>the</strong> slope <strong>of</strong> <strong>the</strong> relationship between X and Y, and e is <strong>the</strong> error. Usually a relationship is established<br />

variable from which Y is predicted, β is <strong>the</strong> intercept, OLS is <strong>the</strong> and case, Reduced variance Major in <strong>the</strong> Axis predicted (RMA) clearly variable showed is that OL<br />

0<br />

between an easy to measure variable (such as SVIs) and a costly to measure variable (such as LAI). The<br />

β is <strong>the</strong> slope <strong>of</strong> <strong>the</strong> relationship between X and Y, specification reduced leading <strong>of</strong> dependent to over or and under independent prediction variables, (Cohen leading to an<br />

1<br />

easy to measure variable is <strong>the</strong>n used to predict <strong>the</strong> costly to measure variable. Although Y is <strong>the</strong><br />

and e is <strong>the</strong> error. Usually a relationship is established<br />

15%<br />

et al.<br />

(FERNANDES<br />

2003). According<br />

et al. 2005).<br />

to Schlerf<br />

Both<br />

et<br />

<strong>the</strong><br />

al.<br />

RMA<br />

(2005)<br />

and<br />

<strong>the</strong><br />

Theil-Sen appr<br />

independent variable, and it would be logical to define LAI as independent from satellite reflectance (and<br />

between an easy to measure variable (such as SVIs)<br />

levels,<br />

degree<br />

but<br />

<strong>of</strong><br />

Theil-Sen<br />

variance attenuation<br />

regression was<br />

is a<br />

superior<br />

linear function<br />

at high error levels.<br />

not <strong>the</strong> o<strong>the</strong>r way around), much <strong>of</strong> <strong>the</strong> remote sensing literature reports <strong>the</strong> LAI being modelled as a<br />

and a costly to measure variable (such as LAI). The <strong>of</strong> <strong>the</strong> correlation between <strong>the</strong> variables under<br />

dependent variable (COHEN et al. 2003). Alternatively <strong>the</strong> coefficients <strong>for</strong> <strong>the</strong> regression model could be<br />

easy to measure variable is <strong>the</strong>n used to predict investigation. The lower <strong>the</strong> correlation, <strong>the</strong> higher is<br />

derived with Y = LAI and X can subsequently be derived by inversion <strong>of</strong> Equation 6.7 so that<br />

<strong>the</strong> costly to measure variable. Although Y is <strong>the</strong> <strong>the</strong> compression <strong>of</strong> variances.<br />

independent Y variable, and it would be logical to<br />

0<br />

X e ,<br />

<br />

(6.8)<br />

1<br />

with <strong>the</strong> error referring to <strong>the</strong> prediction residual <strong>for</strong> each observation.<br />

or if trans<strong>for</strong>mations have to be applied a priori. In <strong>the</strong> simple linea<br />

is<br />

Y X e<br />

0 1<br />

Table 6-4 Relative precision α <strong>for</strong> red, NIR and SWIR spectral bands <strong>for</strong> <strong>the</strong> high resolution sensors used.<br />

One <strong>of</strong> <strong>the</strong> most widely used regression approaches in remote sensing research is <strong>the</strong> ordinary least<br />

Sensor α α α Source<br />

red NIR SWIR squares (OLS) regression. Yet very <strong>of</strong>ten <strong>the</strong> underlying conditions <strong>of</strong> this approach are not fulfilled. First,<br />

one assumption is that X is measured free <strong>of</strong> error (COHEN et al. 2003). In practice this can scarcely be<br />

ASTER 1% (0.15)<br />

and hence <strong>the</strong> predictions <strong>based</strong> on this model. Second, homoscedasticity is assumed, i.e. equal error<br />

variances (FERNANDES et al. 2005). If heteroscedasticity is <strong>the</strong> case, variance in <strong>the</strong> predicted variable is<br />

reduced leading to over or under prediction (COHEN et al. 2003). According to SCHLERF et al. (2005) <strong>the</strong><br />

degree <strong>of</strong> variance attenuation is a linear function <strong>of</strong> <strong>the</strong> correlation between <strong>the</strong> variables under<br />

investigation. The lower <strong>the</strong> correlation, <strong>the</strong> higher is <strong>the</strong> compression <strong>of</strong> variances.<br />

SPOT-4 HRVIR 5.8% 6.4% 5.6% Henry & Meygret (2001)<br />

As an alternative to OLS, FERNANDES et al. (2005) introduced <strong>the</strong> non-parametric Theil-Sen regression to<br />

<strong>the</strong> remote sensing community. It is <strong>based</strong> on <strong>the</strong> median <strong>of</strong> ranked slopes and it thus appears to be more<br />

robust in <strong>the</strong> presence <strong>of</strong> measurement errors. A comparison to several o<strong>the</strong>r regression models including<br />

107<br />

where Y is <strong>the</strong> variable to be predicted, X is <strong>the</strong> variable from whic


108<br />

As an alternative to OLS, Fernandes et al. (2005) <strong>of</strong> LAI values is <strong>for</strong>med including measured LAI<br />

introduced <strong>the</strong> non-parametric Theil-Sen regression values <strong>of</strong> all ESUs whose LAI values fall within this<br />

to <strong>the</strong> remote sensing community. It is <strong>based</strong> on interval. The corresponding surface reflectances are<br />

<strong>the</strong> median <strong>of</strong> ranked slopes and it thus appears grouped accordingly. Mean and standard deviation<br />

to be more robust in <strong>the</strong> presence <strong>of</strong> measurement <strong>for</strong> both groups are recorded. The process is repeated<br />

errors. A comparison to several o<strong>the</strong>r regression <strong>for</strong> case B, but here measurement precision <strong>of</strong> surface<br />

models including OLS and Reduced Major Axis reflectance is taken into account. Mean values <strong>for</strong><br />

(RMA) clearly showed that OLS gives different LAI and surface reflectance <strong>of</strong> cases A and B are<br />

results depending on specification <strong>of</strong> dependent <strong>the</strong>n finally combined in case C. At this point a stable<br />

and independent variables, leading to an average relationship is established, <strong>based</strong> on <strong>the</strong> newly <strong>for</strong>med<br />

difference in predicted LAI <strong>of</strong> 15% (Fernandes et data set that takes measurement and observation<br />

al. 2005). Both <strong>the</strong> RMA and Theil-Sen approaches<br />

per<strong>for</strong>med similarly at low error levels, but Theil-<br />

Sen regression was superior at high error levels.<br />

Fernandes et al. (2005) conclude that <strong>the</strong> rank-<strong>based</strong><br />

errors into account.<br />

8 Theil-Sen estimator should be used instead <strong>of</strong> OLS DERIVATION OF HIGH RESOLUTION LAI MAPS<br />

in situ LAI surface reflectance<br />

or RMA as univariate linear regression when <strong>the</strong><br />

8 DERIVATION OF HIGH RESOLUTION LAI MAPS<br />

<strong>the</strong> sample rank-<strong>based</strong> data is Theil-Sen not error-free. estimator Accordingly should be this used type instead <strong>of</strong> OLS or RMA as univariate linear regression<br />

LAI LAIi<br />

i<br />

r i<br />

<strong>the</strong><br />

when <strong>of</strong> regression rank-<strong>based</strong><br />

<strong>the</strong> sample was Theil-Sen<br />

data used is in not<br />

estimator this error-free. <strong>the</strong>sis. should The Accordingly Theil-Sen be used instead<br />

this type<br />

<strong>of</strong> OLS<br />

<strong>of</strong> regression<br />

or RMA as<br />

was<br />

univariate<br />

used in<br />

linear<br />

this <strong>the</strong>sis.<br />

regression<br />

The<br />

…<br />

…<br />

Theil-Sen<br />

when slope estimator <strong>the</strong><br />

slope<br />

sample is estimator defined data is not<br />

is asdefined<br />

error-free.<br />

as<br />

Accordingly this type <strong>of</strong> regression LAI was used in this <strong>the</strong>sis. The<br />

j<br />

r j<br />

Theil-Sen slope<br />

<br />

estimator<br />

y <br />

is<br />

y<br />

defined as<br />

<br />

j i<br />

1 medians ij s ij , x j x i 1 i j n<br />

,<br />

<br />

x j x i<br />

<br />

y y<br />

<br />

j i<br />

1 medians ij s ij , x j x i 1 i j n<br />

,<br />

with <strong>the</strong> intercept <br />

x x<br />

being j calculated i as (6.9) <br />

case A<br />

(group k)<br />

impact <strong>of</strong> errors in in situ LAI<br />

(6.9)<br />

case B<br />

(group k)<br />

(6.9)<br />

impact <strong>of</strong> errors in surf. refl.<br />

with <strong>the</strong> intercept being calculated as<br />

with <strong>the</strong> median intercept y ybeing<br />

calculated<br />

median x as ... x .<br />

LAIi Li<br />

r i<br />

r i<br />

LAIi<br />

(6.10)<br />

0<br />

<br />

1...<br />

n 1<br />

1<br />

<br />

n<br />

j<br />

j<br />

j<br />

However, 0 median in<strong>for</strong>mation y 1...<br />

y n on<br />

1median<br />

measurement x 1...<br />

x n . errors (6.10) is not explicitly included in Theil-Sen regression. (6.10) To<br />

mean/std. mean/std.<br />

mean/std. mean/std.<br />

evaluate<br />

However,<br />

its<br />

in<strong>for</strong>mation<br />

per<strong>for</strong>mance,<br />

on<br />

results<br />

measurement<br />

derived<br />

errors<br />

<strong>for</strong> <strong>the</strong><br />

is<br />

two<br />

not explicitly<br />

test sites<br />

included<br />

are compared<br />

in Theil-Sen<br />

to ano<strong>the</strong>r<br />

regression.<br />

interesting<br />

To<br />

approach However, proposed in<strong>for</strong>mation by TAN on measurement et al. (2005). errors In this is<br />

choose all LAIj<br />

that fall within<br />

choose all r j that fall within<br />

evaluate its per<strong>for</strong>mance, results derived <strong>for</strong> <strong>the</strong> two<br />

instance,<br />

test<br />

in<strong>for</strong>mation on measurement precision is<br />

<strong>the</strong> sites measurement are error compared <strong>of</strong> LAIi<br />

to ano<strong>the</strong>r <strong>the</strong> measurement interesting error <strong>of</strong> r i<br />

and corresponding r j<br />

and corresponding LAI<br />

not explicitly included in Theil-Sen regression. To<br />

j<br />

approach proposed by<br />

included to <strong>for</strong>m a new database <strong>for</strong> regression<br />

TAN et al. (2005). In this instance, in<strong>for</strong>mation on measurement precision is<br />

evaluate its in situ per<strong>for</strong>mance, LAI results surface derived reflectance <strong>for</strong> <strong>the</strong> analysis.<br />

included<br />

Basically<br />

to <strong>for</strong>m<br />

in<strong>for</strong>mation<br />

a new<br />

on <strong>the</strong> quality <strong>of</strong> in<br />

case database C <strong>for</strong> regression<br />

two test sites LAI are compared to ano<strong>the</strong>r i<br />

interesting<br />

in situ iLAI<br />

surface reflectance<br />

i<br />

situ measurements and satellite (group k)<br />

analysis. Basically in<strong>for</strong>mation<br />

observations<br />

on <strong>the</strong> quality<br />

is<br />

<strong>of</strong><br />

used<br />

in<br />

data set accounting <strong>for</strong> errors<br />

approach proposed … by Tan et al. (2005). … In this to derive new pairs in <strong>of</strong> both in input situ variables LAI and surface<br />

LAI LAIi<br />

i<br />

i<br />

situ measurements and satellite observations is used<br />

instance, in<strong>for</strong>mation LAI j on measurement j<br />

precision reflectance data (cf. Figure 6-1). In case A it is<br />

…<br />

…<br />

to derive new pairs r<strong>of</strong><br />

i in situ LAI<br />

i and surface<br />

is included to LAI <strong>for</strong>m a new database <strong>for</strong><br />

regression assumed that <strong>the</strong> true ... LAI value ... <strong>of</strong> each ESU is<br />

j<br />

j<br />

case A<br />

case B<br />

reflectance data (cf. Figure 6-1). In case A it is<br />

r<br />

analysis. Basically (group k) in<strong>for</strong>mation on (group <strong>the</strong> k) quality <strong>of</strong> in<br />

j<br />

LAI j<br />

found<br />

assumed<br />

within<br />

that <strong>the</strong><br />

a<br />

true<br />

certain<br />

LAI<br />

interval<br />

value <strong>of</strong> each<br />

(defined<br />

ESU<br />

by<br />

is<br />

impact <strong>of</strong> errors in in situ LAI<br />

impact <strong>of</strong> errors in surf. refl.<br />

situ measurements case A and satellite observations case B is used to measurement precision) mean/std. around mean/std. <strong>the</strong> measured LAI<br />

(group k)<br />

(group k)<br />

found within a certain interval (defined by<br />

LAIi Li<br />

i i<br />

LAIi<br />

derive new impact pairs <strong>of</strong> errors in <strong>of</strong> in situ in LAI situ LAI and impact surface <strong>of</strong> errors in surf. reflectance<br />

refl. value. Consequently a new group <strong>of</strong> LAI values is<br />

...<br />

... ...<br />

... measurement precision) around <strong>the</strong> measured LAI<br />

data (cf. Figure LAI <br />

<br />

j 6-3). L j<br />

j<br />

LAI<br />

i In case A it i<br />

is assumed i<br />

i<br />

LAI j that <strong>the</strong><br />

i <strong>for</strong>med<br />

value. Consequently<br />

including measured<br />

a new group<br />

LAI values<br />

<strong>of</strong> LAI<br />

<strong>of</strong> all<br />

values<br />

ESUs<br />

is<br />

...<br />

... ...<br />

...<br />

Figure 6-3 Derivation <strong>of</strong> reference values <strong>for</strong> regression<br />

true LAI mean/std. value <strong>of</strong> mean/std. each ESU is mean/std. found within mean/std. a certain whose<br />

<strong>for</strong>med<br />

LAI<br />

including<br />

values<br />

measured<br />

fall within<br />

LAI<br />

this<br />

values<br />

interval.<br />

<strong>of</strong> all ESUs<br />

The<br />

LAI <br />

j<br />

j<br />

j<br />

LAI j<br />

analysis with measurement errors in field<br />

choose all LAIj<br />

that fall within<br />

choose all j that fall within<br />

interval (defined by measurement precision) around<br />

<strong>the</strong> measurement error <strong>of</strong> LAI<br />

<strong>the</strong> measurement error <strong>of</strong> corresponding surface reflectances are grouped<br />

mean/std. mean/std. i<br />

mean/std. mean/std.<br />

i<br />

and corresponding j<br />

and corresponding LAI whose LAI values measurements fall within and spectral this reflectance interval. The data<br />

j<br />

<strong>the</strong> measured choose all LAI LAI choose all that fall within accordingly.<br />

corresponding<br />

Mean<br />

surface<br />

and standard<br />

reflectances<br />

deviation<br />

are<br />

<strong>for</strong><br />

grouped<br />

both<br />

j that fall value. within Consequently a new group<br />

taken into account (Tan et al. 2005).<br />

j<br />

<strong>the</strong> measurement error <strong>of</strong> LAIi<br />

<strong>the</strong> measurement error <strong>of</strong> i<br />

and corresponding j case C and corresponding LAIj<br />

groups are recorded. The process is repeated <strong>for</strong><br />

(group k)<br />

accordingly. Mean and standard deviation <strong>for</strong> both<br />

data set accounting <strong>for</strong> errors<br />

case B, but here measurement precision <strong>of</strong> surface<br />

in both case input variables C<br />

groups are recorded. The process is repeated <strong>for</strong><br />

(group k)<br />

<br />

reflectance is taken into account. Mean values <strong>for</strong><br />

data set i accounting <strong>for</strong> LAI errors i<br />

case B, but here measurement precision <strong>of</strong> surface<br />

in both input variables<br />

...<br />

...<br />

LAI and surface reflectance <strong>of</strong> cases A and B are<br />

<br />

j<br />

LAI j<br />

reflectance is taken into account. Mean values <strong>for</strong><br />

i<br />

LAIi<br />

... ...<br />

<strong>the</strong>n finally combined in case C. At this point a<br />

mean/std. mean/std.<br />

LAI and surface reflectance <strong>of</strong> cases A and B are<br />

j<br />

LAI j<br />

stable relationship is established, <strong>based</strong> on <strong>the</strong><br />

...<br />

... ...<br />

...<br />

LAI r<br />

r<br />

LAI j


6 Derivation <strong>of</strong> high resolution LAI maps<br />

6.3<br />

Results<br />

The first step in analysing <strong>the</strong> relationship between<br />

SVIs and texture and LAI was a close examination <strong>of</strong><br />

<strong>the</strong> reflectance obtained from different bands <strong>of</strong> high<br />

resolution image data. Irrespective <strong>of</strong> <strong>the</strong> satellite<br />

sensor (and <strong>the</strong> in situ method <strong>for</strong> LAI derivation),<br />

common trends could be observed. In <strong>the</strong> following,<br />

<strong>the</strong> Theil-Sen regression, being per<strong>for</strong>med <strong>for</strong><br />

LAI , LAI and LAI as well as SVIs<br />

e (LAI2000) e (DHP) (DHP)<br />

and texture measures, shed light on <strong>the</strong> relationship<br />

between <strong>the</strong>se variables. LAI measures were declared<br />

as independent, SVIs and texture measures as<br />

dependent variables. In addition, <strong>the</strong> method proposed<br />

by Tan et al. (2005) was applied and compared to <strong>the</strong><br />

results <strong>of</strong> Theil-Sen regression. Finally, <strong>the</strong> derived<br />

relationships are used to produce high resolution LAI<br />

maps <strong>for</strong> <strong>the</strong> two test sites with known accuracies.<br />

The maps are presented at <strong>the</strong> end <strong>of</strong> <strong>the</strong> respective<br />

chapters.<br />

6.3.1<br />

Budongo Forest<br />

Surface reflectance<br />

A first analysis revealed that – as expected – surface<br />

reflectance was highest in <strong>the</strong> NIR (20.9-36.3%) and<br />

lowest in <strong>the</strong> red portion <strong>of</strong> <strong>the</strong> spectrum (2.8-8.1%).<br />

Interestingly, reflectance in ASTER bands was always<br />

slightly higher than in <strong>the</strong> respective bands <strong>of</strong> SPOT-<br />

HRVIR. As mentioned in Chapter 4.2.2, ASTER<br />

had already been acquired as pre-processed surface<br />

reflectance data.<br />

With respect to <strong>the</strong> different <strong>for</strong>est stages, a clear<br />

decrease in ρ , ρ , and ρ <strong>of</strong> ASTER<br />

red SWIR (band 4) SWIR (band 5)<br />

could be observed with increasing <strong>for</strong>est age (cf.<br />

Figure A-7). ρ is highest in intermediate stages<br />

NIR<br />

<strong>of</strong> <strong>the</strong> <strong>for</strong>est, but <strong>the</strong> difference is not significant<br />

with respect to <strong>the</strong> early <strong>for</strong>est stage. Results from a<br />

Mann-Whitney U test however suggest that <strong>the</strong>re is<br />

109<br />

a significant difference between ρ <strong>of</strong> intermediate<br />

NIR<br />

and late <strong>for</strong>est stages (Z=-3.01, p


110<br />

SVIs<br />

In <strong>the</strong> next step <strong>the</strong> relationship between SVIs and<br />

LAI measures was assessed. Here too correlation<br />

analysis and bivariate plots <strong>of</strong> regressor and response<br />

variables served as first indicators. Although<br />

positive correlations (significant at p


6 Derivation <strong>of</strong> high resolution LAI maps<br />

per<strong>for</strong>med best (R 2 =0.89). Generally OLS regression<br />

yielded higher R 2 than Theil-Sen. This can easily be<br />

explained since <strong>the</strong> latter is derived from <strong>the</strong> median<br />

<strong>of</strong> slopes between ranked data points, while OLS<br />

minimizes <strong>the</strong> sum <strong>of</strong> squared errors in <strong>the</strong> variable<br />

to be predicted. As calculation <strong>of</strong> R2 is also <strong>based</strong> on<br />

<strong>the</strong> sum <strong>of</strong> squared errors in relation to total sum <strong>of</strong><br />

squares, OLS always leads to higher R2 . Never<strong>the</strong>less<br />

Theil-Sen regression is regarded as more robust with<br />

respect to outliers.<br />

Table 6-5 Theil-Sen and linear OLS regression models <strong>based</strong> on LAI and SVIs <strong>for</strong> early and intermediate <strong>for</strong>est stages<br />

e (DHP)<br />

in Budongo Forest. Models with <strong>the</strong> highest R2 in Theil-Sen regression are highlighted.<br />

Sensor Y Theil-Sen model R 2 Least squares model R 2<br />

SPOT-HRVIR SR 2.179+0.787 X 0.88 1.949+0.831 X 0.88<br />

NDVI 0.542+0.034 X 0.89 0.523+0.037 X 0.91<br />

NDVI c 0.163+0.065 X 0.85 0.101+0.072 X 0.90<br />

MSR 0.638+0.165 X 0.89 0.593+0.172 X 0.89<br />

RSR -0.082+0.843 X 0.89 -0.591+0.936 X 0.90<br />

NDMI -0.050+0.053 X 0.88 -0.096+0.060 X 0.90<br />

ASTER SR 2.250+0.602 X 0.94 2.407+0.583 X 0.94<br />

NDVI 0.518+0.031 X 0.94 0.516+0.032 X 0.94<br />

NDVI c 0.131+0.077 X 0.86 0.039+0.089 X 0.92<br />

MSR 0.636+0.130 X 0.93 0.651+0.129 X 0.94<br />

RSR -0.027+0.837 X 0.90 -0.672+0.911 X 0.94<br />

NDMI 0. 089+0.038 X 0.91 0.087+0.039 X 0.92<br />

Theil-Sen: R²=0.94<br />

Y=2.250+0.602 X<br />

OLS: R²=0.94<br />

Y=2.407+0.583 X<br />

Figure 6-6 Bivariate plots <strong>of</strong> a) LAIe<br />

(DHP) and SR derived from ASTER and b) LAI (DHP) and NDVI derived from SPOT<br />

c<br />

toge<strong>the</strong>r with OLS (solid line) and Theil-Sen linear regressions (dashed line).<br />

111<br />

Figure 6-6a displays a bivariate plot <strong>of</strong> LAIe (DHP)<br />

and SR derived from ASTER with OSL (solid line)<br />

and Theil-Sen linear regressions (dashed line). It is<br />

obvious that both regression models are quite similar<br />

(as also indicated in Table 6-5). By contrast, a strong<br />

deviance between model equations (and respective<br />

R2 ), as, <strong>for</strong> instance, in <strong>the</strong> case <strong>of</strong> LAI and NDVI (DHP) c<br />

<strong>based</strong> on SPOT data (cf. Figure 6-6b and Table A-11<br />

in <strong>the</strong> Appendix), indicates that distinct data groups<br />

are present. Whereas OLS regression combines both<br />

groups without being influenced by <strong>the</strong> within-group<br />

Theil-Sen: R²=0.60<br />

Y=0.424+0.014 X<br />

OLS: R²=0.79<br />

Y=-0.020+0.061 X


112<br />

correlation, Theil-Sen regression is biased by <strong>the</strong><br />

group containing more data points, in this case data<br />

from intermediate <strong>for</strong>est stages (cf. Equations 6.8 and<br />

6.9).<br />

Although Table 6-5 indicates that ASTER SR<br />

and NDVI retrieve <strong>the</strong> highest R2 in <strong>the</strong> modelled<br />

relationships, a closer analysis reveals that NDVI<br />

is not linearly related to LAI and that a non-<br />

e (DHP)<br />

2<br />

linear quadratic function (NDVI=0.69 LAIe (DHP)<br />

- 0.04 LAI + 0.34) leads to a higher R e (DHP) 2 <strong>of</strong> 0.95<br />

(cf. Figure 6-7a). In order to per<strong>for</strong>m <strong>the</strong> robust, but<br />

linear Theil-Sen regression, <strong>the</strong> variables were log<br />

trans<strong>for</strong>med (cf. Figure 6-7b. The resulting regression<br />

model NDVI =-0.723+0.212 X also retrieves an R log 2<br />

<strong>of</strong> 0.94 (OLS: NDVI =-0.736+0.218 X, R log 2 =0.95)<br />

with X being log trans<strong>for</strong>med LAI . O<strong>the</strong>r SVIs<br />

e (DHP)<br />

were also checked <strong>for</strong> non-linear relations and logtrans<strong>for</strong>med<br />

if necessary, but <strong>the</strong> above-mentioned<br />

ASTER SR and NDVI were still best per<strong>for</strong>ming.<br />

OLS: R²=0.95<br />

Y=0.69 X²-0.04 X+0.34<br />

Texture<br />

As described previously, LAI variables and SVIs<br />

correlated well <strong>for</strong> <strong>the</strong> early and intermediate <strong>for</strong>est<br />

stages. No correlation could in turn be retrieved <strong>for</strong><br />

<strong>the</strong> late <strong>for</strong>est stage. In addition, despite a significant<br />

difference in ρ and in ESU internal variance <strong>of</strong><br />

NIR<br />

ASTER bands 3, 4 and 5, no significant difference<br />

in LAI can be detected (cf. Chapters 5.3.1, 6.3.1<br />

and Figure 5-24). This indicates that besides foliage<br />

amount <strong>the</strong>re is ano<strong>the</strong>r major influence on surface<br />

reflectance in old growth <strong>for</strong>est sites. As discussed<br />

earlier, structural differences can be <strong>the</strong> reason.<br />

Shadow components in particular, resulting from an<br />

uneven canopy with emergent trees (cf. Figure 3-12),<br />

may lead to changes in surface reflectance, especially<br />

in <strong>the</strong> NIR and SWIR.<br />

As structural properties are reflected in image<br />

texture, correlation analyses were per<strong>for</strong>med to<br />

assess <strong>the</strong> relation between LAI and second-order<br />

texture variables (cf. Table 6-1) <strong>for</strong> late <strong>for</strong>est stages.<br />

Significant correlations to LAI variables could be<br />

retrieved <strong>for</strong> several texture measures from both<br />

ASTER and SPOT data, at p


6 Derivation <strong>of</strong> high resolution LAI maps<br />

additional correlations significant at <strong>the</strong> 0.01 level<br />

could be retrieved: <strong>for</strong> LAI and homogeneity<br />

e (LAI2000)<br />

(ASTER band 3, 3x3 kernel), LAI and variance<br />

e (DHP)<br />

(ASTER band 4, 9x9 kernel), as well as contrast<br />

(ASTER band 8, 9x9 kernel) with r <strong>of</strong> -0.88, 0.86<br />

s<br />

and -0.91 respectively. Variance <strong>of</strong> ASTER band<br />

4 had significant correlations with LAI in all<br />

e (DHP)<br />

kernels with r being highest <strong>for</strong> <strong>the</strong> 9x9 window<br />

s<br />

size. Contrast computed from ASTER band 8 and<br />

LAI e (DHP) yielded <strong>the</strong> highest r s with -0.91.<br />

Theil-Sen regression was subsequently applied to all<br />

texture measures with correlations to LAI variables<br />

significant at p


114<br />

First, <strong>the</strong> impact <strong>of</strong> measurement errors in field from newly composed groups as <strong>the</strong>ir estimates, bias,<br />

LAI on <strong>the</strong> retrieved surface reflectance values was precision and accuracy can be assessed according to<br />

assessed (case A). There<strong>for</strong>e groups <strong>of</strong> field LAI, LAIi <strong>the</strong> definitions in Chapter 6.2.1. Normalization by <strong>the</strong><br />

and corresponding surface reflectance were <strong>for</strong>med mean <strong>of</strong> <strong>the</strong> estimates allows <strong>the</strong> later comparison <strong>of</strong><br />

14 according to<br />

<strong>the</strong>se DERIVATION measures OF and HIGH leads RESOLUTION to relative bias, LAI MAPS precision<br />

and accuracy.<br />

LAI LAI LAI c<br />

(6.11)<br />

(6.11)<br />

i<br />

j<br />

i<br />

v<br />

with LAIi being LAIe (DHP) <strong>of</strong> a certain ESU and LAIj referring to LAIe (DHP) <strong>of</strong> all o<strong>the</strong>r ESUs. If LAIj is<br />

within <strong>the</strong> measurement error <strong>of</strong> LAIi, described by cv (cf. Chapter 6.2.1), LAIe (DHP) and corresponding ρred<br />

and ρNIR are added to <strong>the</strong> group <strong>of</strong> LAIi. The resulting groups contain 1 to 6 individual pairs (mean=3.15)<br />

<strong>of</strong> in situ LAIe (DHP) and surface reflectance in <strong>the</strong> red and NIR spectral bands <strong>of</strong> ASTER.<br />

Figure 6-9a shows <strong>the</strong> relationship between observed LAIe (DHP) and mean LAIe (DHP) <strong>of</strong> <strong>the</strong> newly <strong>for</strong>med<br />

groups. As expected, <strong>the</strong> correlation between observed field LAI and associated mean LAIi <strong>of</strong> groups is<br />

very high with R2=1, with LAI being LAI <strong>of</strong> a certain ESU and LAI i e (DHP) j<br />

referring to LAI <strong>of</strong> all o<strong>the</strong>r ESUs. If LAI is<br />

e (DHP) j<br />

within <strong>the</strong> measurement error <strong>of</strong> LAI described by c i, v<br />

(cf. Chapter 6.2.1), LAI and corresponding ρ e (DHP) red<br />

and ρ are added to <strong>the</strong> group <strong>of</strong> LAI . The resulting<br />

NIR i<br />

groups contain 1 to 6 individual pairs (µ=3.15) <strong>of</strong> in<br />

situ LAI and surface slope reflectance 1.00 and in <strong>of</strong>fset <strong>the</strong> red 0.02. and<br />

e (DHP) The differences between observed ρred and ρNIR and<br />

associated NIR spectral mean bands ρred <strong>of</strong> and ASTER. ρNIR are in turn slightly higher as displayed in Figure 6-9b. If observed values are<br />

taken as true values and means from newly composed groups as <strong>the</strong>ir estimates, bias, precision and<br />

accuracy Figure 6-9a can shows be assessed <strong>the</strong> relationship according between to <strong>the</strong> definitions observed in Chapter 6.2.1. Normalization by <strong>the</strong> mean <strong>of</strong> <strong>the</strong><br />

estimates LAI allows and mean <strong>the</strong> later LAIcomparison <strong>of</strong> <strong>the</strong> <strong>of</strong> newly <strong>the</strong>se <strong>for</strong>med<br />

e (DHP) e (DHP) measures and leads to relative bias, precision and accuracy.<br />

groups. As expected, <strong>the</strong> correlation between<br />

observed field LAI and associated mean LAI <strong>of</strong> <strong>the</strong><br />

i<br />

new groups is very high with R2 Correspondingly <strong>the</strong> proximity between observed<br />

ρ and ρ and mean values <strong>of</strong> associated groups<br />

red NIR<br />

is characterized by a relative bias, precision and<br />

accuracy <strong>of</strong>


eflectance <strong>of</strong> 17% in <strong>the</strong> red and 5% in <strong>the</strong> NIR wavelengths, which corresponds to an overall relative<br />

precision <strong>of</strong> 11% (TAN et al. 2005).<br />

6 Derivation <strong>of</strong> high resolution LAI maps<br />

In <strong>the</strong> second step (case B), <strong>the</strong> impact <strong>of</strong> observation errors in surface reflectance on <strong>the</strong> prediction <strong>of</strong><br />

LAIe (DHP) was analysed. Taking into account band specific values (cf. Table 6-4), and following WANG et<br />

al. (2001), two observations i and j are indistinguishable if<br />

2 2<br />

i j<br />

(6.12)<br />

(6.12)<br />

(cf. Equation 6.7). Analogue to <strong>the</strong> previously<br />

described steps, ano<strong>the</strong>r 20 groups <strong>of</strong> data are<br />

<strong>for</strong>med, each containing reflectance values that are<br />

equal within <strong>the</strong> observation precision (i.e. Equation<br />

6.12 holds true) and <strong>the</strong> corresponding LAIe (DHP)<br />

values. The resulting groups contain 1 to 5 data pairs<br />

(mean=2.75). The correlation between observed<br />

surface reflectance and mean surface reflectance <strong>of</strong><br />

<strong>the</strong> associated groups is again very high with R2 =1,<br />

slope 1.00 and <strong>of</strong>fset -0.01 <strong>for</strong> ρ and R red 2 =0.89, slope<br />

0.86 and <strong>of</strong>fset 0.04 <strong>for</strong> ρ (cf. Figure 6-10a).<br />

NIR<br />

Figure 6-10b displays <strong>the</strong> response <strong>of</strong> LAIe (DHP)<br />

to variations in surface reflectance that are<br />

equal to within measurement precision. In situ<br />

LAI and mean values over associated groups<br />

e (DHP)<br />

are well correlated (R2 =0.95) with a slope <strong>of</strong> 0.99<br />

and an intercept <strong>of</strong> 0.05. Yet especially <strong>for</strong> higher<br />

LAI , field data can vary significantly (between<br />

e (DHP)<br />

5.68 and 7.07), with mean values remaining essentially<br />

unchanged (6.22 to 6.35). Relative bias, precision<br />

and accuracy are


116<br />

Assuming that no measurement errors are present<br />

in <strong>the</strong> resulting data set, OLS regression was<br />

per<strong>for</strong>med <strong>the</strong>reafter. As expected, ASTER SR and<br />

log trans<strong>for</strong>med NDVI again produced <strong>the</strong> highest<br />

R2 in <strong>the</strong> modelled relationships, with ASTER SR<br />

per<strong>for</strong>ming slightly better. Figure 6-12 displays<br />

bivariate plots <strong>of</strong> both SVIs and (log trans<strong>for</strong>med)<br />

LAI derived from <strong>the</strong> previously <strong>for</strong>med groups.<br />

e (DHP)<br />

Interestingly, <strong>the</strong> results show that if measurement<br />

errors in both LAI and surface reflectance are<br />

e (DHP)<br />

OLS: R²=0.99<br />

Y=-0.04+1.01 X<br />

taken into account, <strong>the</strong> agreement between <strong>the</strong> two<br />

variables is more precise and more accurate (cf.<br />

Table 6-7). However, modelled relationships derived<br />

from <strong>the</strong> above-described method differ only slightly<br />

from <strong>the</strong> models derived with Theil-Sen regression<br />

in <strong>the</strong> previous chapter, yet yield a higher R2 . Table<br />

6-8 summarizes <strong>the</strong> regression models <strong>for</strong> early and<br />

intermediate stages derived from Theil-Sen regression<br />

and <strong>the</strong> method according to Tan et al. (2005).<br />

The models are very similar with SR per<strong>for</strong>ming<br />

slightly better than log trans<strong>for</strong>med NDVI in <strong>the</strong> Tan<br />

approach.<br />

OLS: R²=1<br />

Y=-0.19+1.00 X<br />

Figure 6-11 Relationship between a) observed LAIe<br />

(DHP) and mean LAI values over associated groups and b) observed surface<br />

e (DHP)<br />

reflectance and mean values over associated groups in <strong>the</strong> red (circles) and NIR (asterisks) spectral bands <strong>of</strong> ASTER,<br />

taking into account observation errors in both field LAI and ASTER data (solid lines represent 1:1 line <strong>for</strong> comparison).<br />

OLS: R²=0.98<br />

Y=2.282+0.604 X<br />

OLS: R²=0.98<br />

Y=-0.742+0.222 X<br />

Figure 6-12 Bivariate plots <strong>of</strong> a) log trans<strong>for</strong>med LAIe<br />

(DHP) and NDVI (ASTER) and b) LAI and SR (ASTER) toge<strong>the</strong>r with OLS<br />

e (DHP)<br />

(solid line) regression derived from case C.


6 Derivation <strong>of</strong> high resolution LAI maps<br />

The SR model is also illustrated in Figure 6-13. Here<br />

<strong>the</strong> original data set <strong>of</strong> LAI and ASTER SR is<br />

e (DHP)<br />

displayed toge<strong>the</strong>r with <strong>the</strong> Theil-Sen regression<br />

model (dashed line) and <strong>the</strong> OLS regression model<br />

(solid line, Tan method). As LAI calculated from<br />

e<br />

both models differs only by about 1%, <strong>the</strong> methods<br />

can be regarded as equal. Yet when OLS regression<br />

is per<strong>for</strong>med on <strong>the</strong> original data set, differences in<br />

LAI <strong>of</strong> up to 7% can occur. Theil-Sen regression can<br />

e<br />

<strong>the</strong>re<strong>for</strong>e be regarded as a robust method to derive<br />

statistically sound relationships between variables<br />

that are not free <strong>of</strong> measurement errors. Theil-Sen<br />

regression models will consequently be used <strong>for</strong> <strong>the</strong><br />

calculation <strong>of</strong> <strong>the</strong> high resolution LAI map.<br />

Table 6-7 Relative bias, precision and accuracies <strong>of</strong> relationships between LAI and surface reflectance derived from<br />

e (DHP)<br />

ASTER data. Groups contain only data from early and intermediate <strong>for</strong>est stages.<br />

Case A Case B Case C<br />

ASTER reflectance [%] LAI e (DHP) [%] LAI e (DHP) [%] ASTER reflectance [%]<br />

ρ red ρ NIR ρ red ρ NIR<br />

Relative bias


SR 2.<br />

250<br />

LAIe , (6.13)<br />

0.<br />

602<br />

118<br />

as used to calculate LAIe <strong>for</strong> early/intermediate stages. LAIe <strong>of</strong> late <strong>for</strong>est stages was modelled according<br />

to<br />

GLCM variance (band 4) 4.912<br />

LAI e <br />

2.<br />

746<br />

(6.14)<br />

(6.14)<br />

As GLCM variance can be relatively high at at <strong>for</strong>est<br />

edges or around large <strong>for</strong>est gaps due to strong<br />

differences edges or around in surface large reflectance <strong>for</strong>est gaps <strong>of</strong> due adjacent to strong pixels, one fur<strong>the</strong>r constraint was included in LAIe<br />

differences in surface reflectance <strong>of</strong> adjacent<br />

pixels, one fur<strong>the</strong>r constraint was included in LAIe calculation: if GLCM variance exceeded <strong>the</strong> data<br />

space shown in Figure 6-8, i.e. it is higher than 18<br />

and lower than 8, Equation 6.13 was applied instead.<br />

Figure 6-14 displays <strong>the</strong> relationship between field<br />

measured LAI values and LAI derived from<br />

e (DHP) e<br />

Equations 6.13 and 6.14. R2 be – at least partially – affected. As no ESU had been<br />

established in <strong>the</strong>se areas as <strong>the</strong>y were mostly not<br />

accessible at <strong>the</strong> end <strong>of</strong> <strong>the</strong> rainy season, no statement<br />

can be made concerning <strong>the</strong> validity <strong>of</strong> LAI in <strong>the</strong><br />

e<br />

affected areas.<br />

Never<strong>the</strong>less – as in <strong>the</strong> areas declared as late <strong>for</strong>est<br />

stages in Figure 6-15 – comparatively high amounts <strong>of</strong><br />

Cynometra alexandri seem to be present (cf. classes<br />

“Cynometra” and “Cynometra mixed” in Figure 2-6).<br />

There<strong>for</strong>e <strong>the</strong> relationship between LAI and SVIs<br />

e<br />

or texture measures is probably not only influenced<br />

by applied logging schemes, but also species<br />

<strong>of</strong> <strong>the</strong> relationship is 0.87 composition. On <strong>the</strong> o<strong>the</strong>r hand compartments W19<br />

with an overall RMSE <strong>of</strong> 0.47 (0.39 <strong>for</strong> early and and W20, which according to Plumptre (1996) should<br />

intermediate, 0.64 <strong>for</strong> late <strong>for</strong>est stages). The relative also be characterized by high Cynometra abundance,<br />

accuracy is 9% (7% and 11% <strong>for</strong> early/intermediate show no significant reduction in LAI compared to<br />

e<br />

and late stages respectively).<br />

neighbouring compartments in <strong>the</strong> south and west.<br />

Consequently Equation 6.14 cannot be applied to<br />

The resulting high resolution LA map <strong>for</strong> Budongo<br />

e compartments with apparently underestimated LAIe. Forest is displayed in Figure 6-15. The dashed Ra<strong>the</strong>r <strong>the</strong>se areas will be excluded from fur<strong>the</strong>r<br />

lines represent those <strong>for</strong>est compartments to which<br />

Equation 6.14 was applied. Due to <strong>the</strong> moving<br />

window size <strong>of</strong> 9x9 pixels those areas appear to be<br />

less detailed. Red squares in Figure 6-15 represent<br />

three focus areas that are displayed in Figures 6-16a,<br />

c and e, toge<strong>the</strong>r with <strong>the</strong> corresponding areas in <strong>the</strong><br />

ASTER image (band combination 4, 3, 2) in Figures<br />

6-16b, d and f. Figure 6-16a and b show <strong>the</strong> area <strong>of</strong><br />

<strong>the</strong> Nature Reserve N15 in <strong>the</strong> southwest <strong>of</strong> Budongo<br />

Forest. The overall LAI pattern <strong>of</strong> <strong>for</strong>est and<br />

e<br />

cleared areas with wooded grassland is well covered<br />

(cf. Figure 6-16a). The latter has LAI values <strong>of</strong> 2.2<br />

e<br />

to 3.7. Inside <strong>the</strong> <strong>for</strong>est LAI variability is higher with<br />

e<br />

minimum and maximum values <strong>of</strong> 3.0 and 8.5 and a<br />

mean <strong>of</strong> 6.0 (±0.9). Figures 6-16c to f display <strong>the</strong> two<br />

analysis, i.e. <strong>MODIS</strong> LAI product validation.<br />

<strong>for</strong>est areas in <strong>the</strong> sou<strong>the</strong>ast and north respectively.<br />

OLS: R²=0.87<br />

Here clear patterns <strong>of</strong> areas with lower and higher<br />

LAI are visible. The areas with lower LAI are<br />

e e<br />

following <strong>the</strong> borders <strong>of</strong> compartments W29, W30,<br />

Y=0.096+1.004 X<br />

W41, W42, and W43. Additionally compartments<br />

Figure 6-14 Relation between LAI estimated from ASTER<br />

e<br />

data with equations 6.12 and 6.13 and in situ<br />

KP2, KP 4, KP7, S4, S5, W41, W42 and W43 seem to<br />

measured LAI.


6 Derivation <strong>of</strong> high resolution LAI maps<br />

Figure 6-15 High resolution LAI map <strong>of</strong> Budongo Forest estimated from ASTER data.<br />

e<br />

119


120<br />

Figure 6-16 Focus areas <strong>of</strong> Budongo Forest displayed in Figure 6-15. a), c) and e) represent LAI calculated from Equations 6.12<br />

e<br />

and 6.13, b), d) and f) show ASTER data <strong>for</strong> <strong>the</strong> corresponding areas (band combination 4-3-2). Dashed areas<br />

represent late <strong>for</strong>est stages.


6 Derivation <strong>of</strong> high resolution LAI maps<br />

6.3.2<br />

Kakamega Forest<br />

Data analysis <strong>for</strong> <strong>the</strong> Kakamega Forest test site<br />

followed <strong>the</strong> same steps as described <strong>for</strong> Budongo<br />

Forest. Once more it must be mentioned that especially<br />

in situ data quality was not comparable. First <strong>of</strong> all,<br />

measurement precision <strong>of</strong> LAI could not be<br />

e (LAI2000)<br />

determined due to <strong>the</strong> lack <strong>of</strong> a third LAI-2000 PCA<br />

device and different filter methods. Consequently<br />

only Theil-Sen regression was per<strong>for</strong>med to model<br />

<strong>the</strong> relationship between LAI variables and SVIs/<br />

texture variables. Second, LAI , which correlated<br />

e (DHP)<br />

best to high resolution satellite data <strong>for</strong> <strong>the</strong> Budongo<br />

Forest test site excluded understorey vegetation in<br />

Kakamega Forest. And last but not least, ASTER data,<br />

which gave better results in Budongo Forest <strong>for</strong> both<br />

SVIs and texture measures, had not been available<br />

<strong>for</strong> Kakamega Forest. Additionally ESUs 11, 13, and<br />

29 had to be excluded from analysis due to clouds<br />

or cloud shadows. The results <strong>of</strong> <strong>the</strong> analyses will be<br />

presented toge<strong>the</strong>r with <strong>the</strong> high spatial resolution<br />

LAI map in <strong>the</strong> following section.<br />

Surface reflectance<br />

The overall surface reflectance pattern <strong>of</strong> <strong>the</strong> different<br />

<strong>for</strong>est stages was comparable to <strong>the</strong> values retrieved<br />

<strong>for</strong> <strong>the</strong> Budongo Forest test site, albeit with mean<br />

surface reflectance per <strong>for</strong>est stage being slightly<br />

lower in <strong>the</strong> red part <strong>of</strong> <strong>the</strong> spectrum (0.5-1%) and<br />

slightly higher in <strong>the</strong> NIR and SWIR bands (1-4%,<br />

cf. Figures 6-4 and A-7. In addition, correlation<br />

analysis revealed similar results as <strong>for</strong> Budongo<br />

Forest. LAI and LAI were again<br />

e (LAI2000) e (DHP)<br />

negatively correlated to ρ green , ρ red and ρ SWIR with r s<br />

ranging between -0.51 and -0.58 (p


122<br />

The corresponding R 2 was, however, lower with 0.65<br />

(OLS) and 0.63 (Theil-Sen).<br />

Figure 6-18 displays a bivariate plot <strong>of</strong> LAIe (LAI2000)<br />

and RSR derived from SPOT data, toge<strong>the</strong>r with <strong>the</strong><br />

Theil-Sen (dashed line) and OLS models (solid line)<br />

given in Table 6-9. The inverted regression equation<br />

retrieved from <strong>the</strong> Theil-Sen model was used <strong>for</strong> LAIe Figure 6-18 Bivariate plot <strong>of</strong> LAI and RSR derived<br />

e (LAI2000)<br />

from SPOT data <strong>for</strong> Kakamega Forest toge<strong>the</strong>r<br />

Table 6-9 Theil-Sen and linear OLS regression models <strong>based</strong> on LAI and SVIs <strong>for</strong> early and intermediate <strong>for</strong>est stages in<br />

e (LAI2000)<br />

Kakamega Forest.<br />

Sensor Y Theil-Sen model R 2 Least squares model R 2<br />

SPOT-HRVIR SR 6.600+0.468 X 0.63 5.326+0.685 X 0.77<br />

Table 6-10<br />

NDVI 0.745+0.011 X 0.48 0.684+0.022 X 0.78<br />

NDVI c 0.306+0.038 X 0.67 0.267+0.044 X 0.72<br />

MSR 3.207+0.089 X 0.58 2.877+0.147 X 0.78<br />

RSR 2.478+0.608 X 0.72 2.002+0.678 X 0.75<br />

NDMI 0.651-0.023 X 0.61 0.687-0.030 X 0.65<br />

Theil-Sen and linear OLS regression models <strong>based</strong> on LAI variables and texture measures <strong>for</strong> late <strong>for</strong>est stages<br />

in Kakamega Forest.<br />

Sensor Kernel Band Texture measure X Theil-Sen model R 2 Least squares model R 2<br />

SPOT-<br />

HRVIR<br />

with OLS (solid line) and Theil-Sen linear<br />

regressions (dashed line).<br />

Theil-Sen: R²=0.72<br />

Y=2.478+0.608 X<br />

OLS (Tan): R²=0.75<br />

Y=2.002+0.678 X<br />

estimation in early and intermediate <strong>for</strong>est stages in<br />

Kakamega Forest.<br />

Texture<br />

With respect to <strong>the</strong> late <strong>for</strong>est stages <strong>of</strong> Kakamega<br />

Forest, high r were retrieved <strong>for</strong> various texture<br />

s<br />

Figure 6-19 Bivariate plot <strong>of</strong> LAI and homogeneity (band<br />

e (DHP)<br />

4, 7x7 kernel) derived from SPOT with OLS<br />

(solid line) and Theil-Sen linear regression<br />

models (dashed line).<br />

Theil-Sen: R²=0.36<br />

Y=0.447-0.047 X<br />

OLS (Tan): R²=0.50<br />

Y=0.468-0.052 X<br />

5x5 4 Homogeneity LAI e (DHP) 0.441-0.045 X 0.35 0.497-0.059 X 0.49<br />

7x7 4 Homogeneity LAI e (DHP) 0.447-0.047 X 0.36 0.468-0.052 X 0.50


6 Derivation <strong>of</strong> high resolution LAI maps<br />

measures at p


124<br />

Figure 6-21 Focus areas <strong>of</strong> Kakamega Forest displayed in Figure 6-22. a) and c) represent LAI calculated from Equation 6.14,<br />

e<br />

b) and d) show SPOT data <strong>of</strong> <strong>the</strong> corresponding areas (band combination 4-3-2).<br />

Overall this model yielded an R 2 <strong>of</strong> 0.53 with<br />

an accuracy <strong>of</strong> 0.8 (relative accuracy 16%). The<br />

resulting high resolution LAI map is displayed in<br />

Figure 6-20. Once more <strong>the</strong> overall LAI pattern<br />

e<br />

is well covered, with slightly higher values in <strong>the</strong><br />

central part <strong>of</strong> <strong>the</strong> <strong>for</strong>est that is assumed to be less<br />

disturbed than <strong>the</strong> peripheral parts. Two focus areas<br />

are displayed in Figure 6-21. Whereas <strong>the</strong> area around<br />

Buyangu Hill (Figure 6-21a and b) shows reasonable<br />

patterns <strong>of</strong> LAI indicating low LAI values (1-5)<br />

e e<br />

in <strong>the</strong> early succession stages <strong>of</strong> Psidium guajava<br />

and high LAI in <strong>the</strong> relatively undisturbed <strong>for</strong>est<br />

e<br />

area <strong>of</strong> Salazar, LAI patterns in Figure 6-21c indicate<br />

e<br />

once more <strong>the</strong> model’s constraint. Very high LAI e <strong>of</strong><br />

>8 is shown <strong>for</strong> areas sou<strong>the</strong>ast <strong>of</strong> Isecheno Forest<br />

Station. For large parts <strong>of</strong> <strong>the</strong>se areas <strong>the</strong> classification<br />

<strong>of</strong> Kakamega Forest (displayed in Figure A-2)<br />

indicates plantation <strong>for</strong>ests <strong>of</strong> Bisch<strong>of</strong>fia javanica, <strong>for</strong><br />

which <strong>the</strong> model is not valid. The same effect occurs<br />

in <strong>the</strong> south and northwest <strong>of</strong> Kakamega Forest.<br />

6.4<br />

Conclusion<br />

The objective <strong>of</strong> <strong>the</strong> analyses described in <strong>the</strong><br />

preceding chapter was <strong>the</strong> production <strong>of</strong> high


6 Derivation <strong>of</strong> high resolution LAI maps<br />

resolution LAI maps <strong>for</strong> <strong>the</strong> two test sites with known<br />

accuracy <strong>based</strong> on in situ measurements and satellite<br />

observations. For that reason empirical transfer<br />

functions were established as <strong>the</strong>y – in contrast to<br />

physical models – are only limited by <strong>the</strong> sampling<br />

distribution. The disadvantage is that empirical<br />

models cannot be transferred in time or space,<br />

but as <strong>the</strong>y serve only as an intermediate step <strong>for</strong><br />

validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product in this <strong>the</strong>sis,<br />

<strong>the</strong> establishment <strong>of</strong> generally valid models was not<br />

necessary. Ideally high resolution LAI maps should<br />

be produced without loss in accuracy in relation to<br />

in situ data. There<strong>for</strong>e, measurement precision <strong>of</strong><br />

<strong>the</strong> input variables and <strong>the</strong>ir impact on <strong>the</strong> resulting<br />

transfer functions must be taken into account.<br />

Whereas precision <strong>of</strong> field measurements in Budongo<br />

Forest was comparable <strong>for</strong> both devices, with 4% and<br />

5% (LAI-2000 PCA and DHP respectively), precision<br />

<strong>of</strong> surface reflectance varies by up to 7%. Once more<br />

it must be emphasized that measurement precision<br />

<strong>of</strong> field LAI as discussed in Chapter 6.2.1 refers to<br />

relative precision, because in<strong>for</strong>mation on true LAI<br />

was not available. Additionally precision <strong>of</strong> surface<br />

reflectance could not be directly determined, but was<br />

retrieved from literature. As Tan et al (2005) state, <strong>the</strong><br />

estimation <strong>of</strong> LAI from spectral surface reflectance<br />

(or derived spectral vegetation indices) is an ill-posed<br />

problem. Whereas small variations in LAI result only<br />

in limited variations in modelled spectral surface<br />

reflectance (i.e. <strong>the</strong> prediction <strong>of</strong> <strong>the</strong> radiation field is<br />

a well-posed problem), <strong>the</strong> inverse calculation results<br />

in an unstable solution, i.e. small variations in input<br />

reflectance due to observation errors can result in a<br />

very low precision <strong>of</strong> LAI.<br />

There<strong>for</strong>e it is necessary to derive stable relationships<br />

that also account <strong>for</strong> errors in input data. To evaluate<br />

<strong>the</strong> per<strong>for</strong>mance <strong>of</strong> Theil-Sen regression in <strong>the</strong><br />

presence <strong>of</strong> measurement errors, <strong>the</strong> results were<br />

compared to a method proposed by Tan et al (2005).<br />

Whereas <strong>the</strong> <strong>for</strong>mer is very robust with respect to<br />

outliers, it does not include in<strong>for</strong>mation on precision<br />

125<br />

<strong>of</strong> input data. Tan et al.’s method incorporates this<br />

in<strong>for</strong>mation. However, a comparison <strong>of</strong> <strong>the</strong> results<br />

showed that <strong>the</strong> derived transfer functions differ only<br />

slightly, with discrepancies in modelled LAI <strong>of</strong> ~1%.<br />

Both approaches are thus assumed to give comparable<br />

and reliable results even in <strong>the</strong> presence <strong>of</strong> errors in<br />

input data.<br />

For Budongo Forest, correlation and regression<br />

analyses revealed that different transfer functions<br />

had to be set up <strong>for</strong> early/intermediate and late <strong>for</strong>est<br />

stages respectively. Whereas SVIs are sensitive to<br />

changes in LAI in early to intermediate <strong>for</strong>est stages<br />

(retrieving an R2 <strong>of</strong> 0.94), no relation could be found<br />

to <strong>the</strong> LAI variables from late <strong>for</strong>est stages. This is<br />

probably caused by a different <strong>for</strong>est structure: in<br />

mature undisturbed tropical rain <strong>for</strong>est emergent<br />

trees are present, creating a very rough canopy<br />

texture. The result is a higher shadow component<br />

than in intermediate <strong>for</strong>est stages, where emergent<br />

trees, in particular, have been logged. Never<strong>the</strong>less<br />

overall LAI can be comparable due to <strong>the</strong> denser<br />

understorey in intermediate <strong>for</strong>est stages. Due to <strong>the</strong><br />

smoo<strong>the</strong>r canopy surface in intermediate <strong>for</strong>est sites,<br />

larger surface areas <strong>of</strong> crowns are exposed, causing a<br />

higher surface reflectance especially in <strong>the</strong> NIR and<br />

SWIR bands. The mutual shadowing <strong>of</strong> crowns in <strong>the</strong><br />

late <strong>for</strong>est stages causes, in turn, a higher variance<br />

in canopy reflection. Consequently a good relation<br />

(R2 =0.71) between GLCM variance <strong>of</strong> ASTER band<br />

4 and LAI could be found <strong>for</strong> late <strong>for</strong>est stages.<br />

e (DHP)<br />

Interestingly in Budongo Forest regressions with<br />

<strong>the</strong> highest R2 were found <strong>for</strong> LAI irrespective<br />

e (DHP),<br />

<strong>of</strong> <strong>the</strong> <strong>for</strong>est stage. The inclusion <strong>of</strong> understorey<br />

(in contrary to LAI , where understorey<br />

e (LAI2000)<br />

vegetation is excluded) improves <strong>the</strong> relation between<br />

LAI variables and SVIs or texture variables. The<br />

correction <strong>for</strong> foliage clumping in LAI , however,<br />

(DHP)<br />

seems to weaken <strong>the</strong> strength <strong>of</strong> <strong>the</strong> modelled<br />

relationship. This has also been reported by Eklundh<br />

et al. (2003) where correlations between LAI and<br />

e


126<br />

SVIs were significantly stronger than between LAI<br />

corrected <strong>for</strong> clumping and SVIs. According to<br />

Chen et al. (2005), SVIs are generally proportional<br />

to foliage cover in <strong>the</strong> view direction. Fur<strong>the</strong>r foliage<br />

clumping affects gap fraction <strong>for</strong> <strong>the</strong> same LAI and<br />

can thus delay <strong>the</strong> occurrence <strong>of</strong> <strong>the</strong> saturation in<br />

reflectance as LAI increases. Consequently LAIe has a stronger relationship to <strong>the</strong> satellite signal. In<br />

this context, Stenberg et al. (2004) suggest applying<br />

clumping corrections a posteriori to LAI estimation<br />

from SVIs.<br />

With respect to satellite sensors, ASTER data<br />

per<strong>for</strong>med better than SPOT data in Budongo Forest.<br />

The reason <strong>for</strong> this remains unclear, but possible<br />

explanations include sensor characteristics (e.g.<br />

slightly different band widths), as well as applied<br />

atmospheric correction algorithms. Last but not least,<br />

<strong>for</strong>est phenology at <strong>the</strong> time <strong>of</strong> ASTER acquisition<br />

might be more similar to phenology at <strong>the</strong> time <strong>of</strong><br />

fieldwork – even though <strong>the</strong> imagery was acquired<br />

one year after.<br />

In general, similar observations can be made <strong>for</strong><br />

Kakamega Forest. Here too <strong>the</strong> exclusion <strong>of</strong> LAI data<br />

from late <strong>for</strong>est stages improved <strong>the</strong> relationships<br />

between SVIs and LAI variables. In this case RSR<br />

was found to be best suited <strong>for</strong> LAI estimation,<br />

however with a lower R2 <strong>of</strong> 0.72. In contrast to<br />

Budongo Forest, <strong>the</strong> best relationship was retrieved<br />

with LAI Apparently <strong>the</strong> fact that understorey<br />

e (LAI2000).<br />

vegetation was not covered with LAI on this test<br />

e (DHP)<br />

site had a negative impact on <strong>the</strong> resulting models. For<br />

late <strong>for</strong>est stages a significant relationship between<br />

effective LAI and a texture variables calculated from<br />

<strong>the</strong> SWIR band was also established, although in<br />

this case <strong>the</strong> relationship was comparatively weak<br />

(R2 =0.36). Interestingly <strong>the</strong> best correlations <strong>for</strong><br />

texture variables were retrieved <strong>for</strong> a 9x9 moving<br />

window <strong>for</strong> ASTER (Budongo Forest) and 7x7 kernel<br />

<strong>for</strong> SPOT (Kakamega Forest), which correspond<br />

to 135 m and 140 m side length respectively. As<br />

individual crowns can have up to 30 m in diameter,<br />

this edge length apparently best captures <strong>the</strong> variation<br />

<strong>of</strong> surface reflectance being reflected by <strong>the</strong> canopy.<br />

In comparison to Broich (2005), who produced<br />

a high resolution LAI map <strong>for</strong> Kakamega Forest<br />

<strong>based</strong> on <strong>the</strong> same LAI-2000 PCA measurements<br />

(but including measurements in south and west<br />

directions, cf. Chapter 5) and Landsat ETM+ SLC-<strong>of</strong>f<br />

data, <strong>the</strong> previously described analyses revealed no<br />

significant relation between SWIR reflectance <strong>of</strong><br />

SPOT and LAI as shown in his study. In this <strong>the</strong>sis<br />

e<br />

better relationships were yielded by <strong>the</strong> calculated<br />

SVIs. The here applied SPOT-4 data has moreover<br />

<strong>the</strong> advantage <strong>of</strong> full spatial coverage in contrast<br />

to <strong>the</strong> Landsat ETM+ SLC-<strong>of</strong>f data. Results from<br />

Budongo Forest fur<strong>the</strong>r suggest that <strong>the</strong> estimation<br />

<strong>of</strong> LAI from remote sensing data might generally<br />

e<br />

be more accurate when different models are set up<br />

<strong>for</strong> different <strong>for</strong>est stages (though in <strong>the</strong> end only one<br />

model was used <strong>for</strong> Kakamega also in this <strong>the</strong>sis.).<br />

For both test sites, care has to be taken to integrate<br />

only those areas in <strong>the</strong> upscaling process, which are<br />

comparable to <strong>the</strong> ESUs on which <strong>the</strong> regression<br />

models were established. In Budongo Forest, <strong>the</strong><br />

dominance <strong>of</strong> certain species (in this case Cynometra<br />

alexandri) may distort <strong>the</strong> modelled relationships.<br />

In Kakamega Forest it is quite obvious that <strong>the</strong><br />

LAI in plantantions is exaggerated. Broich (2005)<br />

encountered similar problems in his analyses.<br />

The resulting high resolution LAI e map <strong>of</strong> Budongo<br />

Forest has a relative accuracy <strong>of</strong> 9%. The upscaling<br />

process thus caused only a slight – yet still acceptable<br />

– degradation in accuracy <strong>of</strong> <strong>the</strong> high resolution<br />

LAI map compared to <strong>the</strong> accuracy <strong>of</strong> in situ<br />

e<br />

measurements and satellite observations. The LAIe map <strong>of</strong> Kakamega Forest is, however, less accurate<br />

at 16% due to inferior in situ data quality. Map<br />

accuracies should be kept in mind when validating<br />

<strong>the</strong> medium resolution LAI product.


7<br />

<strong>Validation</strong> <strong>of</strong> <strong>the</strong><br />

<strong>MODIS</strong> LAI product<br />

As described in Chapter 3.4.4 <strong>the</strong> final steps in <strong>MODIS</strong><br />

LAI product validation involve <strong>the</strong> aggregation <strong>of</strong><br />

<strong>the</strong> high resolution LAI maps and comparison with<br />

e<br />

<strong>the</strong> <strong>MODIS</strong> data sets to be validated. However, care<br />

has to be taken when using <strong>the</strong> high resolution maps<br />

presented in Chapter 6 as <strong>the</strong>y give LAI , which is<br />

e<br />

not corrected <strong>for</strong> foliage clumping. Yet <strong>the</strong> <strong>MODIS</strong><br />

LAI product is supposed to model “true” LAI. In<br />

order to resolve this, a clumping factor was applied<br />

to <strong>the</strong> high resolution maps as suggested by Stenberg<br />

et al. (2004). At least <strong>for</strong> intermediate and late <strong>for</strong>est<br />

stages this can be derived from DHP. For early <strong>for</strong>est<br />

stages, however λ estimated from DHPs seems to be<br />

too small when compared to reference data.<br />

In <strong>the</strong> next step <strong>the</strong> quality <strong>of</strong> input data <strong>for</strong> <strong>the</strong><br />

LAI algorithm and its per<strong>for</strong>mance is assessed to<br />

find <strong>the</strong> right time steps <strong>for</strong> <strong>the</strong> final validation <strong>of</strong><br />

<strong>the</strong> <strong>MODIS</strong> LAI product. Based on uncertainty<br />

measures introduced in chapter 6.2.1, its accuracy<br />

can subsequently be determined <strong>for</strong> <strong>the</strong> respective<br />

study sites and time steps. Here also <strong>the</strong> model<br />

observation precision <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

has to be taken into account, which is reported to<br />

be 0.2 (Tan et al. 2005). Although a true temporal<br />

validation would require field measurements <strong>of</strong> LAI<br />

over several months, at least temporal consistency <strong>of</strong><br />

<strong>the</strong> <strong>MODIS</strong> LAI product is investigated. In<strong>for</strong>mation<br />

on <strong>the</strong> applied LAI algorithm is analysed and LAI<br />

trajectories are derived <strong>for</strong> <strong>the</strong> different <strong>for</strong>est stages.<br />

7.1<br />

Upscaling <strong>of</strong> <strong>the</strong> high resolution<br />

LAI map e<br />

127<br />

To avoid resampling <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product,<br />

Morisette et al. (2006a) suggest comparing <strong>the</strong> high<br />

resolution LAI map with <strong>the</strong> <strong>MODIS</strong> LAI product<br />

e<br />

in <strong>the</strong> projection <strong>of</strong> <strong>the</strong> latter. Since spatial errors are<br />

relative to pixel size, resampling will have less impact<br />

if applied to <strong>the</strong> high spatial resolution LAI maps.<br />

e<br />

The aggregation <strong>of</strong> <strong>the</strong> high spatial resolution LAIe maps should fur<strong>the</strong>r take into account <strong>the</strong> point spread<br />

function <strong>of</strong> <strong>the</strong> high resolution satellite product. This<br />

step is however poorly addressed in current research<br />

and also complicated by various preprocessing<br />

steps and repeated resampling in product generation<br />

(Morisette et al. 2006a). To avoid geolocation<br />

uncertainties and to retrieve a statistically stable<br />

result, Yang et al. (2006) recommend comparison <strong>of</strong><br />

both products at multipixel (patch) scale (cf. Chapter<br />

3.4.5).<br />

7.2<br />

Correction <strong>for</strong> foliage clumping<br />

As previously mentioned, <strong>the</strong> <strong>MODIS</strong> LAI algorithm<br />

is designed to give true LAI. On <strong>the</strong> o<strong>the</strong>r hand <strong>the</strong><br />

regression models developed in Chapter 6 are <strong>based</strong><br />

on LAI and LAI , which obviously<br />

e (LAI2000) e (DHP)<br />

correlated better to SVIs and texture measures<br />

derived from high resolution satellite data. Yet<br />

according to Stenberg et al. (2004), a correction <strong>for</strong><br />

foliage clumping can also be applied to <strong>the</strong> final LAIe maps, if <strong>the</strong> clumping factor λ is known.<br />

Table 7-1 lists mean clumping factors derived<br />

from hemispherical photography <strong>for</strong> <strong>the</strong> different<br />

<strong>for</strong>est stages <strong>of</strong> both study sites. A comparison with<br />

clumping indices derived on a global scale by Chen<br />

et al. (2005) from multi-angular POLDER 1 data<br />

reveals that <strong>the</strong> clumping factor λ <strong>for</strong> intermediate<br />

and late <strong>for</strong>est stages is well within <strong>the</strong> expected range<br />

<strong>for</strong> both test sites (cf. Table 7.1). Chen et al. (2005)


128<br />

estimated <strong>the</strong> clumping index <strong>for</strong> evergreen broadleaf<br />

<strong>for</strong>ests to be 0.63 on global average, with minimum<br />

and maximum values <strong>of</strong> 0.59 to 0.68. Mean λ derived<br />

from DHP <strong>for</strong> intermediate and late <strong>for</strong>est stages are<br />

0.67 and 0.68 respectively and thus lie on <strong>the</strong> upper<br />

end <strong>of</strong> that range. For early <strong>for</strong>est stages, however,<br />

λ derived from DHP seems to be unreasonably<br />

low. This was also observed in Chapter 5.4, when<br />

LAI <strong>of</strong> early <strong>for</strong>est stages were discussed. This<br />

(DHP)<br />

was due to relatively low λ values derived from in<br />

situ measurements: <strong>for</strong> Budongo and Kakamega<br />

Forests λ values <strong>of</strong> 0.51 and 0.52 were determined.<br />

The corresponding λ given by Chen et al. (2005) are<br />

however much higher, with 0.74 on average (0.64 to<br />

0.83) <strong>for</strong> herbaceous cover, and 0.73 (0.62 to 0.80) <strong>for</strong><br />

deciduous open to closed shrub cover. The apparent<br />

exaggeration <strong>of</strong> LAI due to <strong>the</strong> low λ derived<br />

(DHP)<br />

from DHP <strong>for</strong> early <strong>for</strong>est stages probably weakened<br />

<strong>the</strong> per<strong>for</strong>mance <strong>of</strong> regression models <strong>based</strong> on this<br />

parameter (cf. Chapter 6.3). λ values derived <strong>for</strong><br />

Kakamega Forest are also not within <strong>the</strong> range given<br />

by Chen et al. (2005) <strong>for</strong> early <strong>for</strong>est stages: <strong>for</strong> <strong>the</strong><br />

latter consequently no correction <strong>of</strong> foliage clumping<br />

can be applied.<br />

7.3<br />

7.3.1<br />

Results<br />

Budongo Forest<br />

Ideally <strong>the</strong> <strong>MODIS</strong> LAI algorithm should produce<br />

data that is accurate at <strong>the</strong> biome or regional level.<br />

Fur<strong>the</strong>r it should respond correctly to biomelevel<br />

LAI trajectories associated with interannual<br />

climate variability and be sensitive to meaningful<br />

perturbations to LAI, such as from significant<br />

disturbances (Cohen et al. 2006). Whereas <strong>the</strong> two<br />

latter objectives can only be confirmed with time<br />

series data, <strong>the</strong> spatial validation <strong>of</strong> product accuracy<br />

at <strong>the</strong> regional level requires a thorough analysis <strong>of</strong><br />

algorithm input and output. Since <strong>the</strong> <strong>MODIS</strong> LAI<br />

algorithm is <strong>based</strong> on daily surface reflectance data,<br />

in<strong>for</strong>mation on cloud and land cover as well as biome<br />

specific LUTs (Figure 4-2) errors in <strong>the</strong>se data sets,<br />

i.e. uncertainties in input data and uncertainties in<br />

<strong>the</strong> applied model, obviously influence <strong>the</strong> quality <strong>of</strong><br />

<strong>the</strong> resulting LAI product. For that reason <strong>the</strong> quality<br />

<strong>of</strong> input data will be analysed in a first step be<strong>for</strong>e<br />

<strong>the</strong> resulting algorithm output is validated in terms<br />

<strong>of</strong> accuracy in comparison to <strong>the</strong> up-scaled high<br />

resolution LAI maps. Note that all overview figures<br />

e<br />

in this chapter refer to <strong>the</strong> area shown in Figure<br />

2-2, <strong>the</strong> final product comparison (cf. Figure 7-5) is<br />

displayed <strong>for</strong> <strong>the</strong> extent <strong>of</strong> Figure 6-16).<br />

Spatial validation<br />

For <strong>the</strong> generation <strong>of</strong> <strong>the</strong> 8-day LAI composites<br />

<strong>the</strong>re is no in<strong>for</strong>mation determining which surface<br />

reflectance data sets are used <strong>for</strong> product generation.<br />

Consequently acquisition date and viewing geometry<br />

remain unclear. As <strong>the</strong> quality flags <strong>of</strong> <strong>the</strong> input<br />

reflectance data are, however, passed through from<br />

<strong>the</strong> original data sets to <strong>the</strong> MOD15A2 product (cf.<br />

Table A-2), <strong>the</strong> quality <strong>of</strong> input data can be judged<br />

indirectly.<br />

Table 7-1 Comparison <strong>of</strong> λ derived from DHP <strong>for</strong> <strong>for</strong>est stages <strong>of</strong> <strong>the</strong> respective study sites with λ given in Chen et al. (2005) <strong>for</strong><br />

<strong>the</strong> classes “deciduous open to closed shrub cover” and “evergreen broadleaf <strong>for</strong>est”.<br />

Data source λ (early stage) λ (intermediate stage) λ (late stage)<br />

In situ data, Budongo Forest 0.51 0.67 0.68<br />

In situ data, Kakamega Forest 0.52 0.64 0.65<br />

Chen et al. (2005) 0.73 (min. 0.62, max. 0.80) 0.63 (min. 0.59, max. 0.68)


7 <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

As <strong>for</strong> high resolution optical data, <strong>the</strong> main quality<br />

constraint <strong>for</strong> <strong>the</strong> MOD15A2 product is cloud cover.<br />

Although <strong>the</strong> product is generated as an 8-day<br />

composite, cloud contamination and subsequent<br />

errors in algorithm results are still major obstacles.<br />

Figure 7-1 summarizes <strong>the</strong> quality in<strong>for</strong>mation <strong>of</strong><br />

<strong>MODIS</strong> LAI products <strong>for</strong> <strong>the</strong> Budongo Forest region<br />

with respect to cloud cover and aerosols. Data sets<br />

<strong>of</strong> Julian Days (JD) 273 to 353 (30 September to 19<br />

December, 2005) served as input, which corresponds<br />

to all <strong>MODIS</strong> LAI data sets acquired during <strong>the</strong> time<br />

<strong>of</strong> field work as well as two time steps be<strong>for</strong>e and<br />

after. Black pixels resemble areas, where no clouds,<br />

cirrus clouds or aerosols could be detected at any<br />

time step. Red pixels are ei<strong>the</strong>r water (Lake Albert)<br />

Figure 7-1<br />

Figure 7-2<br />

Quality in<strong>for</strong>mation on cloud cover, cirrus clouds<br />

and aerosols <strong>of</strong> <strong>the</strong> MOD15A2 data sets<br />

JD 273-353 2005.<br />

Percentage <strong>of</strong> invalid (i.e. cloud contaminated) pixels <strong>for</strong> JD 273 to 353 in 2005.<br />

129<br />

or pixels where clouds, cirrus clouds or aerosols were<br />

detected at each <strong>of</strong> <strong>the</strong> 11 time steps. It is obvious<br />

that <strong>the</strong> <strong>for</strong>ested region in <strong>the</strong> middle is more prone<br />

to cloud contamination than <strong>the</strong> surrounding areas.<br />

This is understandable, as <strong>the</strong> aerial content <strong>of</strong> water<br />

vapour is higher above <strong>for</strong>ested regions. Local cloud<br />

convection and rainfall over those areas could also be<br />

observed during field work.<br />

Figure 7-2 shows <strong>the</strong> same in<strong>for</strong>mation as Figure 7-1,<br />

but plotted temporally <strong>for</strong> <strong>the</strong> observed time steps.<br />

Obviously <strong>the</strong> data sets <strong>of</strong> JD 305, 313, 321 and 337<br />

had <strong>the</strong> best data quality, each with less than 40%<br />

invalid pixels (caused by cloud cover/aerosols). This<br />

value seems to be quite large, but needs to be seen<br />

in relation to <strong>the</strong> water pixels <strong>of</strong> Lake Albert. These<br />

water surfaces make up approximately 12% <strong>of</strong> <strong>the</strong><br />

image extent and are invalid per definitionem.<br />

The second aspect influencing product quality is<br />

<strong>the</strong> accuracy <strong>of</strong> <strong>the</strong> applied land cover map. As <strong>the</strong><br />

8-biome map used in <strong>the</strong> <strong>MODIS</strong> C5 algorithm is<br />

not available to <strong>the</strong> public, level 3 <strong>of</strong> <strong>the</strong> MOD12Q1<br />

land cover product – as used to produce C4 LAI data<br />

– is displayed instead in Figure 7-3. Theoretically<br />

<strong>the</strong> only difference between both biome maps is <strong>the</strong><br />

distinction between evergreen and deciduous <strong>for</strong> <strong>the</strong><br />

classes “broadleaf <strong>for</strong>est” and “needleleaf <strong>for</strong>est”.<br />

As shown in Figure 7-3, <strong>the</strong>re is variation in <strong>for</strong>est<br />

area <strong>of</strong> <strong>the</strong> MOD12Q1 land cover product over


130<br />

time, which is evident since <strong>the</strong> MOD12Q1 product<br />

is produced on an annual basis. From 2001 on, <strong>the</strong><br />

percentage <strong>of</strong> pixels classified as broadleaf <strong>for</strong>est<br />

increases from 13.0% to 16.9% in 2002 and to<br />

21.7% in 2003. For 2004 a slight decrease to 20.9%<br />

can be observed. A visual assessment revealed that<br />

correspondence between <strong>the</strong> <strong>for</strong>ested areas in ASTER<br />

data and <strong>the</strong> MOD12Q1 product was best <strong>for</strong> <strong>the</strong> year<br />

2001. Whereas <strong>the</strong> extent <strong>of</strong> <strong>the</strong> main <strong>for</strong>est block is<br />

covered well in all MOD12Q1 land cover data sets,<br />

misclassifications <strong>of</strong> <strong>for</strong>est area are mainly visible<br />

to <strong>the</strong> south and west <strong>of</strong> <strong>the</strong> main <strong>for</strong>est block in <strong>the</strong><br />

2001<br />

2003<br />

<strong>MODIS</strong> Land Cover (MOD12Q1, Type 3)<br />

Figure 7-3<br />

water<br />

savanna shrubs<br />

2002<br />

2004<br />

broadleaf crops broadleaf <strong>for</strong>est unvegetated<br />

Type 3 <strong>of</strong> MOD12Q1 data sets 2001-2004. An update is not available.<br />

years 2002-2004. These areas are densely populated<br />

and remaining <strong>for</strong>est patches hardly exist in reality.<br />

Consequently <strong>the</strong>y should ra<strong>the</strong>r be mapped as<br />

“broadleaf crops”. Since an update <strong>of</strong> <strong>the</strong> 2004 land<br />

cover map has not been available, 2004 land cover<br />

data has been used also <strong>for</strong> LAI calculation from<br />

2005 onward.<br />

Since in <strong>the</strong> <strong>MODIS</strong> LAI algorithm biome specific<br />

LUT are used <strong>for</strong> <strong>the</strong> comparison <strong>of</strong> observed with<br />

modelled BRFs (cf. Chapter 4.3.3), misclassification<br />

<strong>of</strong> land cover results in <strong>the</strong> application <strong>of</strong> <strong>the</strong> wrong<br />

grasses/cereal crops


7 <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

LUT to a specific pixel. Consequently resulting LAI<br />

values rely on erroneous assumptions. Whereas this<br />

does not affect <strong>the</strong> outcomes <strong>of</strong> this <strong>the</strong>sis, as <strong>the</strong>se<br />

focus on <strong>the</strong> correctly classified <strong>for</strong>est area only,<br />

this aspect must be taken into account in subsequent<br />

regional applications.<br />

Also <strong>the</strong> applied algorithm has a major influence on<br />

product quality. As Shabanov et al. (2005) stated, <strong>the</strong><br />

C4 LAI algorithm suffered from <strong>the</strong> dominance <strong>of</strong><br />

poor-quality backup algorithm retrievals over woody<br />

vegetation. According to Yang et al. (2006) this was<br />

due to a mismatch between modelled and observed<br />

surface reflectances resulting in a failure <strong>of</strong> <strong>the</strong> main<br />

algorithm (cf. Equation 4.2). The backup algorithm,<br />

which is <strong>based</strong> on NDVI, is apparently consistent<br />

with <strong>the</strong> main algorithm at low NDVI. For high<br />

NDVI values (0.82-1) however, <strong>the</strong> backup algorithm<br />

is out <strong>of</strong> <strong>the</strong> retrieval domain <strong>of</strong> <strong>the</strong> main algorithm<br />

and reports single LAI values <strong>of</strong> 6.1. In C5 this issue<br />

seems to be solved, as consistency between simulated<br />

and measured <strong>MODIS</strong> surface reflectances is ensured<br />

(NASA 2007b).<br />

JD 289 (C5)<br />

JD 289 (C5)<br />

Applied LAI algorithm (MOD15A2)<br />

Figure 7-4<br />

physical perfect<br />

Quality in<strong>for</strong>mation on <strong>the</strong> applied <strong>MODIS</strong> LAI algorithm (C5 data) <strong>for</strong> JD 289 and 329.<br />

131<br />

Figure 7-4 shows <strong>the</strong> algorithm per<strong>for</strong>mance <strong>for</strong> <strong>the</strong><br />

Budongo Forest study site <strong>for</strong> two time steps: JD 289<br />

and 329. It is obvious that <strong>for</strong> both 8-day composites<br />

<strong>the</strong> derivation <strong>of</strong> LAI <strong>based</strong> on <strong>the</strong> “physical<br />

saturated” algorithm prevails <strong>for</strong> <strong>the</strong> <strong>for</strong>ested regions.<br />

“Physical saturated” means that LAI is derived <strong>based</strong><br />

on <strong>the</strong> physical model (i.e. <strong>the</strong> main algorithm),<br />

but under saturated conditions (i.e. increasing LAI<br />

influences less modelled surface reflectances). LAI <strong>of</strong><br />

areas surrounding <strong>the</strong> <strong>for</strong>est is in turn mainly retrieved<br />

through <strong>the</strong> main algorithm without saturation<br />

(“physical perfect”). The major difference between<br />

both time steps is <strong>the</strong> amount <strong>of</strong> pixels calculated<br />

using <strong>the</strong> empirical method. Whereas <strong>the</strong> NDVI<strong>based</strong><br />

backup algorithm is used <strong>for</strong> 8% <strong>of</strong> pixels on<br />

JD 289 (16 October 2005), hardly any retrieval can<br />

be found <strong>for</strong> JD 329 (25 November 2005). Generally,<br />

overall data quality (no clouds, aerosols or shadow<br />

detected) was best on JD 329, 337, and 361.<br />

Finally, algorithm outputs must be validated. After<br />

reprojection to <strong>the</strong> Sinusoidal Projection, <strong>the</strong> high<br />

resolution LAI map presented in Figure 7-5a (cf.<br />

e<br />

also Chapter 6.3.4) is aggregated to 1 km spatial<br />

resolution (Figure 7-5b). Subsequently clumping<br />

JD 329 (C5)<br />

JD 329 (C5)<br />

physical saturated empirical not retrieved


132<br />

factors as noted in Table 7-1 are applied to <strong>the</strong><br />

different <strong>for</strong>est stages (Figure 7-5c). Due to <strong>the</strong><br />

above described quality analyses, <strong>the</strong> <strong>MODIS</strong> LAI<br />

data set <strong>of</strong> JD 329 (Figure 7-5d) is chosen <strong>for</strong> spatial<br />

validation as it had good data quality and was close to<br />

<strong>the</strong> main time <strong>of</strong> field work. As already noted, areas<br />

in <strong>the</strong> north and south west <strong>of</strong> Budongo Forest, where<br />

LAI is apparently underestimated, are excluded from<br />

e<br />

validation. In addition LAI from early <strong>for</strong>est stages<br />

LAIe, 15 m (ASTER)<br />

LAIe, 15 m (ASTER)<br />

LAI, 1000 m (ASTER)<br />

LAI, 1000 m (ASTER)<br />

(effective) LAI<br />

is not taken into account, as λ derived from DHP is<br />

apparently underestimated.<br />

Frequency distributions <strong>for</strong> <strong>the</strong> <strong>for</strong>est areas <strong>of</strong> <strong>the</strong><br />

four (effective) LAI products presented in Figure<br />

7-5 are displayed in Figure 7-6 toge<strong>the</strong>r with <strong>the</strong><br />

respective normal distributions. Whereas <strong>the</strong> original<br />

high resolution LAI map (Figure 7-6a) is clearly<br />

e<br />

following a normal distribution, deviation is obvious<br />

< 1 1-2 2-3 3-4 4-5 5-6 6-7 7-8<br />

8-9 >9 no data<br />

LAIe, 1000 m (ASTER)<br />

LAIe, 1000 m (ASTER)<br />

LAI, 1000 m (<strong>MODIS</strong>)<br />

LAI, 1000 m (<strong>MODIS</strong>)<br />

Figure 7-5 (Effective) LAI maps <strong>for</strong> <strong>the</strong> Budongo Forest test site. a) High resolution LAI map derived from ASTER data,<br />

e<br />

b) same map aggregated to 1000 m spatial resolution, c) clumping correction applied to b) and d) <strong>MODIS</strong> LAI<br />

product <strong>of</strong> JD 329.


7 <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

<strong>for</strong> <strong>the</strong> aggregated high resolution LAI e map (Figure<br />

7-6b). Yet <strong>the</strong> low LAI e at <strong>the</strong> <strong>for</strong>est edges are ra<strong>the</strong>r<br />

an artefact <strong>of</strong> <strong>the</strong> aggregation process. The effect is <strong>the</strong><br />

same <strong>for</strong> <strong>the</strong> clumping corrected LAI shown in Figure<br />

7-6c as it is <strong>based</strong> on <strong>the</strong> aggregated high resolution<br />

LAI map. For <strong>the</strong> following analyses, and to avoid<br />

e<br />

errors introduced by different spatial resolution,<br />

<strong>for</strong>est areas with a distance <strong>of</strong> less than 1000 m to <strong>the</strong><br />

<strong>for</strong>est border are consequently not taken into account.<br />

In contrast to Figures 7-6a to c, <strong>MODIS</strong> LAI data (cf.<br />

Figure 7-6d) shows two local maxima in its frequency<br />

distribution. Although minimum values and mean<br />

values are comparable to <strong>the</strong> original high resolution<br />

LAI map, <strong>the</strong>re is a clear dominance <strong>of</strong> LAI values<br />

e<br />

Figure 7-6 Frequency distribution <strong>of</strong> (effective) LAI values derived from a) ASTER LAIe<br />

(15 m spatial resolution), b) ASTER LAIe (1000 m), c) ASTER LAI (1000 m) and d) <strong>MODIS</strong> LAI (1000 m) data sets. For each histogram <strong>the</strong> normal distribution<br />

is indicated.<br />

133<br />

between 5.5 and 6.0 LAI derived from ASTER is in<br />

e<br />

turn more continuous. Interestingly, <strong>the</strong> <strong>MODIS</strong> LAI<br />

product does not capture higher LAI values (>6) in<br />

<strong>the</strong> centre <strong>of</strong> <strong>the</strong> <strong>for</strong>est, but on <strong>the</strong> o<strong>the</strong>r hand records<br />

higher LAI values <strong>for</strong> <strong>the</strong> Cynometra-dominated<br />

areas in <strong>the</strong> southwest <strong>of</strong> Budongo.<br />

Table 7-2 displays mean (effective) LAI and standard<br />

deviation averaged <strong>for</strong> <strong>the</strong> different <strong>for</strong>est stages.<br />

As expected, mean LAI values per <strong>for</strong>est stage <strong>of</strong><br />

e<br />

original and aggregated high resolution map are<br />

almost identical, with standard deviations being<br />

slightly higher <strong>for</strong> <strong>the</strong> 15 m spatial resolution map.<br />

In turn, <strong>MODIS</strong> LAI values underestimate ASTER


134<br />

derived LAI by ~3 <strong>for</strong> intermediate and late <strong>for</strong>est<br />

stages. Obviously <strong>MODIS</strong> LAI is closer to ASTER<br />

derived LAI than to LAI, which is corrected <strong>for</strong><br />

e<br />

clumping.<br />

For a more detailed comparison <strong>of</strong> <strong>the</strong> different LAI<br />

products a patch-by-patch comparison is per<strong>for</strong>med.<br />

Mean (effective) LAI and standard deviation are<br />

computed <strong>for</strong> each <strong>for</strong>est compartment included in<br />

<strong>the</strong> analysis (N=48). With regards to <strong>the</strong> polygons<br />

lying at <strong>for</strong>est edges only those areas were included<br />

that were more than 1000 m away from <strong>the</strong> <strong>for</strong>est<br />

border. Figure A-9 shows that mean LAI (ASTER)<br />

e<br />

in 15 m and 1000 m spatial resolution derived <strong>for</strong> <strong>the</strong><br />

respective polygons vary slightly due to <strong>the</strong> different<br />

scales. Consequently only LAI products aggregated<br />

to 1000 m are included in <strong>the</strong> validation process.<br />

Table 7-2<br />

Figure 7-7 shows <strong>the</strong> resulting scatter plots <strong>of</strong><br />

regressions between <strong>the</strong> <strong>MODIS</strong> LAI product <strong>of</strong> JD<br />

329 and ASTER derived LAI and LAI, respectively.<br />

e<br />

Obviously <strong>MODIS</strong> LAI is closer to LAI than LAI<br />

e<br />

(dashed lines represent <strong>the</strong> 1:1 relationship). Whereas<br />

<strong>the</strong> relationship between <strong>MODIS</strong> LAI and ASTER<br />

LAI is characterized by an accuracy <strong>of</strong> 0.53 (relative<br />

e<br />

accuracy <strong>of</strong> 9%, cf. Figure 7-7a), ASTER LAI is<br />

poorly represented by <strong>MODIS</strong> LAI with an RMSE <strong>of</strong><br />

3.3 (relative accuracy <strong>of</strong> 60%, cf. Figure 7-7b). This<br />

can mainly be attributed to <strong>the</strong> large bias <strong>of</strong> <strong>the</strong> latter,<br />

which is -3.28 with respect to ASTER LAI. Although<br />

<strong>MODIS</strong> LAI also underestimates LAI , <strong>the</strong> respective<br />

e<br />

bias is much lower at -0.35 (corresponding to relative<br />

biases <strong>of</strong> -38% and -6% respectively). The fact that<br />

both regression models yield an R2 <strong>of</strong> 0.64 puts <strong>the</strong><br />

significance <strong>of</strong> this measure into perspective.<br />

Mean (and standard deviation) <strong>of</strong> (effective) LAI derived from ASTER and <strong>MODIS</strong> <strong>for</strong> different <strong>for</strong>est stages.<br />

<strong>Product</strong> Early stage Intermediate stage Late stage<br />

LAI e , 15 m (ASTER) 4.6 (±1.5) 6.1 (±0.9) 5.9 (±0.9)<br />

LAI e , 1000 m (ASTER) 4.6 (±0.9) 6.0 (±0.7) 5.9 (±0.4)<br />

LAI, 1000 m (ASTER) n/a 8.8 (±1.5) 8.7 (±0.6)<br />

LAI, 1000 m (<strong>MODIS</strong>) 4.4 (±1.6) 5.8 (±0.2) 5.6 (±0.2)<br />

OLS: R²=0.64<br />

Y=-0.577+1.032 X<br />

OLS: R²=0.64<br />

Y=-0.577+0.691 X<br />

Figure 7-7 Scatter plots showing <strong>the</strong> regression between a) ASTER LAI (1000 m) and <strong>the</strong> C5 <strong>MODIS</strong> LAI product and<br />

e<br />

b) ASTER LAI (1000 m) and <strong>the</strong> <strong>MODIS</strong> LAI product <strong>for</strong> those <strong>for</strong>est compartments included in <strong>the</strong> analysis (N=48).<br />

The dashed line represents <strong>the</strong> 1:1 relationship, <strong>the</strong> solid line OLS regression.


7 <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

The <strong>MODIS</strong> LAI product thus represents LAI e <strong>of</strong><br />

Budongo Forest with a relative accuracy <strong>of</strong> 9%. This<br />

corresponds to <strong>the</strong> accuracy <strong>of</strong> <strong>the</strong> high resolution<br />

LAI map and is only a slight degradation in accuracy<br />

e<br />

compared to in situ measurements and satellite<br />

observations.<br />

Although a validation <strong>of</strong> <strong>the</strong> (outdated) C4 <strong>MODIS</strong><br />

LAI algorithm was not per<strong>for</strong>med, quality in<strong>for</strong>mation<br />

on <strong>the</strong> applied algorithm as well as <strong>the</strong> final LAI<br />

product <strong>for</strong> JD 329 (2005) is shown <strong>for</strong> comparison in<br />

Figure 7-8. The main difference is a higher retrieval<br />

rate <strong>of</strong> <strong>the</strong> physical perfect algorithm, but a resulting<br />

lower LAI quality. The <strong>for</strong>ested area appears to be<br />

less well outlined, with some unreasonably low LAI<br />

values resulting from <strong>the</strong> physical perfect algorithm<br />

QC, JD 329 (C4)<br />

QC, JD 329 (C4)<br />

Figure 7-8<br />

Table 7-3<br />

LAI, 1000 m (<strong>MODIS</strong>)<br />

LAI, 1000 m (<strong>MODIS</strong>)<br />

Quality in<strong>for</strong>mation on <strong>the</strong> applied <strong>MODIS</strong> LAI algorithm (C4 data) <strong>for</strong> JD 329 and <strong>the</strong> respective <strong>MODIS</strong> LAI product.<br />

For legends please refer to Figures 7-4 and 7-5.<br />

Relative frequency <strong>of</strong> C4 and C5 LAI algorithm usage <strong>for</strong> <strong>the</strong> Budongo Forest test site.<br />

Collection Applied algorithm 2000 2001 2002 2003 2004 2005 2006 Average<br />

C4 Physical perfect 37% 36% 46% 37% 31% 47% 45% 40%<br />

Physical saturated 17% 17% 22% 25% 15% 21% 18% 19%<br />

Empirical 26% 31% 29% 32% 28% 23% 34% 29%<br />

Not retrieved 22% 16% 3% 6% 26% 9% 2% 12%<br />

C5 Physical perfect 22% 25% 26% 29% 28% 23% n/a 24%<br />

Physical saturated 44% 50% 53% 51% 55% 65% n/a 50%<br />

Empirical 14% 20% 20% 17% 16% 12% n/a 17%<br />

Not retrieved 20% 4% 2% 7% 0% 17% n/a 9%<br />

135<br />

within <strong>the</strong> main <strong>for</strong>est area. Retrievals from <strong>the</strong><br />

physical saturated algorithm on <strong>the</strong> o<strong>the</strong>r hand<br />

capture mean LAI <strong>of</strong> <strong>the</strong> <strong>for</strong>est area quite well with<br />

LAI values comparable to those <strong>of</strong> <strong>the</strong> high resolution<br />

LAI map. e<br />

Temporal consistency<br />

In order to evaluate <strong>the</strong> temporal consistency <strong>of</strong><br />

<strong>MODIS</strong> LAI data, <strong>the</strong> generated time series are<br />

fur<strong>the</strong>r analysed with respect to data quality. A<br />

comparison between <strong>MODIS</strong> LAI products <strong>of</strong> C4 (cf.<br />

Figure A-10) and C5 (cf. Figure 7-9) revealed major<br />

differences. Note that only <strong>the</strong> <strong>for</strong>est area (without<br />

surroundings) was included in this analysis.


136<br />

2004 2005<br />

Julian Day<br />

Figure 7-9<br />

1<br />

33<br />

65<br />

97<br />

129<br />

161<br />

193<br />

225<br />

257<br />

289<br />

321<br />

353<br />

17<br />

49<br />

81<br />

113<br />

145<br />

177<br />

209<br />

241<br />

273<br />

305<br />

337<br />

0<br />

10<br />

20<br />

30<br />

40<br />

not retrieved<br />

empirical<br />

physical saturated<br />

physical perfect<br />

relative frequency in [%]<br />

50<br />

60<br />

70<br />

80<br />

90<br />

Relative frequencies <strong>of</strong> C5 <strong>MODIS</strong> LAI algorithm<br />

usage <strong>for</strong> <strong>the</strong> years 2000-2005 over <strong>the</strong> Budongo<br />

Forest test site.<br />

100<br />

2000 2001 2002 2003<br />

Julian Day<br />

1<br />

33<br />

65<br />

97<br />

129<br />

161<br />

193<br />

225<br />

257<br />

289<br />

321<br />

353<br />

17<br />

49<br />

81<br />

113<br />

145<br />

177<br />

209<br />

241<br />

273<br />

305<br />

337<br />

1<br />

33<br />

65<br />

97<br />

129<br />

161<br />

193<br />

225<br />

257<br />

289<br />

321<br />

353<br />

17<br />

49<br />

81<br />

113<br />

145<br />

177<br />

209<br />

241<br />

273<br />

305<br />

337<br />

0<br />

10<br />

20<br />

relative frequency in [%]<br />

30<br />

40<br />

50<br />

60<br />

70<br />

80<br />

90<br />

100


7 <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

Due to algorithm changes <strong>the</strong> percentage <strong>of</strong> pixels<br />

with <strong>the</strong> highest algorithm quality “physical perfect”<br />

is clearly reduced from 40% in C4 to 24% in C5 on<br />

average (cf. Table 7-3). In contrast <strong>the</strong> amount <strong>of</strong><br />

pixels retrieved with <strong>the</strong> physical algorithm under<br />

supposedly saturated conditions increases from 19%<br />

to 50% (cf. Figure 7-4). As this algorithm seems to<br />

per<strong>for</strong>m better over <strong>for</strong>ested areas than <strong>the</strong> physical<br />

perfect method, this is a clear increase in quality. Also<br />

<strong>the</strong> fact that in C5 fewer pixels are retrieved with <strong>the</strong><br />

empirical algorithm or not retrieved at all is resulting<br />

in higher overall quality <strong>of</strong> <strong>the</strong> product. Interestingly,<br />

<strong>the</strong> data sets that could not be retrieved at all differ in<br />

C4 and C5 with respect to JD (i.e. <strong>the</strong> data set <strong>of</strong> JD<br />

1 in 2001 could not be retrieved at all in C4, but in<br />

C5 it is available). This can only be due to processing<br />

problems, not input data quality.<br />

As expected <strong>the</strong>re is also a dynamic in <strong>the</strong> quality<br />

data corresponding to <strong>the</strong> annual time course: <strong>for</strong><br />

both C4 and C5 data <strong>the</strong> highest data quality is<br />

usually retrieved in <strong>the</strong> dry season (beginning and<br />

end <strong>of</strong> year). With respect to C5 data, in some time<br />

steps all pixels could be retrieved using <strong>the</strong> physical<br />

method (ei<strong>the</strong>r under saturation or not). During <strong>the</strong><br />

rainy seasons, <strong>the</strong> amount <strong>of</strong> pixels processed using<br />

<strong>the</strong> empirical method clearly increases.<br />

Figure 7-10<br />

Mean <strong>MODIS</strong> LAI retrieved <strong>for</strong> a 5x5 pixel window in <strong>the</strong> centre <strong>of</strong> <strong>the</strong> Budongo Forest test site <strong>for</strong> <strong>the</strong> year 2005. Red<br />

indicates original and blue interpolated data, solid lines represent mean LAI, dashed lines respective standard<br />

deviations. Note that slight deviations between original and interpolated LAI <strong>for</strong> a certain time step are due to<br />

averaging over <strong>the</strong> 5x5 window.<br />

137<br />

However, <strong>the</strong> applied algorithm has a strong effect<br />

on <strong>the</strong> resulting LAI and thus on data quality. Figure<br />

7-10 displays an annual LAI trajectory <strong>of</strong> <strong>the</strong> year<br />

2005 derived from <strong>MODIS</strong> data <strong>for</strong> a 5x5 pixel area<br />

in <strong>the</strong> centre <strong>of</strong> Budongo Forest. Red lines indicate<br />

<strong>the</strong> original data, with <strong>the</strong> mean displayed as a solid<br />

and <strong>the</strong> standard deviations displayed as dashed lines.<br />

Over <strong>the</strong> year, several extreme minima are visible<br />

<strong>for</strong> different time steps that correspond to aerosol or<br />

could detection over <strong>the</strong> <strong>for</strong>est areas and mostly to <strong>the</strong><br />

subsequent usage <strong>of</strong> <strong>the</strong> empirical algorithm. In both<br />

cases unrealistically low LAI values are produced.<br />

It becomes quite clear that reliable LAI values can<br />

only be derived <strong>for</strong> very few time steps annually.<br />

Though seasonal variation in LAI exists in tropical<br />

rain <strong>for</strong>ests, this biophysical parameter is more stable<br />

than in o<strong>the</strong>r ecosystems. Consequently temporal<br />

interpolation may be applied. For <strong>the</strong> analysis, all<br />

pixels with bad data quality (i.e. processed with<br />

<strong>the</strong> empirical algorithm or not processed at all, and<br />

those in which aerosols, clouds or cirrus clouds were<br />

detected) are excluded. To fill <strong>the</strong> data gaps, linear<br />

interpolation is per<strong>for</strong>med with TiSeG (Colditz et<br />

al. 2008) as indicated in Figure 7-10 (blue line). The<br />

resulting LAI trajectory is much more stable, but is –<br />

in <strong>the</strong> worst case <strong>for</strong> this test site – only <strong>based</strong> on 7<br />

valid (out <strong>of</strong> 46) time steps.


138<br />

In order to evaluate <strong>the</strong> temporal signature <strong>of</strong> <strong>the</strong><br />

<strong>MODIS</strong> LAI product, Figure 7-11 shows <strong>the</strong> resulting<br />

LAI curves plotted toge<strong>the</strong>r with <strong>the</strong> respective<br />

standard deviations <strong>for</strong> <strong>the</strong> different <strong>for</strong>est stages<br />

and averaged over <strong>the</strong> years 2000-2005. Whereas<br />

a seasonal variation is clearly visible <strong>for</strong> <strong>the</strong> early<br />

<strong>for</strong>est stage with LAI maxima <strong>of</strong> up to 4.7 during <strong>the</strong><br />

rainy season and minima <strong>of</strong> 2.4 in <strong>the</strong> dry season,<br />

LAI curves <strong>for</strong> intermediate and late <strong>for</strong>est stages<br />

remain more stable over <strong>the</strong> year. For both a range <strong>of</strong><br />

0.4 could be observed with minima at <strong>the</strong> end <strong>of</strong> <strong>the</strong><br />

dry season and maxima at <strong>the</strong> end <strong>of</strong> <strong>the</strong> rainy season.<br />

It is fur<strong>the</strong>r obvious that mean LAI <strong>of</strong> intermediate<br />

and late <strong>for</strong>est stages are quite similar with respect to<br />

<strong>the</strong>ir large standard deviations (6.1±1.5 and 5.9±2.2),<br />

so that a distinction <strong>of</strong> <strong>for</strong>est stages <strong>based</strong> on LAI is<br />

not possible. This emphasises once more <strong>the</strong> results<br />

retrieved in Chapter 5.3.1.<br />

7.3.2<br />

Kakamega Forest<br />

For <strong>the</strong> spatial validation <strong>of</strong> <strong>MODIS</strong> LAI <strong>for</strong> <strong>the</strong><br />

Kakamega Forest test site <strong>the</strong> quality <strong>of</strong> input data to<br />

<strong>the</strong> <strong>MODIS</strong> LAI algorithm was, once again, analysed<br />

first. Here <strong>the</strong> overview figures refer to <strong>the</strong> area<br />

shown in Figure 2-11, <strong>the</strong> product comparison (cf.<br />

Figure 7-15) is given <strong>for</strong> <strong>the</strong> subset <strong>of</strong> Figure 6-20.<br />

Figure 7-12 displays <strong>the</strong> quality in<strong>for</strong>mation <strong>of</strong><br />

C5 <strong>MODIS</strong> LAI <strong>of</strong> JD 281-361 (8 October to<br />

27 December) <strong>of</strong> <strong>the</strong> year 2004 with respect to<br />

cloud cover. Although <strong>the</strong> time period taken into<br />

consideration is almost <strong>the</strong> same <strong>for</strong> both test sites,<br />

<strong>the</strong> overall data quality here is slightly better than<br />

<strong>for</strong> Budongo Forest. This might be attributed to <strong>the</strong><br />

fact that Budongo Forest receives 30% more rainfall<br />

on average during November and thus probably<br />

experiences higher cloud cover (cf. Figures 2-3 and<br />

2-13). Similar to Budongo Forest, <strong>the</strong> <strong>for</strong>ested areas<br />

in Kakamega Forest are more prone to cloud cover<br />

than surrounding areas.<br />

Figure 7-11<br />

Figure 7-12<br />

Annual <strong>MODIS</strong> LAI trajectories <strong>for</strong> <strong>the</strong> different<br />

<strong>for</strong>est stages <strong>of</strong> Budongo Forest, averaged over<br />

<strong>the</strong> years 2000-2005. Light green indicates early,<br />

dark green late <strong>for</strong>est stages. Means are plotted<br />

as solid lines, standard deviations as dashed lines.<br />

Quality in<strong>for</strong>mation <strong>of</strong> <strong>the</strong> MOD15A2 data sets<br />

JD 281-361 2004. Invalid pixels are ei<strong>the</strong>r<br />

processed with <strong>the</strong> empirical algorithm or clouds/<br />

aerosols are detected.<br />

The better data quality is also obvious in Figure 7-13.<br />

Only JD 297 has more than 40% invalid pixels. Time<br />

steps 281, 329 and 361 show <strong>the</strong> best data quality<br />

corresponding to <strong>the</strong> least cloud coverage. Since <strong>the</strong><br />

data set <strong>of</strong> JD 281 was acquired shortly be<strong>for</strong>e <strong>the</strong><br />

beginning <strong>of</strong> field work in Kakamega Forest at <strong>the</strong><br />

end <strong>of</strong> <strong>the</strong> rainy season, it was chosen <strong>for</strong> product<br />

validation. At this time, LAI is assumed to be at its


7 <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

yearly peak. By contrast, a slight decrease in LAI<br />

is supposed to correspond to <strong>the</strong> dry season, which<br />

begins around <strong>the</strong> end <strong>of</strong> November.<br />

The class “broadleaf <strong>for</strong>est” <strong>of</strong> <strong>the</strong> MOD12Q1 land<br />

cover map (cf. Figure 7-14a) corresponds well to <strong>the</strong><br />

<strong>for</strong>ested areas in <strong>the</strong> SPOT image, with some minor<br />

misclassifications towards <strong>the</strong> west and south <strong>of</strong> <strong>the</strong><br />

main <strong>for</strong>est block. As <strong>for</strong> Budongo Forest, LAI <strong>of</strong><br />

Figure 7-13<br />

2004<br />

Percentage <strong>of</strong> invalid (i.e. cloud contaminated) pixels <strong>for</strong> JD 281 to 361 in 2004 <strong>for</strong> <strong>the</strong> Kakamega Forest test site.<br />

<strong>MODIS</strong> Land Cover (MOD12Q1, Type 3)<br />

Figure 7-14<br />

broadleaf <strong>for</strong>est<br />

broadleaf crops<br />

savanna<br />

139<br />

those areas was mainly retrieved with <strong>the</strong> physical<br />

saturated algorithm (cf. Figure 7-14b). Outside <strong>the</strong><br />

<strong>for</strong>est <strong>the</strong> physical perfect method prevails.<br />

LAI products are displayed in Figure 7-15 with <strong>the</strong><br />

high resolution LA map (<strong>based</strong> on SPOT data) and<br />

e<br />

its aggregated version above and <strong>the</strong> aggregated map<br />

corrected <strong>for</strong> foliage clumping as well as C5 <strong>MODIS</strong><br />

LAI data JD 281 (2004) below. Mean <strong>MODIS</strong> LAI<br />

JD 281 (C5)<br />

JD 281 (C5)<br />

Applied LAI algorithm (MOD15A2)<br />

physical perfect<br />

grasses/cereal crops unvegetated<br />

empirical<br />

physical saturated<br />

not retrieved<br />

a) Type 3 <strong>of</strong> <strong>MODIS</strong> land cover (MOD12Q1) <strong>of</strong> 2004 and b) <strong>MODIS</strong> LAI algorithm applied to <strong>the</strong> data set JD 281<br />

(2005) <strong>for</strong> <strong>the</strong> same area.


140<br />

effective LAI (SPOT), 20 m<br />

effective LAI (SPOT), 20 m<br />

LAI (SPOT), 1000 m<br />

LAI (SPOT), 1000 m<br />

(effective) LAI<br />

Figure 7-15 (Effective) LAI maps <strong>for</strong> <strong>the</strong> Kakamega Forest test site. a) High resolution LAI map derived from SPOT data, b) same<br />

e<br />

map aggregated to 1000 m spatial resolution, c) clumping correction applied to b) and d) <strong>MODIS</strong> LAI product <strong>of</strong><br />

JD 281 (2004).<br />

effective LAI (SPOT), 1000 m<br />

effective LAI (SPOT), 1000 m<br />

LAI (<strong>MODIS</strong>), 1000 m<br />

LAI (<strong>MODIS</strong>), 1000 m<br />

< 1 1-2 2-3 3-4 4-5 5-6 6-7 7-8<br />

8-9 >9 no data


7 <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

<strong>of</strong> <strong>the</strong> <strong>for</strong>est area is slightly higher than in Budongo<br />

Forest. Similar to Budongo Forest, <strong>for</strong>est areas with a<br />

distance <strong>of</strong> less than 1000 m to <strong>the</strong> <strong>for</strong>est border are<br />

not considered in <strong>the</strong> validation process.<br />

The frequency distributions <strong>of</strong> <strong>the</strong> maps are given<br />

in Figure 7-16 toge<strong>the</strong>r with <strong>the</strong>ir normal distributions.<br />

As in Budongo Forest, <strong>the</strong> original high<br />

resolution LAI map (Figure 7-15a) follows a normal<br />

e<br />

distribution, yet with some extremely high LAIe values on <strong>the</strong> upper end <strong>of</strong> <strong>the</strong> range. This is due<br />

to <strong>the</strong> a<strong>for</strong>ementioned plantation areas, where <strong>the</strong><br />

applied regression model is not valid (cf. Chapter<br />

6.3.2). These areas will be excluded from spatial<br />

Figure 7-16 Frequency distribution <strong>of</strong> (effective) LAI values derived from a) SPOT LAIe<br />

(20 m spatial resolution), b) SPOT LAIe (1000 m), c) SPOT LAI (1000 m) and d) <strong>MODIS</strong> LAI (1000 m) data sets. For each histogram <strong>the</strong> normal distribution<br />

is indicated.<br />

141<br />

validation. Additionally, a number <strong>of</strong> relatively low<br />

LAI values between 0 and 2 can be observed due to<br />

e<br />

<strong>for</strong>est glades and grasslands. The aggregated LAIe and <strong>the</strong> clumping corrected LAI maps generally show<br />

a similar frequency distribution. LAI derived from<br />

<strong>MODIS</strong> data, however, again has a relatively low<br />

variation <strong>of</strong> LAI values <strong>for</strong> <strong>the</strong> <strong>for</strong>ested areas and is<br />

concentrated between 5.5 and 7. LAI values between<br />

1 and 4 can also be observed <strong>for</strong> more fragmented<br />

areas and <strong>for</strong>est glades. For <strong>the</strong> following analyses<br />

only <strong>the</strong> area north <strong>of</strong> Isecheno Forest Station<br />

is included in order to exclude plantation areas.<br />

Additionally a buffer area <strong>of</strong> 1000 m to <strong>the</strong> <strong>for</strong>est<br />

border is not taken into account <strong>for</strong> fur<strong>the</strong>r analyses.


142<br />

Table 7-4 displays mean LAI (LAI e ) and standard<br />

deviations <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> <strong>the</strong> main <strong>for</strong>est<br />

block. Different <strong>for</strong>est stages are not distinguished<br />

here, as - in contrast to Budongo Forest - no larger<br />

areas (compartments) belonging to <strong>the</strong> same stage are<br />

present. Disturbances <strong>of</strong> near-natural <strong>for</strong>est are ra<strong>the</strong>r<br />

local, but <strong>the</strong>y are also dispersed over <strong>the</strong> whole<br />

<strong>for</strong>est, so that an upscaling <strong>of</strong> <strong>the</strong> existing 30 m land<br />

cover classification (cf. Figure A-2) to 1000 m spatial<br />

resolution would involve scaling errors.<br />

Once more mean LAI e values <strong>for</strong> <strong>the</strong> original and<br />

aggregated high resolution map are identical. Their<br />

standard deviations are comparable to those derived<br />

<strong>for</strong> Budongo Forest. Mean LAI is lower than in<br />

e<br />

Budongo as some <strong>for</strong>est glades with comparably low<br />

Table 7-4<br />

Mean (and standard deviation) <strong>of</strong> (effective)<br />

LAI derived from SPOT and <strong>MODIS</strong> <strong>for</strong> <strong>the</strong><br />

nor<strong>the</strong>rn part <strong>of</strong> Kakamega Forest.<br />

<strong>Product</strong> Mean (effective) LAI<br />

LAI e , 20 m (SPOT) 5.1 (±0.8)<br />

LAI e , 1000 m ( SPOT ) 5.1 (±0.8)<br />

LAI, 1000 m ( SPOT ) 8.0 (±1.3)<br />

LAI, 1000 m (<strong>MODIS</strong>) 6.4 (±0.3)<br />

OLS: R²=0.11<br />

Y=0.705-0.017 X<br />

LAI are included. In contrary to Budongo Forest,<br />

e<br />

<strong>MODIS</strong> LAI is higher than LAI . e<br />

In addition a patch-by-patch comparison is per<strong>for</strong>med<br />

<strong>for</strong> Kakamega Forest. Consequently 15 polygons<br />

are distributed randomly over <strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong><br />

Kakamega Forest and mean and standard deviation<br />

are computed <strong>for</strong> all (effective) LAI products at<br />

1000 m spatial resolution. Figure 7-17 shows <strong>the</strong><br />

resulting scatter plots <strong>of</strong> regressions between <strong>the</strong><br />

<strong>MODIS</strong> LAI product <strong>of</strong> JD 281 and SPOT derived<br />

LAI and LAI. As already indicated in Figure 7-16<br />

e<br />

and in <strong>the</strong> low standard deviation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI<br />

product given in Table 7-4, <strong>the</strong> variability <strong>of</strong> <strong>MODIS</strong><br />

LAI is very low compared to <strong>the</strong> SPOT derived<br />

products (dashed lines represent <strong>the</strong> 1:1 relationship).<br />

Consequently <strong>MODIS</strong> LAI is characterized by a<br />

comparatively low accuracy <strong>of</strong> 25% (RMSE <strong>of</strong> 1.5)<br />

with respect to SPOT LAI and an accuracy <strong>of</strong> 33%<br />

e<br />

(RMSE <strong>of</strong> 2.0) <strong>for</strong> SPOT LAI respectively. The biases<br />

are 1.2 and -1.6 (corresponding to relative biases <strong>of</strong><br />

18% and -26% respectively).<br />

The <strong>MODIS</strong> LAI product thus represents nei<strong>the</strong>r<br />

LAI nor LAI <strong>of</strong> Kakamega Forest with satisfactory<br />

e<br />

accuracy. Compared to <strong>the</strong> relative accuracy <strong>of</strong> 16%<br />

OLS: R²=0.11<br />

Y=0.705-0.011 X<br />

Figure 7-17 Scatter plots showing <strong>the</strong> regression between a) SPOT LAI (1000 m) and <strong>the</strong> C5 <strong>MODIS</strong> LAI product and<br />

e<br />

b) SPOT LAI (1000 m) and <strong>the</strong> <strong>MODIS</strong> LAI product <strong>for</strong> those <strong>for</strong>est compartments included in <strong>the</strong> analysis (N=15).<br />

The dashed line represents <strong>the</strong> 1:1 relationship.


7 <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

<strong>of</strong> <strong>the</strong> high resolution LAI e map presented in Chapter<br />

6.3.2, this is a fur<strong>the</strong>r degradation in accuracy that<br />

might partly be attributed to <strong>the</strong> bad in situ data<br />

quality.<br />

Figure 7-18 shows <strong>the</strong> resulting LAI curve plotted<br />

toge<strong>the</strong>r with its standard deviation <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn<br />

part <strong>of</strong> Kakamega Forest and averaged over<br />

<strong>the</strong> years 2000-2005. Once more <strong>the</strong> LAI curve<br />

remains relatively stable over <strong>the</strong> year with its<br />

standard deviation being slightly higher than that<br />

<strong>of</strong> intermediate and late <strong>for</strong>est stages in Budongo<br />

Forest. This is obviously <strong>the</strong> result <strong>of</strong> <strong>the</strong> missing<br />

subdivision in <strong>for</strong>est stages <strong>for</strong> Kakamega Forest.<br />

However mean LAI is comparable to that retrieved<br />

<strong>for</strong> Budongo Forest.<br />

Figure 7-18<br />

7.4<br />

Annual <strong>MODIS</strong> LAI trajectory <strong>for</strong> <strong>the</strong> nor<strong>the</strong>rn<br />

part <strong>of</strong> Kakamega Forest averaged over <strong>the</strong><br />

years 2000-2005. Mean LAI is plotted as solid<br />

line, standard deviations as dashed lines.<br />

Conclusion<br />

The preceding chapter completes <strong>the</strong> validation <strong>of</strong> <strong>the</strong><br />

<strong>MODIS</strong> LAI product <strong>for</strong> both Budongo and Kakamega<br />

Forests. Based on <strong>the</strong> analyses and correction <strong>of</strong> in situ<br />

data presented in Chapter 5 and <strong>the</strong> high resolution<br />

LAIe maps created in Chapter 6, two explicit <strong>MODIS</strong><br />

LAI products could be validated: JD 329 <strong>for</strong> <strong>the</strong> year<br />

143<br />

2005 <strong>for</strong> Budongo Forest and JD 281 <strong>for</strong> 2004 <strong>for</strong><br />

Kakamega Forest. The products were chosen due to<br />

<strong>the</strong>ir good input data quality in terms <strong>of</strong> cloud cover<br />

and algorithm per<strong>for</strong>mance. Obviously cloud cover<br />

has a major influence on <strong>the</strong> applied algorithm and<br />

on product quality as a consequence. This means<br />

that <strong>the</strong> derived product accuracies are only valid <strong>for</strong><br />

<strong>the</strong> considered time steps and quality levels. LAI <strong>of</strong><br />

<strong>for</strong>ested areas was mostly derived with <strong>the</strong> “physical<br />

saturated algorithm”, which gave a good per<strong>for</strong>mance<br />

over <strong>the</strong> observed broadleaf <strong>for</strong>ests. The new LUTs<br />

associated to <strong>the</strong> C5 LAI algorithm seem to have a<br />

good correspondence to observed surface reflectance,<br />

even though LAI retrievals over broadleaf <strong>for</strong>ests are<br />

mostly per<strong>for</strong>med over a small area in <strong>the</strong> spectral<br />

space (Shabanov et al. 2005). Yet it remains unclear,<br />

why <strong>the</strong> main algorithm is not always per<strong>for</strong>med in<br />

<strong>the</strong> saturation domain over broadleaf <strong>for</strong>est areas.<br />

Temporal LAI trajectories fur<strong>the</strong>r revealed that <strong>the</strong><br />

back-up algorithm has major problems in terms <strong>of</strong><br />

LAI stability. Sufficient stability can only be derived<br />

when pixels with deficient data quality are excluded<br />

and temporal interpolation is per<strong>for</strong>med.<br />

Overall <strong>the</strong> spatial validation <strong>for</strong> <strong>the</strong> Budongo<br />

Forest test site revealed a product accuracy <strong>of</strong> 0.53<br />

corresponding to a relative accuracy <strong>of</strong> 9%, which<br />

is exactly <strong>the</strong> same as <strong>the</strong> relative accuracy <strong>of</strong> <strong>the</strong><br />

resulting high resolution LAI map <strong>of</strong> Budongo<br />

e<br />

Forest. It can thus be concluded that <strong>MODIS</strong> LAI<br />

well represents (up-scaled) in situ LAI on a regional<br />

level. However, it must be emphasised that product<br />

accuracy was determined with respect to LAI and not<br />

e<br />

LAI. The latter is underestimated by <strong>MODIS</strong> LAI,<br />

with a bias <strong>of</strong> -3.28 with respect to ASTER LAI.<br />

For Kakamega Forest <strong>MODIS</strong> LAI product validation<br />

led to a comparatively low accuracy <strong>of</strong> 1.5<br />

with respect to LAI , which corresponds to a relative<br />

e<br />

accuracy <strong>of</strong> 25%. In this case <strong>MODIS</strong> LAI is higher<br />

than LAI derived from upscaled field data (bias <strong>of</strong><br />

e<br />

1.2) and lower than reference LAI (bias <strong>of</strong> -1.6).


144<br />

However, this is mainly a result <strong>of</strong> <strong>the</strong> inferior in<br />

situ data quality as described in Chapter 5.3.2, <strong>the</strong><br />

resulting degradation <strong>of</strong> accuracy in <strong>the</strong> upscaling<br />

process and <strong>the</strong> low relative accuracy <strong>of</strong> <strong>the</strong> produced<br />

high resolution LAI map with 16%. This problem<br />

e<br />

emphasises once more that accurate field data is<br />

necessary <strong>for</strong> validation. <strong>Validation</strong> results <strong>for</strong><br />

Kakamega Forest must thus be handled with care.<br />

Interestingly mean <strong>MODIS</strong> LAI retrieved <strong>for</strong> <strong>the</strong><br />

<strong>for</strong>est areas <strong>of</strong> two study sites and <strong>the</strong> respective<br />

validation time steps differ quite a bit. Whereas mean<br />

<strong>MODIS</strong> LAI <strong>for</strong> intermediate and late <strong>for</strong>est stages<br />

on JD 329 in Budongo Forest is 5.8 (±0.2) and 5.6<br />

(±0.2) respectively, mean <strong>MODIS</strong> LAI on JD 281 <strong>for</strong><br />

<strong>the</strong> nor<strong>the</strong>rn part <strong>of</strong> Kakamega Forest is 6.4 (±0.3).<br />

This is, however, <strong>the</strong> result <strong>of</strong> seasonal LAI variation<br />

ra<strong>the</strong>r than a true difference <strong>of</strong> leaf area between<br />

<strong>the</strong> two <strong>for</strong>ests as <strong>the</strong> validation time steps are more<br />

than 1 ½ months apart. If mean LAI <strong>for</strong> JD 281 and<br />

329 <strong>for</strong> <strong>the</strong> respective years (2004 and 2005) are<br />

compared directly, differences become inessential.<br />

For JD 281 (2004) mean LAI <strong>of</strong> intermediate and late<br />

<strong>for</strong>est stages in Budongo Forest are 6.3 (±0.4) and 6.2<br />

(±0.4) respectively. Mean LAI <strong>for</strong> Kakamega Forest<br />

<strong>for</strong> JD 329 (2005) is calculated as 5.7(±0.5). This<br />

corresponds to seasonal climatic variations (onset<br />

<strong>of</strong> dry season) and suggests that even in <strong>the</strong> case <strong>of</strong><br />

tropical rain <strong>for</strong>est stands - which are supposed to<br />

remain relatively stable over <strong>the</strong> year - <strong>the</strong> time <strong>of</strong><br />

field work should not exceed a few weeks in transition<br />

seasons (unless seasonal variability is expressly<br />

to be monitored). Fur<strong>the</strong>r it should coincide as far<br />

as possible with <strong>the</strong> acquisition <strong>of</strong> high resolution<br />

satellite data used <strong>for</strong> upscaling and acquisition <strong>of</strong> <strong>the</strong><br />

validated product.<br />

Temporal analyses <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

<strong>for</strong> Budongo Forest fur<strong>the</strong>r revealed that LAI<br />

at <strong>the</strong> early <strong>for</strong>est stage corresponds strongly to<br />

interannual climate changes with a range in LAI <strong>of</strong><br />

2.3 over <strong>the</strong> year. LAI <strong>of</strong> mature <strong>for</strong>est sites shows a<br />

comparatively low annual LAI variability with 0.4 <strong>for</strong><br />

intermediate and late <strong>for</strong>est stages. This corresponds<br />

to <strong>the</strong> seasonal change in LAI <strong>of</strong> 0.34 found by De<br />

Wasseige et al. (2003) <strong>for</strong> a semi-deciduous <strong>for</strong>est in<br />

<strong>the</strong> CAR.<br />

In contrast to <strong>the</strong> validation <strong>of</strong> <strong>the</strong> C4 <strong>MODIS</strong> LAI<br />

product <strong>for</strong> a tropical rain <strong>for</strong>est in Brazil published<br />

by Aragão et al. (2006), some major improvements<br />

can be observed <strong>for</strong> <strong>the</strong> C5 <strong>MODIS</strong> LAI products<br />

analysed in this <strong>the</strong>sis. Whereas Aragão et al. (2006)<br />

found a predominance <strong>of</strong> <strong>the</strong> backup algorithm<br />

over <strong>the</strong> <strong>for</strong>ested region due to a failure <strong>of</strong> <strong>the</strong> main<br />

algorithm over vegetation areas with high NDVI<br />

values, this issue can be considered as resolved <strong>for</strong><br />

C5. The majority <strong>of</strong> pixels representing <strong>the</strong> <strong>for</strong>est<br />

areas <strong>of</strong> Budongo and Kakamega Forest are now<br />

generated with <strong>the</strong> “physical saturated” algorithm,<br />

at least under cloud free conditions. If clouds are<br />

present, <strong>the</strong> back-up algorithm prevails - owing to a<br />

failure <strong>of</strong> <strong>the</strong> main algorithm – but in this case <strong>the</strong><br />

respective pixels should be excluded from fur<strong>the</strong>r<br />

analysis anyway. The general overprediction <strong>of</strong><br />

<strong>MODIS</strong> LAI with respect to field LAI <strong>of</strong> 1-2 found by<br />

Aragão et al. (2006) cannot be confirmed <strong>for</strong> <strong>the</strong> test<br />

sites in this <strong>the</strong>sis (note that <strong>the</strong>ir field measurement<br />

also refer to LAI ). e<br />

The sensitivity <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product to<br />

structural changes must, however, be viewed<br />

critically. Temporal analyses mainly confirmed<br />

<strong>the</strong> results <strong>of</strong> monotemporal field data: although<br />

structural differences are present in intermediate and<br />

late <strong>for</strong>est stages, <strong>the</strong> corresponding LAI is relatively<br />

similar. Internal LAI variations in <strong>the</strong> respective<br />

<strong>for</strong>est stages exceed by far <strong>the</strong> differences to mean<br />

LAI <strong>of</strong> <strong>the</strong> o<strong>the</strong>r <strong>for</strong>est stages. However this does not<br />

necessarily imply that structural differences <strong>of</strong> <strong>for</strong>est<br />

stages due to degradation and human disturbances<br />

cannot be monitored with <strong>the</strong> <strong>MODIS</strong> LAI. However<br />

<strong>the</strong> perturbation <strong>of</strong> LAI must be strong enough that it<br />

both affects in situ and remotely measured LAI data.


8 Discussion and Outlook<br />

As discussed in <strong>the</strong> introduction <strong>of</strong> this <strong>the</strong>sis,<br />

<strong>the</strong>re has been a clear lack <strong>of</strong> validation sites <strong>for</strong><br />

<strong>the</strong> <strong>MODIS</strong> LAI product in tropical rain <strong>for</strong>est<br />

environments. Especially in Africa, no test sites<br />

have been available so far. Consequently, this <strong>the</strong>sis<br />

dealt with <strong>the</strong> validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product<br />

<strong>for</strong> two test sites in East Africa, Budongo Forest in<br />

Uganda and Kakamega Forest in Kenya. Both sites<br />

were accessible within <strong>the</strong> framework <strong>of</strong> <strong>the</strong> BIOTA<br />

East Africa project. In <strong>the</strong> following, <strong>the</strong> main results<br />

<strong>of</strong> this <strong>the</strong>sis will be summarised and discussed with<br />

respect to <strong>the</strong>ir contribution to current scientific<br />

discussions. An outlook concerning fur<strong>the</strong>r research<br />

questions is also given.<br />

The first part <strong>of</strong> <strong>the</strong> <strong>the</strong>sis dealt with ground-<strong>based</strong><br />

measurements <strong>of</strong> LAI. The establishment <strong>of</strong> a valid<br />

and representative sampling scheme is an issue that has<br />

<strong>of</strong>ten not been sufficiently addressed in many studies.<br />

However it is <strong>of</strong> major importance <strong>for</strong> <strong>the</strong> accuracy<br />

and representativeness <strong>of</strong> in situ measurements,<br />

especially if <strong>the</strong> ga<strong>the</strong>red data is subsequently used<br />

<strong>for</strong> large scale validation ef<strong>for</strong>ts toge<strong>the</strong>r with field<br />

data from o<strong>the</strong>r studies. CEOS-LPV and VALERI<br />

guidelines were found to be very helpful here, yet<br />

an adaptation to <strong>the</strong> tropical rain <strong>for</strong>est environment<br />

was still required. Especially constraints related to<br />

geolocation errors in satellite and field data as well<br />

as to canopy structure in tropical rain <strong>for</strong>ests were<br />

looked at.<br />

145<br />

The methods proposed by CEOS-LPV and VALERI<br />

are well suited <strong>for</strong> study sites that are characterised<br />

by a structurally homogeneous canopy. Internal<br />

variability <strong>of</strong> <strong>the</strong> vegetation here occurs at scales that<br />

are much smaller than <strong>the</strong> ESU area (usually defined<br />

by <strong>the</strong> pixel size <strong>of</strong> <strong>the</strong> high resolution satellite<br />

data used <strong>for</strong> upscaling). In tropical rain <strong>for</strong>ests,<br />

however, canopy structural variability can only be<br />

captured at larger scales. As fur<strong>the</strong>r constraints are<br />

put on <strong>the</strong> geolocation <strong>of</strong> ESUs and geocorrection <strong>of</strong><br />

satellite data in this environment, ESU size had to be<br />

significantly enlarged. In addition, sampling distance<br />

was increased to guarantee a spatially independent<br />

field database. Interestingly, <strong>the</strong>se aspects have so<br />

far not been addressed in CEOS-LPV and VALERI<br />

recommendations, although validation work thus<br />

far comprised studies also in temperate and boreal<br />

<strong>for</strong>ests.<br />

For <strong>the</strong> standardisation <strong>of</strong> validation ef<strong>for</strong>ts – which<br />

is <strong>the</strong> goal <strong>of</strong> CEOS-LPV – appropriate sampling<br />

strategies should thus be proposed <strong>for</strong> different<br />

groups <strong>of</strong> validation sites. The outcome <strong>of</strong> this <strong>the</strong>sis<br />

suggests that <strong>the</strong>se groups should be characterized by<br />

similar land cover properties (such as canopy height)<br />

and scale <strong>of</strong> internal vegetation variability. These, in<br />

turn, define appropriate ESU sizes, sampling schemes<br />

and <strong>the</strong> number <strong>of</strong> individual measurements. This<br />

standardisation is needed to assure valid sampling<br />

approaches <strong>for</strong> all studies and to facilitate <strong>the</strong> eventual<br />

comparison <strong>of</strong> validation results. As a consequence,<br />

“best practice” protocols <strong>for</strong> data collection and<br />

description should be developed. With respect to<br />

ESU size <strong>the</strong>se could also be transferred to o<strong>the</strong>r in<br />

situ methods, only <strong>the</strong> distance between individual<br />

sampling points depends on <strong>the</strong> used instruments.<br />

Concerning in situ methods <strong>of</strong> LAI estimation, a<br />

review revealed that only indirect optical instruments<br />

were applicable <strong>for</strong> field data sampling on <strong>the</strong> two<br />

study sites. O<strong>the</strong>r methods were ei<strong>the</strong>r not transferable<br />

(e.g. leaf litter collection) or too time-consuming,


146<br />

with <strong>the</strong> result that a representative amount <strong>of</strong> ESUs<br />

could not have been sampled in <strong>the</strong> available time<br />

frame. Consequently, measurements were conducted<br />

with <strong>the</strong> LAI-2000 PCA and, <strong>for</strong> comparison, with<br />

DHPs. These instruments have been rarely applied<br />

to tropical rain <strong>for</strong>ests be<strong>for</strong>e, so that <strong>the</strong> effect <strong>of</strong><br />

non-ideal measurement conditions (e.g. no diffuse<br />

radiation present over longer time periods and high<br />

vegetation stands resulting in mixed pixels) were not<br />

known. A thorough analysis <strong>of</strong> <strong>the</strong> results <strong>of</strong> this <strong>the</strong>sis<br />

led a) to <strong>the</strong> development <strong>of</strong> a novel sampling scheme<br />

<strong>for</strong> LAI-2000 PCA data and b) to a quantification <strong>of</strong><br />

<strong>the</strong> measurement precision <strong>of</strong> both instruments.<br />

With respect to LAI-2000 PCA, three devices were<br />

deployed to monitor <strong>the</strong> influence <strong>of</strong> changing<br />

illumination conditions and direct radiation on <strong>the</strong><br />

measurements. For <strong>the</strong> latter, a correction scheme<br />

could be developed, so that LAI-2000 PCA devices<br />

may now also be used <strong>for</strong> large test sites under<br />

conditions in which indirect radiation is hardly<br />

present over <strong>the</strong> time period needed to conduct <strong>the</strong><br />

measurements. As <strong>the</strong> influence <strong>of</strong> direct radiation<br />

led to an underestimation <strong>of</strong> LAI <strong>of</strong> up to<br />

e (LAI2000)<br />

10% <strong>for</strong> <strong>the</strong> ESUs measured in this <strong>the</strong>sis, this effect<br />

should not be disregarded. Although several authors<br />

have discussed this problem be<strong>for</strong>e (De Wasseige<br />

et al. 2003, Leblanc & Chen 2001), a comparable<br />

correction scheme had not been available prior<br />

to this. Based on data <strong>of</strong> <strong>the</strong> third LAI-2000 PCA<br />

device measurement precision <strong>for</strong> each ESU could be<br />

quantified (up to 10% under non-ideal conditions).<br />

The measurement precision <strong>of</strong> DHPs was determined<br />

at 5%. Whereas DHPs have <strong>the</strong> advantage <strong>of</strong> lower<br />

instrument cost and <strong>the</strong> availability <strong>of</strong> a permanent<br />

illustration <strong>of</strong> <strong>the</strong> sampling site, LAI-2000 PCA<br />

measurements <strong>of</strong>fer continuous monitoring <strong>of</strong> light<br />

intensity (in <strong>the</strong> case <strong>of</strong> A and B1 sensors) and thus<br />

allow a more accurate determination <strong>of</strong> measurement<br />

precision if <strong>the</strong> recommendations <strong>for</strong> instrument setup<br />

from this <strong>the</strong>sis are followed. Especially under<br />

completely direct or diffuse illumination conditions,<br />

<strong>the</strong> relative measurement precision can be even be<br />

higher than that <strong>of</strong> DHPs (cf. Table 6-2).<br />

A comparison <strong>of</strong> <strong>the</strong> LAI values resulting from<br />

both instruments revealed <strong>the</strong> advantages and<br />

disadvantages <strong>of</strong> both approaches. Whereas<br />

understorey vegetation, which comprises up to 14%<br />

<strong>of</strong> site-specific LAI at <strong>the</strong> two test sites considered,<br />

e<br />

may be included in DHP, <strong>the</strong> analysis and processing<br />

<strong>of</strong> <strong>the</strong> photographs requires much more time than<br />

<strong>the</strong> analysis <strong>of</strong> LAI-2000 PCA measurements. On<br />

<strong>the</strong> o<strong>the</strong>r hand, <strong>the</strong> set-up <strong>of</strong> three LAI-2000 PCA<br />

devices in a tropical rain<strong>for</strong>est (with one <strong>of</strong> <strong>the</strong>m<br />

being placed in a nearby <strong>for</strong>est gap) requires greater<br />

logistical ef<strong>for</strong>t. Although LAI values retrieved<br />

e<br />

from both instruments seem reasonable, no general<br />

conclusion can be made to determine which method<br />

leads to more accurate results. More studies like <strong>the</strong><br />

one described by Clark et al. (2008) <strong>for</strong> an old-growth<br />

tropical rain <strong>for</strong>est around La Selva Biological Station<br />

in Costa Rica are needed to compare <strong>the</strong> results <strong>of</strong><br />

indirect optical measurements to destructive ones.<br />

Although <strong>the</strong> ef<strong>for</strong>t <strong>of</strong> setting up modular towers and<br />

<strong>the</strong> accomplishment <strong>of</strong> direct measurements on more<br />

than 55 vertical transects can only be undertaken<br />

as part <strong>of</strong> a larger project network, a combination<br />

<strong>of</strong> indirect measurements be<strong>for</strong>e harvesting and a<br />

comparison to <strong>the</strong> results <strong>of</strong> direct measurements<br />

afterwards would improve <strong>the</strong> understanding <strong>of</strong> <strong>the</strong><br />

relationship between LAI and LAI in tropical rain<br />

e<br />

<strong>for</strong>ests. Un<strong>for</strong>tunately indirect measurements were<br />

not acquired in La Selva.<br />

A comparison <strong>of</strong> <strong>the</strong> results retrieved by Clark et<br />

al. (2008) and <strong>the</strong> outcomes <strong>of</strong> this <strong>the</strong>sis however<br />

indicate a good correspondence. Mean LAIe (LAI2000)<br />

and LAI values calculated <strong>for</strong> <strong>the</strong> intermediate<br />

e (DHP)<br />

and late <strong>for</strong>est stages <strong>of</strong> Budongo Forest are – at 6.14<br />

and 6.25 (intermediate) and 6.35 and 6.14 (late) –<br />

only slightly higher than mean landscape LAI <strong>of</strong> 6.00<br />

derived from direct measurements in <strong>the</strong> old-growth


8 Discussion and Outlook<br />

tropical rain <strong>for</strong>est around La Selva Biological<br />

Station. Mean LAI and LAI values <strong>for</strong><br />

e (LAI2000) e (DHP)<br />

Kakamega Forest were lower, at 5.13 and 4.07 <strong>for</strong><br />

intermediate and 5.88 and 4.17 <strong>for</strong> late <strong>for</strong>est stages,<br />

respectively. This should, however, be interpreted<br />

as a result <strong>of</strong> inferior in situ data quality (as e.g.<br />

understorey vegetation was not included <strong>for</strong> LAIe (DHP)<br />

<strong>for</strong> <strong>the</strong> reasons discussed in Chapter 5), ra<strong>the</strong>r than a<br />

true difference in LAI <strong>of</strong> <strong>the</strong> two <strong>for</strong>ests.<br />

e<br />

This comparison with <strong>the</strong> results <strong>of</strong> Clark et al. (2008)<br />

also puts <strong>the</strong> attempt to derive true LAI values through<br />

correction <strong>for</strong> foliage clumping into perspective.<br />

Although <strong>the</strong> clumping factors derived from DHP<br />

processing are in line with <strong>the</strong> findings <strong>of</strong> Chen et al.<br />

(2005) except <strong>for</strong> early <strong>for</strong>est stages, <strong>the</strong> clumping<br />

correction was applied without an attempt to correct <strong>the</strong><br />

overestimation <strong>of</strong> LAI through non-foliage elements.<br />

In this context several authors state that, depending<br />

on <strong>the</strong> used instruments and applied corrections,<br />

LAI can be very close to true LAI (Eriksson et al.<br />

e<br />

2005, Planchais and Pontailler 1999). This might<br />

e.g. be attributed to <strong>the</strong> fact that <strong>the</strong> correction <strong>for</strong><br />

foliage clumping and <strong>the</strong> correction <strong>for</strong> <strong>the</strong> influence<br />

<strong>of</strong> woody area may cancel each o<strong>the</strong>r out. Yet <strong>for</strong> <strong>the</strong><br />

latter, no satisfactory correction approach exists so<br />

far. First, <strong>the</strong> sometimes applied calculation <strong>of</strong> WAI<br />

through DHP or LAI-2000 PCA measurements during<br />

leafless times in deciduous <strong>for</strong>ests is not transferable<br />

to (semi) evergreen rain <strong>for</strong>ests; second <strong>the</strong> calculation<br />

overestimates <strong>the</strong> real contribution <strong>of</strong> WAI to PAI as<br />

woody elements are at least partly covered by foliage<br />

during <strong>the</strong> vegetation period. There<strong>for</strong>e, fur<strong>the</strong>r ef<strong>for</strong>ts<br />

to develop <strong>the</strong> correct estimation <strong>of</strong> clumping factors<br />

and especially <strong>the</strong> quantification <strong>of</strong> WAI are needed.<br />

Approaches that incorporate <strong>the</strong> possibilities <strong>of</strong> nearinfrared<br />

photography as e.g. proposed by Kucharik et<br />

al. (1998) and Chapman (2007) have a large potential<br />

in this context and should be fur<strong>the</strong>r advanced.<br />

It can be concluded that both instruments, LAI-2000<br />

PCA and DHP, are applicable to <strong>the</strong> tropical <strong>for</strong>est<br />

147<br />

environment if <strong>the</strong> recommendations concerning<br />

instrument set-up and <strong>the</strong> avoidance <strong>of</strong> certain<br />

radiation conditions are followed. This emphasises<br />

once again <strong>the</strong> need <strong>for</strong> a carefully planned field<br />

campaign, as severe errors in <strong>the</strong> data basis have an<br />

effect on all fur<strong>the</strong>r steps in <strong>the</strong> validation process.<br />

This problem is currently only addressed in a very<br />

small number <strong>of</strong> studies (e.g. Cohen et al. 2003,<br />

Huang et al. 2006, Tan et al. 2005). Never<strong>the</strong>less, <strong>the</strong><br />

results from in situ measurements described in this<br />

<strong>the</strong>sis provide <strong>the</strong> first quantitative ground-<strong>based</strong><br />

assessment <strong>of</strong> LAI in semi-deciduous rain <strong>for</strong>ests in<br />

East Africa.<br />

The second part <strong>of</strong> this <strong>the</strong>sis describes <strong>the</strong> generation<br />

<strong>of</strong> high resolution LAI maps <strong>based</strong> on in situ<br />

e<br />

measurements and high resolution satellite data from<br />

ASTER and SPOT. Acknowledging that errors are still<br />

present in <strong>the</strong> data sets, a robust regression method<br />

was sought in order to retrieve stable regression<br />

models. Theil-Sen regression, as introduced by<br />

Fernandes et al. (2005) fulfilled <strong>the</strong>se needs, so that<br />

transfer functions could be established <strong>for</strong> both study<br />

sites. Once more it must be emphasised that OLS<br />

regression, which is widely applied in remote sensing<br />

studies, may lead to biased results in <strong>the</strong> presence <strong>of</strong><br />

measurement errors. This aspect should also be taken<br />

into account when recommendations concerning <strong>the</strong><br />

validation <strong>of</strong> satellite products are to be made.<br />

For Budongo Forest, where better quality <strong>of</strong> in situ<br />

data could be retrieved and more detailed ancillary<br />

in<strong>for</strong>mation was available, a thorough analysis<br />

revealed that different transfer functions must be<br />

established <strong>for</strong> different <strong>for</strong>est stages. Although<br />

mean LAI retrieved <strong>for</strong> intermediate and late <strong>for</strong>est<br />

e<br />

stages was similar, structural differences mainly in<br />

<strong>the</strong> upper canopy led to significantly different surface<br />

reflectances. Especially <strong>for</strong> late <strong>for</strong>est stages, a more<br />

complex and thus very uneven upper canopy led to<br />

reduced surface reflectances in <strong>the</strong> NIR and SWIR<br />

bands compared to those <strong>for</strong> intermediate <strong>for</strong>est


148<br />

stages. Turner et al (1999) states in this context, that<br />

early overflight times in <strong>the</strong> morning promote this<br />

effect, since shadows are still perceptible owing to<br />

<strong>the</strong> large θ . sun<br />

Depending on <strong>the</strong> structural differences <strong>of</strong> <strong>for</strong>est<br />

stages, ei<strong>the</strong>r SVIs or texture variables were better<br />

suited to model LAI from high resolution satellite<br />

e<br />

data. In Budongo Forest, SR per<strong>for</strong>med best <strong>for</strong> early<br />

and intermediate <strong>for</strong>est stages with an R2 <strong>of</strong> 0.94.<br />

Generally SR seems to be better suited to model LAI<br />

<strong>of</strong> high biomass stands than NDVI, as it has a better<br />

dynamic range under <strong>the</strong>se conditions. However,<br />

<strong>the</strong> variability <strong>of</strong> SVIs was comparably low due to<br />

high internal heterogeneity in <strong>for</strong>ests; this was also<br />

observed by Colombo et al. (2003). In contrast to<br />

<strong>the</strong> latter’s study, multilinear regression did not yield<br />

higher R2 . For late <strong>for</strong>est stages a texture measure<br />

was better suited to model LAI . The GLCM variance<br />

e<br />

<strong>of</strong> ASTER band 4 retrieved an R2 <strong>of</strong> 0.71. Based on<br />

similar observations <strong>of</strong> structural differences in <strong>for</strong>est<br />

canopies, correlations between LAI and <strong>the</strong> shade<br />

e<br />

fraction were observed by Seed & King (2003) <strong>for</strong><br />

temperate <strong>for</strong>ests with a canopy closure <strong>of</strong> more than<br />

80% and by Aragão et al. (2006) <strong>for</strong> a primary rain<br />

<strong>for</strong>est in Brazil. Generally, <strong>the</strong> potential <strong>of</strong> texture<br />

measures applied to high and very high resolution<br />

optical satellite data should be fur<strong>the</strong>r exploited.<br />

Transfer functions between LAI e (DHP) and satellite<br />

data led to better results in Budongo Forest <strong>for</strong> all<br />

<strong>for</strong>est stages. Though <strong>the</strong> derived relationships are<br />

only valid <strong>for</strong> <strong>the</strong> respective satellite scenes with<br />

<strong>the</strong>ir characteristic atmospheric conditions and<br />

phenological vegetation states, <strong>the</strong>y transfer <strong>the</strong> field<br />

<strong>based</strong> point measurements to a high spatial resolution<br />

validation surface with <strong>the</strong> highest possible accuracy.<br />

Consequently, a high resolution LAI map could be<br />

e<br />

produced <strong>for</strong> Budongo Forest with a relative accuracy<br />

<strong>of</strong> 9%. This is only a slight degradation in accuracy<br />

relative to field measurements.<br />

For Kakamega Forest <strong>the</strong> previously mentioned<br />

inferior in situ data quality led to reduced quality <strong>of</strong><br />

<strong>the</strong> high resolution LAI map compared to Budongo<br />

e<br />

Forest. For example, DHPs <strong>of</strong> understorey vegetation<br />

had not been taken in Kakamega Forest, so that<br />

transfer functions <strong>based</strong> on LAI generally<br />

e (LAI2000)<br />

per<strong>for</strong>med better. Fur<strong>the</strong>r, unlike Budongo, Kakamega<br />

Forest had not been managed in compartments.<br />

As a result, no dedicated areas <strong>of</strong> homogenous<br />

intermediate or late <strong>for</strong>est stages are perceptible.<br />

Forest stages are ra<strong>the</strong>r dispersed evenly over <strong>the</strong><br />

<strong>for</strong>est. Consequently texture in<strong>for</strong>mation did not lead<br />

to similar results as in Budongo. Ra<strong>the</strong>r a regression<br />

model <strong>based</strong> on RSR was used to estimate LAI <strong>for</strong> e<br />

<strong>the</strong> whole test site. Similar models have so far only<br />

been applied to coniferous, deciduous or mixed<br />

temperate <strong>for</strong>ests (e.g. Chen et al. 2002, Huang et al<br />

2006, and Stenberg et al. 2004). Overall this model<br />

yielded an R2 <strong>of</strong> 0.53 with a relative accuracy <strong>of</strong><br />

16% (RMSE <strong>of</strong> 0.8). Although <strong>the</strong>se results are not<br />

as good as <strong>the</strong> ones retrieved <strong>for</strong> Budongo Forest,<br />

<strong>the</strong>y are comparable to those <strong>of</strong> Cohen et al. (2006).<br />

In <strong>the</strong>ir study Landsat-<strong>based</strong> LAI maps were created<br />

<strong>for</strong> a moist tropical <strong>for</strong>est in Brazil with a correlation<br />

<strong>of</strong> 0.54 and an RMSE <strong>of</strong> 0.83. These maps were also<br />

used <strong>for</strong> C4 <strong>MODIS</strong> LAI product validation.<br />

The third part <strong>of</strong> this <strong>the</strong>sis involved <strong>the</strong> final<br />

validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product <strong>for</strong> Budongo<br />

and Kakamega Forest <strong>based</strong> on <strong>the</strong> high resolution<br />

LAI maps. Here, algorithm input data was analyzed<br />

e<br />

first. As observed by Cohen et al. (2006) <strong>for</strong> <strong>the</strong>ir study<br />

area in Brazil, <strong>the</strong> <strong>MODIS</strong> land cover class “broadleaf<br />

<strong>for</strong>est” corresponded well to real land cover <strong>of</strong> <strong>the</strong><br />

two test sites (although it must be noted once more<br />

that <strong>the</strong> new eight biome map as implemented in <strong>the</strong><br />

C5 <strong>MODIS</strong> LAI algorithm had not been available<br />

and C4 land cover data was analyzed instead).<br />

The predominance <strong>of</strong> back-up algorithm retrievals<br />

as reported by Aragão et al. (2006) <strong>for</strong> C4 <strong>MODIS</strong><br />

LAI data could not be confirmed in this <strong>the</strong>sis. At


8 Discussion and Outlook<br />

least <strong>for</strong> <strong>the</strong> dry seasons and <strong>the</strong> validated time steps<br />

<strong>the</strong> physical saturated algorithm prevailed over <strong>the</strong><br />

<strong>for</strong>ested areas. This is probably <strong>the</strong> result <strong>of</strong> algorithm<br />

improvements in C5 (cf. NASA 2007b), which<br />

reduced <strong>the</strong> failure rate <strong>of</strong> <strong>the</strong> physical algorithm.<br />

If applied, <strong>the</strong> back-up algorithm still suffered from<br />

severe problems. In tropical regions probably cloud<br />

cover is one <strong>of</strong> <strong>the</strong> major issues <strong>for</strong> failure <strong>of</strong> <strong>the</strong> main<br />

algorithm. This might partly explain <strong>the</strong> bad data<br />

quality <strong>of</strong> back-up algorithm retrievals. To solve this<br />

problem, <strong>the</strong> operational production <strong>of</strong> composites<br />

over more than 8-days is suggested to ensure a cloudfree<br />

data basis even during <strong>the</strong> rainy season. Monthly<br />

composites <strong>of</strong> <strong>the</strong> C5 <strong>MODIS</strong> LAI product are, <strong>for</strong><br />

instance, <strong>of</strong>fered <strong>for</strong> download by <strong>the</strong> “Climate and<br />

Vegetation Research Group” at Boston University.<br />

However, only data sets covering <strong>the</strong> years 2000,<br />

2001, 2002 and 2007 are currently available.<br />

The spatial validation <strong>of</strong> <strong>the</strong> C5 <strong>MODIS</strong> LAI product<br />

<strong>for</strong> Budongo Forest was finally per<strong>for</strong>med as a patchby-patch<br />

comparison. It revealed that <strong>the</strong> C5 <strong>MODIS</strong><br />

LAI product represented <strong>the</strong> up-scaled in situ LAIe with an accuracy <strong>of</strong> 0.53. This corresponds to a<br />

relative accuracy <strong>of</strong> 9%, which is identical with <strong>the</strong><br />

accuracy <strong>of</strong> <strong>the</strong> high resolution LAI maps. Those in<br />

e<br />

turn were generated with only a slight degradation<br />

<strong>of</strong> field data accuracy (taking into account <strong>the</strong><br />

above-mentioned precision <strong>of</strong> 5% <strong>for</strong> LAI and<br />

e (DHP)<br />

assuming a negligible bias). These findings once<br />

more indicate an improvement <strong>of</strong> <strong>the</strong> C5 algorithm<br />

compared to C4. Aragão et al. (2006) reported,<br />

<strong>for</strong> instance, a significant overprediction <strong>of</strong> fieldmeasured<br />

LAI by <strong>the</strong> C4 <strong>MODIS</strong> LAI product <strong>for</strong> <strong>the</strong><br />

previously mentioned Brazilian rain <strong>for</strong>est site. An<br />

overprediction <strong>of</strong> LAI in <strong>for</strong>ested biomes was also<br />

reported by Cohen et al. (2006), who validated <strong>the</strong> C4<br />

<strong>MODIS</strong> LAI product <strong>for</strong> five boreal, temperate and<br />

tropical <strong>for</strong>est sites across <strong>the</strong> Western Hemisphere.<br />

By contrast, C5 <strong>MODIS</strong> LAI, as analysed within this<br />

<strong>the</strong>sis, underestimated <strong>the</strong> reference LAI slightly with<br />

e<br />

a bias <strong>of</strong> -0.35 (relative bias <strong>of</strong> -6%). Interestingly,<br />

149<br />

an even larger bias (-3.28) was present <strong>for</strong> reference<br />

LAI, which was calculated by applying <strong>the</strong> clumping<br />

factors derived from DHP to <strong>the</strong> high resolution LAIe maps (again no correction <strong>for</strong> WAI was per<strong>for</strong>med).<br />

This underscores <strong>the</strong> above-mentioned assumption<br />

that LAI might be close to “real” LAI (note that <strong>the</strong><br />

e<br />

o<strong>the</strong>r studies also compared <strong>MODIS</strong> LAI to LAI <strong>of</strong> e<br />

in situ measurements).<br />

For Kakamega Forest, <strong>the</strong> spatial validation <strong>of</strong> <strong>the</strong><br />

<strong>MODIS</strong> LAI product led to a comparatively low<br />

accuracy <strong>of</strong> 1.5 (corresponding to a relative accuracy<br />

<strong>of</strong> 25%). This is most likely <strong>the</strong> result <strong>of</strong> inferior<br />

field data quality (as mentioned above) <strong>for</strong> this test<br />

site and <strong>the</strong> resulting degradation <strong>of</strong> accuracy in <strong>the</strong><br />

upscaling process. Consequently, <strong>the</strong> results must<br />

be interpreted carefully, as <strong>the</strong>y may be <strong>based</strong> on<br />

inaccurate assumptions. The fact that mean in situ<br />

LAI data <strong>for</strong> intermediate and late <strong>for</strong>est stages in<br />

e<br />

Kakamega Forest differs by up to 2 LAI units from <strong>the</strong><br />

respective data sampled in Budongo Forest indicates<br />

<strong>the</strong> reliability (or non-reliability) <strong>of</strong> field data in<br />

Kakamega. Once again this problem emphasises <strong>the</strong><br />

need <strong>for</strong> accurate field data (and sampling protocols)<br />

<strong>for</strong> validation purposes.<br />

Although a real temporal validation – which would<br />

require in situ LAI measurements over several<br />

months – was beyond <strong>the</strong> scope <strong>of</strong> this <strong>the</strong>sis, at<br />

least <strong>the</strong> temporal consistency <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI<br />

product was investigated. An analysis <strong>of</strong> time series<br />

<strong>for</strong> <strong>the</strong> years 2000-2005 showed that <strong>the</strong> data was<br />

reliable and stable, but only if temporal interpolation<br />

was applied to bad data quality pixels. The observed<br />

variability in LAI (0.4 <strong>for</strong> intermediate and late<br />

<strong>for</strong>est stages) fur<strong>the</strong>r corresponds to in situ measured<br />

seasonal LAI trajectories found by De Wasseige et al.<br />

(2003) <strong>for</strong> a semi-deciduous <strong>for</strong>est in <strong>the</strong> CAR and<br />

by Wirth et al. (2001) <strong>for</strong> a tropical moist rain<strong>for</strong>est<br />

in Panama. Maximum LAI values are associated<br />

with <strong>the</strong> end <strong>of</strong> <strong>the</strong> rainy season, minimum LAI<br />

values to <strong>the</strong> dry season. It can thus be assumed


150<br />

that <strong>the</strong> <strong>MODIS</strong> LAI product responds correctly to<br />

biome-level LAI changes associated with interannual<br />

climate variability.<br />

Interestingly, <strong>the</strong> above-mentioned results contradict<br />

<strong>the</strong> findings <strong>of</strong> Myneni et al (2007), who discovered<br />

large seasonal swings in leaf area <strong>for</strong> <strong>the</strong> Amazonian<br />

rain <strong>for</strong>est. In <strong>the</strong>ir study, which analysed <strong>MODIS</strong><br />

LAI data from 2000-2005, maximum leaf area was<br />

always observed during <strong>the</strong> dry season and decreasing<br />

approximately 25% during <strong>the</strong> rainy season. According<br />

to <strong>the</strong>ir reasoning, <strong>the</strong>re is a correspondence between<br />

maximum leaf area and maximum solar radiation. To<br />

shed light on <strong>the</strong>se relationships and to facilitate future<br />

temporal validation campaigns, new sensor systems<br />

as e.g described by Baret (2007) could be employed.<br />

They consist <strong>of</strong> small independent sensor units that<br />

record above canopy photosyn<strong>the</strong>tically active<br />

radiation (PAR) and below canopy transmittance.<br />

These sensors communicate to a master unit that can<br />

store <strong>the</strong> collected data. According to Baret (2007)<br />

<strong>the</strong>se sensors can be produced at low cost and are<br />

autonomous <strong>for</strong> up to six months. The sensors <strong>of</strong>fer<br />

continuous monitoring <strong>of</strong> PAR balance, with LAI<br />

being subsequently calculated from transmitted PAR<br />

(analogous to LAI-2000 PCA measurements). As with<br />

o<strong>the</strong>r indirect optical methods, however, limitations<br />

in terms <strong>of</strong> <strong>the</strong> estimation <strong>of</strong> WAI or clumping factors<br />

would persist.<br />

To summarise, this <strong>the</strong>sis describes <strong>the</strong> first validation<br />

study <strong>for</strong> <strong>the</strong> <strong>MODIS</strong> LAI product in African rain<br />

<strong>for</strong>ests. Additionally, it has validated <strong>the</strong> C5 <strong>MODIS</strong><br />

LAI algorithm <strong>for</strong> <strong>the</strong> first time ever <strong>for</strong> <strong>the</strong> tropical<br />

<strong>for</strong>est biome. The outcomes are already expected<br />

to contribute to CEOS-LPV LAI inter-comparison<br />

ef<strong>for</strong>ts by Garrigues et al. (2006). Field data and<br />

derived high resolution LAI maps will fur<strong>the</strong>r be<br />

made available within <strong>the</strong> VALERI and CEOS-LPV<br />

networks <strong>for</strong> <strong>the</strong> validation <strong>of</strong> biophysical products<br />

derived from o<strong>the</strong>r medium resolution satellite<br />

sensors.<br />

Although <strong>the</strong> hypo<strong>the</strong>sis that subtle structural<br />

changes in rain <strong>for</strong>ests can be monitored through<br />

repeated measurements <strong>of</strong> LAI could not be verified<br />

– as differences in LAI <strong>of</strong> intermediate and late <strong>for</strong>est<br />

changes were not significant – <strong>the</strong> outcomes <strong>of</strong> this<br />

<strong>the</strong>sis can never<strong>the</strong>less help to improve knowledge<br />

on tropical rain<strong>for</strong>ests and <strong>the</strong>ir response to climatic<br />

changes and human disturbance. As <strong>the</strong> differences <strong>of</strong><br />

<strong>the</strong> temporal LAI signature <strong>of</strong> intermediate and early<br />

<strong>for</strong>est stages in this <strong>the</strong>sis were tremendous, fur<strong>the</strong>r<br />

research should be carried out to better understand<br />

<strong>the</strong> link between <strong>for</strong>est degradation and temporal<br />

signature <strong>of</strong> biophysical parameters.<br />

The outcomes <strong>of</strong> this <strong>the</strong>sis may fur<strong>the</strong>r help to<br />

reduce predictive uncertainties in biophysical process<br />

models, as e.g. <strong>the</strong> CoupModel used by Miao (2008)<br />

to model <strong>the</strong> carbon dynamics <strong>of</strong> Kakamega Forest.<br />

<strong>MODIS</strong> LAI data may here serve as supplementary<br />

data <strong>for</strong> model calibration. Knowledge about <strong>the</strong><br />

accuracy <strong>of</strong> this data set is, however, crucial, as it<br />

is taken as an absolute reference. Also numerical<br />

wea<strong>the</strong>r prediction models need LAI data <strong>for</strong> <strong>the</strong><br />

calculation <strong>of</strong> e.g. evapotranspiration processes.<br />

There<strong>for</strong>e LAI was defined as an “essential climate<br />

variable” by <strong>the</strong> Global Climate Observing System.<br />

<strong>Validation</strong> <strong>for</strong> different biomes also helps to improve<br />

later versions <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product algorithm<br />

and thus leads to an increase <strong>of</strong> derivative product<br />

accuracy. Errors in <strong>the</strong> <strong>MODIS</strong> LAI product do not<br />

only extend into water and energy balance <strong>of</strong> climate<br />

models (atmosphere-biosphere exchange component<br />

<strong>of</strong> general circulation climate models), but also in<br />

higher level <strong>MODIS</strong> products as e.g. <strong>the</strong> <strong>MODIS</strong><br />

GPP (MOD17A2) product. Fur<strong>the</strong>r improvements<br />

<strong>of</strong> <strong>the</strong> LAI algorithm can be expected with <strong>the</strong> new<br />

version (Collection 6). According to Didan (2007)<br />

an optimisation <strong>of</strong> retrievals with SWIR data is<br />

envisaged, which might possibly lead to a better LAI<br />

retrieval <strong>for</strong> broadleaf evergreen <strong>for</strong>est in Amazonia<br />

and Central Africa.


References<br />

Abrams, M. and S. Hook (2002). ASTER User Handbook, Version 2. Accessed 19 October, 2007.<br />

.<br />

Abuelgasim, A.A., Fernandes, R.A. and S.G. Leblanc (2006). Evaluation <strong>of</strong> national and global LAI products<br />

derived from optical remote sensing instruments over Canada. – In: IEEE Transactions on Geoscience<br />

and Remote Sensing, 44, pp. 1872−1884.<br />

Alexandre, D. (1981). L’indice foliaire des <strong>for</strong>êts tropicales. Analyse bibliographique. – In: Acta Oecologica,<br />

Oecologia Generalis, 2 (4), pp. 299–312.<br />

Alth<strong>of</strong>, A. (2005). Human Impact on Flora and Vegetation <strong>of</strong> Kakamega Forest, Kenya: structure, distribution<br />

and disturbance <strong>of</strong> plant communities in an East African rain<strong>for</strong>est. Accessed 5 September, 2007.<br />

.<br />

Anderson, M.C. (1964). Studies <strong>of</strong> <strong>the</strong> woodland climate. I. The photographic computation <strong>of</strong> light conditions.<br />

– In: Journal <strong>of</strong> Ecology, 52, pp. 27-41.<br />

Andrua, H. J. (2003). Tropical Secondary Forest Management in Africa: Reality and perspectives. Uganda<br />

Country Paper. – In: Proceedings <strong>of</strong> <strong>the</strong> “Workshop on Tropical Secondary Forest Management in Africa:<br />

Reality and Perspectives”, pp. 9-13. December 2002, Nairobi, Kenya.<br />

Aragão, L.E., Shimabukuro, Y.E., Espírito Santo, F.D.B. and M. Williams (2005). Landscape pattern and spatial<br />

variability <strong>of</strong> leaf area index in Eastern Amazonia. – In: Forest Ecology and Management, 211, pp. 240-<br />

256.<br />

Aragão, L.E., Shimabukuro, Y.E., Espírito Santo, F.D.B. and M. Williams (2006). Spatial validation <strong>of</strong> <strong>the</strong><br />

collection 4 <strong>MODIS</strong> LAI product in eastern Amazonia. – In: IEEE Transactions on Geoscience and<br />

Remote Sensing, 43 (11), pp. 2526-2534.<br />

Arvidson, T., Gasch, J. and S.N. Goward (2001). Landsat 7’s long-term acquisition plan – an innovative<br />

approach to building a global imagery archive. – In: Remote Sensing <strong>of</strong> Environment, 78, pp. 13-26.<br />

Aryal, B. (2002). Are trees <strong>for</strong> <strong>the</strong> poor? A study from Budongo Forest, Uganda. Unpublished Master Thesis,<br />

Department <strong>of</strong> Economics and Social Science, Agricultural University <strong>of</strong> Norway, Ǻs.<br />

Asner, G.P., Scurlock, J.M.O. and J.A. Hicke (2003). Global syn<strong>the</strong>sis <strong>of</strong> leaf area index observations:<br />

implications <strong>for</strong> ecological and remote sensing studies. – In: Global Ecology and Biogeography, 12, pp.<br />

191-205.<br />

Atzberger, C., Jarmer, T., Kötz, B., Schlerf, M., and W. Werner (2003). Spectroradiometric determination<br />

<strong>of</strong> wheat canopy biophysical variables. Comparison <strong>of</strong> several empirical-statistical methods. – In: R.<br />

Goossens (Ed.), Remote sensing in transition. Proceedings <strong>of</strong> <strong>the</strong> 23rd EARSEeL Symposium on Remote<br />

Sensing in Transition, Ghent, Belgium, 2-5 June 2003, pp. 463-470.<br />

Bacour, C., Baret, F., Béal, D., Weiss, M. and K. Pavageau (2006). Neural network estimation <strong>of</strong> LAI, fAPAR,<br />

fCover and LAI×Cab, from top <strong>of</strong> canopy MERIS reflectance data: Principles and validation. – In:<br />

Remote Sensing <strong>of</strong> Environment, 105, pp. 313-325.<br />

Bahrenberg, G., Giese, E. and J. Nipper (2003). Statistische Methoden in der Geographie Band 2. Multivariate<br />

Statistik. Berlin, Stuttgart.<br />

Baret, F. (1995). Use <strong>of</strong> spectral reflectance variation to retrieve canopy biophysical characteristics. – In: Darson,<br />

M. and S. Plummer (eds.). Advances in environmental remote sensing, Wiley and Sons, pp. 33-51.<br />

151


152<br />

Baret, F., Clevers, J.G.P.W. and M.D. Steven (1995). The robustness <strong>of</strong> canopy gap fraction estimates from red<br />

and near-infrared reflectances: a comparison <strong>of</strong> approaches. – In: Remote Sensing <strong>of</strong> Environment, 54,<br />

pp. 141-151.<br />

Baret, F. (2004a). Can-Eye s<strong>of</strong>tware. Accessed 21 September, 2007. .<br />

Baret, F. (2004b). A simple method to calibrate hemispherical photographs. Accessed 21 September, 2007.<br />

.<br />

Baret, F., Morisette, J.T., Fernandes, R.A., Champeaux, J.L., Myneni, R.B., Chen, J., Plummer, S., Weiss, M.,<br />

Bacour, C., Garrigues, S. and J.E. Nickeson (2006). Evaluation <strong>of</strong> <strong>the</strong> Representativeness <strong>of</strong> Networks<br />

<strong>of</strong> Sites <strong>for</strong> <strong>the</strong> Global <strong>Validation</strong> and Intercomparison <strong>of</strong> Land Biophysical <strong>Product</strong>s: Proposition <strong>of</strong><br />

<strong>the</strong> CEOS-BELMANIP. – In: IEEE Transactions on Geoscience and Remote Sensing, 44 (7), pp. 1794-<br />

1802.<br />

Baret, F. (2007). Sample results <strong>for</strong> LAI estimates from Hemisphotos and CAN_EYE. Accessed 2 July, 2008.<br />

<br />

Baret, F. and P. Rosello (2007). <strong>Validation</strong> <strong>of</strong> Land European Remote Sensing Instruments. Accessed 15 August,<br />

2007. .<br />

Baret, F., Hagolle, O., Geiger, B., Bicheron, P., Miras, B., Huc, M., Ber<strong>the</strong>lot, B., Niño, F., Weiss, M., Samain,<br />

O., Roujean, J.L. and M. Leroy (2007). LAI, fAPAR and fCover CYCLOPES global products derived<br />

from VEGETATION Part 1: Principles <strong>of</strong> <strong>the</strong> algorithm. – In: Remote Sensing <strong>of</strong> Environment, 110 (3),<br />

pp. 275-286.<br />

Baret, F., Weiss, M., Garrigues, S., Allard, D., Leroy, M., Jeanjean, H., Fernandes, R., Myneni, R.B., Morisette,<br />

J.T., Privette, J., Bohbot, H., Bosseno, R., Dedieu, G., Di Bella, C., Espana, M., Gond, V., Gu, X.F., Guyon,<br />

D., Lelong, C., Maisongrande, P., Mougin, E., Nilson, T., Veroustraete, F. and R. Vintilla (submitted).<br />

VALERI: A network <strong>of</strong> sites and a methodology <strong>for</strong> <strong>the</strong> validation <strong>of</strong> medium spatial resolution satellite<br />

products. – In: Remote Sensing <strong>of</strong> Environment, submitted <strong>for</strong> publication.<br />

Barnes, W.L., Pagano, T.S. and V.V. Salomonson (1998). Prelaunch characteristics <strong>of</strong> <strong>the</strong> Moderate Resolution<br />

Imaging Spectroradiometer (<strong>MODIS</strong>) on EOS-AM1. – In: IEEE Transactions on Geoscience and Remote<br />

Sensing, 36 (4), pp. 1088-1100.<br />

Bartelink, H.H. (1997). Allometric relationships <strong>for</strong> biomass and leaf area <strong>of</strong> beech (Fagus sylvatica L).<br />

– In: Annales des Sciences Forestieres, 54, pp. 39-50.<br />

Beentje, H.J. (1990). The <strong>for</strong>ests <strong>of</strong> Kenya. – In: Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg,<br />

23a, pp. 265-286.<br />

Beentje, H.J. (1994). Kenya trees, shrubs and lianas, National Museum <strong>of</strong> Kenya, Nairobi.<br />

Belward, A.S. (1999). International co-operation in satellite sensor calibration; <strong>the</strong> role <strong>of</strong> <strong>the</strong> CEOS working<br />

group on calibration and validation. – In: Advances in Space Research, 23 (8), pp. 1443-1448.<br />

Berk, A., Bernstein L.S., Anderson G.P., Acharya P.K., Robertson D.C., Chetwynd J.H. and S.M. Adler-Golden<br />

(1998). MODTRAN cloud and multiple scattering upgrades with application to AVIRIS. – In: Remote<br />

Sensing <strong>of</strong> Environment, 65, pp. 367-375.<br />

Berterretche, M. Hudak, A.T., Cohen, W.B., Maiersperger, T.K., Gower, S.T. and J. Dungan (2005). Comparison<br />

<strong>of</strong> regression and geostatistical methods <strong>for</strong> mapping <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> (LAI) with Landsat ETM+ data<br />

over a boreal <strong>for</strong>est. – In: Remote Sensing <strong>of</strong> Environment, 96, pp. 49-61.


References<br />

Bicheron, P., Leroy, M. and O. Hautecoeur (1998). LAI and fAPAR mapping at global scale by model inversion<br />

against spaceborne POLDER data. – In: Proceedings <strong>of</strong> IGARSS, 3, 1998, pp. 1228-1230.<br />

Bicheron, P. and M. Leroy (1999). A Method <strong>of</strong> Biophysical Parameter Retrieval at Global Scale by Inversion<br />

<strong>of</strong> a Vegetation Reflectance Model. – In: Remote Sensing <strong>of</strong> Environment, 67, pp. 251-266.<br />

Birth, G. S. and G. McVey (1968). Measuring <strong>the</strong> color <strong>of</strong> growing turf with a reflectance spectrophotometer.<br />

– In: Agronomy Journal, 60, pp. 640-643.<br />

Bjørlykke, K.O. (1973). Glacial Conglomerates <strong>of</strong> late Precambrian age from <strong>the</strong> Bunyoro Series, W. Uganda.<br />

– In: International Journal <strong>of</strong> Earth Sciences, 62 (3), pp. 938-947.<br />

Black, T.A., Chen, J.M., Lee, X. and R.M. Sagar (1991). Characteristics <strong>of</strong> shortwave and longwave irradiances<br />

under a Douglas-fir <strong>for</strong>est stand. – In: Canadian Journal <strong>of</strong> Forest Research, 21, pp. 1020-1028.<br />

Blackett, H.L. (1994). Forest Inventory Report No. 3: Kakamega. Kenya Indigenous Forest Conservation<br />

Programme (KIFCON), Nairobi, Kenya.<br />

Bleher, B., Uster, D. and T. Bergsdorf (2006). Assessment <strong>of</strong> threat status and management effectiveness in<br />

Kakamega Forest, Kenya. – In: Biodiversity and Conservation, 15, pp. 1159-1177.<br />

Bonhomme, R., Varlet-Grancher, C. and P. Chartier (1974). The use <strong>of</strong> hemispherical photographs <strong>for</strong><br />

determining <strong>the</strong> leaf area index <strong>of</strong> young crops. – In: Photosyn<strong>the</strong>tica, 8 (3), pp. 299-301.<br />

Bréda, N. (2003). <strong>Ground</strong>-<strong>based</strong> measurements <strong>of</strong> leaf area index: a review <strong>of</strong> methods, instruments and current<br />

controversies. – In: Journal <strong>of</strong> Experimental Botany, 54 (392), pp. 2403-2417.<br />

Broich, M. (2005). Ableitung des Blattflächenindex aus Landsat ETM+ SLC-<strong>of</strong>f Daten für ostafrikanische<br />

Regenwaldökosysteme am Beispiel von Kakamega Forest (Westkenia). Unveröffentlichte Diplomarbeit,<br />

eingereicht am Geographischen Institut der RWTH Aachen.<br />

Brooks, T.M., Mittermeier, R.A., Mittermeier, C.G., da Fonseca, G.A.B., Rylands, A.B., Konstant, W.R., Flick,<br />

P., Pilgrim, J., Oldfield, S., Magin, G. and C. Hilton-Taylor (2002). Habitat loss and extinction in <strong>the</strong><br />

hotspots <strong>of</strong> biodiversity. – In: Conservation Biology, 16, pp. 909-923.<br />

Brown, L. J., Chen, J. M., Leblanc, S. G. and J. Cihlar (2000). Short wave infrared correction to <strong>the</strong> simple<br />

ratio: an image and model analysis. – In: Remote Sensing <strong>of</strong> Environment, 71, pp. 16-25.<br />

Burgess, N., D’Amico Hales, J., Underwood, E., Dinerstein, E., Olson, D., Itoua, I., Schipper, J., Ricketts, T.<br />

and K. Newman. (2004). Terrestrial ecoregions <strong>of</strong> Africa and Madagascar: a conservation assessment.<br />

Island Press, Washington, D.C.<br />

Butson, R. and R.A. Fernandes (2004). A consistency analysis <strong>of</strong> surface reflectance and leaf area index retrieval<br />

from overlapping clear-sky Landsat ETM+ imagery. – In: Remote Sensing <strong>of</strong> Environment, 89, pp. 369-<br />

380.<br />

Cahen, L., Snelling, N.J., Delhal, J. and J.R. Vail (1984). The geochronology and evolution <strong>of</strong> Africa. Ox<strong>for</strong>d.<br />

Casa, R. and H.G. Jones (2005). LAI retrieval from multiangular image classification and inversion <strong>of</strong> a ray<br />

tracing model. – In: Remote Sensing <strong>of</strong> Environment, 98, pp. 414-428.<br />

Chapman, L. (2007). Potential applications <strong>of</strong> near infra-red hemispherical imagery in <strong>for</strong>est environments.<br />

– In: Agricultural and Forest Meteorology, 143, pp. 151-156.<br />

153


154<br />

Chason, J.W., Baldocchi, D.D. and M.A. Huston (1991). A comparison <strong>of</strong> direct and indirect methods <strong>for</strong><br />

estimating <strong>for</strong>est canopy leaf area. – In: Agricultural and Forest Meteorology, 57, pp. 107-128.<br />

Chen, J.M., Black, T.A. and R.S. Adams (1991). Evaluation <strong>of</strong> hemispherical photography <strong>for</strong> determining plant<br />

area index and geometry <strong>of</strong> a <strong>for</strong>est stand. – In: Agriculture and Forest Meteorology, 56, pp. 129-143.<br />

Chen, J.M. and T.A. Black (1992). Defining leaf area index <strong>for</strong> non-flat leaves. – In: Plant Cell Environment,<br />

15, pp. 421-429.<br />

Chen, J.M. and J. Cihlar (1995a). Plant canopy gap-size analysis <strong>the</strong>ory <strong>for</strong> improving optical measurements <strong>of</strong><br />

leaf-area index. – In: Applied Optics, 34 (27), pp. 6211-6222.<br />

Chen, J.M. and J. Cihlar (1995b). Quantifying <strong>the</strong> effect <strong>of</strong> canopy architecture on optical measurements <strong>of</strong><br />

leaf area index using two gap size analysis methods. – In: IEEE Transactions on Geoscience and Remote<br />

Sensing, 33 (3), pp. 777-787.<br />

Chen, J. M. (1996). Evaluation <strong>of</strong> vegetation indices and a modified simple ratio <strong>for</strong> Boreal applications.<br />

– In: Canadian Journal <strong>of</strong> Remote Sensing, 22, 229-241.<br />

Chen, J. M., and J. Cihlar (1996). Retrieving leaf area index <strong>of</strong> boreal conifer <strong>for</strong>ests using Landsat TM images.<br />

– In: Remote Sensing <strong>of</strong> Environment, 55, pp. 153-162.<br />

Chen, J.M., Rich P.M., Gower, S.T., Norman J.M. and S. Plummer (1997). <strong>Leaf</strong> area index <strong>of</strong> boreal <strong>for</strong>ests:<br />

Theory, techniques, and measurements. – In: Journal <strong>of</strong> Geophysical Research, 102 (D24), p. 29429-<br />

29443.<br />

Chen, J.M., Pavlic, G., Brown, L., Cihlar, J., Leblanc, S.G., White, H.P., Hall, R.J., Peddle, D.R., King, D.J.,<br />

Tr<strong>of</strong>ymow, J.A., Swift, E., Van der Sanden, J. and P.K.E. Pellikka (2002). Derivation and validation <strong>of</strong><br />

Canada-wide coarse-resolution leaf area index maps using high-resolution satellite imagery and ground<br />

measurements. – In: Remote Sensing <strong>of</strong> Environment, 80, 165-184.<br />

Chen, J.M., Menges, C.H. and S.G. Leblanc (2005). Global derivation <strong>of</strong> <strong>the</strong> vegetation clumping index from<br />

multi-angular satellite data. – In: Remote Sensing <strong>of</strong> Environment, 97, 447-457.<br />

Clark, D.A. (2007). Detecting Tropical Forests’ Responses to Global Climatic and Atmospheric Change: Current<br />

Challenges and a Way Forward. – In: Biotropica, 39 (1), pp. 4-19.<br />

Clark, D.B, Olivas, P.C., Oberbauer, S.F., Clark, D.A. and M.G. Ryan (2008). First direct landscape-scale<br />

measurement <strong>of</strong> tropical rain <strong>for</strong>est <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong>, a key driver <strong>of</strong> global primary productivity.<br />

– In: Ecology Letters, 11 (2), pp. 163-172.<br />

Cohen, W.B., Maiersperger, T.K., Gower, S.T. and D.P. Turner (2003). An improved strategy <strong>for</strong> regression <strong>of</strong><br />

biophysical variables and Landsat ETM+ data. – In: Remote Sensing <strong>of</strong> Environment, 84, pp. 561-571.<br />

Cohen, W.B., Maiersperger, T.K., Turner, D.P., Ritts, W.D., Pflugmacher, D., Kennedy, R.E., Kirschbaum,<br />

A., Running, S.W., Costa, M. and S.T. Gower (2006). <strong>MODIS</strong> land cover and LAI collection 4 product<br />

quality across nine sites in <strong>the</strong> Western Hemisphere. – In: IEEE Transactions on Geoscience and Remote<br />

Sensing, 44, pp. 1843-1857.<br />

Colditz, R.R., Conrad, C., Wehrmann, T., Schmidt, M. and S.W. Dech (2008). TiSeG – A flexible s<strong>of</strong>tware tool<br />

<strong>for</strong> time series generation <strong>of</strong> <strong>MODIS</strong> data utilizing <strong>the</strong> quality assessment science data set. – In: IEEE<br />

Transactions on Geoscience and Remote Sensing (accepted).<br />

Colombo, R., Bellingeri, D., Fasolini, D. and C. Marino (2003). Retrieval <strong>of</strong> leaf area index in different<br />

vegetation types using high resolution satellite data. – In: Remote Sensing <strong>of</strong> Environment, 86, pp. 120-131.


References<br />

Cuarón, A.D. (2000). A global perspective on habitat disturbance and tropical rain<strong>for</strong>est mammals.<br />

– In: Conservation Biology, 14 (6), pp. 1574-1579.<br />

Curran, P.J., and A. Hay (1986). The importance <strong>of</strong> measurement error <strong>for</strong> certain procedures in remote sensing<br />

at optical wavelengths. – In: Photogrammetric Engineering and Remote Sensing, 52, pp. 229-241.<br />

Curran, P.J., Dungan, J.L. and H.L. Gholz (1992). Seasonal LAI in Slash Pine Estimated with Landsat TM.<br />

– In: Remote Sensing <strong>of</strong> Environment, 39, pp. 3-13.<br />

Cutini, A., Matteucci, G. and G.S. Mugnozza (1998). Estimation <strong>of</strong> leaf area index with <strong>the</strong> Li-Cor LAI 2000 in<br />

deciduous <strong>for</strong>ests. – In: Forest Ecology and Management, 105, pp. 55-65.<br />

Deng, F., Chen, J.M., Plummer, S. and M. Chen (2006). Algorithm <strong>for</strong> global leaf area index retrieval using<br />

satellite imagery. – In: IEEE Transactions on Geoscience and Remote Sensing, 44, pp. 2219-2229.<br />

De Wasseige, C., Bastin, D. and P. Defourny (2003). Seasonal variation <strong>of</strong> tropical <strong>for</strong>est LAI <strong>based</strong> on field<br />

measurements in Central African Republic. – In: Agricultural and Forest Meteorology, 119, pp. 181-194.<br />

Dech, S.W., Tungalagsaikhan, P., Preusser, C. and R.E. Meisner (1998). Operational value-adding to AVHRR<br />

data over Europe: methods, results, and prospects. – In: Aerospace Science and Technology, 2 (5),<br />

pp. 335-346.<br />

Delta-T Devices Ltd. (1999). HemiView Canopy Analysis S<strong>of</strong>tware. Version 2.1 SR1. Copyright by Delta-T<br />

Devices Ltd.<br />

Didan, K. (2007). <strong>MODIS</strong> Land Team. Collection 5.0 and beyond. Accessed 3 July 2008 .<br />

Disney, M. (2001). Improved estimation <strong>of</strong> surface biophysical parameters through inversion <strong>of</strong> linear BRDF<br />

models. PhD <strong>the</strong>sis. Accessed 15 December, 2007. .<br />

Dufrêne, E. and N. Bréda (1995). Estimation <strong>of</strong> deciduous <strong>for</strong>est leaf area index using direct and indirect<br />

methods. – In: Oecologia, 104, pp. 156-162.<br />

Eckert, S. (2006). A Contribution to Sustainable Forest Management in Patagonia. – In: Remote Sensing Series,<br />

45, 154 p.<br />

Eggeling, W.J. (1947). Observations on <strong>the</strong> ecology <strong>of</strong> Budongo Rain Forest, Uganda. – In: Journal <strong>of</strong> Ecology,<br />

34, pp. 20-87.<br />

Eklundh, L., Hall, K., Eriksson, H., Ardö, J. and P. Pilesjö, 2003. Investigating <strong>the</strong> use <strong>of</strong> Landsat <strong>the</strong>matic<br />

mapper data <strong>for</strong> estimation <strong>of</strong> <strong>for</strong>est leaf area index in sou<strong>the</strong>rn Sweden. Canadian Journal <strong>of</strong> Remote<br />

Sensing, 29, 349–362.<br />

Eriksson, H., Eklundh, L., Hall, K. and A. Lindroth (2005). Estimating LAI in deciduous <strong>for</strong>est stands.<br />

– In: Agricultural and Forest Meteorology, 129, pp. 27-37.<br />

Eriksson, H., Eklundh, L., Kuusk, A. and T. Nilson (2006). Impact <strong>of</strong> understory vegetation on <strong>for</strong>est canopy<br />

reflectance and remotely sensed LAI estimates. – In: Remote Sensing <strong>of</strong> Environment, 103, pp. 408-418.<br />

Ernsting, A. and D. Rughani (2007). ‚Reduced Emissions From De<strong>for</strong>estation‘: Can Carbon Trading Save Our<br />

Ecosystems? Accessed 12 July, 2008. <br />

155


156<br />

FAO (2006). Global Forest Resources Assessment 2005. Progress towards sustainable <strong>for</strong>est management. FAO<br />

FORESTRY PAPER 147, Rome.<br />

Fashing, P.J., Forrestel, A., Scully, C. and M. Cords (2004). Long-term tree population dynamics and <strong>the</strong>ir<br />

implications <strong>for</strong> <strong>the</strong> conservation <strong>of</strong> <strong>the</strong> Kakamega Forest, Kenya. – In: Biodiversity and Conservation,<br />

13, pp. 753-771.<br />

Fashing, P.J and J. Mwangi Gathual (2004). Spatial variability in <strong>the</strong> vegetation structure and composition <strong>of</strong> an<br />

East African rain <strong>for</strong>est. – In: African Journal <strong>of</strong> Ecology, 42, pp. 189-197.<br />

Fassnacht, K.S., Gower, S.T., Norman, J.M. and R.E. McMurtrie (1994). A comparison <strong>of</strong> optical and direct<br />

methods <strong>for</strong> estimating foliage surface area index in <strong>for</strong>ests. – In: Agricultural and Forest Meteorology,<br />

71, pp. 183-207.<br />

Fassnacht, K.S., Gower, S.T., MacKenzie, M.D., Nordheim, E.V. and T.M. Lillesand (1997). Estimating <strong>the</strong> leaf<br />

area index <strong>of</strong> North Central Wisconsin <strong>for</strong>ests. – In: Remote Sensing <strong>of</strong> Environment, 61, pp. 229-245.<br />

Fensholt, R., Sandholt, I. and M. Schultz Rasmussen (2004). Evaluation <strong>of</strong> <strong>MODIS</strong> LAI, fAPAR and <strong>the</strong> relation<br />

between fAPAR and NDVI in a semi-arid environment using in situ measurements. – In: Remote Sensing<br />

<strong>of</strong> Environment, 91, pp. 490-507.<br />

Ferment, A., Picard, N., Gourlet-Fleury, S. and C. Baraloto (2001). A comparison <strong>of</strong> five indirect methods <strong>for</strong><br />

characterizing <strong>the</strong> light environment in a tropical <strong>for</strong>est. – In: Annals <strong>of</strong> Forest Science, 58, pp. 877-891.<br />

Fernandes, R.A., Butson, C., Leblanc, S.G. and R. Latifovic (2003). Landsat-5 and Landsat-7 ETM+ <strong>based</strong><br />

accuracy assessment <strong>of</strong> leaf area index products <strong>for</strong> Canada derived from SPOT-4 VEGETATION data.<br />

– In: Canadian Journal <strong>of</strong> Remote Sensing, 29 (2), pp. 241-258.<br />

Fernandes, R., and S.G. Leblanc (2005). Parametric (modified least squares) and non-parametric (Theil–<br />

Sen) linear regressions <strong>for</strong> predicting biophysical parameters in <strong>the</strong> presence <strong>of</strong> measurement errors.<br />

– In: Remote Sensing <strong>of</strong> Environment, 95, pp. 303-316.<br />

Foody, G.M. and P.M. Atkinson (2002). Uncertainty in remote sensing and GIS. Chichester, UK, Wiley and<br />

Sons, 326 p..<br />

Frazer, G.W., Canham, C.D. and K.P. Lertzman (1999). Gap Light Analyzer (GLA), Version 2.0: Imaging<br />

s<strong>of</strong>tware to extract canopy structure and gap light transmittance indices from true-colour fisheye<br />

photographs, users manual and program documentation. Copyright © 1999: Simon Fraser University,<br />

Burnaby, British Columbia, and <strong>the</strong> Institute <strong>of</strong> Ecosystem Studies, Millbrook, New York.<br />

Frazer, G.W., Fournier, R.A., Tr<strong>of</strong>ymow, J.A. and R.J. Hall (2001). A comparison <strong>of</strong> digital and film fisheye<br />

photography <strong>for</strong> analysis <strong>of</strong> <strong>for</strong>est canopy structure and gap light transmission. – In: Agricultural and<br />

Forest Meteorology, 109, pp. 249-263.<br />

Gao, X., Huete, A.R. and K. Didan (2003). Multisensor Comparisons and <strong>Validation</strong> <strong>of</strong> <strong>MODIS</strong> Vegetation<br />

Indices at <strong>the</strong> Semiarid Jornada Experimental Range. – In: IEEE Transactions on Geoscience and Remote<br />

Sensing, 41 (10), pp. 2368-2381.<br />

Garrigues, S., Allard, D., Weiss, M. and F. Baret (2002). Comparing VALERI sampling schemes to better<br />

represent high spatial resolution satellite pixel from ground measurements: How to characterize an ESU.<br />

Accessed 30 December, 2007. .<br />

Gascon, F., Gastellu-Etchegorry, J.P., Lefevre-Fonollosa, M.-J. and E. Dufrene (2004). Retrieval <strong>of</strong> <strong>for</strong>est<br />

biophysical variables by inverting a 3-D radiative transfer model and using high and very high resolution<br />

imagery. – In: International Journal <strong>of</strong> Remote Sensing, 25 (24), pp. 5601-5616.


References<br />

Gastellu-Etchegorry, J.P, Zagolski, F. and J. Romier (1996a). A simple anisotropic reflectance model <strong>for</strong><br />

homogeneous multilayer canopies. – In: Remote Sensing <strong>of</strong> Environment, 57, pp. 22-38.<br />

Gastellu-Etchegorry, J.P., Demarez, V., Pinel, V. and F. Zagolski (1996b). Modeling radiative transfer in<br />

heterogeneous 3-D vegetation canopies. – In: Remote Sensing <strong>of</strong> Environment, 58, pp. 131-156.<br />

Gobron, N., Pinty, B. and M.M. Verstraete (1997). Theoretical limits to <strong>the</strong> estimation <strong>of</strong> <strong>the</strong> leaf area index on<br />

<strong>the</strong> basis <strong>of</strong> visible and near-infrared remote sensing data. – In: IEEE Transactions on Geoscience and<br />

Remote Sensing, 35, pp. 1438-1445.<br />

Goel, N.S. and D.E. Strebel (1983). Inversion <strong>of</strong> vegetation canopy reflectance models <strong>for</strong> estimating agronomic<br />

variables. I. Problem definition and initial results using Suits model. – In: Remote Sensing <strong>of</strong> <strong>the</strong><br />

Environment, 13, pp. 487-507.<br />

Goel, N.S. and R.L. Thompson (1984a). Inversion <strong>of</strong> vegetation canopy reflectance models <strong>for</strong> estimating<br />

agronomic variables. III. Estimation using only canopy reflectance data as illustrated by <strong>the</strong> Suits model.<br />

– In: Remote Sensing <strong>of</strong> <strong>the</strong> Environment, 15, pp. 223-236.<br />

Goel, N.S. and R.L. Thompson (1984b). Inversion <strong>of</strong> Vegetation Canopy Reflectance Models <strong>for</strong> Estimating<br />

Agronomic Variables. V. Estimation <strong>of</strong> <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> and Average <strong>Leaf</strong> Angle Using Measured Canopy<br />

Reflectances. – In: Remote Sensing <strong>of</strong> <strong>the</strong> Environment, 16, pp. 69-85.<br />

Goel, N.S. and T. Grier (1988). Estimation <strong>of</strong> Canopy Parameters <strong>for</strong> Inhomogeneous Vegetation Canopies<br />

from Reflectance Data: III. TRIM: A Model <strong>for</strong> Radiative Transfer in Heterogeneous Three-Dimensional<br />

Canopies. – In: Remote Sensing <strong>of</strong> <strong>the</strong> Environment, 25, pp. 255-293.<br />

Gower, S.T., Kucharik, C.J. and J.M. Norman (1999). Direct and indirect estimation <strong>of</strong> leaf area index, fAPAR,<br />

and net primary production <strong>of</strong> terrestrial ecosystems. – In: Remote Sensing <strong>of</strong> Environment, 70, pp. 29-51.<br />

Guen<strong>the</strong>r, B., Xiong, X., Salomonson, V.V., Barnes, W.L. and J. Young (2002). On-orbit per<strong>for</strong>mance <strong>of</strong> <strong>the</strong><br />

Earth Observing System Moderate Resolution Imaging Spectroradiometer; first year <strong>of</strong> data. – In: Remote<br />

Sensing <strong>of</strong> Environment, 83 (1-2), pp. 16-30.<br />

Gullison, R.E., Frumh<strong>of</strong>f, P.C., Canadell, J.G., Field, C.B., Nepstad, D.C., Hayhoe, K., Avissar, R., Curran,<br />

L.M., Friedlingstein, P., Jones, C.D. and C. Nobre (2007). Tropical Forests and Climate Policy.<br />

– In: Science, 316, pp. 985-986<br />

Hale, S.E. and C. Edwards (2002). Comparison <strong>of</strong> film and digital hemispherical photography across a wide<br />

range <strong>of</strong> canopy densities. – In: Agricultural and Forest Meteorology, 112, pp. 51-56.<br />

Hall-Beyer, M. (2007). The GLCM Tutorial Home Page. Version 2.10. Accessed 26 October, 2007.<br />

.<br />

Hamilton, A.C. (1974). Distribution patterns <strong>of</strong> <strong>for</strong>est trees in Uganda and <strong>the</strong>ir historical significance.<br />

– In: Vegetatio, 29, pp. 21-35.<br />

Hamilton, A.C. (1981). The quaternary history <strong>of</strong> African Forests: its relevance to conservation. – In: African<br />

Journal <strong>of</strong> Ecology, 19, pp. 1-6.<br />

Hamilton, A.C. (1982). Environmental history <strong>of</strong> East Africa. A study <strong>of</strong> <strong>the</strong> Quaternary. London.<br />

Hansen, M.C. and R. DeFries (2004). Detecting long-term global <strong>for</strong>est change using continuous fields <strong>of</strong> treecover<br />

maps from 8-km advanced very high resolution radiometer (AVHRR) data <strong>for</strong> <strong>the</strong> years 1982-99.<br />

– In: Ecosystems, 7, pp. 695-716.<br />

157


158<br />

Haralick, R.M., Shanmugan, K. and I. Dinstein (1973). Textural features <strong>for</strong> image classification. – In: IEEE<br />

Transactions on Systems, Man, and Cybernetics, 3 (6), pp. 610–621.<br />

Hardisky, M.A., Klemas, V. and R. M. Smart (1983). The influence <strong>of</strong> soil salinity, growth <strong>for</strong>m, and leaf<br />

moisture on <strong>the</strong> spectral radiance <strong>of</strong> Spartina alterniflora canopies. – In: Photogrammetric Engineering<br />

and Remote Sensing, 49, pp. 77-83.<br />

Heiskanen, J. (2006). Estimating aboveground tree biomass and leaf area index in a mountain birch <strong>for</strong>est using<br />

ASTER satellite data. – In: International Journal <strong>of</strong> Remote Sensing, 27 (6), pp. 1135-1158.<br />

Henry, P. and A. Meygret (2001). Calibration <strong>of</strong> HRVIR and VEGETATION Cameras on SPOT4. – In: Advances<br />

in Space Research, 28 (19), pp. 49-58.<br />

Herwitz, S.R., Peterson, D.L. and J.R. Eastman (1990). Thematic Mapper detection <strong>of</strong> changes in <strong>the</strong> leaf area<br />

<strong>of</strong> closed canopy pine plantations in central Massachusetts. – In: Remote Sensing <strong>of</strong> Environment, 29,<br />

pp. 129-140.<br />

Hester, B. and W. Boberg (2006). Opportunities <strong>for</strong> Mining Investment. Department <strong>of</strong> Geological Survey and<br />

Mines, Republic <strong>of</strong> Uganda.<br />

Hill, M.J, Senarath, U., Lee, A, Zeppel, M., Nightingale, J.M., Williams, R.J. and T.R. McVicar (2006).<br />

Assessment <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product <strong>for</strong> Australian ecosystems. – In: Remote Sensing <strong>of</strong> Environment,<br />

101, pp. 495-518.<br />

Horler, D.N.H. and F. J. Ahern (1986). Forestry in<strong>for</strong>mation content <strong>of</strong> <strong>the</strong>matic mapper data. – In: International<br />

Journal <strong>of</strong> Remote Sensing, 7(3), pp. 405-428.<br />

Howard, P.C. (1991). Nature Conservation in Uganda’s Tropical Forest Reserves. IUCN, Gland, Switzerland<br />

and Cambridge, UK.<br />

Howard, P., Davenport, T. and R. Mat<strong>the</strong>ws (1996). Budongo Forest Reserve, Biodiversity Report, Report No.<br />

3, Forest Department, Kampala, Uganda.<br />

Howard, P., Davenport, T. and F. Kigenyi (1997). Planning conservation areas in Uganda’s natural <strong>for</strong>ests.<br />

– In: Oryx, 31 (4), pp. 253-264.<br />

Huang, D., Yang, W., Tan, B., Rautiainen, M., Zhang, P., Hu, J., Shabanov, N.V., Linder, S., Knyazikhin, Y. and<br />

R.B. Myneni (2006): The Importance <strong>of</strong> Measurement Errors <strong>for</strong> Deriving Accurate Reference <strong>Leaf</strong> <strong>Area</strong><br />

<strong>Index</strong> Maps <strong>for</strong> <strong>Validation</strong> <strong>of</strong> Moderate-Resolution Satellite LAI <strong>Product</strong>s. – In: IEEE Transactions on<br />

Geoscience and Remote Sensing, 44, pp. 1866-1871.<br />

Huemmrich, K.F., Privette, J.L., Mukelabai, M., Myneni, R.B. and Y. Knyazikhin (2005). Time-series validation<br />

<strong>of</strong> <strong>MODIS</strong> land biophysical products in a Kalahari woodland, Africa. – In: International Journal <strong>of</strong><br />

Remote Sensing, 26 (19), pp. 4381-4398.<br />

Hyer, E.J. and S.J. Goetz (2004). Comparison and sensitivity analysis <strong>of</strong> instruments and radiometric methods<br />

<strong>for</strong> LAI estimation: assessments from a boreal <strong>for</strong>est site. – In: Agricultural and Forest Meteorology, 122,<br />

pp. 157-174.<br />

Inoue, A., Yamamoto, K., Mizoue, N. and Y. Kawahara (2004). Effects <strong>of</strong> image quality, size and camera type<br />

on <strong>for</strong>est light environment estimates using digital hemispherical photography. – In: Agricultural and<br />

Forest Meteorology, 126, pp. 89-97.<br />

Jacquemoud, S., Baret, F., Andrieu, B., Danson, F.M. and K. Jaggard (1995). Extraction <strong>of</strong> vegetation<br />

biophysical parameters by inversion <strong>of</strong> <strong>the</strong> PROSPECT + SAIL models on sugar beet canopy reflectance<br />

data. Application to TM and AVIRIS sensors. – In: Remote Sensing <strong>of</strong> Environment, 52, pp. 163-172.


References<br />

Jacquemoud, S., Bacour, C., Poilvé, H. and J.-P. Frangi (2000). Comparison <strong>of</strong> Four Radiative Transfer Models<br />

to Simulate Plant Canopies Reflectance: Direct and Inverse Mode. – In: Remote Sensing <strong>of</strong> Environment,<br />

74, pp. 471-481.<br />

Jensen, R.R. and M.W. Bin<strong>for</strong>d (2004). Measurement and comparison <strong>of</strong> <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> estimators derived<br />

from satellite remote sensing techniques. – In: International Journal <strong>of</strong> Remote Sensing, 25 (20),<br />

pp. 4251-4265.<br />

Jensen, J.R. (2007). Remote sensing <strong>of</strong> <strong>the</strong> environment: an Earth resource perspective, 2nd ed., Prentice Hall.<br />

Jin, M. and D.-L. Zhang (2002). Observed variations <strong>of</strong> leaf area index and its relationship with surface<br />

temperatures during warm seasons. – In: Meteorology and Atmospheric Physics, 80, pp. 117-129.<br />

Jonckheere, I., Fleck, S., Nackaerts, K., Muys, B., Coppin, P., Weiss, M. and F. Baret (2004). Review <strong>of</strong><br />

methods <strong>for</strong> in situ leaf area index determination Part I. Theories, sensors and hemispherical photography.<br />

– In: Agricultural and Forest Meteorology, 121, pp. 19-35.<br />

Jonckheere I., Nackaerts, K. Muys, B. and P. Coppin (2005). Assessment <strong>of</strong> automatic gap fraction estimation<br />

<strong>of</strong> <strong>for</strong>ests from digital hemispherical photography. – In: Agricultural and Forest Meteorology, 132, pp. 96-114.<br />

JPL (2001). ASTER Higher-Level <strong>Product</strong> User Guide. Version 2.0. JPL D-20062, 80 p. Accessed<br />

6 January, 2008. <br />

JPL (2004). ASTER <strong>Product</strong> description AST07. Accessed 10 January, 2008. <br />

Justice, C.O. and J.R.G. Townshend (1981). A comparison <strong>of</strong> unsupervised classification procedures using<br />

Landsat MSS data <strong>for</strong> an area <strong>of</strong> complex surface conditions in Basilicata, sou<strong>the</strong>rn Italy. – In: Remote<br />

Sensing <strong>of</strong> Environment, 12, pp. 407-420.<br />

Justice, C., Belward, A., Morisette, J., Lewis, P., Privette, J. and F. Baret (2000). Developments in <strong>the</strong> ‘validation’<br />

<strong>of</strong> satellite sensor products <strong>for</strong> <strong>the</strong> study <strong>of</strong> <strong>the</strong> land surface. – In: International Journal <strong>of</strong> Remote Sensing,<br />

21 (17), pp. 3383-3390.<br />

Kalácska, M., Sánchez-Az<strong>of</strong>eifa, G.A., Rivard, B., Calvo-Alvarado, J.C., Journet, A.R.P., Arroyo-Mora, J.P.<br />

and D. Ortiz-Ortiz (2004). <strong>Leaf</strong> area index measurements in a tropical moist <strong>for</strong>est: A case study from<br />

Costa Rica. – In: Remote Sensing <strong>of</strong> Environment, 91, pp. 134-152.<br />

Kalácska, M., Calvo-Alvarado, J.C. and G.A. Sánchez-Az<strong>of</strong>eifa (2005). Calibration and assessment <strong>of</strong> seasonal<br />

leaf area index <strong>of</strong> a tropical dry <strong>for</strong>est in different stages <strong>of</strong> succession. – In: Tree physiology, 25, pp.<br />

733-744.<br />

Kanemasu, E.T., Niblett, C.L., Manges, H., Lenhert, D. and M.A. Newman (1974). Wheat: Its growth and<br />

disease severity as deduced from ERTS-1. – In: Remote Sensing <strong>of</strong> Environment, 3, pp. 43-47.<br />

Kanemasu, E.T., Heilman, J.L., Bagley, J.O. and W.L. Powers (1977). Using LANDSAT data to estimate<br />

evapotranspiration <strong>of</strong> winter wheat. – In: Environmental Management, 1, pp. 515-520.<br />

Kang, S., Running, S.W., Lim, J.-H., Zhao, M., Park, C.R. and R. Loehman (2003). A regional phenology model<br />

<strong>for</strong> detecting onset <strong>of</strong> greenness in temperate mixed <strong>for</strong>ests, Korea: an application <strong>of</strong> <strong>MODIS</strong> leaf area<br />

index. – In: Remote Sensing <strong>of</strong> Environment, 86, pp. 232-242.<br />

Kaufmann, M. and C. Troendle (1981) The relationship <strong>of</strong> leaf area and foliage biomass to sapwood conducting<br />

area in four subalpine <strong>for</strong>est tree species. – In: Forest Science, 27 (3), pp. 477-482.<br />

159


160<br />

Kendall, R.L. (1969). An Ecological history <strong>of</strong> <strong>the</strong> Lake Victoria Basin. – In: Ecological Monographs, 39,<br />

pp. 121-176.<br />

Kimes, D.S., Knyazikhin, Y., Privette, J.L., Abuelgasim, A.A. and F. Gao (2000). Inversion methods <strong>for</strong><br />

physically <strong>based</strong> models. – In: Remote Sensing Reviews, 18, pp. 381-439.<br />

Kucharik, C.J., Norman, J.M. and S.T. Gower (1998). Measurements <strong>of</strong> branch area and adjusting leaf area<br />

index indirect measurements. – In: Agricultural and Forest Meteorology, 91, pp. 69-88<br />

Knyazikhin, Y., Martonchik, J.V, Myneni, R.B., Diner, D.J. and S.W. Running (1998a). Synergistic algorithm <strong>for</strong><br />

estimating vegetation canopy leaf area index and fraction <strong>of</strong> absorbed photosyn<strong>the</strong>tically active radiation<br />

from <strong>MODIS</strong> and MISR data. – In: Journal <strong>of</strong> Geophysical Research, 103 (D24), pp. 32,257-32,276.<br />

Knyazikhin, Y., Martonchik, J.V., Diner, D.J., Myneni, R.B., Verstraete, M.M., Pinty, B. and N. Gobron.<br />

(1998b). Estimation <strong>of</strong> vegetation canopy leaf area index and fraction <strong>of</strong> absorbed photosyn<strong>the</strong>tically<br />

active radiation from atmosphere-corrected MISR data. – In: Journal <strong>of</strong> Geophysical Research, 103,<br />

pp. 32,239-32,256.<br />

Knyazikhin, Y., Glassy, J., Privette, J.L., Tian, Y., Lotsch, A., Zhang, Y., Wang, Y., Morisette,<br />

J.T., Votava, P., Myneni, R.B., Nemani, R.R. and S.W. Running (1999). <strong>MODIS</strong> <strong>Leaf</strong> <strong>Area</strong><br />

<strong>Index</strong> (LAI) and Fraction <strong>of</strong> Photosyn<strong>the</strong>tically Active Radiation Absorbed by Vegetation<br />

(FPAR) <strong>Product</strong> (MOD15) Algorithm Theoretical Basis Document. Accessed 8 August, 2007.<br />

< http://modis.gsfc.nasa.gov/data/atbd/atbd_mod15.pdf >.<br />

Kowkaro, J.O. (1988). Conservation status <strong>of</strong> <strong>the</strong> Kakamega Forest in Kenya: <strong>the</strong> easternmost relict <strong>of</strong> <strong>the</strong><br />

equatorial rain <strong>for</strong>ests <strong>of</strong> Africa. – In: Monographs in Systematic Botany from <strong>the</strong> Missouri Botanical<br />

Garden, 25, pp. 471-489.<br />

Kucharik, C.J., Norman, J.M. and S.T. Gower (1998). Measurements <strong>of</strong> branch area and adjusting leaf area<br />

index indirect measurements. – In: Agricultural and Forest Meteorology, 91, 69-88.<br />

Kucharik, C.J., Norman, J.M. and S.T. Gower (1999). Characterization <strong>of</strong> radiation regimes in nonrandom<br />

<strong>for</strong>est canopies: Theory, measurements and a simplified modeling approach. – In: Tree Physiology, 19,<br />

pp. 695-706.<br />

Kuusk, A. (1995). A fast, invertible canopy reflectance model. – In: Remote Sensing <strong>of</strong> Environment, 51, pp.<br />

342-350.<br />

Kuusk, A., Nilson, T. and Paas, M. (2002). Angular distribution <strong>of</strong> radiation beneath <strong>for</strong>est canopies using a<br />

CCD-radiometer. – In: Agricultural and Forest Meteorology, 110, pp. 259-273.<br />

Lambin, E.F. (1999). Monitoring <strong>for</strong>est degradation in tropical regions by remote sensing: some methodological issues.<br />

– In: Global Ecology and Biogeography, 8, pp. 191-198.<br />

Lang, A.R.G., Xiang, Y. and J.M. Norman (1985). Crop structure and <strong>the</strong> penetration <strong>of</strong> direct sunlight.<br />

– In: Agricultural and Forest Meteorology, Vol. 35, pp. 83-101.<br />

Lang, A.R.G. (1986). <strong>Leaf</strong> area and average leaf angle from direct transmission <strong>of</strong> sunlight. – In: Australian<br />

Journal <strong>of</strong> Botany, 34, pp. 349-355.<br />

Lang, A.R.G. and Y. Xiang (1986). Estimation <strong>of</strong> leaf area index from transmission <strong>of</strong> direct sunlight in<br />

discontinuous canopies. – In: Agricultural and Forest Meteorology, 37, pp. 229-243.<br />

Lang, A.R.G., McMurtrie, R.E. and M. Benson (1991). Validity <strong>of</strong> <strong>the</strong> surface area indices <strong>of</strong> Pinus radiata<br />

estimated from transmittance <strong>of</strong> <strong>the</strong> sun‘s beam. – In: Agricultural and Forest Meteorology, 57, pp. 157-170.


References<br />

Langdale-Brown, I., Osmaston, H.A. and J.G. Wilson (1964). The Vegetation <strong>of</strong> Uganda and its bearing on<br />

land-use, Entebbe.<br />

Laumonier, Y., Trichon, V., Gastellu-Etchegorry, J.P. and P. Birnbaum (1994). Reflectance, <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> and<br />

structural studies <strong>of</strong> a rain <strong>for</strong>est canopy using <strong>the</strong> “Opération Canopée” Association’s tree top raft-hot air<br />

airship combination. – In: Geocarto International, 9 (4), pp. 17-29.<br />

Leboeuf, A., Beaudoin, A., Fournier, R.A., Guindon, L., Lu<strong>the</strong>r, J.E. and M.-C. Lambert (2007). A shadow<br />

fraction method <strong>for</strong> mapping biomass <strong>of</strong> nor<strong>the</strong>rn boreal black spruce <strong>for</strong>ests using QuickBird imagery. –<br />

In: Remote Sensing <strong>of</strong> Environment, 110 (4), pp. 488-500.<br />

Leblanc, S.G. and J.M. Chen (2001). A practical scheme <strong>for</strong> correcting multiple scattering effects on optical<br />

LAI measurements. – In: Agricultural and Forest Meteorology, 110, pp. 125-139.<br />

Leblanc, S.G., Fernandes, R. and J.M. Chen (2002). Recent advancements in optical field leaf area index, foliage<br />

heterogeneity, and foliage angular distribution measurements. – In: IEEE International Geoscience and<br />

Remote Sensing Symposium 2002, IGARSS ’02, 5, pp. 2902-2904.<br />

Leblanc, S.G., Chen, J.M., Fernandes, R., Deering, D.W. and A. Conley (2005). Methodology comparison<br />

<strong>for</strong> canopy structure parameters extraction from digital hemispherical photography in boreal <strong>for</strong>ests.<br />

– In: Agricultural and Forest Meteorology, 129, pp. 187-207.<br />

Le Dantec, V., Dufrêne, E. and B. Saugier (2000). Interannual and spatial variation in maximum leaf area index<br />

<strong>of</strong> temperate deciduous stands. – In: Forest Ecology and Management, 134, pp. 71-81.<br />

Lee, K.S., Park, Y.I., Kim, S.H., Park, J.H., Woo, C.S. and K.C. Jang (2004). Remote sensing estimation <strong>of</strong><br />

<strong>for</strong>est LAI in close canopy situation. – In: Proceedings <strong>of</strong> <strong>the</strong> XXth ISPRS Congress, 12-23 July 2004,<br />

Istanbul, Turkey, pp. 400-405.<br />

Li, X. and A.H. Strahler (1985). Geometric-optical modelling <strong>of</strong> a conifer <strong>for</strong>est. – In: IEEE Transactions on<br />

Geoscience and Remote Sensing, 23, pp. 705-721.<br />

Li, X. and A.H. Strahler (1986). Geometric-optical bidirectional reflectance modeling <strong>of</strong> a conifer <strong>for</strong>est canopy.<br />

– In: IEEE Transactions on Geoscience and Remote Sensing, 24, pp. 906-919.<br />

Liang, S., Fang, H., Chen, M., Shuey, C.J., Walthall, C., Daughtry, C., Morisette, J., Schaaf, C. and A. Strahler<br />

(2002). Validating <strong>MODIS</strong> land surface reflectance and albedo products: methods and preliminary results.<br />

– In: Remote Sensing <strong>of</strong> Environment, 83, pp. 149-162.<br />

Liang, S. (2004). Quantitative Remote Sensing <strong>of</strong> Land Surfaces, Wiley and Sons, 534 p.<br />

LI-COR (1992). LAI-2000 plant canopy analyzer operating manual. LI-COR, Lincoln, NE.<br />

LI-COR (2005). LAI-2000 File Viewer. Version 1.04. Copyright by LI-COR, Lincoln, NE.<br />

Lieth, H. and M.J.A. Werger (Eds.) (1989). Tropical rain <strong>for</strong>est ecosystems. Ecosystems <strong>of</strong> <strong>the</strong> world, Bd. 14.<br />

Amsterdam.<br />

Lind, E.M. and M.E.S. Morrison (1974). East African Vegetation. London.<br />

Lu, L., Li, X., Ma, M.G., Che, T., Huang, C.L., Veroustraete, F., Dong, Q.H., Bogaert, J. and R. Ceulemans<br />

(2004). Investigating relationship between Landsat ETM+ data and LAI in a semi-arid grassland <strong>of</strong><br />

Northwest China. – In: IEEE International Geoscience and Remote Sensing Symposium proceedings, 20-<br />

24 September 2004, Anchorage, Alaska.<br />

161


162<br />

Lung, T. (2004). Landbedeckungsänderungen im Gebiet “Kakamega Forest und assoziierte Waldgebiete”<br />

(Westkenia) - Multispektrale Klassifikation von Landsat-Satellitenbilddaten und Auswertung mittels<br />

Methoden im Raster-GIS. – In: Karlsruher Geowissenschaftliche Schriften, A (15), 114 p.<br />

MacDonald, R. (1966). Uganda - Geology: from „The Atlas <strong>of</strong> Uganda, 2nd Edition, 1969“, Dept. <strong>of</strong> Lands and<br />

Surveys, Government <strong>of</strong> Uganda, Entebbe, Uganda.<br />

Macfarlane, C., H<strong>of</strong>fman, M., Eamus, D., Kerp, N., Higginson, S., McMurtie, R. and M. Adams (2007).<br />

Estimation <strong>of</strong> leaf area index in eucalypt <strong>for</strong>est using digital hemispherical photography. – In: Agricultural<br />

and Forest Meteorology, 143, pp. 176-188.<br />

Masuoka, E., Fleig, A.J., Wolfe, R.E. and F.S. Patt (1998). Key characteristics <strong>of</strong> <strong>MODIS</strong> data products.<br />

– In: IEEE Transactions on Geoscience and Remote Sensing, 36 (4), pp. 1313-1323.<br />

Masson, V., Champeaux, J.-L., Chauvin, F., Meriguet, C. and R. Lacaze (2003). A global database <strong>of</strong> Land<br />

Surface Parameters at 1-km Resolution in Meteorological and Climate Models. – In: Journal <strong>of</strong> Climate,<br />

16 (9), pp. 1261-1282.<br />

McCoy, R.M. (2005). Field Methods in Remote Sensing. New York. 159 p.<br />

Miao, G. (2008). Modeling <strong>of</strong> carbon dynamics in an African rain<strong>for</strong>est. A case study with CoupModel in<br />

Kakamega Forest, Kenya. Master <strong>of</strong> Science Thesis at <strong>the</strong> Department <strong>of</strong> Land and Water Resources Engineering,<br />

Royal Institute <strong>of</strong> Technology (KTH), Stockholm, Sweden, 58 p.<br />

Miller, J.B. (1967). A <strong>for</strong>mula <strong>for</strong> average foliage density. – In: Australian Journal <strong>of</strong> Botany, 15, pp. 141-144.<br />

Mitchell, N. and G. Schaab (in print). Developing a disturbance index <strong>for</strong> five East African <strong>for</strong>ests using GIS<br />

to analyse historical <strong>for</strong>est use as an important driver <strong>of</strong> current land use/cover. – In: African Journal <strong>of</strong><br />

Ecology.<br />

Mitchell, N. (2004). The exploitation and disturbance history <strong>of</strong> Kakamega Forest, Western Kenya.<br />

– In: Bielefelder Ökologische Beiträge, 20, 77 p.<br />

Mitchell, N., Lung, T. and G. Schaab (2006). Tracing significant losses and limited gains in <strong>for</strong>est cover <strong>for</strong><br />

<strong>the</strong> Kakamega-Nandi complex in western Kenya across 90 years by use <strong>of</strong> satellite imagery, aerial<br />

photography and maps. – In: Proceedings <strong>of</strong> <strong>the</strong> ISPRS (TC7) Mid-Term Symposium „Remote Sensing:<br />

From Pixels to Processes“, 8-11 May 2006, ITC Enschede (The Ne<strong>the</strong>rlands); ed. by N. Kerle and A.K.<br />

Skidmore, pp. 778-785.<br />

Monsi, M. and T. Saeki (1953). Uber den Lichtfaktor in den Pflanzengesellschaften und seine Bedeutung für die<br />

St<strong>of</strong>fproduktion. – In: Japanese Journal <strong>of</strong> Botany, 14, pp. 22-52.<br />

Monteith, J. (1976). Vegetation and <strong>the</strong> atmosphere. Volume 2 – Case Studies, London, 439 p.<br />

Moran, P.A.P. (1950). Notes on continuous stochastic phenomena. – In: Biometrika, 37, pp. 17-23.<br />

Morisette, J.T., Baret, F., Privette, J.L., Myneni, R.B., Nickeson, J.E., Garrigues, S., Shabanov, N.V., Weiss,<br />

M., Fernandes, R.A., Leblanc, S.G., Kalacska, M., Sánchez-Az<strong>of</strong>eifa, G.A., Chubey, M., Rivard, B.,<br />

Stenberg, P., Rautiainen, M., Voipio, P., Manninen, T., Pilant, A.N., Lewis, T.E., Iiames, J.S., Colombo,<br />

R., Meroni, M., Busetto, L., Cohen, W.B., Turner, D.P., Warner, E.D., Petersen, G.W., Seufert, G. and R.<br />

Cook (2006a). <strong>Validation</strong> <strong>of</strong> global moderate-resolution LAI products: A framework proposed within <strong>the</strong><br />

CEOS Land <strong>Product</strong> <strong>Validation</strong> Subgroup. – In: IEEE Transactions on Geoscience and Remote Sensing,<br />

44, pp. 1804-1817.<br />

Morisette, J.T., Baret, F. and S. Liang (2006b). Special Issue on Global Land <strong>Product</strong> <strong>Validation</strong>.<br />

– In: IEEE Transactions on Geoscience and Remote Sensing, 44 (7), pp. 1695-1698.


References<br />

Musila, W., Kamoni, P.T., Albrecht, A., Todt, H., Uster, D. and H. Dalitz (2004). Soil characteristics <strong>of</strong> Kakamega<br />

Forest, Kenya. – In: H. Dalitz (ed). BIOTA report no. 2, Bielefelder Ökologische Beiträge, 21.<br />

Mussche, S., Samson, R., Nachtergale, L., De Schrijver, A., Lemeur, R. and N. Lust (2001). A comparison <strong>of</strong><br />

optical and direct methods <strong>for</strong> monitoring <strong>the</strong> seasonal dynamics <strong>of</strong> leaf area index in deciduous <strong>for</strong>ests.<br />

– In: Silva Fennica, 35 (4), pp. 373-384.<br />

Mutangah, J.G. (1996). An Investigation <strong>of</strong> Vegetation Status and Process in Relation to Human Disturbance in<br />

Kakamega Forest, Western Kenya, Ph.D. <strong>the</strong>sis, University <strong>of</strong> Wales Aberystwyth, Great Britain.<br />

Myers, N., Mittermeier, R.A., Mittermeier, C.G., da Fonseca, G.A.B. and J. Kent (2000). Biodiversity hotspots<br />

<strong>for</strong> conservation priorities. – In: Nature, 403, pp. 853-858.<br />

Myneni, R.B., Asrar, G. and S.A.W. Gerstl (1990). Radiative transfer in three dimensional leaf canopies.<br />

– In: Transport Theory and Statistical Physics, 19(3-5), pp. 205-250.<br />

Myneni, R.B. and J. Ross (1991). Photon-Vegetation Interactions: Applications in Optical Remote Sensing and<br />

Plant Ecology, Springer-Verlag, Heidelberg, Germany, 560 p.<br />

Myneni, R.B., Maggion, S., Iaquinto, J., Privette, J.L., Gobron, N., Pinty, B., Kimes, D.S., Verstraete, M.M.<br />

and D.L. Williams (1995). Optical remote-sensing <strong>of</strong> vegetation-modeling, caveats, and algorithms.<br />

– In: Remote Sensing <strong>of</strong> Environment, 51, pp. 169-188.<br />

Myneni, R.B., Nemani, R.R. and S.W. Running (1997). Algorithm <strong>for</strong> <strong>the</strong> estimation <strong>of</strong> global land cover,<br />

LAI and FPAR <strong>based</strong> on radiative transfer models. – In: IEEE Transactions on Geoscience and Remote<br />

Sensing, 35, pp. 1380-1393.<br />

Myneni, R.B., H<strong>of</strong>fman, S., Knyazikhin, Y., Privette, J.L., Glassy, J., Tian, Y., Wang, Y., Song, X., Zhang, Y.,<br />

Smith, Y., Lotsch, A., Friedl, M., Morisette, J.T., Votava, P., Nemani, R.R. and S.W. Running (2002).<br />

Global products <strong>of</strong> vegetation leaf area and fraction absorbed PAR from year one <strong>of</strong> <strong>MODIS</strong> data.<br />

– In: Remote Sensing <strong>of</strong> <strong>the</strong> Environment, 83, pp. 214-231.<br />

Myneni, R.B., Shabanov, N., Yang, W., Tan, B., Dong, H., Knyazikhin, Y., Votava, P., Nemani, R. and S.W.<br />

Running (2004). <strong>MODIS</strong> LAI and FPAR: Project Status and <strong>Validation</strong>. Presentation held at LAI<br />

Intercomparison Group Meeting, Missoula, MT, 16 August 2004.<br />

Myneni, R.B., Yang, W., Nemani, R., Huete, R., Dickinson, R., Knyazikhin, Y., Didan, K., Fu, R., Negrón<br />

Juárez, R., Saatchi, S., Hashimoto, H., Ichii, K., Shabanov, N., Tan, B., Ratana, P., Privette, J., Morisette,<br />

J., Vermote, E., Roy, D., Wolfe, R., Friedl, M., Running, S., Votava, P., El-Saleous, N., Devadiga, S., Su,<br />

Y., and V. Salomonson (2007): Large seasonal swings in leaf area <strong>of</strong> Amazon rain<strong>for</strong>ests. – In: PNAS,<br />

104 (12), pp. 4820–4823.<br />

NASA (2007a). Specifications <strong>of</strong> <strong>MODIS</strong> sensor. Accessed 19 November, 2007. .<br />

NASA (2007b). Changes in <strong>MODIS</strong> LAI/FPAR algorithm in collection 5. Accessed 21 November, 2007.<br />

.<br />

Nemani, R., Pierce, L., Running, S., and L. Band (1993). Forest ecosystem processes at <strong>the</strong> watershed scale:<br />

Sensitivity to remotely-sensed leaf area index estimates. – In: International Journal <strong>of</strong> Remote Sensing,<br />

14 (13), 2519-2534.<br />

Neumann, H.H., Hartog, G.D. and R.H. Shaw (1989). <strong>Leaf</strong> area measurements <strong>based</strong> on hemispheric<br />

photographs and leaf-littercollection in a deciduous <strong>for</strong>est during autumn leaf-fall. – In: Agricultural and<br />

Forest Meteorology, 45, pp. 325-345.<br />

163


164<br />

NIES (2007). Tropical De<strong>for</strong>estation. National Institute <strong>for</strong> Environmental Studies, Japan. Accessed 26 October,<br />

2007. <br />

Nilson, T. (1971). A <strong>the</strong>oretical analysis <strong>of</strong> <strong>the</strong> frequency <strong>of</strong> gaps in plant stands. – In: Agricultural Meteorology,<br />

8, pp. 25-38.<br />

Nilson, T. (1999). Inversion <strong>of</strong> gap frequency data in <strong>for</strong>est stands. – In: Agricultural and Forest Meteorology,<br />

98-99, pp. 437-448.<br />

Nilson, T. and A. Kuusk (2005). Improved algorithm <strong>for</strong> estimating canopy indices from gap fraction data in<br />

<strong>for</strong>est canopies. – In: Agricultural and Forest Meteorology, 124, pp. 157-169.<br />

O’Hara, K.L. and N.I. Valappil, (1995). Sapwood-leaf are prediction equations <strong>for</strong> multi-aged ponderosa pine<br />

stands in western Montana and central Oregon. – In: Canadian Journal <strong>of</strong> Forest Research, 25, pp. 1553-<br />

1557.<br />

Ogawa, H., Yoda, K. and T. Kira (1961). A preliminary survey on <strong>the</strong> vegetation <strong>of</strong> Thailand. – In: Kira, T. and<br />

T. Umesao (eds.). Nature and life in sou<strong>the</strong>ast Asia, 1, pp. 21-159.<br />

Pandya, M.R., Singh, R.P., Chaudhari, K.N., Bairagi, G.D., Sharma, R., Dadhwal, V.K. and J.S. Parihar (2006).<br />

<strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> Retrieval Using IRS LISS-III Sensor Data and <strong>Validation</strong> <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI <strong>Product</strong><br />

Over Central India. – In: IEEE Transactions on Geoscience and Remote Sensing, 44 (7), pp. 1858-1866.<br />

Panferov, O., Knyazikhin, Y., Myneni, R.B., Szarzynski, J., Engwald, S., Schnitzler, K.G. and G. Gravenhorst<br />

(2001). The role <strong>of</strong> canopy structure in <strong>the</strong> spectral variation <strong>of</strong> transmission and absorption <strong>of</strong> solar<br />

radiation in vegetation canopies. – In: IEEE Transactions on Geoscience and Remote Sensing, 39,<br />

pp. 241-253.<br />

Parmesan, C. and G. Yohe (2003). A globally coherent fingerprint <strong>of</strong> climate change impacts across natural<br />

systems. – In: Nature, 421, pp. 37-42.<br />

Peddle, D.R. and R.L. Johnson (2000). Spectral mixture analysis <strong>of</strong> airborne remote sensing imagery <strong>for</strong><br />

improved prediction <strong>of</strong> leaf area index in mountainous terrain, Kananaskis, Alberta. – In: Canadian<br />

Journal <strong>of</strong> Remote Sensing, 26, pp. 177-188.<br />

Peterson, D.L., Spanner, M.A., Running, S.W. and K.B. Teuber (1987). Relationship <strong>of</strong> Thematic Mapper<br />

Simulator Data to <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> <strong>of</strong> Temperate Coniferous Forests. – In: Remote Sensing <strong>of</strong><br />

Environment, 22, pp. 323-341.<br />

Pinty, B., Gobron, N., Widlowski, J.-L., Gerstl, S.A.W., Verstraete, M.M., Antunes, M., Bacour, C., Gascon, F.,<br />

Gastellu, J.-P., Goel, N., Jacquemoud, S., North, P., Qin, W. and R. Thompson (2001). Radiation transfer<br />

model intercomparison (RAMI) exercise. – In: Journal <strong>of</strong> Geophysical Research, 106 (D11), pp. 11937-<br />

11956.<br />

Pisek, J. and J.M. Chen (2007). Comparison and validation <strong>of</strong> <strong>MODIS</strong> and VEGETATION global LAI products<br />

over four BigFoot sites in North America. – In: Remote Sensing <strong>of</strong> Environment, 109 (1), pp. 81-94.<br />

Planchais, I. and J.-Y. Pontailler (1999). Validity <strong>of</strong> leaf areas and angles estimated in a beech <strong>for</strong>est from<br />

analysis <strong>of</strong> gap frequencies, using hemispherical photographs and a plant canopy analyser. – In: Annals<br />

<strong>of</strong> Forest Science, 56, pp. 1–10.<br />

Plummer, S., Arino, O., Ranera, F., Chen, J., Dedieu, G., Simon, M., Ciais, P., Bondeau, A., Quegan, S., Schultz,<br />

M. and M. Raupach (2005). The GLOBCARBON initiative: multi-sensor estimation <strong>of</strong> global biophysical<br />

products <strong>for</strong> global terrestrial carbon studies. – In: Proceedings <strong>of</strong> <strong>the</strong> MERIS (A)ATSR Workshop 2005<br />

(ESA SP-597). 26-30 September 2005, ESRIN, Frascati, Italy. Editor: H. Lacoste. Published on CD-ROM.


References<br />

Plumptre, A.J. (1995). The importance <strong>of</strong> “seed trees” <strong>for</strong> <strong>the</strong> natural regeneration <strong>of</strong> selectively logged tropical <strong>for</strong>est.<br />

– In: Commonwealth Forestry Review, 74, pp. 253-258.<br />

Plumptre, A.J. (1996). Changes following 60 years <strong>of</strong> selective timber harvesting in <strong>the</strong> Budongo Forest Reserve,<br />

Uganda. – In: Forest Ecology and Management, 89, pp. 101-113.<br />

Privette, J.L., Myneni, R.B., Knyazikhin, Y., Mukelabai, M., Roberts, G., Tian, Y., Wang, Y. and S.G. Leblanc<br />

(2002). Early spatial and temporal validation <strong>of</strong> <strong>MODIS</strong> LAI product in <strong>the</strong> Sou<strong>the</strong>rn Africa Kalahari.<br />

– In: Remote Sensing <strong>of</strong> Environment, 83, pp. 232-243.<br />

Pollock, R.B. and E.T. Kanemasu (1979). Estimating leaf-area index <strong>of</strong> wheat with Landsat data. – In: Remote<br />

Sensing <strong>of</strong> Environment, 8, pp. 307-312.<br />

Rautiainen, M. (2005). Retrieval <strong>of</strong> leaf area index <strong>for</strong> a coniferous <strong>for</strong>est by inverting a <strong>for</strong>est reflectance<br />

model. – In: Remote Sensing <strong>of</strong> Environment, 99 (3), pp. 295-303.<br />

Reynolds, V. (2005). The Chimpanzees <strong>of</strong> Budongo Forest. Ecology, Behaviour and Conservation. Ox<strong>for</strong>d<br />

University Press. 312 p.<br />

Rich, P.M. (1988). Video image analysis <strong>of</strong> hemispherical canopy photography. – In: Mausel, P.W. (Ed.),<br />

Proceedings <strong>of</strong> <strong>the</strong> First Special Workshop on Videography. American Society <strong>for</strong> Photogrammetry and<br />

Remote Sensing, Terre Haute, IN, 19-20 May, pp. 84–95.<br />

Rich, P.M., (1989). A manual <strong>for</strong> analysis <strong>of</strong> hemispherical canopy photography. Los Alamos National<br />

Laboratory Report LA-11732-M.<br />

Rich, P.M., (1990). Characterizing plant canopies with hemispherical photographs. – In: Remote Sensing<br />

Reviews, 5, pp. 13-29.<br />

Rich, P.M., Clark, D.B., Clark, D.A. and S.F. Oberbauer (1993). Long-term study <strong>of</strong> solar radiation regimes in<br />

a tropical wet <strong>for</strong>est using quantum sensors and hemispherical photography. – In: Agriculture and Forest<br />

Meteorology, 65, pp. 107-127.<br />

Rich, P.M., Wood, J., Vieglais, D.A., Burek, K. and N. Webb (1999). Hemiview User Manual Version 2.1.<br />

Delta-T Devices Ltd., Cambridge, UK.<br />

Richards, J.A. and X. Jia (2006). Remote Sensing Digital Image Analysis: An Introduction. 4 th edition. Springer.<br />

469 p.<br />

Richter, R. (1998). Correction <strong>of</strong> satellite imagery over mountainous terrain. – In: Applied Optics, 37 (18),<br />

pp. 4004-4015.<br />

Richter, R. (2007). Atmospheric / Topographic Correction <strong>for</strong> Satellite Imagery. ATCOR-2/3 User Guide,<br />

Version 6.3, DLR-IB 565-01/07, Wessling, Germany, 134 pp.<br />

Rogers, R. and T.M. Hinckley (1979). Foliar weight and area related to current sapwood area in oak.<br />

– In: Forest Science, 25 (2), pp. 298-303.<br />

Root, T.L., Price, J.T., Hall, K.R., Schneider, S.H., Rosenzweigk, C. and J.A. Pounds (2003). Fingerprints <strong>of</strong><br />

global warming on wild animals and plants. – In: Nature, 421, pp. 57-60.<br />

Rouse, J.W., Haas, R.H., and J.A. Schell (1974). Monitoring <strong>the</strong> vernal advancement <strong>of</strong> retrogradiation <strong>of</strong><br />

natural vegetation. NASA/GSFC, type III, final report, 371 p.<br />

Ross, J. (1981). The Radiation Regime and Architecture <strong>of</strong> Plant Stands, Den Haag, Ne<strong>the</strong>rlands, 391 p.<br />

165


166<br />

Rossello, P. (2005). <strong>Ground</strong> data processing & production <strong>of</strong> <strong>the</strong> level 1 high resolution maps <strong>for</strong> Counami.<br />

Accessed 3 July, 2008. <br />

Rozhnyatovsky, I.I. (1954). Determination <strong>of</strong> leaf area with <strong>the</strong> help <strong>of</strong> light sensitive paper.<br />

– In: Soviet Botanical Journal, 39, 3, pp. 419-420.<br />

Russell, G., Jarvis, P.G. and J.L. Monteith (1989). Absorption <strong>of</strong> radiance by canopies and stand growth.<br />

– In: Russell, G.; Marshall, B.; Jarvis, P.G. (Ed.). Plant canopies: <strong>the</strong>ir growth, <strong>for</strong>m and function.<br />

Cambridge: Cambridge University Press, 1989, pp. 21-40.<br />

Sachs, L. (2004). Angewandte Statistik: Anwendung statistischer Methoden. Springer, Berlin, Deutschland, 890 p.<br />

Saeki, T. (1975). Distribution <strong>of</strong> radiant energy and CO2 in terrestrial communities. – In: Cooper, J.P. (Ed.)<br />

(1975). Photosyn<strong>the</strong>sis and productivity in different environments, Cambridge, pp. 297–322.<br />

Schlerf, M., Atzberger, C. and J. Hill (2005). Remote sensing <strong>of</strong> <strong>for</strong>est biophysical variables using HyMap<br />

imaging spectrometer data. – In: Remote Sensing <strong>of</strong> Environment, 95, pp. 177-194.<br />

Schlüter, T. (1997). Geology <strong>of</strong> East Africa, Berlin, 484 p.<br />

Schmidt, R. (1992). Degradation <strong>of</strong> a Forest in <strong>the</strong> Shimba Hills National Reserve, Kenya, as Initiated by Man<br />

and Maintained by Wildlife. – In: J.G. Goldammer (Ed.) (1992). Tropical Forests in Transition, Basel,<br />

pp. 85-104.<br />

Schneider, S.H., Semenov, S., Patwardhan, A., Burton, I., Magadza, C.H.D., Oppenheimer, M., Pittock, A.B.,<br />

Rahman, A., Smith, J.B., Suarez A. and F. Yamin (2007). Assessing key vulnerabilities and <strong>the</strong> risk from<br />

climate change. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution <strong>of</strong> Working<br />

Group II to <strong>the</strong> Fourth Assessment Report <strong>of</strong> <strong>the</strong> Intergovernmental Panel on Climate Change, M.L. Parry,<br />

O.F. Canziani, J.P. Palutik<strong>of</strong>, P.J. van der Linden and C.E. Hanson (Eds.), Cambridge University Press,<br />

Cambridge, UK, pp. 779-810.<br />

Scholes, R.J., Frost, P.G.H. and Y. Tian (2004). Canopy structure in savannas along a moisture gradient on<br />

Kalahari sands. – In: Global Change Biology, 10 (3), pp. 292-302.<br />

Schultka, W. (1974). Möglichkeiten der Gebüsch- und Waldentwicklung in Grasland- und Feuchtsavannen-<br />

Gebieten von Kenia. Dissertation, Universität Giessen.<br />

Schultka, W. (1975). Gehölz-Sukzessionen in Regenwald- und Feuchtsavannen-Gebieten bei Kakamega (West-<br />

Kenia), Oberhessische Naturwissenschaftliche Zeitschrift, 42, pp. 35-48.<br />

Schulz, A. and T. Wagner (2002). Influence <strong>of</strong> <strong>for</strong>est type and tree species on canopy ants (Hymenoptera:<br />

Formicidae) in Budongo Forest, Uganda. – In: Oecologia, 133, pp. 224-232.<br />

Scurlock, J.M.O., Asner, G.P. and S.T. Gower (2001). Global <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> Data from Field Measurements,<br />

1932-2000. Data set. Available on-line from <strong>the</strong> Oak Ridge National Laboratory Distributed Active<br />

Archive Center, Oak Ridge, Tennessee, U.S.A. Accessed 16 July, 2008. .<br />

Seed, E.D. and D.J. King (2003). Shadow brightness and shadow fraction relations with effective leaf area<br />

index: importance <strong>of</strong> canopy closure and view angle in mixedwood boreal <strong>for</strong>est. – In: Canadian Journal<br />

<strong>of</strong> Remote Sensing, 29 (3), pp. 324-335.<br />

Sellers, P.J., Meeson, B.W., Hall, F.G., Asrar, G., Murphy, R.E., Schiffer, R.A., Bre<strong>the</strong>rton, F.P., Dickinson,<br />

R.E., Ellingson, R.G., Field, C.B., Huemmrich, K.F., Justice, C.O., Melack, J.M., Roulet, N.T., Schimel,<br />

D.S. and P.D. Try (1995). Remote Sensing <strong>of</strong> <strong>the</strong> Land Surface <strong>for</strong> Studies <strong>of</strong> Global Change: Models<br />

Algorithms Experiments. – In: Remote Sensing <strong>of</strong> Environment, 51, pp. 3-26.


References<br />

Sellers, P.J., Tucker, C.J., Collatz, G.J., Los, S.O., Justice, C.O., Dazlich, D.A. and D.A. Randallet (1994).<br />

A global l°x1° NDVI data set <strong>for</strong> climate studies. Part 2: The generation <strong>of</strong> global fields <strong>of</strong> terrestrial<br />

biophysical parameters from <strong>the</strong> NDVI. – In: International Journal <strong>of</strong> Remote Sensing, 15 (17), pp. 3519-<br />

3545.<br />

Sellers, P.J., Los, S.O., Tucker, C.J., Justice, C.O., Dazlich, D.A., Collatz, G.J. and D.A. Randall (1996). A<br />

revised land surface parameterization (SiB2) <strong>for</strong> atmospheric GCM’s. Part II: The generation <strong>of</strong> global<br />

fields <strong>of</strong> terrestrial biophysical parameters from satellite data. – In: Journal <strong>of</strong> Climate, 9, pp. 706-737.<br />

Shabanov, N.V., Knyazikhin, Y., Baret, F. and R.B. Myneni (2000). Stochastic Modeling <strong>of</strong> Radiation Regime<br />

in Discontinuous Vegetation Canopies. – In: Remote Sensing <strong>of</strong> Environment, 74, pp. 125-144.<br />

Shabanov, N.V., Wang, Y., Buermann, W., Dong, J., H<strong>of</strong>fman, S., Smith, G.R., Tian, Y., Knyazikhin, Y. and<br />

R.B. Myneni (2003). Effect <strong>of</strong> foliage spatial heterogeneity in <strong>the</strong> <strong>MODIS</strong> LAI and FPAR algorithm over<br />

broadleaf <strong>for</strong>ests. – In: Remote Sensing <strong>of</strong> Environment, 85, pp. 410-423.<br />

Shabanov, N.V., Huang, D., Yang, W., Tan, B., Knyazikhin, Y., Myneni, R.B., Ahl, D.E., Gower, S.T., Huete,<br />

A.R., Aragão, L.E. and Y.E. Shimabukuro (2005). Optimization <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI and FPAR algorithm<br />

per<strong>for</strong>mance over broadleaf <strong>for</strong>ests. – In: IEEE Transactions on Geoscience and Remote Sensing, 43 (8),<br />

pp. 1855-1865.<br />

Shabanov, N.V., Huang, D., Knjazikhin, Y., Dickinson, R.E. and R.B. Myneni (2007). Stochastic radiative transfer<br />

model <strong>for</strong> mixture <strong>of</strong> discontinuous vegetation canopies. – In: Journal <strong>of</strong> Quantitative Spectroscopy and<br />

Radiative Transfer, 107, pp. 236-262.<br />

Sheil, D. (1995). A critique <strong>of</strong> permanent plot methods and analysis with examples from Budongo Forest,<br />

Uganda. – In: Forest Ecology and Management, 77, pp. 11-34.<br />

Sheil, D. (1998). A half century <strong>of</strong> permanent plot observation in Budongo Forest, Uganda: histories, highlights<br />

and hypo<strong>the</strong>ses. – In: Forest Biodiversity Research, Monitoring and Modeling: Conceptual background<br />

and Old World case studies. Proceedings from <strong>the</strong> 1995 Smithsonian MAB Washington Symposium,<br />

M.A.B. UNESCO, Paris, pp. 399-428.<br />

Sheil, D. (1999). Developing tests <strong>of</strong> successional hypo<strong>the</strong>ses with size-structured populations, and an<br />

assessment using long-term data from a Ugandan rain <strong>for</strong>est. – In: Vegetatio, 140 (1), pp. 117-127.<br />

Smith F.W., Sampson A.D. and N.J. Long (1991). Comparison <strong>of</strong> leaf area index estimates from tree allometrics<br />

and measured light interception. – In: Forest Science, 37, pp. 1682 -1688.<br />

Smith, N.J. (1993). Estimating plant area index and light extinction coefficients in stands <strong>of</strong> Douglas-fir<br />

(Pseudotsuga Menziesii). – In: Canadian Journal <strong>of</strong> Forest Research, 23, pp. 317-321.<br />

Smolander, H. and P. Stenberg (1996). Response <strong>of</strong> LAI-2000 estimates to changes in plant surface area index<br />

in a Scots pine stand. – In: Tree Physiology, 16, pp. 345-349.<br />

Song, C., Woodcock, C.E., Pax-Lenny, M. and S.A. Macomber (2001). Classification and change detection using<br />

Landsat TM date – when and how to correct atmospheric effects. – In: Remote Sensing <strong>of</strong> Environment,<br />

75 (2), pp. 230-244.<br />

Spanner, M.A., Pierce, L.L., Peterson, D.L. and S.W. Running (1990a). Remote sensing <strong>of</strong> temperate coniferous<br />

<strong>for</strong>est leaf area index – <strong>the</strong> influence <strong>of</strong> canopy closure, understory vegetation and background reflectance.<br />

– In: International Journal <strong>of</strong> Remote Sensing, 11, pp. 95-111.<br />

Spanner, M.A., Pierce, L.L., Running, S.W. and D.L. Peterson (1990b). The seasonality <strong>of</strong> AVHRR data <strong>of</strong><br />

temperate coniferous <strong>for</strong>ests: Relationship with leaf area index. – In: Remote Sensing <strong>of</strong> Environment,<br />

33, pp. 97-112<br />

167


168<br />

Spanner, M.A., Johnson, L. and J. Miller (1994). Remote sensing <strong>of</strong> seasonal leaf area index across <strong>the</strong> Oregon<br />

Transect. - In: Ecological Applications, 4 (2), pp. 258-271.<br />

Steinbrecher, R. (2004). Impact <strong>of</strong> changing biodiversity in tropical agro-<strong>for</strong>est ecosystems in Africa on <strong>the</strong><br />

exchange potential <strong>of</strong> trace gases between ecosystems and atmosphere. – In: BIOTA East Africa, Final<br />

report phase I: 2001-2004<br />

Steininger, M.K. (2000). Satellite estimation <strong>of</strong> tropical secondary <strong>for</strong>est above-ground biomass: data from<br />

Brazil and Bolivia. – In: International Journal <strong>of</strong> Remote Sensing, 21 (6 & 7), pp. 1139–1157.<br />

Stenberg, P., Linder, S., Smolander, H. and J. Flower-Ellis (1994). Per<strong>for</strong>mance <strong>of</strong> <strong>the</strong> LAI-2000 plant canopy<br />

analyzer in estimating leaf area index <strong>of</strong> some Scots pine stands. – In: Tree Physiology, 14, pp. 981-995.<br />

Stenberg, P., Rautiainen, M., Manninen, T., Voipio, P. and H. Smolander (2004). Reduced Simple Ratio Better<br />

than NDVI <strong>for</strong> Estimating LAI in Finnish Pine and Spruce Stands. – In: Silva Fennica, 38 (1), pp. 3-15.<br />

Stork, N.E., Balston, J., Farquhar, G.D., Franks, P.J., Holtum, J.A.M. and M.J. Liddell (2007). Tropical rain<strong>for</strong>est<br />

canopies and climate change. – In: Australian Ecology, 32, pp. 105-112.<br />

Strachan, I.B. and J.H. McCaughey (1996). Spatial and Vertical <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> <strong>of</strong> a Decidouous Forest<br />

Resolved Using <strong>the</strong> LAI-2000 Plant Canopy Analyzer. – In: Forest Science, 42 (2), pp. 176-182.<br />

Strachan, I.B. and J.H. McCaughey (1996). Spatial and vertical leaf area index <strong>of</strong> a deciduous <strong>for</strong>est resolved<br />

using <strong>the</strong> LAI-2000 Plant Canopy Analyzer. – In: Forest Science, 42, pp. 176-181.<br />

Suits, G.H. (1972). The Calculation <strong>of</strong> <strong>the</strong> Directional Reflectance <strong>of</strong> a Vegetative Canopy. – In: Remote Sensing<br />

<strong>of</strong> Environment, 2, pp. 183-192.<br />

Tan, B., Hu, J., Zhang, P., Huang, D. and N. Shabanov (2005). <strong>Validation</strong> <strong>of</strong> Moderate Resolution Imaging<br />

Spectroradiometer leaf area index product in croplands <strong>of</strong> Alpilles, France. – In: Journal <strong>of</strong> Geophysical<br />

Research, 110, D01107, doi:10.1029/2004JD004860.<br />

Thomas, S.C. and W.E. Winner (2000). A rotated ellipsoidal angle density function improves estimation <strong>of</strong> foliage<br />

inclination distributions in <strong>for</strong>est canopies. – In: Agricultural and Forest Meteorology, 100, pp. 19-24.<br />

Tian, Y., Woodcock, C.E., Wang, Y., Privette, J.L., Shabanov, N.V., Zhou, L., Zhang, Y., Buermann, W., Dong,<br />

J., Veikkanen, B., Häme, T., Andersson, K., Ozdogan, M., Knyazikhin, Y. and R.B. Myneni (2002a).<br />

Multiscale analysis and validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product over Maun, Botswana, I. Uncertainty<br />

Assessment. – In: Remote Sensing <strong>of</strong> Environment, 83, pp. 414-430.<br />

Tian, Y., Zhang, Y., Knyazikhin, J., Myneni, R.B., Glassy, J., Dedieu, G. and S.W. Running (2000). Prototyping<br />

<strong>of</strong> <strong>MODIS</strong> LAI and FPAR algorithm with LASUR and LANDSAT data. – In: IEEE Transactions on<br />

Geoscience and Remote Sensing, 38 (5), pp. 2387-2401.<br />

Tian, Y., Wang, Y., Zhang, Y., Knyazikhin, Y., Bogaert, J. and R.B. Myneni (2002a). Radiative transfer <strong>based</strong><br />

scaling <strong>of</strong> LAI/FPAR retrievals from reflectance data <strong>of</strong> different resolutions. – In: Remote Sensing <strong>of</strong> <strong>the</strong><br />

Environment, 84, pp. 143-159.<br />

Tian, Y., Woodcock, C.E., Wang, Y., Privette, J.L., Shabanov, N.V., Zhou, L., Zhang, Y., Buermann, W., Dong,<br />

J., Veikkanen, B., Häme, T., Andersson, K., Ozdogan, M., Knyazikhin, Y. and R.B. Myneni (2002b).<br />

Multiscale analysis and validation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI product over Maun, Botswana, II. Sampling<br />

strategy. – In: Remote Sensing <strong>of</strong> Environment, 83, pp. 431-441.<br />

Turner, D.P., Cohen, W.B., Kennedy,R.E., Fassnacht, K.S. and J.M. Briggs (1999). Relationships between <strong>Leaf</strong><br />

<strong>Area</strong> <strong>Index</strong> and Landsat TM Spectral Vegetation Indices across Three Temperate Zone Sites. – In: Remote<br />

Sensing <strong>of</strong> <strong>the</strong> Environment, 70, pp. 52-68.


References<br />

USGS (2005). Landsat: A Global Land-Observation Project. Accessed 18 December, 2007.<br />

.<br />

USGS (2006a). ASTER Shortwave Infrared (SWIR) Crosstalk-Corrected <strong>Product</strong>s. Accessed 16 November,<br />

2007. .<br />

USGS (2006b). Data Set characteristics <strong>of</strong> <strong>MODIS</strong>/Terra <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong>/FPAR 8-day L4 Global 1km SIN<br />

Grid. Accessed 19 November, 2007. .<br />

Van Gardingen, P.R., Jackson, G.E., Hernandez-Daumas, S., Russell, G. and L. Sharp (1999). <strong>Leaf</strong> area index<br />

estimates obtained <strong>for</strong> clumped canopies using hemispherical photography. – In: Agricultural and Forest<br />

Meteorology, 94, pp. 243-257.<br />

Verhoef, W. and N.J.J. Bunnik (1981). Influence <strong>of</strong> crop geometry on multispectral reflectance determined<br />

by <strong>the</strong> use <strong>of</strong> canopy reflectance models. – In: NLR MP 81042 U / Proceedings <strong>of</strong> <strong>the</strong> International<br />

Colloquium on spectral signatures <strong>of</strong> objects in remote sensing, Avignon (France), 8-11 September 1981,<br />

pp. 273-289.<br />

Verhoef, W. (1984). Light scattering by leaf layers with application to canopy reflectance. – In: Remote Sensing<br />

<strong>of</strong> Environment, 16, pp. 125-141.<br />

Vermote, E.F., Tanré, D., Deuze, J.L., Herman, M. and J.J. Morcrette (1997). Second simulation <strong>of</strong> <strong>the</strong> satellite<br />

signal in <strong>the</strong> solar spectrum, 6S: an overview. – In: IEEE Transactions on Geoscience and Remote<br />

Sensing, 35, pp. 675-686.<br />

Vertessy, R.A., Benyon, R.G., O’Sullivan, S.K. and P.R. Gribben (1995), Relationships between stem diameter,<br />

sapwood area, leaf area and transpiration in a young mountain ash <strong>for</strong>est. – In: Tree Physiology, 15,<br />

pp. 559-567.<br />

Walter, J.-M.N., Fournier, R.A., Soudani, K. and E. Meyer (2003). Integrating clumping effects in <strong>for</strong>est canopy<br />

structure: an assessment through hemispherical photographs. – In: Canadian Journal <strong>of</strong> Remote Sensing,<br />

29 (3), pp. 388-410.<br />

Wang, Y.S. and D.R. Miller (1987). Calibration <strong>of</strong> <strong>the</strong> hemispherical photographic technique to measure leaf<br />

area index distributions in hardwood <strong>for</strong>ests. – In: Forest Science 33, pp. 210-226.<br />

Wang, Y., Tian, Y., Zhang, Y., El-Saleous, N., Knyazikhin, Y., Vermote, E. and R.B. Myneni (2001). Investigation<br />

<strong>of</strong> product accuracy as a function <strong>of</strong> input and model uncertainties: Case study with SeaWiFS and <strong>MODIS</strong><br />

LAI/FPAR algorithm. – In: Remote Sensing <strong>of</strong> Environment, 78, pp. 296-311.<br />

Wang, Y., Buermann, W., Stenberg, P., Smolander, H., Häme, T., Tian, Y., Hu, J., Knyazikhin, Y. and R.B.<br />

Myneni (2003). A new parameterization <strong>of</strong> canopy spectral response to incident solar radiation: Case<br />

study with hyperspectral data from pine dominant <strong>for</strong>est. – In: Remote Sensing <strong>of</strong> <strong>the</strong> Environment, 85<br />

(3), pp. 304-315.<br />

Wang, Y., Woodcock, C.E., Buermann, W., Stenberg, P., Voipio, P., Smolander, H. Häme, T., Tian, Y., Hu, J., Knyazikhin,<br />

Y. and R.B. Myneni (2004). Evaluation <strong>of</strong> <strong>the</strong> <strong>MODIS</strong> LAI algorithm at a coniferous <strong>for</strong>est site in Finland.<br />

– In: Remote Sensing <strong>of</strong> Environment, 91, pp. 114-127.<br />

Wang, Q., Tenhunen, J., Quoc Dinh, N., Reichstein, M., Otieno, D., Granier, A. and K. Pilegarrd (2005).<br />

Evaluation <strong>of</strong> seasonal variation <strong>of</strong> <strong>MODIS</strong> derived leaf area index at two European deciduous broadleaf<br />

<strong>for</strong>est sites. – In: Remote Sensing <strong>of</strong> Environment, 96, pp. 475-484.<br />

Warren Wilson, J. (1959). Analysis <strong>of</strong> <strong>the</strong> spatial distribution <strong>of</strong> foliage by two-dimensional point quadrats.<br />

– In: New Phytology, 58, pp. 92-101.<br />

169


170<br />

Warren Wilson, J. (1960). Inclined point quadrats. – In: New Phytology, 59, pp. 1-8.<br />

Warren Wilson, J. (1963). Estimation <strong>of</strong> foliage denseness and foliage angle by inclined point quadrats.<br />

– In: Australian Journal <strong>of</strong> Botany, 11, pp. 95-105.<br />

Watson, D.J. (1947). Comparative physiological studies in <strong>the</strong> growth <strong>of</strong> field crops. I. Variation in net<br />

assimilation rate and leaf area between species and varieties, and within and between years. – In: Annals<br />

<strong>of</strong> Botany, 11, pp. 41-76.<br />

Weiss, M. (2006). CAN-EYE output variables. Definitions and <strong>the</strong>oretical background. Accessed 21 September,<br />

2007. .<br />

Weiss, M. and F. Baret (1999). Evaluation <strong>of</strong> canopy biophysical variable retrieval per<strong>for</strong>mances from <strong>the</strong><br />

accumulation <strong>of</strong> large swath satellite data. – In: Remote Sensing <strong>of</strong> Environment, 70, pp. 293-306.<br />

Weiss, M., Baret, F., Myneni, R.B., Pragnère, A. and Y. Knyazikhin (2000). Investigation <strong>of</strong> a model inversion<br />

technique to estimate canopy biophysical variables from spectral and directional reflectance data.<br />

– In: Agronomie, 20, pp. 3-22.<br />

Weiss, M., Baret, F., Smith, G. J., Jockheere, I. and P. Coppin (2004). Review <strong>of</strong> methods <strong>for</strong> in situ leaf area<br />

index (LAI) determination: Part II. Estimation <strong>of</strong> LAI, errors and sampling. – In: Forest Ecology and<br />

Management, 121, pp. 37-53.<br />

Welles, J.M. (1990). Some indirect methods <strong>of</strong> estimating canopy structure. – In: Remote Sensing Reviews, 5,<br />

pp. 31-43.<br />

Welles, J.M. and J.M. Norman (1991). Instrument <strong>for</strong> indirect measurement <strong>of</strong> canopy architecture.<br />

– In: Agronomy Journal, 83 (5), pp. 818-825.<br />

WGCW (2007). Official Website <strong>of</strong> <strong>the</strong> Working Group on Calibration and <strong>Validation</strong>. Accessed 15 August,<br />

2007. .<br />

White, F. (1983). The vegetation <strong>of</strong> Africa. A descriptive memoir to accompany <strong>the</strong> Unesco/AETFAT/UNSO<br />

vegetation map <strong>of</strong> Africa. Natural resources research (XX). UNESCO, Switzerland, 356 p.<br />

White, J., Running, S., Nemani, R., Keane, R. and K. Ryan (1997). Measurement and remote sensing <strong>of</strong> LAI in<br />

Rocky Mountain montane ecosystems. – In: Canadian Journal <strong>of</strong> Forest Research, 27, pp. 1714-1727.<br />

White, M.A., Asner, G.P., Nemani, R.R., Privette, J.L. and S.W. Running (2000). Measuring Fractional Cover<br />

and <strong>Leaf</strong> <strong>Area</strong> <strong>Index</strong> in Arid Ecosystems: Digital Camera, Radiation Transmittance, and Laser Altimetry<br />

Methods. – In: Remote Sensing <strong>of</strong> <strong>the</strong> Environment, 74, pp. 45-57.<br />

Whitmore, T.C. (2003). An Introduction to Tropical Rain Forests. Ox<strong>for</strong>d University Press, Ox<strong>for</strong>d, 296 p.<br />

Wiegand, C.L., Gausman, H.W., Cuellar, J.A., Gerbermann, A.H. and A.J. Richardson (1974). Vegetation<br />

density as deduced from ERTS-1 MSS response. – In: Third ERTS Symposium, NASA-SP-351, I (A),<br />

pp. 93-116.<br />

Wilson, D., Alm, J., Riutta, T., Laine, J., Byrne, K.A., Farrell, E.P. and E.-S. Tuittila (2007). A high resolution<br />

green area index <strong>for</strong> modelling <strong>the</strong> seasonal dynamics <strong>of</strong> CO2 exchange in peatland vascular plant<br />

communities. – In: Plant Ecology, 190 (1), pp. 37-51.<br />

Wirth, R., Weber, B. and R.J. Ryel (2001). Spatial and temporal variability <strong>of</strong> canopy structure in a tropical<br />

moist <strong>for</strong>est. – In: Acta Oecologica, 22, pp. 235-244.


References<br />

Wright, S.J. (2005). Tropical <strong>for</strong>ests in a changing environment. – In: Trends in Ecology and Evolution, 20 (10),<br />

pp. 553-560.<br />

Wright, S.J. and H.C. Muller-Landau (2006). The future <strong>of</strong> tropical <strong>for</strong>est species. – In: Biotropica, 38 (3),<br />

pp. 287-301.<br />

Wolfe, R.E., Nishihama, M., Fleig, A.J., Kuyper, J.A., Roy, D.P., Storey, J.C. and F.S. Patt (2002). Achieving<br />

sub-pixel geolocation accuracy in support <strong>of</strong> <strong>MODIS</strong> land science. – In: Remote Sensing <strong>of</strong> Environment,<br />

83 (1-2), pp. 31-49.<br />

Wulder, M.A., LeDrew, E.F., Franklin, S.E. and M.B. Lavigne (1998). Aerial image texture in<strong>for</strong>mation in <strong>the</strong><br />

estimation <strong>of</strong> nor<strong>the</strong>rn deciduous and mixed wood <strong>for</strong>est leaf area index (LAI). – In: Remote Sensing <strong>of</strong><br />

<strong>the</strong> Environment, 64, pp. 64-76.<br />

Wulder, M.A. and S.E. Franklin (2003). Remote Sensing <strong>of</strong> Forest Environments: Concepts and Case Studies.<br />

Kluwer Academic Publishers, 552 p.<br />

Xiao, X., Zhang, Q., Saleska, S., Hutyra, L., De Camargo, P., W<strong>of</strong>sy, S., Frolking, S., Boles, S., Keller, M. and<br />

B. Moore (2005). Satellite-<strong>based</strong> modeling <strong>of</strong> gross primary production in a seasonally moist tropical<br />

evergreen <strong>for</strong>est. – In: Remote Sensing <strong>of</strong> Environment, 94, pp. 105-122.<br />

Yamashita, M., Yoshimura, M. and T. Nakashizuka (2004). Cloud cover estimation using multitemporal<br />

hemisphere images. – In: Proceedings <strong>of</strong> <strong>the</strong> XXth ISPRS Congress, Geo-Imagery Bridging Continents,<br />

12-23 July 2004, Istanbul, Turkey. Accessed 2 January, 2008. < http://www.isprs.org/istanbul2004/<br />

comm7/papers/162.pdf>.<br />

Yang, W., Tan, B., Huang, D., Rautiainen, M., Shabanov, N.V., Wang, Y., Privette, J.L., Huemmrich, K.F.,<br />

Fensholt, I., Weiss, M., Ahl, D.E., Gower, S.T., Nemani, Y., Knyazikhin, Y. and R.B. Myneni (2006).<br />

<strong>MODIS</strong> leaf area index products: From validation to algorithm improvement. – In: IEEE Transactions on<br />

Geoscience and Remote Sensing, 44, pp. 1885-1898.<br />

Zhang, Y., Tian, Y., Knyazikhin, J., Martonchik, J.V., Diner, D.J., Leroy, M. and R.B. Myneni (2000).<br />

Prototyping <strong>of</strong> MISR LAI and FPAR algorithm with POLDER data over Africa. – In: IEEE Transactions<br />

on Geoscience and Remote Sensing, 38 (5), pp. 2402-2418.<br />

Zhang, X., Friedl, M.A., Schaaf, C.B. and A.H. Strahler (2004). Climate controls on vegetation phenological<br />

patterns in nor<strong>the</strong>rn mid- and high latitudes inferred from <strong>MODIS</strong> data. – In: Global Change Biology, 10,<br />

pp. 1133-1145.<br />

Zhang, Q., Xiao, X, Braswell, B., Linder, E., Baret, F. and B. Moore III (2005). Estimating light absorption by<br />

chlorophyll, leaf and canopy in a deciduous broadleaf <strong>for</strong>est using <strong>MODIS</strong> data and a radiative transfer<br />

model. – In: Remote Sensing <strong>of</strong> Environment, 99, pp. 357-371.<br />

Zhang, Q., Xiao, X., Braswell, B., Linder, E., Ollinger, S., Smith, M.-L., Jenkins, J.P., Baret, F., Richardson,<br />

A.D., Moore, B. and R. Minocha (2006). Characterization <strong>of</strong> seasonal variation <strong>of</strong> <strong>for</strong>est canopy in a<br />

temperate deciduous broadleaf <strong>for</strong>est, using daily <strong>MODIS</strong> data. – In: Remote Sensing <strong>of</strong> Environment,<br />

105, pp. 189-203.<br />

171


172


Appendix <strong>of</strong> Figures<br />

Figure A-1<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA with 45° view cap in a) vertical projection and b) horizontal projection (solid<br />

lines represent field <strong>of</strong> view assuming an average canopy height <strong>of</strong> 20 m, dashed lines represent average canopy<br />

height <strong>of</strong> 10 m).<br />

173


174<br />

Figure A-2<br />

Land cover classification, 2003<br />

Near natural & old secondary <strong>for</strong>est<br />

Secondary <strong>for</strong>est<br />

Bushland / shrubs<br />

Secondary bushland - Psidium guajava<br />

Grassland with scattered trees<br />

Grassland<br />

Plantation <strong>for</strong>est - Pinus patula<br />

Plantation <strong>for</strong>est - Bisch<strong>of</strong>fia javanica<br />

Tea plantation<br />

Agricultural land<br />

Water<br />

O<strong>the</strong>rs<br />

T. Lung, 2007<br />

18/05/2003 Landsat 7 (ETM+)<br />

10/01/2003 Landsat 7 (ETM+)<br />

0 2.55km Land cover classification <strong>for</strong> Kakamega Forest, <strong>based</strong> on Landsat 7 ETM+ data <strong>of</strong> 2003 (courtesy <strong>of</strong> T. Lung, BIOTA<br />

E02, <strong>for</strong> methodology refer to Lung 2004).


Appendix <strong>of</strong> Figures<br />

Figure A-3<br />

Interactive classification process in CAN-EYE. a) unclassified and b) classified hemispherical photographs covering<br />

one transect.<br />

175


176<br />

Left late<br />

Figure 5. Schema <strong>of</strong> <strong>the</strong> experimental design used.<br />

Diameter <strong>of</strong> <strong>the</strong> image<br />

Left corner<br />

Left lateral ruler<br />

Camera<br />

Background alignment<br />

Front Nail<br />

alignement<br />

<br />

Optical centre<br />

Perpandicular ruler<br />

Right corner<br />

Right lateral ruler<br />

Figure 6. Example <strong>of</strong> an image <strong>of</strong> <strong>the</strong> experimental design taken with <strong>the</strong> hemispherical camera and<br />

used <strong>for</strong> <strong>the</strong> calibration <strong>of</strong> <strong>the</strong> projection function. The horizontal doted yellow line corresponds to <strong>the</strong><br />

diameter <strong>of</strong> <strong>the</strong> image passing through <strong>the</strong> optical centre (defined by its coordinates as measured<br />

previously). The camera is aligned Background thanks to <strong>the</strong> front nail and background line.<br />

alignement<br />

A dedicated matlab code ‘calib_project.m’ was developed to display <strong>the</strong> results.<br />

Front nail<br />

alignement<br />

Axis <strong>of</strong> <strong>the</strong> design<br />

<br />

Right la<br />

Perpandicular ruler<br />

Figure A-4 Figure 5. Schema <strong>of</strong> <strong>the</strong> experimental design used.<br />

Diameter <strong>of</strong> <strong>the</strong> image<br />

y<br />

Left lateral ruler<br />

W<br />

y<br />

x<br />

H<br />

Background alignment<br />

x<br />

Camera<br />

Experimental design to assess <strong>the</strong> projection quality <strong>of</strong> <strong>the</strong> used camera systems and to retrieve <strong>the</strong> projection<br />

function (Baret 2004b).<br />

Right lateral ruler


Appendix <strong>of</strong> Figures<br />

Figure A-5<br />

Figure A-6<br />

Results <strong>of</strong> different LAI measures <strong>for</strong> each ESU in Budongo Forest.<br />

Results <strong>of</strong> different LAI measures <strong>for</strong> each ESU in Kakamega Forest.<br />

177


178<br />

Figure A-7 a) ρred<br />

, b) ρNIR and c) ρSWIR <strong>of</strong> band 4 and d) ρ <strong>of</strong> band 5 in [%] derived from ASTER data and plotted per<br />

SWIR<br />

<strong>for</strong>est stage. Upper and lower box boundaries indicate 25th and 75th percentiles, whiskers indicate minimum and<br />

maximum values except outliers (defined as more than 3 box lengths from 25 th or 75 th percentile, displayed as point).<br />

Mean is shown as solid line.


Appendix <strong>of</strong> Figures<br />

Figure A-8 a) ρred<br />

, b) ρNIR and c) ρ in [%] derived from SPOT data and plotted per <strong>for</strong>est stage <strong>for</strong> Kakamega Forest. Upper<br />

SWIR<br />

and lower box boundaries indicate 25th and 75th percentiles, whiskers indicate minimum and maximum values except<br />

outliers (defined as more than 3 box lengths from 25 th or 75 th percentile, displayed as points). Mean is shown as solid<br />

line.<br />

Bivariate plot <strong>of</strong> LAI<br />

Figure A-9 e (ASTER) in 15 m and 1000 m spatial resolution <strong>for</strong> <strong>the</strong> <strong>for</strong>est compartments included in LAI<br />

product validation (N=48).<br />

179


180<br />

Figure A-10<br />

not retrieved<br />

empirical<br />

physical saturated<br />

physical perfect<br />

Relative frequencies <strong>of</strong> C4 <strong>MODIS</strong> LAI algorithm<br />

usage <strong>for</strong> <strong>the</strong> years 2000-2006 over <strong>the</strong> Budongo<br />

Forest test site.


Appendix <strong>of</strong> Tables<br />

Table A-1<br />

Logging history <strong>of</strong> Budongo Forest compartments according to Forest Department Offices, logging return sheets <strong>of</strong><br />

sawmills and old working plans (Plumptre 1996). Note that <strong>the</strong> Budongo <strong>for</strong>est block is fur<strong>the</strong>r split up in <strong>the</strong><br />

Waibira, Nyakafunjo and Biiso blocks (from east to west). The table has not been updated. Various additional logging<br />

activities could be observed during <strong>the</strong> time <strong>of</strong> field work.<br />

Forest Block Comp. <strong>Area</strong><br />

(ha)<br />

Date<br />

logged<br />

Total<br />

(m 3 ha -1 )<br />

Mahogany<br />

(m 3 ha -1 )<br />

Date<br />

treated<br />

181<br />

Arboricide<br />

Biiso B1 582 1935 19.9 12.3 1957-58 n/a<br />

(l ha -1 )<br />

n/a 1981-82 21.5 12.9 None n/a<br />

B2 600 1936-38 40.2 33.9 1958-59 20.4<br />

B3 867 1939-40 34.9 23.7 1959-60 n/a<br />

B4 748 1941-42 34.8 19.8 1955-57 20.0<br />

n/a 1985-92 Some pitsawing n/a n/a<br />

B5 871 1943-44 34.2 25.2 1960-61 42.6<br />

B6 394 1943 14.0 10.4 1961-63 42.6<br />

B7 288 1944 20.2 14.7 n/a n/a<br />

Nyakafunjo N1 412 1945 58.7 47.1 1962-63 n/a<br />

N2 630 1945-47 46.2 33.1 1955-56 11-1<br />

N3 620 1947-52 80.0 39.1 1959-61 22.0<br />

N4 341 1952-54 94.0 48.0 1960-62 n/a<br />

N5 568 1954 51.5 41.2 1963-64 38.1<br />

N6 600 1956-57 59.8 49.4 1956-58 41.6<br />

N7 474 1958-59 31.9 26.1 1957-58 44.8<br />

N8 498 1957-58 25.7 20.8 1957 47.0<br />

N9 498 1958-59 15.2 10.7 1958 42.3<br />

N10 618 1959-60 30.2 23.5 1958-60 50.4<br />

N11 564 1960 26.5 21.7 1958-60 42.6<br />

N12 389 1960 3.0 2.5 1959-60 32.9<br />

N13 610 1960 13.0 8.5 1960 40.3<br />

N14 535 1961-62 n/a n/a None n/a<br />

N15 777 Unlogged n/a n/a None n/a


182<br />

Forest Block Comp. <strong>Area</strong><br />

(ha)<br />

Date<br />

logged<br />

Total<br />

(m 3 ha -1 )<br />

Mahogany<br />

(m 3 ha -1 )<br />

Date<br />

treated<br />

Arboricide<br />

WAIBIRA W16 543 1961-62 21.7 14.3 1961-62 n/a<br />

(l ha -1 )<br />

W17 609 Unlogged n/a n/a None n/a<br />

W18 681 1960-61 28.8 19.3 1960-62 43.6<br />

W19 886 1962-63 25.6 16.3 1961-63 41.4<br />

W20 569 1963/64 30.8 20.3 1962-63 41.4<br />

W21 1116 1963-64 36.1 23.9 1963-64 29.1<br />

W22 1036 1965-67 35.9 23.0 1966 n/a<br />

W23 736 1966-68 26.4 18.5 1966 n/a<br />

W24 710 1967-69 41.2 9.3 1966-67 n/a<br />

W25 1181 1968-69 9.6 5.4 1970-71 n/a<br />

W26 781 1970-71 30.6 19.1 1971-72 36.0<br />

W27 791 1970-71 n/a n/a 1970-71 46.5<br />

W28 1010 1972-73 51.4 29.1 1972-73 29.5<br />

W29 744 1972-74 64.0 40.5 None n/a<br />

W30 580 1974-75 8.1 7.5 None n/a<br />

W31 689 Unlogged n/a n/a None n/a<br />

W32 968 Unlogged n/a n/a None n/a<br />

W33 731 Unlogged n/a n/a None n/a<br />

W34 751 Unlogged n/a n/a None n/a<br />

W35 680 Pitsawn n/a n/a None n/a<br />

W36 591 Pitsawn n/a n/a None n/a<br />

W37 566 1978-83 22.2 10.7 None n/a<br />

W38 422 1983 (Partially pitsawn) None n/a<br />

W39 667 1974-76 43.5 25.5 None n/a<br />

W40 996 Pitsawn n/a n/a None n/a<br />

W41 691 Pitsawn n/a n/a None n/a<br />

W42 804 Pitsawn n/a n/a None n/a<br />

W43 829 Pitsawn n/a n/a None n/a<br />

KANIYO-PABIDI K1 177 1970-72 11.7 9.4 n/a<br />

K2 533 1970-72 3.1 2.3 n/a<br />

K3 330 1977 n/a n/a None<br />

K4 630 1987-92 20.7 12.6 None n/a<br />

K5 311 1985-87 25.6 17.2 None n/a<br />

K7 526 1987-89 11.1 9.0 None n/a<br />

K8 169 1986-89 31.5 15.1 None n/a<br />

K11 420 Unlogged n/a n/a None n/a<br />

K12 1097 Unlogged n/a n/a None n/a<br />

K13 625 Unlogged n/a n/a None n/a


Appendix <strong>of</strong> Tables<br />

Forest Block Comp. <strong>Area</strong><br />

(ha)<br />

Date<br />

logged<br />

Total<br />

(m 3 ha -1 )<br />

Mahogany<br />

(m 3 ha -1 )<br />

Date<br />

treated<br />

183<br />

Arboricide<br />

SIBA S1 810 1963-69 40.9 23.5 1972-73 34.4<br />

Table A-2<br />

Quality control FPARLAI_QC SDS <strong>of</strong> MOD15A2 (USGS 2006b).<br />

Bit range Bit setting Description<br />

0 - 1 MODLAND QA<br />

(l ha -1 )<br />

S2 602 1969-70 n/a n/a 1971-72 40.0<br />

S3 528 1966-70 n/a n/a n/a n/a<br />

S4 799 1972-77 28.9 n/a n/a n/a<br />

S5 829 1971 n/a n/a n/a n/a<br />

S6 588 1971 n/a n/a 1972-75 n/a<br />

S7 686 1990 n/a n/a None n/a<br />

S8 446 1979/90 48.3 26.3 None n/a<br />

00 Best possible<br />

01 OK, but not <strong>the</strong> best<br />

10 Not produced due to cloud<br />

11 Not produced due to o<strong>the</strong>r reasons<br />

2 DEAD DETECTOR<br />

3 - 4 CLOUD STATE<br />

5 - 7 SCF_QC<br />

00 Detectors apparently fine <strong>for</strong> up to 50% <strong>of</strong> channels 1,2<br />

01 Dead detectors caused >50% adjacent detector retrieval<br />

00 Significant clouds NOT present (clear)<br />

01 Significant clouds WERE present<br />

10 Mixed cloud present on pixel<br />

11 cloud state not defined, assumed clear<br />

000 Main(RT) method used with <strong>the</strong> best possible results<br />

001 Main(RT) method used with saturation<br />

010 Main(RT) method failed due to geometry problems, empirical method used<br />

011 Main(RT) method failed due to problems o<strong>the</strong>r than geometry, empirical method used<br />

100 Pixel could not be retrieved


184<br />

Table A-3<br />

Additional FparExtra_QC SDS <strong>of</strong> MOD15A2 (USGS 2006b).<br />

Bit range Bit setting Description<br />

0-1 SCF_QC<br />

Table A-4<br />

00 Land<br />

01 Shore<br />

10 Freshwater<br />

11 Ocean<br />

2 SNOW_ICE<br />

0 No snow, ice detected<br />

1 Snow, ice detected<br />

3 AEROSOL<br />

4 CIRRUS<br />

0 No or low atmospheric aerosol levels detected<br />

1 Average or high aerosol levels detected<br />

0 No cirrus detected<br />

1 Cirrus was detected<br />

5 INTERNAL_CLOUD_MASK<br />

0 No clouds detected<br />

1 Clouds WERE detected<br />

6 CLOUD_SHADOW<br />

0 No cloud shadow detected<br />

7 SCF_MASK<br />

1 Cloud shadow was detected<br />

0 Custom SCF mask, EXCLUDE this pixel<br />

1 Custom SCF mask, INCLUDE this pixel<br />

Fill values <strong>for</strong> non-vegetated surfaces in MOD15A2 (USGS 2006b).<br />

DN Non-Terrestrial Value <strong>Index</strong><br />

255 Fill Value<br />

254 Perennial salt or inland fresh water<br />

253 Barren, sparse vegetation (rock, tundra, desert)<br />

252 Perennial snow, ice<br />

251 Permanent wetlands/inundated marshland<br />

250 Urban/built-up<br />

249 Unclassified/not able to determine


Appendix <strong>of</strong> Tables<br />

Table A-5<br />

Image parameters <strong>for</strong> camera systems used in Budongo and Kakamega Forest.<br />

Camera Image size (lines, rows) Optical center (line, row) Diameter<br />

Nikon Coolpix 4300 1704, 2272 1180, 862 1593<br />

Nikon Coolpix 4500 1704, 2272 1152, 831 1584<br />

Table A-6 Spearman’s rank correlation coefficient r <strong>for</strong> LAI measures and surface reflectance derived from SPOT (N=29)<br />

s<br />

and ASTER data (N=26) <strong>for</strong> all <strong>for</strong>est stages (** significant at p=0.01, * significant at p=0.05) <strong>of</strong> Budongo Forest.<br />

Sensor Band LAI e (LAI2000) LAI e (DHP) LAI (DHP)<br />

SPOT-HRVIR 1 -0.54(**) -0.46(*) -0.34<br />

2 -0.48(**) -0.44(*) -0.34<br />

3 0.10 0.25 0.11<br />

4 -0.54(**) -0.38(*) -0.33<br />

ASTER 1 -0.56(**) -0.50(**) -0.33<br />

2 -0.58(**) -0.51(**) -0.32<br />

3 -0.14 0.15 -0.01<br />

4 -0.61(**) -0.46(*) -0.46(*)<br />

5 -0.58(**) -0.51(**) -0.33<br />

6 -0.59(**) -0.54(**) -0.35<br />

7 -0.48(*) -0.49(*) -0.33<br />

8 -0.59(**) -0.51(**) -0.29<br />

9 -0.46(*) -0.53(**) -0.35<br />

Table A-7 Spearman’s rank correlation coefficient r <strong>for</strong> LAI measures and SVIs derived from SPOT (N=29) and ASTER data<br />

s<br />

(N=26) <strong>for</strong> all <strong>for</strong>est stages (** significant at p=0.01, * significant at p=0.05) <strong>of</strong> Budongo Forest.<br />

Sensor SVI LAI e (LAI2000) LAI e (DHP) LAI (DHP)<br />

SPOT-HRVIR SR 0.45(*) 0.48(**) 0.37(*)<br />

NDVI 0.44(*) 0.46(*) 0.36<br />

NDVI C 0.45(*) 0.42(*) 0.31<br />

MSR 0.45(*) 0.48(**) 0.37(*)<br />

RSR 0.43(*) 0.43(*) 0.32<br />

NDMI 0.42(*) 0.50(**) 0.37<br />

ASTER SR 0.48(*) 0.62(**) 0.45(*)<br />

NDVI 0.48(*) 0.62(**) 0.44(*)<br />

NDVI C 0.53(**) 0.50(**) 0.36<br />

MSR 0.48(*) 0.60(**) 0.45(*)<br />

RSR 0.51(**) 0.59(**) 0.40(*)<br />

NDMI 0.43(*) 0.62(**) 0.46(*)<br />

185


186<br />

Table A-8 Spearman’s rank correlation coefficient r <strong>for</strong> LAI measures and SVIs derived from SPOT (N=10) and ASTER data<br />

s<br />

(N=8) <strong>for</strong> <strong>the</strong> late <strong>for</strong>est stage (** significant at p=0.01, * significant at p=0.05) <strong>of</strong> Budongo Forest.<br />

Sensor SVI LAI e (LAI2000) LAI e (DHP) LAI (DHP)<br />

SPOT-HRVIR SR -0.53 -0.44 -0.54<br />

NDVI -0.55 -0.58 -0.60<br />

NDVI C -0.60 -0.70(*) -0.72(*)<br />

MSR -0.53 -0.47 -0.50<br />

RSR -0.59 -0.60 -0.66(*)<br />

NDMI -0.62 -0.61 -0.65(*)<br />

ASTER SR -0.21 -0.10 -0.45<br />

NDVI -0.14 -0.14 -0.52<br />

NDVI C -0.52 -0.29 -0.24<br />

MSR -0.21 -0.10 -0.45<br />

RSR -0.33 -0.31 -0.62<br />

NDMI -0.21 -0.10 -0.45<br />

Table A-9 Spearman’s rank correlation coefficient r <strong>for</strong> LAI measures and SVIs derived from SPOT (N=19) and ASTER data<br />

s<br />

(N=18) <strong>for</strong> early and intermediate <strong>for</strong>est stages (** significant at p< 0.01) <strong>of</strong> Budongo Forest.<br />

Sensor SVI LAI e (LAI2000) LAI e (DHP) LAI (DHP)<br />

SPOT SR 0.81 (**) 0.80 (**) 0.64 (**)<br />

NDVI 0.80 (**) 0.79 (**) 0.64 (**)<br />

NDVI C 0.75 (**) 0.76 (**) 0.64 (**)<br />

MSR 0.81 (**) 0.80 (**) 0.65 (**)<br />

RSR 0.78 (**) 0.78 (**) 0.63 (**)<br />

NDMI 0.81 (**) 0.84 (**) 0.66 (**)<br />

ASTER SR 0.84 (**) 0.86 (**) 0.65 (**)<br />

NDVI 0.84 (**) 0.86 (**) 0.66 (**)<br />

NDVI C 0.67 (**) 0.72 (**) 0.63 (**)<br />

MSR 0.84 (**) 0.86 (**) 0.65 (**)<br />

RSR 0.84 (**) 0.84 (**) 0.62 (**)<br />

NDMI 0.76 (**) 0.88 (**) 0.71 (**)


Appendix <strong>of</strong> Tables<br />

Table A-10 Theil-Sen and linear OLS regression models <strong>based</strong> on LAI and SVIs <strong>for</strong> early and intermediate <strong>for</strong>est stages<br />

e (LAI2000)<br />

<strong>of</strong> Budongo Forest toge<strong>the</strong>r with <strong>the</strong> respective R2 .<br />

Sensor Y Theil-Sen model R 2 Least squares model R 2<br />

SPOT-HRVIR SR 3.427+0. 593 X 0.88 3.036+0.669 X 0.86<br />

NDVI 0. 594+0.026 X 0.86 0.571+0.029 X 0.88<br />

NDVI c 0. 258+0.050 X 0.83 0.193+0.059 X 0.90<br />

MSR 0. 897+0.124 X 0.86 0.818+0.139 X 0.87<br />

RSR 0.912+0.693 X 0.88 0.606+0.759 X 0.89<br />

NDMI 0. 018+0.043 X 0.86 -0.017+0.049 X 0.88<br />

ASTER SR 3.009+0.487 X 0.88 3.168+0.468 X 0.89<br />

NDVI 0.560+0.025 X 0.90 0.557+0.026 X 0.90<br />

NDVI c 0.207+0.065 X 0.87 0.148+0.073 X 0.91<br />

MSR 0.787+0.108 X 0.89 0.104+0.819 X 0.90<br />

RSR 0.849+0.707 X 0.89 0.465+0.743 X 0.92<br />

NDMI 0.147+0.029 X 0.85 0.139+0.031 X 0.86<br />

Table A-11 Theil-Sen and linear OLS regression models <strong>based</strong> on LAI and SVIs <strong>for</strong> early and intermediate <strong>for</strong>est stages <strong>of</strong><br />

(DHP)<br />

Budongo Forest toge<strong>the</strong>r with <strong>the</strong> respective R2 .<br />

Sensor Y Theil-Sen model R 2 Least squares model R 2<br />

SPOT-HRVIR SR 2.073+0.550 X 0.66 0.695+0.683 X 0.74<br />

NDVI 0.122+0.049 X 0.64 0.463+0.031 X 0.79<br />

NDVI c 0.424+0.014 X 0.60 -0.020+0.061 X 0.79<br />

MSR 0.645+0.112 X 0.65 0.325+0.142 X 0.75<br />

RSR -0.545+0.628 X 0.68 -2.037+0.773 0.76<br />

NDMI 0.092+0.041 X 0.67 -0.191+0.50 X 0.77<br />

ASTER SR 2.121+0.426 X 0.76 1.491+0.485 X 0.80<br />

NDVI 0.503+0.023 X 0.76 0.463+0.027 X 0.83<br />

NDVI c 0.054+0.061 X 0.65 -0.110+0.075 X 0.81<br />

MSR 0.605+0.093 X 0.76 0.445+0.108 X 0.81<br />

RSR -0.523+0.023 X 0.74 -2.133+0.761 X 0.81<br />

NDMI 0.069+0.028 X 0.76 0.024+0.033 X 0.80<br />

187


188<br />

Table A-12<br />

Spearman’s rank correlation coefficient rs <strong>for</strong> LAI and texture measures derived from ASTER data (N=8) <strong>for</strong> <strong>the</strong> late<br />

<strong>for</strong>est stage (** significant at p=0.01, * significant at p=0.05) <strong>of</strong> <strong>the</strong> Budongo Forest test site. Due to <strong>the</strong> large amount<br />

<strong>of</strong> data only texture measures with at least one significant correlation to any LAI measure are shown.<br />

Sensor Kernel Band Texture measure LAI e (LAI2000) LAI e (DHP) LAI (DHP)<br />

ASTER 3*3 3 Homogeneity -0.88(**) -0.67 -0.48<br />

3 Entropy 0.64 0.76 (*) 0.29<br />

4 Variance 0.31 0.81 (*) 0.60<br />

7 Correlation 0.38 0.81 (*) 0.60<br />

5*5 3 Homogeneity -0.79 (*) -0.50 -0.29<br />

4 Variance 0.41 0.83 (*) 0.55<br />

7*7 3 Homogeneity -0.79 (*) -0.50 -0.29<br />

4 Variance 0.41 0.83 (*) 0.55<br />

8 Variance -0.71(*) -0.74 (*) -0.57<br />

8 Contrast -0.52 -0.71 (*) -0.50<br />

9*9 4 Variance 0.38 0.86 (**) 0.64<br />

4 Entropy 0.24 0.74 (*) 0.57<br />

8 Contrast -0.45 -0.91 (**) -0.69<br />

11*11 3 Homogeneity -0.81 (*) -0.48 -0.26<br />

4 Variance 0.41 0.83 (*) 0.55<br />

8 Variance -0.71(*) -0.74 (*) -0.57<br />

SPOT-HRVIR 3*3 1 Entropy 0.62 0.71 (*) 0.45<br />

3 Variance 0.43 0.81 (*) 0.67<br />

5*5 3 Contrast 0.38 0.71 (*) 0.69<br />

3 Dissimilarity 0.38 0.71 (*) 0.69<br />

7*7 3 Contrast 0.38 0.71 (*) 0.69<br />

3 Dissimilarity 0.38 0.71 (*) 0.69<br />

9*9 3 Contrast 0.43 0.74 (*) 0.71 (*)<br />

3 Dissimilarity 0.38 0.81 (*) 0.60<br />

11*11 3 Contrast 0.38 0.76 (*) 0.62<br />

Table A-13 Spearman’s rank correlation coefficient r <strong>for</strong> LAI measures and surface reflectance derived from SPOT data (N=27,<br />

s<br />

<strong>for</strong> LAI N= 25) <strong>for</strong> all <strong>for</strong>est stages (** significant at p=0.01, * significant at p=0.05) <strong>of</strong> Kakamega Forest.<br />

e (LAI2000)<br />

Sensor Band LAI e (LAI2000) LAI e (DHP) LAI (DHP)<br />

SPOT-HRVIR 1 0.51(*) -0.58(**) -0.39(*)<br />

2 -0.57(**) -0.56(**) -0.32<br />

3 -0.46(*) -0.24 -0.17<br />

4 -0.53(**) -0.55(**) -0.36


Appendix <strong>of</strong> Tables<br />

Table A-14 Spearman’s rank correlation coefficient r <strong>for</strong> LAI measures and SVIs derived from SPOT data (* significant at p=0.05,<br />

s<br />

N =27, <strong>for</strong> LAI N= 25, all <strong>for</strong>est stages) <strong>for</strong> Kakamega Forest.<br />

e (LAI2000)<br />

Sensor SVI LAI e (LAI2000) LAI e (DHP) LAI (DHP)<br />

SPOT-HRVIR SR 0.17 0.46(*) 0.43(*)<br />

NDVI 0.15 0.46(*) 0.41(*)<br />

NDVI C 0.46(*) 0.55(**) 0.38<br />

MSR 0.16 0.46(*) 0.42(*)<br />

RSR 0.36 0.55(**) 0.45(*)<br />

NDMI 0.12 0.33 0.32<br />

Table A-15 Spearman’s rank correlation coefficient r <strong>for</strong> LAI measures and SVIs derived from SPOT data (* significant at p=0.05,<br />

s<br />

N =16, early and intermediate <strong>for</strong>est stages) <strong>for</strong> Kakamega Forest.<br />

Sensor SVI LAI e (LAI2000) LAI e (DHP) LAI (DHP)<br />

SPOT-HRVIR SR 0.65(**) 0.75(**) 0.64(*)<br />

NDVI 0.63(*) 0.75(**) 0.59(*)<br />

NDVI C 0.66(**) 0.62(*) 0.48<br />

MSR 0.65(**) 0.75(**) 0.60(*)<br />

RSR 0.68(**) 0.67(**) 0.55(*)<br />

NDMI 0.50 0.56(*) 0.42<br />

Table A-16 Theil-Sen and linear OLS regression models <strong>based</strong> on LAI and SVIs <strong>for</strong> early and intermediate <strong>for</strong>est stages<br />

e (DHP)<br />

in Kakamega Forest.<br />

Sensor Y Theil-Sen model R 2 Least squares model R 2<br />

SPOT-HRVIR SR 6.421+0.645 X 0.36 3.542+1.295 X 0.68<br />

NDVI 0.739+0.015 X 0.22 0.635+0.039 X 0.63<br />

NDVI c 0.268+0.059 X 0.51 0.133+0.088 X 0.72<br />

MSR 3.164+0.125 X 0.32 2.512+0.273 X 0.66<br />

RSR 1.836+0.949 X 0.58 1.362-0.069 X 0.74<br />

NDMI 0.643-0.027 X 0.37 0.765-0.057 X 0.57<br />

189


190<br />

Table A-17 Spearman’s rank correlation coefficient r <strong>for</strong> LAI and texture measures derived from SPOT data (N=10) <strong>for</strong> <strong>the</strong> late<br />

s<br />

<strong>for</strong>est stage (* significant at p=0.05) <strong>of</strong> <strong>the</strong> Kakamega Forest test site. Due to <strong>the</strong> large amount <strong>of</strong> data only texture<br />

measures with at least one significant correlation to any LAI measure are shown.<br />

Sensor Kernel Band Texture measure LAI e (LAI2000) LAI e (DHP) LAI (DHP)<br />

SPOT-HRVIR 3*3 1 Variance 0.45 0.73 (*) 0.75 (*)<br />

1 Entropy 0.50 0.66 (*) 0.72 (*)<br />

4 Homogeneity -0.43 -0.67 (*) -0.69 (*)<br />

5*5 1 Entropy 0.21 0.67 (*) 0.75 (*)<br />

4 Homogeneity -0.57 -0.83 (**) -0.79 (**)<br />

7*7 4 Homogeneity -0.60 -0.84 (**) -0.75 (*)<br />

9*9 4 Homogeneity -0.76 (*) -0.67 (*) -0.58


191<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse solar<br />

radiation in rugged terrain, ρ can be approximated iteratively starting with ρ terrain<br />

0 =0.1 by<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

(A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

1 APPENDIX OF EQUATIONS<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

path radiance <strong>for</strong> elevation z and viewing geometry<br />

1 APPENDIX OF EQUATIONS<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

;<br />

1 APPENDIX OF EQUATIONS<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle<br />

1 APPENDIX OF EQUATIONS<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

(direct plus diffuse components);<br />

1 APPENDIX OF EQUATIONS<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

a sun-to-ground transmittance (direct beam);<br />

1 APPENDIX OF EQUATIONS<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

E s : extraterrestrial solar irradiance;<br />

E d (x, y, z): diffuse solar flux ;<br />

E g (z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

1 APPENDIX OF EQUATIONS<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment, whose<br />

radius can be calculated as<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

(A.2)<br />

where cosθ is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong> average<br />

canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1° (ring 4) <strong>the</strong><br />

field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal direction (cf.<br />

Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The horizontal projection in<br />

Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on <strong>the</strong> average canopy height. It<br />

is represented by a circular sector whose radius r can be calculated by equation A.1. Its chord (i.e. is maximum<br />

width) is <strong>the</strong>n equivalent to<br />

Appendix <strong>of</strong> Equations<br />

Atmospheric correction (cf. Chapter 4.2.2)<br />

Disregarding <strong>the</strong> adjacency effect and accounting <strong>for</strong> <strong>the</strong> directional dependence <strong>of</strong> direct and diffuse<br />

solar radiation in rugged terrain, can be approximated iteratively starting with<br />

0<br />

terrain<br />

=0.1 by<br />

<br />

<br />

<br />

<br />

<br />

y<br />

x<br />

v<br />

z<br />

E<br />

z<br />

y<br />

x<br />

E<br />

y<br />

x<br />

z<br />

E<br />

y<br />

x<br />

b<br />

z<br />

z<br />

L<br />

y<br />

x<br />

DN<br />

c<br />

c<br />

d<br />

y<br />

x<br />

terrain<br />

i<br />

terrain<br />

ground<br />

d<br />

s<br />

s<br />

v<br />

v<br />

v<br />

,<br />

,<br />

,<br />

,<br />

cos<br />

,<br />

,<br />

,<br />

,<br />

,<br />

)<br />

,<br />

(<br />

2<br />

1<br />

0<br />

2<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

, (A.1)<br />

with<br />

x, y: horizontal coordinates, corresponding to <strong>the</strong> georeferenced pixel positions;<br />

z: elevation in m derived from <strong>the</strong> DEM;<br />

<br />

<br />

,<br />

,<br />

2 v<br />

z<br />

L : path radiance <strong>for</strong> elevation z and viewing geometry <br />

<br />

,<br />

v<br />

;<br />

<br />

v<br />

v z <br />

, : ground-to-sensor transmittance <strong>for</strong> elevation z and viewing angle v<br />

(direct plus diffuse<br />

components);<br />

<br />

v<br />

v z <br />

, : sun-to-ground transmittance (direct beam);<br />

<br />

y<br />

x,<br />

: angle between <strong>the</strong> solar ray and <strong>the</strong> surface normal (illumination angle);<br />

b(x,y): binary factor: b=1 if pixel receives direct solar beam, o<strong>the</strong>rwise b=0;<br />

Es: extraterrestrial solar irradiance;<br />

Ed(x, y, z): diffuse solar flux ;<br />

Eg(z): global flux (direct plus diffuse solar flux on a horizontal surface);<br />

terrain<br />

(i) (x, y): locally varying average terrain reflectance, calculated iteratively (i=1,2,3) and<br />

terrain<br />

v (x, y): terrain view factor (range 0-1) calculated from local slope.<br />

Theoretical field <strong>of</strong> view <strong>of</strong> LAI-2000 PCA (cf. Chapter 5.1.1)<br />

The <strong>the</strong>oretical field <strong>of</strong> view <strong>of</strong> a single LAI-2000 PCA measurement is represented by a sphere segment,<br />

whose radius can be calculated as<br />

h<br />

r<br />

<br />

cos<br />

1<br />

, (A.2)<br />

where <br />

cos is <strong>the</strong> maximum zenith angle <strong>of</strong> <strong>the</strong> outmost ring included in LAI calculation and h is <strong>the</strong><br />

average canopy height. Assuming an average canopy height <strong>of</strong> 20 m and a maximum zenith angle <strong>of</strong> 58.1°<br />

(ring 4) <strong>the</strong> field <strong>of</strong> view would thus <strong>the</strong>oretically include all canopy elements within 37.8 m in horizontal<br />

direction (cf. Figure A-1a). At 10 m height <strong>the</strong> <strong>the</strong>oretical field <strong>of</strong> view decreases to 18.9 m. The<br />

horizontal projection in Figure A-1b shows that also <strong>the</strong> maximum width <strong>of</strong> a measurement depends on<br />

<strong>the</strong> average canopy height. It is represented by a circular sector whose radius r can be calculated by<br />

equation A.1. Its chord (i.e. is maximum width) is <strong>the</strong>n equivalent to<br />

)<br />

2<br />

(sin<br />

2 r<br />

<br />

, (A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.<br />

(A.3)<br />

which in Figure A-1b equals 28.9 m at an average canopy height <strong>of</strong> 20 m and 14.5 m at an average canopy<br />

height <strong>of</strong> 10 m. Theoretically <strong>the</strong> distance between two measurements should be at least equivalent to r to<br />

guarantee spatial independence.


192<br />

Curriculum Vitae<br />

Personal in<strong>for</strong>mation<br />

Employment record<br />

Education<br />

Publications<br />

Name<br />

Date <strong>of</strong> birth<br />

Place <strong>of</strong> birth<br />

Marital status<br />

Address<br />

since 03/2008<br />

01/2007-02/2008<br />

10/2002-12/2006<br />

08/2002<br />

04/1997-08/2002<br />

08/1996-02/1997<br />

1987-1996<br />

1984-1987<br />

Tanja Kraus<br />

May 11, 1977<br />

Nürnberg, Germany<br />

married<br />

Klugstr. 33<br />

80638 München<br />

Germany<br />

tanja@kraus.net<br />

Scientific Assistant to <strong>the</strong> Director (Wissenschaftliche Referentin des Direktors)<br />

at <strong>the</strong> German Remote Sensing Data Center (DFD) <strong>of</strong> <strong>the</strong> German Aerospace<br />

Center (DLR), Oberpfaffenh<strong>of</strong>en, Germany<br />

Research assistant at <strong>the</strong> Chair <strong>for</strong> Remote Sensing, University <strong>of</strong> Würzburg<br />

Employed at DFD-DLR in <strong>the</strong> following projects:<br />

01/2006-12/2006 Research assistant in GSE Forest Monitoring<br />

04/2005-12/2005 Assistant in <strong>the</strong> proposal phase <strong>of</strong> GITEWS (German-<br />

Indonesian Tsunami Early Warning System)<br />

06/2004-12/2006 PhD student in BIOTA East Africa, subproject E02<br />

10/2002-05/2004 Research assistant in BIOTA East Africa, subproject E02<br />

Degree in Geography (Diplom-Geographin), Thesis on “Quantifizierung von<br />

Biomasse in Savannenökosystemen mit Landsat-ETM+”<br />

Studies <strong>of</strong> Geography at <strong>the</strong> Friedrich-Alexander-University Erlangen-Nürnberg<br />

with minors Geology and Biology<br />

Au Pair in Washington, D.C.<br />

Studies <strong>of</strong> English at Georgetown University, Washington, D.C.<br />

Geschwister-Scholl-Gymnasium, Rö<strong>the</strong>nbach a. d. Pegnitz<br />

Primary School (Grundschule am Forstersberg, Rö<strong>the</strong>nbach a. d. Pegnitz)<br />

Kraus, T., Schmidt, M., Dech, S.W. and C. Samimi. The potential <strong>of</strong> optical high resolution data <strong>for</strong><br />

<strong>the</strong> assessment <strong>of</strong> leaf area index in East African rain <strong>for</strong>est ecosystems. – In: Special Issue <strong>of</strong><br />

<strong>the</strong> International Journal <strong>of</strong> Remote Sensing (in review).<br />

Eisfelder, C., Kraus, T., Bock, M., Werner, M., Buchroithner, M.F. and G. Strunz. Towards automated<br />

<strong>for</strong>est-type mapping – a service within GSE Forest Monitoring <strong>based</strong> on SPOT-5 and IKONOS<br />

data. – In: Special Issue <strong>of</strong> <strong>the</strong> International Journal <strong>of</strong> Remote Sensing (in review).


Curriculum Vitae<br />

Publications (continued)<br />

Kraus, T., Bock, M. and G. Strunz (2007). Forest type mapping <strong>based</strong> on SPOT-5 and IKONOS data<br />

as a service within GSE Forest Monitoring. – In: Proceedings <strong>of</strong> <strong>the</strong> ForestSAT Conference<br />

2007, Montpellier, France, November 5-7, 2007, 5 p.<br />

Kraus, T., Schmidt, M., Dech, S. and C. Samimi (2007). The potential <strong>of</strong> optical high resolution data<br />

<strong>for</strong> <strong>the</strong> assessment <strong>of</strong> leaf area index in East African rain <strong>for</strong>est ecosystems. – In: Proceedings<br />

<strong>of</strong> <strong>the</strong> ForestSAT Conference 2007, Montpellier, France, November 5-7, 2007, 5 p.<br />

Werner, M., Kraus, T., Bock, M., Strunz, G. and K.-F. Wetzel (2007). Objektbasierte<br />

Waldflächenkartierung in Schleswig-Holstein mit SPOT-5-Daten. – In: J. Strobl, T. Blaschke<br />

und G. Griesebner (Hrsg.) 2007: Geoin<strong>for</strong>matik 2007. Beiträge zum 19. AGIT- Symposium<br />

Salzburg, pp. 852-857.<br />

Kraus, T. (2005). LAI measurements in a tropical rain<strong>for</strong>est: Kakamega Forest, Kenya.<br />

VALERI Workshop, Avignon, March 10, 2005. Available online<br />

<br />

Kraus, T. and R. Ressl (2005). DEMMIN - Testsite <strong>for</strong> environmental monitoring and validation.<br />

VALERI Workshop, Avignon, March 10, 2005. Available online<br />

<br />

Müller U., Conrad C. and T. Kraus (2004). Analyse der raum-zeitlichen Vegetationsdynamik in<br />

Ostafrika unter Verwendung von <strong>MODIS</strong>-Daten. – In: J. Strobl, T. Blaschke und G. Griesebner<br />

(Hrsg.) 2004: Angewandte Geographische In<strong>for</strong>mationsverarbeitung XVI. Beiträge zum<br />

AGIT-Symposium Salzburg 2004, pp. 484-489.<br />

Samimi, C. and T. Kraus (2004). Biomass estimation using Landsat-TM and -ETM+. Towards<br />

a regional model <strong>for</strong> Sou<strong>the</strong>rn Africa? – In: GeoJournal Special Issue: Systems Modelling<br />

Across Geography´s Interface, 59 (3), pp. 177-187.<br />

Schaab, G., Kraus, T. and G. Strunz (2004). GIS and remote sensing activities as an integrating<br />

link within <strong>the</strong> BIOTA-East Africa project. Sustainable use and conservation <strong>of</strong> biological<br />

diversity – A challenge <strong>for</strong> society. – In: Proceedings <strong>of</strong> <strong>the</strong> International Symposium Berlin,<br />

December 1-4, 2003, Berlin, pp. 161-168.<br />

Riedlinger, T., Voigt, S., Gün<strong>the</strong>r, K.P., Gesell, G., Künzer, C., Zwing, A., Kraus, T., Kiefl, R., Tetzlaff,<br />

A., Zhang, J., Hasenauer, S., Reinartz, P., Sundermann, D. and H. Mehl (2003). Multisource<br />

satellite data facilitating disaster management during <strong>the</strong> 2003 <strong>for</strong>est fires in Portugal.<br />

– In: Proceedings <strong>of</strong> <strong>the</strong> Mediterranean seminar on new technologies applied to <strong>the</strong> management<br />

<strong>of</strong> disasters risks, Madrid, October 3-6, 2003. Available online at<br />

<br />

Kraus T. and G. Schaab (2003). Biodiversitäts<strong>for</strong>schung in Westkenia – Mit GIS und Fernerkundung<br />

von lokalen Beobachtungen zu Aussagen in Zeit und Raum. – In: J. Strobl, T. Blaschke und<br />

G. Griesebner (Hrsg.) 2003: Angewandte Geographische In<strong>for</strong>mationsverarbeitung XV.<br />

Beiträge zum AGIT-Symposium Salzburg 2003, pp. 244-249.<br />

Kraus, T. and C. Samimi (2002). Biomass estimation <strong>for</strong> land use management and fire management<br />

using Landsat-TM and -ETM+. – In: Erdkunde, 56 (2), pp. 130-143.<br />

193

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!