+ All Categories
Home > Documents > Carbon storage and nutrient shifts along an altitudinal...

Carbon storage and nutrient shifts along an altitudinal...

Date post: 21-Jul-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
68
Faculty of Bioscience Engineering Isotope Bioscience Laboratory - ISOFYS Academic year 2015 – 2016 Carbon storage and nutrient shifts along an altitudinal gradient in Nyungwe forest, Rwanda Dries Van der Heyden Promotors: Prof. dr. ir. Pascal Boeckx and prof. dr. Landry Cizungu Ntaboba Tutor: ir. Marijn Bauters Master dissertation submitted in partial fulfilment of the requirements for the degree of Master of Science in Bioscience Engineering: Forest and Nature Management
Transcript
Page 1: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

Faculty of Bioscience Engineering

Isotope Bioscience Laboratory - ISOFYS

Academic year 2015 – 2016

Carbon storage and nutrient shifts along an altitudinal

gradient in Nyungwe forest, Rwanda

Dries Van der Heyden

Promotors: Prof. dr. ir. Pascal Boeckx and prof. dr. Landry Cizungu Ntaboba

Tutor: ir. Marijn Bauters

Master dissertation submitted in partial fulfilment of the requirements for the

degree of Master of Science in Bioscience Engineering: Forest and Nature

Management

Page 2: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional
Page 3: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

Copyright

The author and promoters give the permission to use this thesis for consultation and to copy

parts of it for personal use. Every other use is subject to the copyright laws, more specifically the

source must be extensively specified when using results from this thesis.

De auteurs en promotors geven de toelating deze scriptie voor consultatie beschikbaar te stellen en delen ervan te

kopiëren voor persoonlijk gebruik. Elk ander gebruik valt onder de beperkingen van het auteursrecht, in het

bijzonder met betrekking tot de verplichting uitdrukkelijk de bron te vermelden bij het aanhalen van resultaten uit

deze scriptie.

Ghent, August 2016

Prof. Dr. Ir. Pascal Boeckx Dries Van der Heyden

Page 4: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional
Page 5: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

Acknowledgments

First of all I would like to express my gratitude to prof. Pascal Boeckx and prof. Landry Cizungu

for giving me the opportunity to participate in this research and for all their practical assistance. I

also want to thank Marijn Bauters for his professional help and personal guidance. Thanks to

Cys, for being a great companion and an efficient fieldworker.

Further, I would like to convey my appreciation to the Rwandese and Congolese co-workers:

James, ‘mon frère’, Fidele, Jean-Baptiste, Sartié and the very helpful guards and the staff in

Nyungwe. From them I learned the word ‘ubushakashatsi’, which is Kinyarwanda for ‘research’, a

word that expresses well the labour, envy and love I’ve put into this thesis.

I would also like to acknowledge the Belgian staff of Isofys and Geert Sioen of the INBO and

thank them for their help.

Lastly, I am very grateful for the support of my parents, brother, sister, friends and family during

this year and at university.

Page 6: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

Table of contents

Abbreviations I

Abstract II

Résumé III

Samenvatting IV

1 Introduction 1

2 Literature review 2

2.1 Tropical forests 2

2.1.1 What and where? 2

2.1.2 Value and importance 3

2.2 Carbon cycling 5

2.3 Tropical montane forests 8

2.3.1 Characteristics 8

2.3.2 Altitudinal transects 11

2.3.3 Nutrient cycling and isotopes 13

2.3.4 Albertine rift montane forests 16

3 Research questions 17

4 Materials and methods 18

4.1 Study area 18

4.2 Altitudinal transect choice and site selection 19

4.3 Establishment and measurement of permanent sample plots 21

4.4 Sampling and processing of soil, litter, wood and leafs 22

4.5 Planimetric plot areas 23

4.6 Estimation of carbon stocks 24

4.6.1 Aboveground carbon 24

4.6.2 Carbon in litter and belowground carbon 25

4.7 Statistical analyses 25

5 Results 26

5.1 Planimetric plot area 26

5.2 Forest structure 26

5.3 Carbon stocks 28

5.4 Litter, fine roots and the O-horizon 31

5.5 Mineral soil 33

Page 7: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

6 Discussion 36

6.1 The correction for planimetric plot areas and steep slopes 36

6.2 Changes in forest structure with altitude 36

6.3 Changing carbon stocks along an altitudinal gradient 36

6.3.1 Underestimated AGC in the lower strata 37

6.4 Mineral soil nutrient concentrations along an altitudinal gradient 37

6.5 How do the isotopic compositions change with altitude and what do they tell us? 38

7 General conclusion and future perspectives 39

8 List of references 40

9 Appendices 51

A. RAINFOR field work database codes for living trees 51

B. Lab protocol for soil bioavailable P 52

C. Graphs of the additional carbon stocks 53

D. Graphs of N & C in litter, O-horizon and roots 54

E. Graphs of δ13C and δ 15N in the soil at different depths 55

Page 8: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional
Page 9: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

I

Abbreviations

AGB Aboveground Biomass

AGC Aboveground Carbon

BGC Belowground Carbon

DBH Diameter at Breast Height

FAO Food and Agriculture Organization of the United Nations

GPP Gross Primary Productivity

masl meters above sea level

NBP Net Biome Productivity

NEP Net Ecosystem Productivity

NPP Net Primary Productivity

POM Point of Measurement

RAINFOR Red Amazónica de Inventarios Forestales (Amazon Forest Inventory

Network)

SOM Soil Organic Matter

TMCF Tropical Montane Cloud Forest

TMF Tropical Montane Forest

UNEP United Nations Environment Programme

WCMC World Conservation Monitoring Centre

WRB World Reference Base for soil resources

WWF World Wildlife Fund/World Wide Fund for Nature

Page 10: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

II

Abstract

Tropical montane forests sustain high biodiversity and deliver important ecosystem services such

as the provision of fresh water and carbon sequestration. Yet their functioning is understudied,

especially in Africa. The current process of climate warming encourages studying these

ecosystems which play a significant role in climate change mitigation but are under threat. An

altitudinal transect study was performed in a tropical montane forest in Rwanda. Twenty

permanent sample plots of 40x40m were established in 4 altitudinal strata (1800m, 2200m,

2500m and 2800m). Forest structure and carbon stocks were quantified and linked with

properties of soil, litter and roots. In accordance with the prevailing literature aboveground

carbon tended to decrease with increasing altitude, while carbon in the roots and the upper soil

layer demonstrated an inverse relation. . Increasing altitude was associated with a significant

increase in the C:N-ratio of leaves and litter and a significant decrease of the isotopic

composition δ15N of leaves, litter and soil, which indicates that at a higher altitude there is more

nitrogen limitation and an altered N cycle .These findings confirm nutrient availability to be a key

regulator of forest carbon cycling processes.

Page 11: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

III

Résumé

Les forêts tropicales montagnardes assurent une biodiversité substantielle et livres des services

écosystémiques importantes, comme la provision d’eau pure et le stockage de carbone. Pourtant

le nombre d’études sur ce sujet est limité. Il y a surtout une manque d’études sur le

fonctionnement des forêts montagnardes en Afrique. Le réchauffement climatique actuel incite à

l’étude de ces écosystèmes, qui jouent un rôle majeur dans la mitigation, mais qui sont menacés.

Une étude transect de hauteur a été conduite dans une forêt tropicale montagnarde au Ruanda.

Vingt zones d’altitudes ont été établies à quatre hauteurs différentes (1800m, 2200m, 2500 en

2800m). La structure de la forêt et le stockage du carbone ont été quantifiés et les liens avec

certains caractéristiques du sol, la litière et les racines ont été établis. En ligne avec la littérature

existante la carbone dans la biomasse vivante au-dessus du sol tendait à diminuer avec l’altitude,

tandis qu’une relation inverse était observée pour la carbone dans les racines et la couche

supérieure du sol. L’augmentation de l’altitude était associée à l’augmentation significative du

ratio C-N des feuilles et de la litière, et la diminution significative de la composition isotopique

δ15N des feuilles, de la litière et du sol. Ceci indique qu’en grande altitude la quantité d’azote est

plus limitée et le cycle de l’azote est modifié. Ces observations confirment que la disponibilité des

nutriments jouent un rôle clé dans le cycle du carbone au sein des forêts

Page 12: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

IV

Samenvatting

Tropische bergbossen waarborgen een hoge biodiversiteit en leveren belangrijke

ecosysteemdiensten zoals de provisie van zuiver water en koolstofopslag. Desondanks wordt hier

weinig onderzoek naar gedaan. Vooral het functioneren van de bergbossen in Afrika is nog

weinig onderzocht. De huidige klimaatopwarming zet aan tot het bestuderen van deze

ecosystemen, die een belangrijke rol spelen in de mitigatie van klimaatverandering maar wel

bedreigd zijn. Een hoogtetransect-onderzoek werd uitgevoerd in een tropisch bergbos in

Rwanda. Twintig permanente proefvlakken werden uitgezet in vier verschillende hoogtezones

(1800m, 2200m, 2500 en 2800m). De structuur van het bos en de koolstofstocks werden

gekwantificeerd en gelinkt aan bepaalde eigenschappen van de bodem, de strooisellaag en de

wortels. In overeenstemming met de meeste literatuur vertoonde de bovengrondse koolstof een

dalende trend met stijgende hoogte, waar de koolstof in de bovenste bodemlaag en in de wortels

net een omgekeerde relatie vertoonden. Toenemende hoogte was geassocieerd met een

significante stijging van de C:N-ratio van bladeren en strooisel en een significante daling van de

isotopensamenstelling δ15N van bladeren, strooisel en bodem, wat wijst op meer stikstoflimitatie

en een veranderde stikstofcylus naarmate de hoogte toeneemt. Deze bevindingen bevestigen dat

de beschikbaarheid van nutriënten een sleutelrol vervult in de koolstofcyclus van bossen.

Page 13: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

1

1 Introduction

With the Paris climate conference (COP21) in December 2015, the huge threat of recent climate

change received renewed attention. In this respect, the important role of tropical forests is widely

acknowledged. Despite extensive research on the topic, the process-level understanding of

carbon and forest dynamics still needs to be improved (Bustamante et al., 2016). Furthermore it

remains largely unknown how tropical forests will respond and feedback to climate change at

ecosystem level (Zhou et al., 2013).

Advancements in the understanding of ecosystem ecology and functioning are often achieved

slowly and step by step. Therefore in a collaboration between CAVElab1 and ISOFYS2, Ghent

University, four master-students in Bioscience Engineering conducted research in tropical

montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional diversity

while the other two studied carbon storage and nutrient shifts. Permanent sample plots were

established on an altitudinal transect since the environmental gradient offers a great experimental

setup. The same protocol was used, allowing for a prudent comparison between a central African

and an Andean TMF.

In this framework, the present thesis deals with carbon storage and nutrient shifts in a montane

tropical rainforest in Rwanda. As the study site was located near the border with the Democratic

Republic of Congo and a partnership between ISOFYS and the Catholic University of Bukavu

(UCB) in eastern Congo already existed, a local collaboration was evident, resulting in the

exchange of knowledge and practical assistance.

1 Computational & Applied Vegetation Ecology laboratory 2 Isotope Bioscience Laboratory

Page 14: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

2

2 Literature review

2.1 Tropical forests

2.1.1 What and where?

A definition of tropical forest explains its two components: tropical, the climatic or geographic

aspect and forest, a land cover type. The geographic delineation of ‘tropical’ indicates the location

between the Tropic of Cancer and the Tropic of Capricorn, the tropical belt, whereas the climatic

delineation points to the regions where certain long-term meteorological conditions are fulfilled.

For instance according to the Köppen-Geiger classification system tropical climates where

allocated if a minimum average temperature of 18 °C was reached (at sea level or at low

elevations) during all twelve months of the year. Looking at the land cover type, numerous

definitions of forest exist. A clear and internationally accepted example, which is also used in this

literature review, is the one by the Food and Agriculture Organization of the United Nations

(FAO): Forest is a land spanning more than 0.5 hectares with trees higher than 5 meters and a canopy cover of

more than 10 percent, or trees able to reach these thresholds in situ. It does not include land that is predominantly

under agricultural or urban land use.

The FAO Global Forest Resources Assessment 2015 shows that forest nowadays covers around

3999 M ha of our planet’s surface (Keenan et al., 2015). 44% of this area is covered by tropical

forests, mainly in South America, West-Central Africa and Southeast Asia (Figure 1).

Figure 1. Global forest distribution. Tropical forests are indicated with orange to dark purplish red, according to the tree cover density. Map adopted from the Global Forest Resources Assessment (FAO, 2010).

Page 15: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

3

2.1.2 Value and importance

A first major value of tropical forests is their biodiversity. Of all terrestrial biomes they contain

the highest species diversity (Myers et al., 2000), with one-half to two-thirds of the Earth’s

terrestrial biodiversity to be found in these ecosystems (Gardner et al., 2010). Slik et al. (2015)

recently estimated the minimum number of tropical forest tree species to fall between ∼40.000

and ∼53.000, which is in sharp contrast with the 1166 species that make up the northern

temperate tree flora (Latham and Ricklefs, 1993). An additional value related to biodiversity is the

intactness, the way in which an area is unmodified by human impacts and how it is retaining its

natural character and influence. Although genuine pristine ecosystems are really scarce, in the

tropics there is still a quite extended area of old-growth forests that has experienced little to no

recent human disturbance. Important high-biodiversity ‘wilderness areas’ are Amazonia, the

Congo forests of Central Africa and the tropical rainforest of New Guinea (Mittermeier et al.,

2003). Gibson et al. (2011) found that biodiversity values are clearly lower in degraded forests

which encourages the protection of primary forests.

There are different ethical viewpoints on the value of nature and its biodiversity, with the three

main approaches being biocentrism, ecocentrism and anthropocentrism (Callicott, 2004). The

first two approaches respectively enfranchise morality to all living beings and to non-individual

environmental entities whereas the latter grants moral standing exclusively to human beings. The

anthropocentric environmental ethic considers nonhuman natural entities and nature as a whole

to be only a means for human ends, so the value of nature is derived from what people assign as

beneficial for human well-being. This approach is standing on traditional Western moral

philosophy and it is the most convenient and clear way to explain people the importance of

conservation and environmental protection. A quite recent anthropocentric approach that aims

to increase the appreciation and value of natural systems is the concept of ecosystem services.

Ecosystem services are the direct and indirect contributions of ecosystems that benefit human

wellbeing. It is a clear concept that is increasingly used on a global scale by policy makers as well

as scientists. It became widespread with the famous article of Constanza et al. (1997), ‘The value

of the world’s ecosystem services and natural capital’, and after the Millennium Ecosystem

Assessment of the United Nations in 2005 (MA 2005). The services ecosystems provide, are

generally classified in four groups: provisioning services, regulating services, supporting services

and cultural services. Forests, for instance, can provide wood and fresh water, regulate climate

and weather patterns, have a supporting function through soil formation and deliver a big

amount of cultural services such as recreation and beautiful landscapes (aesthetic value). It must

also be acknowledged that ecosystems can have a negative impact on people, the so-called

disservices. An example of a tropical forest disservice is crop damage from wild animals

(Sandbrook and Burgess, 2015). It is clear however that the benefits are generally much bigger.

Services can be of local importance and of global importance. E.g. wild bee populations in

tropical forests deliver an important value for local farmers through crop pollination (Ricketts et

al., 2004) and climate regulation via carbon sequestration can be seen as a more generic global

service.

Page 16: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

4

Figure 2. Terrestrial organic carbon in soil and vegetation. Map adopted from the UNEP-WCMC report ‘Towards a global map of natural capital: key ecosystem assets’ (Dickson et al., 2014).

In 2014 the United Nations Environment Programme World Conservation Monitoring Centre

(UNEP-WCMC) created a report with the first attempt to give an overview of the global

distribution of ecosystem assets that have the capacity to generate a basket of ecosystem services

(Dickson et al., 2014). They identified and mapped five key assets of which two are predominant

in tropical forests: terrestrial biodiversity (adjusted for intactness) and terrestrial organic carbon in

vegetation and soil. The latter is depicted in Figure 2 and when compared with the distribution of

tropical forest in Figure 1, the overlap is quite clear. The role of tropical forests in the global

carbon cycle is more elaborately discussed in section 2.2.

Despite all the positive services and goods tropical forests deliver, the forest area in the tropics

decreased from 1966 M ha in 1990 to 1770 M ha in 2015 (Keenan et al., 2015). Another huge

problem in the tropical forests is forest degradation, an item that is even more difficult to

quantify but is also responsible for a lot of biodiversity and carbon stock losses. From pan-

tropical assessments of the importance of different deforestation and forest degradation drivers it

is clear that agriculture is the most prevalent deforestation driver whereas timber extraction and

logging is the most serious forest degradation driver (Geist and Lambin, 2002; Hosonuma et al.,

2012). An international programme which tries to tackle these problems by creating a financial

value to tropical forests and their carbon stocks is the United Nations Framework Convention on

Climate Change (UNFCCC) mechanism for Reducing Emissions from Deforestation and

Degradation in developing countries (REDD+). It also claims to go beyond deforestation and

forest degradation, by including the role of conservation, sustainable management of forests and

enhancement of forest carbon stocks. The well-functioning and results of this programme are

however subject of a lot of discussion.

Page 17: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

5

2.2 Carbon cycling

Carbon is in constant exchange between the oceans, the atmospheric pool and terrestrial

ecosystems via complex interactions that are increasingly understood (Falkowski et al., 2000).

Minor changes in one of these pools or of process determining parameters could have a huge

influence on both our planet’s biosphere or atmosphere. Just like other biogeochemical and

climatological processes, the global carbon cycle is heavily altered by human activities.

Anthropogenic carbon emissions lead to an atmospheric CO2 level unprecedented for at least 800

000 years (Pachauri and Meyer, 2014). Around 50% of these emissions between 1750 and 2011

originated in the last 40 years. Together with the emissions of other anthropogenic greenhouse

gasses it is extremely likely that they have been the dominant cause of the observed climate

warming since the mid-20th century. With the current understanding of the ocean carbon cycle, it

looks like the sink strength (the magnitude of net accumulation and storage of carbon from

atmospheric CO2) will probably weaken, leaving a potentially more important role to terrestrial

ecosystems (Falkowski et al., 2000).

In the terrestrial biosphere carbon is stored in vegetation, detritus and soil. Forests hold the

greatest share of these terrestrial C stocks and hence exert strong control on the evolution of

atmospheric CO2. The process at the basis of carbon sequestration is photosynthesis at the leaf-

level, where atmospheric CO2 is used to assimilate sugars or plant biomass by the use of solar

irradiance. In this process also water is used and oxygen gas is generated. Next to the basic

requirements for photosynthesis different abiotic and biotic controls determine forest

productivity (Hopkins and Hüner, 2004; Schulze et al., 2002):

the amount and light-use efficiency of foliage

water availability

ambient temperature

presence and availability of soil nutrients

adaptations of species to extreme temperatures

efficient use of water and nutrients

The photosynthetic input in ecosystems, the Gross Primary Productivity (GPP), is partially

compensated by the main natural carbon outputs: autotrophic respiration (Ra, i.e. from plants)

and heterotrophic respiration (Rh, i.e. from microbes and animals). In addition carbon is also

released from ecosystems by harvest and fire. The resultant quantity from photosynthesis minus

plant respiration, is termed Net Primary Productivity (NPP) and further subtraction of the

heterotrophic respiration results in the Net Ecosystem Productivity (NEP). NEP generally

accounts for a whole ecosystem in which the accumulation of carbon is calculated over a whole

season or other time period (IPCC, 2003). Carbon losses due to disturbances such as fire and

harvest are relatively infrequent and difficult to quantify at ecosystem scale (Boisvenue and

Running, 2006). However at the biome scale these disturbances can be considered processes that

occur on a regular basis in one area or another. The term used to refer to the net production of

organic matter in a region containing a range of ecosystems, including disturbances is the Net

Biome Productivity (NBP). It is calculated by summing NEPs over the region and subtracting the

losses due to disturbances. NBP seems to be the best way for analysing long-term, large-scale

changes in carbon but distinction among NPP, NBP and NEP is often unclear in literature.

Page 18: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

6

Forests throughout the world are known to be a large and persistent sink. Pan et al. (2011)

estimated a total forest sink of 2.4 ± 0.4 Pg C year-1 for the period from 1990 to 2007. This

number has been contested by some scientists, e.g. by S. Joseph Wright (2013), but in general

there is little doubt about the occurrence and approximate magnitude of this substantial carbon

sink. The sink strength is highest in tropical forests, presumably because conditions for

photosynthesis are very favourable during most time of the year (Grace et al., 2014). The

difference with temperate forests is nevertheless not that big, because respiration is also very high

in the tropics and most temperate forests are in a relatively juvenile phase, being managed to be

productive with trees in their most active growth phase (Luyssaert et al., 2008). But then again the

significance of tropical forests also comes from the fact that they extent a vast area of ~1770 M

ha (Keenan et al., 2015). Approximately 45% of all terrestrial C is estimated to be stored in

tropical forests, accounting for a carbon stock as biomass that adds up to 271 ± 16 petagrams

(Pg) and a quantity of carbon as soil organic matter that is possibly even higher (Anderson-

Teixeira et al., 2016; Grace et al., 2014). However although they constitute the largest component

of the terrestrial C sink, carbon loss from deforestation, forest degradation, harvesting and fires

in tropical forests is estimated higher than the carbon gained from forest and woodland growth.

Grace et al. (2014) reported this source to be 2.01 ± 1.1 petagrams of carbon per year (Pg C year-

1) and a tropical forest sink of 1.85 ± 0.09 Pg C year-1 for the period from 2005 to 2010.

Despite the major importance of tropical forests to the global C cycle, their ecosystem-level C

cycling is not as well understood as that of extra-tropical forests (Anderson-Teixeira et al., 2016).

Especially on the African continent information about the carbon balance is still sparse (Ciais et

al., 2009; Williams et al., 2007). Nevertheless progress has been made with the creation of

AfriTRON (the African Tropical Rainforest Observatory Network), a network of monitoring

plots which already lead to the first large-scale comparative study of above-ground biomass of

African tropical forests published in 2013 (Lewis et al., 2009, 2013).

Data on above-ground biomass and their spatial patterns acquired via long-term monitoring plots

are very valuable. They provide insights into how tropical forests function, deliver information

for estimating carbon losses from deforestation and forest degradation, assist calibrating and

validating carbon mapping exercises and are necessary for modelling tropical forests and their

response to a changing environment.

CO2

Plant

biomass

Medium- term

carbon storage

Long-

term

carbon

storage

Figure 3. Terrestrial ecosystems carbon uptake and storage. Figure adopted from Walker and Steffen (1997).

Page 19: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

7

Notwithstanding important work by scientists such as Jobbágy & Jackson (2000), Schwartz &

Namri (2002) and Batjes (2008), robust estimates of soil organic carbon (SOC) pools and its

spatial and temporal variability continue to be lacking (Dieleman et al., 2013; Don et al., 2011).

Models that simulate the carbon cycle and vegetation dynamics are crucial to have an idea what to

expect in future global change scenarios and to identify possible tipping points. In tropical forests

anthropogenic deforestation will play a primary role in the evolution of the carbon cycle but

other carbon releases driven by climate change may not be neglected. Results for the twenty-first

century by Cramer et al. (2004), who applied a dynamic global vegetation model, confirm large

losses of carbon despite the uncertainty about deforestation rates. Additionally some climate

models produced large carbon fluxes due to increased drought stress caused by rising

temperatures and decreasing rainfall. However one climate model produced an additional carbon

sink. The magnitude of the effect of climate change on tropical forests remains quite uncertain

(Malhi et al., 2008) but coupled climate-carbon-cycle models often agree on two counteracting

effects on terrestrial carbon storage (Cox et al., 2013):

1. an increase due to the simultaneous enhancement of plant photosynthesis and water use

efficiency under higher atmospheric CO2 concentrations

2. a decrease due to higher soil and plant respiration rates associated with warming

temperatures

The balance between these effects varies enormously - up to a range of 330 gigatonnes in the

projected change by 2100 - according to the use of different models. Across multiple models Cox

et al. (2013) found an emergent linear relationship between the sensitivity of tropical land carbon

storage to warming and the sensitivity of annual growth rate of atmospheric CO2 to tropical

temperature anomalies. This provided a tight constraint on the sensitivity of tropical carbon to

climate change leading to an estimation of 53 ± 17 gigatonnes of carbon per degree Celsius

released over tropical land by warming alone. However these models depart from a (large) CO2

fertilization effect, increasing terrestrial carbon uptake due to increasing atmospheric CO2

concentrations, which is still being heavily debated (Battipaglia et al., 2015; van der Sleen et al.,

2014). This again demonstrates the need of more data on carbon cycling in tropical forests.

Page 20: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

8

2.3 Tropical montane forests

Tropical montane forest (TMF), being evolutionarily and ecologically distinct from tropical

lowland forest, is a type of forest that has been underrepresented in studies concerning ecosystem

ecology and biogeochemical cycles. In fact, TMFs are even more poorly researched than the

lowland tropics (Bubb et al., 2004). To better understand their role in the global carbon cycle,

there is a need of extra studies aimed at quantifying the overall carbon balance of TMFs

(Bruijnzeel and Veneklaas, 1998). A study by Spracklen and Righelato (2014) suggests that

current regional carbon stock estimates are too small since they are usually reported for the

planimetric area while the real surface area is 40 % greater due to the prevalence of steep slopes.

Malhi et al. (2010) clearly identified five major justifications for investing effort in studying

TMFs. Tropical elevations are for instance excellent sites for understanding past climate change

in the tropics and tropical mountains are potential arks for biodiversity in a century of rapid

warming. Furthermore environmental responses to global change drivers in TMFs may provide

an early indication of what the future holds for many of the world’s ecosystems (González et al.,

2013).

2.3.1 Characteristics

Altitudinal zonation

The large heterogeneity of the tropical forest biome encourages classification. Different

classification systems exist but an obvious recurrent distinction is the elevation. Elevation

influences climatic factors, such as temperature, incoming solar radiation, humidity and the

presence and frequency of clouds and fog, which are in its turn responsible for a lot of botanical

characteristics of the forest. In natural systems TMFs are delineated by the ecotone between

lowland tropical forest and lower montane tropical forest as lower limit and the treeline, where

the mountains are high enough, as upper limit. Ideally the lower mountain limit should be

defined climatically or according to the vegetation for every region over the whole world. Since

this is not feasible conventions are used. In the tropics the lower limit is usually set at 1000

m.a.s.l. (Körner et al., 2005).

While increasing in altitude different vegetation zones occur. This phenomenon, called altitudinal

zonation, quite early aroused the interest of several scientists. The famous Alexander von

Humboldt even wrote about it in the beginning of the 19th century (Humboldt as cited in Bach

and Gradstein, 2011). Different approaches to define this altitudinal belts have been used, based

upon abiotic as well as biotic factors. Troll (1959) for instance suggested a zonation based on

thermohygric conditions but most zonation schemes are contingent on biotic factors such as

vegetation structure and floristics (Bach and Gradstein, 2011). Four general forest zones on

tropical mountains have been recognised worldwide: lowland, lower montane, upper montane

and subalpine (Ashton, 2003). Classically the emphasis of research on altitudinal zonation was on

structural and physiognomic criteria, e.g. Richards (1952) and Grubb et al. (1963) but recent

studies focus mainly on floristics. The dependence on altitude of some physiognomic or

structural features such as the tree height and leaf size, are quite straightforward whereas others

such as the number of strata (vegetation layers) are much more discussable. Described characters

of structure and physiognomy for the altitudinal zones in tropical rain forest are presented in

Table 1.

Page 21: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

9

Table 1. Characters of structure and physiognomy of the four general altitudinal forest zones (adopted from Ashton (2003))

An illustration of a more recent study on altitudinal zonation is the one by Andreas Hemp (2006).

Hemp produced the first statistically significant floristically defined altitudinal zonation of the

vegetation of Mt Kilimanjaro. On 600 plots over a range of 2400 m the entire vegetation (trees,

shrubs, epiphytes, lianas and herbs) was surveyed and classified by the phytosociological method

of Braun-Blanquet (1964). With his study Hemp showed clearly distinguishable community units

to be a result of stable and distinct ecological altitudinal differences. Although the term

“zonation” actually implies these altitudinal belts and distinct ecotones, whether a continuous

change over the whole range occurs or whether these changes are discontinuous, is still a matter

of debate. For instance Hamilton (1975) and Lovett (1998) did not observe a zonation of forests

on different East African mountains. Hemp on the other hand suggested that distinct differences

occur in all tropical montane forests and stated that not observing an altitudinal zonation of

forests may have resulted from different sampling methods and intensities or the mix of data

from different slopes. Obviously the transition between successive altitudinal belts can be much

more gradual in certain cases than in others but at least some show abrupt boundaries of forest

vegetation belts with very distinctive characteristics, e.g. bamboo zones.

Boundaries of altitudinal zones are mainly determined by factors closely related to altitude but

other factors (which are not always strongly correlated with altitude) such as rainfall and edaphic

soil conditions can be of major importance as well (see section 2.3.2). Soil conditions and

altitudinal nutrient shifts possibly responsible for vegetation changes will be elaborated in section

2.3.3.

Tropical montane cloud forests

A specific forest formation which also typically occurs in a relatively narrow altitudinal belt is

tropical montane cloud forest (TMCF) (Aldrich et al., 1997). However these cloud forests can be

quite extensive as well and therefore they are not seen as a zone occurring in tropical montane

forest but rather as a subtype of the latter (Bruijnzeel and Veneklaas, 1998). Numerous

definitions of this specific forest type exist but a broad generally accepted definition describes

TMCFs as tropical montane forests that are frequently covered in clouds or mist (Bruijnzeel and

Formation Tropical lowland evergreen rain forest

Tropical lower montane rain forest

Tropical upper montane rain forest

Tropical sub-alpine forest

Canopy height, emergent trees 24–45 m 15–33 m 1.5–18 m 1.5–9(15) m

Number of strata 3 2 1 (1)

Principal leaf size-class of woody plants Mesophyll Notophyll Microphyll Nanophyll

Pinnate leaves Frequent Rare Very rare Very rare

Buttresses Usually frequent & large

Uncommon or small

Usually absent None

Cauliflory Frequent Rare Absent (Absent)

Woody climbers Abundant Usually none None None

Bole climbers Often abundant Frequent to abundant

Very few Few

Vascular epiphytes Frequent Abundant Frequent Frequent

Non-vascular epiphytes Occasional Occasional to abundant

Often abundant Abundant

Page 22: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

10

Hamilton, 2000). So in addition to rainfall, water droplets are captured directly from the low

cloud cover by condensation on the vegetation, strongly influencing the hydrology, ecology and

soil properties of these forests (Bubb et al., 2004). Other consequences of frequent cloud cover

are a reduced total solar radiation (Hamilton, 1995) but an increased insolation of UV-B light due

to the reflection from clouds (Flenley, 1995). All climatic conditions together are responsible for

the distinctive floristic and structured form of TMCF, e.g. small xerophyllic leafs, gnarled trunks,

distinctive species and a higher proportion of biomass as epiphytes.

The altitude at which cloud forests in tropical mountains occur, is influenced by the location. At a

greater distance from the equator, condensation will occur at lower altitudes due to lower average

temperatures (Bruijnzeel and Hamilton, 2000). At similar latitudes altitude will be influenced by

local effects such as the size and orientation of the mountain, sea surface temperatures and

exposure to the prevailing winds. The effect of the size of the mountain deserves special

attention. Cloud forests on small (isolated) mountains, especially on islands or in coastal areas,

generally occur at reduced altitudes in comparison with the altitude of cloud forest on larger

mountains away from the coast (Bruijnzeel and Hamilton, 2000; Bruijnzeel et al., 2011;

Bruijnzeel, 2001; Martin et al., 2011). This effect is called the ‘telescoping’, ‘Massenerhebung’ or

mass elevation effect. The first main reason to explain this phenomenon is that bigger mountains

have higher (overlying air) temperatures because their landmasses absorb more solar radiation.

The second main raison is the higher atmospheric moisture content of oceanic air, leading to

lower cloud formation.

Distribution

Few accurate data on the distribution of tropical montane forests exist but Kapos et al. (2000)

estimated an area of 3 257 275 km2 which included altitudinal ranges from 300 m to above 4 500

m. More effort has been put in the distribution of cloud forests. Global forest cover assessments

together with global elevation data and regionally specific altitudinal limits were used to estimate

potential as well as actual areas of TMCF (Bruijnzeel et al., 2011). The potential area, the entire

area between the specified altitudinal limits regardless of actual forest cover, was estimated

around 380 000 km² whereas the actual area was estimated at 215 000 km². The actual area

estimated in this way constitutes 1.4% of the world’s tropical forest area and 6.6% of the area

covered by TMF as defined above. An indication of the locations of TMCFs is given by a map

with more than 560 sites with confirmed cloud forest presence (Figure 4). Of all TMCFs 46%

can be found in Asia and northern Australia, 43% in the Americas and the Hawaiian archipelago

and 16% in Africa.

Another approach in which a hydro-climatic definition of tropical montane cloud forest was

used, yielded a much greater global extent of 2.21 M km² (Bruijnzeel et al., 2011). This can be

explained because fragmented areas of forest and sparse woodland at 500-m resolution were

included and a wider range of conditions and structural features, e.g. certain lowland forests that

are significantly cloud-affected, was covered.

Both approaches compared actual distributions with potential distributions, showing great losses

due to deforestation. The first approach suggests that ~56% of the original forest still remains

whereas the second suggests that ~45% of its original distribution still remains. TMCFs are

undeniably rare and threatened ecosystems.

Page 23: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

11

Figure 4. Tropical montane cloud forest sites by UNEP-WCMC.

Biodiversity and ecosystem functions

Tropical montane forests have unique characteristics of biodiversity and high endemism (Aldrich

et al., 1997; Spracklen and Righelato, 2014). Many areas serve as refugia for species which are

endangered by the transformation or destruction of ecosystems at lower elevations. Furthermore

TMFs are much appreciated for the provision of hydrological services and freshwater supply. For

instance the montane forests of Kenya are known as “Kenya’s Water Towers”, yielding possibly

more than 15800 million cubic meters of water per year (Crafford et al., 2012).

2.3.2 Altitudinal transects

From a scientific viewpoint TMFs are interesting to study because the mountainous area in which

they grow, displays a rapid rate of change in environmental characteristics with relatively short

geographic distances (González et al., 2013). The responses of ecosystems on these

environmental gradients are adequately studied by means of altitudinal transects. To illustrate the

diversity of altitudinal transects studies in TMFs some examples are presented in Table 2.

Furthermore TMFs show recurrent broad-scale ecological patterns that can be detected with

reasonable power and contrasted between different regions. But caution is required in studying

environmental conditions and its influence on the vegetation: it is very important to distinguish

between environmental changes that are physically tied to meters above sea level and those that

are not generally altitude specific (Körner, 2007).

Page 24: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

12

Table 2. Different biological and ecological altitudinal transect studies in tropical montane forests. LAI= leaf area index.

Topic Country of study site Source

LAI and stand leaf biomass Ecuador Moser et al. (2007) Microbial biomass, fungal biomass and microbial community structure

Ecuador Krashevska et al. (2008)

Forest structure and AGB Brazil Alves et al. (2010) Moth diversity Australia Ashton et al. (2016) Vascular plant species richness Malaysia (Borneo) Grytnes & Beaman (2006)

A first altitudinal driver mainly impacting biodiversity and its evolution is the available land area.

With increasing altitude the land area shrinks, narrowing opportunities for life in concurrent

fragmented climatic mountain ‘islands’. The habitat diversity and spatial isolation as a

consequence are also enhancing segregation of populations and thus potentially causing

speciation. Next to this global altitude-related phenomenon four major climatic changes are

associated with altitude:

a decreasing total atmospheric pressure and partial pressure of all atmospheric gases,

a reduction of the atmospheric temperature,

an increasing radiation under a cloudless sky and

a higher fraction of UV-B radiation at any given total solar radiation.

Environmental changes that are not generally altitude specific exert strong influences in TMFs as

well and they have introduced confusion in scientific literature on altitude phenomena (Körner,

2007). Examples of the latter are precipitation, wind velocity, hours of sunshine, geology and

human land use. However the first two are discussable since orographic precipitation and wind

speed often significantly rise with increasing altitude although they are also very dependent on the

mountain topography (Hodkinson, 2005). Data collected along altitudinal gradients will always

reflect the combined effect of regional peculiarities and general altitude phenomena but the

distinction is crucial when the results of different studies are compared and actual physical

conditions must be documented as much as possible.

Temperature, being closely and inversely correlated with elevation, has been suggested as the first

and foremost explanatory variable in changing ecological and biogeochemical processes along an

altitudinal gradient. Vitousek et al. (1992) even used elevation as a surrogate for mean annual

temperature on the Mauna Loa in Hawaii. In the free atmosphere the temperature decrease is

generally taken to lie between 5.5 and 6.5 °C for each km of ascent, although it can vary

according to local conditions, topography and meteorological circumstances (Anslow and Shawn

as cited in Hodkinson, 2005). Raich et al. (2006) discussed the effect of a decreasing atmospheric

pressure and partial pressure of CO2 on forest carbon cycling processes but argued that

elevational gradients provide a useful basis to identify the effect of decreasing temperature on

forest ecosystems.

The temperature decrease with increasing altitude has also been put forward to explain the

existing paradigm that with increasing altitude aboveground carbon decreases while belowground

carbon increases (Leuschner et al., 2013). The reduced carbon stocks at higher altitude are a

consequence of a reduced NPP which was for instance observed in studies from Kitayama and

Aiba (2002) and Waide et al. (1998). An increasing belowground carbon is for example found in

the studies from Dieleman et al. (2013), Kitayama and Aiba (2002) and Leuschner et al. (2007).

Page 25: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

13

The effect of temperature on NPP could be directly or indirectly due to the effect of temperature

on nutrient availability, driven by the rate of decomposition and nutrient mineralisation (Myers,

1975). Moreover Fernández-Martínez et al. (2014) recently demonstrated the crucial importance

of nutrient availability to carbon cycling. This aspect is elaborated in the next section. Other

explanations for a reduced NPP that have been suggested are: persistent cloud cover, high wind

velocity and waterlogging (Bruijnzeel and Veneklaas, 1998).

2.3.3 Nutrient cycling and isotopes

Nutrient cycling is an important aspect that influences the physiognomy of TMFs. While the

carbon cycle has been discussed in section 2.2, the link with nitrogen (N) and phosphorus (P)

availability needs some further explanation. Additionally the role of some cationic elements will

be touched upon and the use of isotopic ratios in biogeochemical and ecological research will be

addressed.

Carbon is taken out of the atmosphere under the form of CO2 (by photosynthesis, see earlier) but

other nutrients necessary for plant growth are absorbed by the roots. Nutrients are subsequently

transported and incorporated in plant tissue. When dead plant material falls on the forest floor,

decomposition makes recycling of these nutrients possible. In lowland tropical rainforest the hot

and humid conditions allow for a rapid decomposition and due to the high demand from the

vegetation most nutrients do not remain in the soil for long (Jones et al., 2013). Thanks to this

rapid consumption nutrients are also concentrated near the soil surface. In other climates and at

higher altitudes temperature, sunlight and/or water are much more limited (or too abundant, e.g.

waterlogging), influencing nutrient cycling and plant growth (Jolly et al., 2005). Nutrient cycling

of a certain element is also dependent upon the availability of other nutrients. Any part in the

biogeochemical process in which availability of one or more nutrients constraints the rate at

which another nutrient cycles, is referred to as nutrient limitation (Townsend et al., 2011). The

most important macronutriens that sustain life on earth are N and P. Their cycling within natural

ecosystems is explained more detailed below. Information on these processes were largely

retrieved from two handbooks: Global Biogeochemical Cycles in the Climate System, edited by

Schulze et al. (2001) and Soil Conditions and Plant Growth, edited by Gregory and Nortcliff

(2013).

Nitrogen and phosphorus cycling

In natural ecosystems atmospheric N enters the soil predominantly through biological fixation by

free-living bacteria or bacteria living in association with N-fixing plants such as legumes

(Fabaceae). Dinitrogen gas (N2), the most abundant atmospheric N form, is reduced to ammonia

(NH3) and incorporated into amino acids for protein synthesis. Smaller quantities enter the

system directly under the form of NH3, ammonium (NH4+), nitrate (NO3

-) and particulate and

organic N. These depositions originate from a variety of sources, most noteworthy for TMFs:

biomass burning, e.g. Velescu et al. (2016). Forest fire in the system itself, on the other hand,

causes an output of N through volatilization.

The largest pool in the soil is found as nitrogenous compounds in the organic matter as a result

from the decomposition of plant material and microorganisms. This organic N is made available

for plants through mineralisation by microbes and soil fauna, resulting in the release of NH4+. On

the contrary, microbial immobilisation reduces the availability of N to plants. Thus the rates of

these processes strongly control the soil N cycle and govern N transfer to the wider environment.

The released NH4+ is converted to nitrite (NO2

-) and subsequently to NO3- in a process called

Page 26: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

14

nitrification which is carried out by bacteria and fungi. NO3- is bio-available and highly mobile.

When it is not taken up by plants or microbes, NO3- can be lost through leaching in the water

and through denitrification in the air. Denitrification is the microbial reduction of NO3- to

gaseous nitric oxide (NO), nitrous oxide (N2O) and N2. Plants are known to take up organic N

forms but mostly take up NH4+ and NO3

-.

In grazed ecosystems a substantial amount of the N taken up by plants is returned in the form of

dung and urine, often dislocated from the area of uptake. This sometimes leads to profound

effects on nutrient dynamics and bioavailability. In forests this role is rather small and not

accounted for on an ecosystem level.

In contrast to atmospheric N inputs, those of P are very limited. P in natural ecosystems is

almost exclusively derived from soluble primary minerals originating from soil parent material. In

the soil P processes are controlled by

the form in which P occurs (organic or inorganic),

inorganic adsorption and desorption reactions,

dissolution and precipitation reactions and

mobilisation and immobilisation by the microbial community.

Recycling from dead plant materials (and dung and urine) as explained for the N cycle also

accounts for P. The detailed explanation of these processes however is out of the scope of this

thesis.

More relevant is the link between the availability of P and N and carbon cycling. Studies on

nutrient cycling in forests span more than 100 years and have been dominated by the role of N as

a limiting factor, especially in temperate forests (Attiwill and Adams, 1993). However in his paper

ca. 30 years ago Vitousek (1984) first discussed at length the idea that phosphorus specifically

limits productivity of many tropical rainforests. It is clear that this idea cannot be generally

applied for all moist tropical forests but his statement seemed to be correct for lowland forests

on older soils (Herbert and Fownes, 1995; Vitousek and Farrington, 1997). These soils are highly

weathered and P has been leached, resulting in the observed limitation. In montane forests, in

contrast, where soils are less weathered and nitrogen mineralization occurs at slower rates, it is

suggested by Tanner et al. (1998) (among others) that nitrogen, rather than phosphorus, limits

productivity. This hypothesis has been adopted a lot and has been confirmed by some

experiments although not all transect studies looking at N limitation in TMFs provided much

evidence (Fisher et al., 2013; van de Weg et al., 2009). While a lot of studies looked at limitation

of single nutrients and compared between ecosystems, recent attention shifted somewhat towards

studying multiple nutrients that could limit growth and acknowledge that even within an

ecosystem different nutrients could be limiting (Velescu et al., 2016).

Cationic elements

Mineral elements needed for plant growth and fecundity are taken up at the root surface from the

soil solution in their cationic forms. Kaspari et al. (2008) showed that their role in ecosystems

may not be overlooked when studying nutrient limitation and biogeochemical processes in

ecosystems.

Page 27: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

15

Isotopic ratios

First the principles and terminology concerning isotopes in this thesis will be explained. The

isotopic ratio R is the molar ratio of a minor (heavier) to a major (lighter) isotope. The relevant

isotopic ratios in this study are 15N/14N for nitrogen and 13C/12C for carbon. The isotopic

composition δ is the normalized departure of R from R*, its value in a standard reference

material, so: δ=R/R* - 1. The reference materials for the isotopic ratios of N and C are

respectively air, so atmospheric molecular N which is by convention set at 0 ‰, and Pee Dee

Belemnite (PDB). From here on the isotopic compositions will be noted as δ15N and δ13C.

Due to fractionation processes, δ15N of plants or soils can be used to infer N cycling but this has

to be done with great caution since it responds on multiple drivers. Our understanding

concerning δ15N in plants and soil is still incomplete but has improved a lot compared to a

decade ago and is extensively reviewed by Craine et al. (2015b). I refer to this article for the full

explanation, but some important influences on δ15N are given here:

Plant δ15N is greatly influenced by the role of mycorrhizal fungi

Plant δ15N as well as soil δ15N are said to be influenced by the role of climate but this is

discussable

Soil δ15N changes occur with microbial processing

Different forms of N have different signatures of δ15N

Higher δ15N in hot/dry systems refers to an open N cycle, while smaller δ15N in cold/wet

systems refer to a closed N cycle

Soil δ15N changes among soil fractions

δ15N is higher when more N is available and decreases when N is limited

δ13C in plants has been used extensively in studies on the carbon-water balance of plants and

ecosystems, being effected by short-term as well as longer term environmental conditions (Seibt

et al., 2008). For instance with independent estimates of gas exchange or environmental

conditions it can be used as proxy for water use efficiency. Furthermore δ13C in soil organic

matter (SOM) is useful as a constraint in carbon cycle models, although our understanding of the

mechanisms leading to 13C enrichments in SOM still needs to be improved (Ehleringer et al.,

2000). Additionally δ13C values can be very valuable in geochemical research (e.g. Watanabe et al.,

(2000)).

Page 28: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

16

2.3.4 Albertine rift montane forests

The Albertine Rift montane forests are tropical and subtropical moist broadleaf forests of a

mountain chain with exceptional faunal and moderate floral endemism (WWF, 2011). The area is

a priority ecoregion for the World Wildlife Fund (WWF) and it supports many endangered

species such as the mountain gorilla (Gorilla beringei beringei) and the eastern lowland gorilla (Gorilla

beringei graueri). These afromontane forests, stretching over the Democratic Republic of Congo

(70%), Uganda (20%), Rwanda (6%), Burundi (3%) and Tanzania (1%), form a part of the

Albertine Rift that extends from the northern tip of Lake Albert to the southern tip of Lake

Tanganyika. Containing the highest number of vertebrate species on the African continent, the

Albertine Rift is one of the most important regions for conservation in Africa (Plumptre et al.,

2007). Top priority sites for conservation in the Albertine Rift based upon the numbers of

endemic and globally threatened species are shown in Table 3.

Table 3. Top priority sites for conservation in the Albertine Rift based upon the paper by Plumptre et al. (2007); NP=National Park

Priority site Country

Virunga NP Democratic Republic of Congo Kahuzi Biega NP Democratic Republic of Congo

Itombwe Massif Democratic Republic of Congo

Bwindi Impenetrable NP Uganda Nyungwe NP Rwanda

Next to its biodiversity, the Albertine Rift montane forests are also very important for the

delivery of other ecosystem services, especially to the people living in the area. However despite

the forests’ high importance, much of them remain poorly studied. Furthermore civil wars and

international conflicts in the region hampered research and tourism development in the past

decades and in more peaceful areas the high density of people (up to 500-600 km-2) threatens

nature through farming activities and poaching.

Page 29: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

17

3 Research questions

Altitudinal transect studies in TMFs concerning biogeochemical processes such as carbon storage

are relatively numerous in America and Australasia but largely lack in Africa. Yet the literature

review clearly illustrated its importance. This study aimed to estimate carbon stocks along an

altitudinal gradient in a montane tropical rainforest in Central-Africa while also assessing forest

structure and nutrient cycling to a certain extent. A particularly interesting approach was applied

by including the isotopic compositions of nitrogen and carbon. More specifically the study tried

to determine significant differences between altitudinal strata and examined if linear trends could

be found with increasing altitude. At its core this study can be brought to one question: how do

the carbon stocks change with the altitude and can a trend in nutrient availability be found?

Figure 5. Images of Nyungwe National Park. Upper: a landscape at the top of Bigugu, the highest mountain of the national park, with large heather vegetation occurring above the treeline; left corner: forests on hillsides at midlevel altitudes; right corner: a closer look at the vegetation at lower to midlevel altitudes (the pictures were taken in August at the end of the dry season)

Page 30: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

18

4 Materials and methods

The study for this thesis is based on data collected during a field campaign in Nyungwe National

Park in August and September 2015.

4.1 Study area

Nyungwe National Park is located in southwestern Rwanda (2°17' – 2°49' S, 29°03'– 29°29' E)

(Figure 6,a) (Gharahi Ghehi et al., 2012). The National Park covers an area of about 1000 km²

with altitudes ranging from 1485 to 2950 masl. (Figure 6,b). In Nyungwe vast stretches of forest

are interrupted by two large swamps. It was gazetted as a forest reserve in 1933 and became a

National Park in 2004 (About Rwanda, 2016). Together with the contiguous Kibira National Park

in neighbouring country Burundi it forms one of the largest blocks of montane tropical forest in

Africa (Plumptre et al., 2002).

Figure 6. Location of the Nyungwe forest in Rwanda. (a) seven climate stations around the forest are presented by red stars, the recent climate station in the Nyungwe forest is presented by a yellow star; (b) elevation map; (c) parent material map of the Nyungwe forest: Q = quartzite; QI = quartzite intercalated with schists; IQm = quartzite intercalated with micaschists; GQ = granitic and quartzitic rocks; G = acid rocks (granite); Gm = micaceous acid rocks (granitoide); Im = micaschists; I = schists (figure adopted from Gharahi Ghehi et al. (2012)).

Page 31: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

19

Climatologically Nyungwe is characterised as a humid tropical climate with a major dry season

between July and August and a small dry season occurring in December and January (Cizungu,

2014). Recent weather data for the surrounding region of the National Park are not easily

available but based upon seven climate stations in the vicinity of Nyungwe (Figure 6, a) useful

data for the period from 1974 to 1989 is at hand (Gharahi Ghehi et al., 2012). Annual

precipitation ranged from 1308 to 2071 mm with an average of 1660 mm and average annual

temperature was 17°C, with small seasonal variations. The average monthly minimum was 11°C

and the average monthly maximum 23°C. However one automatic climate station has been

established inside the forest near Uwinka visitor centre at 2465 m.a.s.l. (2°28′S, 29°12′E) in 2007

(Figure 6, a), demonstrating an average annual mean temperature of 14.5°C and an annual

precipitation of 1824.7 mm.

The main parent materials from which soils in Nyungwe have been developed are schists,

micaschists, quartzitic schists and granites (Figure 6, c). Micaschists are dominant in the eastern

part of the forest, mostly exceeding 2000 m.a.s.l., whereas schists are dominant in the western

part, featuring a larger area of lower altitudes. Dominating soil groups as defined by the World

Reference Base for soil resources (WRB; FAO, 2014) are Acrisols, Alisols and Cambisols (Van

Ranst et al., 2000a, 2000b, 2000c, 2000d). In addition to these Luvisols, Podzols, Regosols,

Leptosols and Ferralsols are associated with sloping terrain in Nyungwe and Histosols and

Gleysols occur in the swamps, small valleys and depressions.

In terms of biodiversity Nyungwe contains one of the most important montane tropical

rainforests in central Africa, especially regarding endemicity (Plumptre et al., 2002, 2007). For

instance 1105 different plants and more than 230 species of trees have been found in Nyungwe

and 13 species of primates are known to inhabit the forest, including chimpansees (Pan troglodytes

schweinfurthii) and the rare owl-faced guenons (Cercopithecus hamlyni). Furthermore the National

Park is also one of the most important sites for bird conservation in Africa with a total of 280

bird species of which 26 are endemic to the Albertine Rift. Concerning the distribution of this

biodiversity surveys by Plumptre et al. (2002) clearly showed that both species richness and

species diversity for most taxa are significantly higher in the western part of the forest than in the

eastern part. Few is known about other ecosystem services of Nyungwe National Park but as a

part of the Congo-Nile watershed, is it definitely also an important water catchment area,

delivering water for about 70% of Rwanda (USAID & WCS 2013).

4.2 Altitudinal transect choice and site selection

Since there was ample prior knowledge on the physical factors directly related with altitude, this

study tried to span the entire altitudinal range of Nyungwe forest in 4 more or less evenly

distributed strata (Table 4, Figure 7). Site location was selected on a certain altitudinal level based

on practical feasibility to work as well as basic forest characteristics. From a practical viewpoint

for instance slopes could not be too steep and too far from hiking trails or roads. Regarding the

forest characteristics we searched for relatively undisturbed old-growth forests with full canopy

closure. The latter was rather difficult, because a great part of Nyungwe forest seemed to show a

quite open structure, with large patches of open terrain dominated by a native but invasive liana,

Sericostachys scandens. As a consequence some plots contain small open places.

Page 32: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

20

Table 4. Basic stratum characteristics. Targeted altitudes where chosen as rough guidelines for the plot altitudes based upon an approximate elevation map of Nyungwe National Park. The number of tree species per stratum was retrieved from Taveirne (2016) who elaborated on functional diversity in the same study site. PSP stands for permanent sample plot. Refecence soil groups were retrieved from Van Ranst et al. (2000a, 2000b, 2000c, 2000d).

Stratum 1 Stratum 2 Stratum 3 Stratum 4

PSP numbers 9,10,11,12,13 6,7,17,18,21 19,20,22,23,24 1,2,3,4,5 Targeted altitude (masl) 1800 2200 2500 2800 Mean altitude (masl) 1760 2200 2512 2865 Min. and max. altitude (masl) 1659, 1835 2141, 2293 2456, 2557 2767, 2937 Number of tree species 46 37 20 15 Reference soil groups (WRB) Cambisol Cambisol,

Alisol Cambisol,

Alisol Regosol, Leptosol (1,2)

Cambisol (3,4,5)

Figure 7. Map of the altitudinal transect in Nyungwe National Park with the dots representing the locations of the PSPs and the boxes giving details of the measured altitude per plot. The grey star is the location of the visitor’s center Uwinka (adopted from Taveirne (2016)).

Page 33: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

21

4.3 Establishment and measurement of permanent sample plots

To make sure that the results can be compared with other studies, the establishment and

measuring of the permanent sample plots (PSPs) was done following a modified worldwide

standardised method drawn up by the Amazon Forest Inventory Network (RAINFOR, Y. Malhi

et al., 2002). The 2016 revised version of this protocol, the ‘Field Manual for Plot Establishment

and Remeasurement’ (Phillips et al., 2016) is available on the RAINFOR website.

Site conditions caused restrictions in

following the protocol and some adoptions

were made. Due to the extreme topography

an area smaller than the standard size of 1 ha

recommended by RAINFOR was chosen.

On the selected sites square plots of 40x40m

were accurately delineated by compass, for

the ease of working with the sides parallel

and perpendicular to the slope. The plot

consisted of four square subplots with sides

of 20m (Figure 8) and the slopes of all sides

where measured with a laser rangefinder

(Forestry Pro, Nikon, Japan). The plots were

made permanent by digging out the corners

and placing a brick in the centre. On every

corner and in the centre coordinates and

altitude were recorded by GPS (GPSMAP®

64s, Garmin, Taiwan).

Within these plots all living trees with a minimum diameter at reference height of 0.1 m were

tagged and identified by a local botanist after which diameters were measured. Each tree was also

given a status code according to the RAINFOR protocol (Appendix A). A tree was included in

the plot if more than 50% of the root system was situated within the plot boundaries. The

reference height and the point of measurement (POM) were 1.3 m, also referred to as diameter at

breast height (DBH), and tags were nailed in the stem at 1.6 m, or 0.3 m above the POM. If 1.3

m was not used as the POM in order to avoid abnormalities such as trunk deformities or buttress

roots, the height of the alternate POM was carefully recorded. The height at which diameters

were measured is not the vertical height above the ground but the straight line distance along the

trunk. On slopes the downhill side of the tree was used to measure 1.3 m and for leaning trees

this length was measured along the side of the stem closest to the ground. In case of climbing

vines or roots from other plants growing against the stem at DBH the measuring tape was passed

underneath. In the rare cases where lianas or stranglers were to firmly attached to the tree stem,

the diameter was estimated by holding the tape straight to the stem at the POM to optically

determine an approximation. If a tree had multiple stems, all stems with a DBH greater than or

equal to 0.1 m were measured, tagged and recorded. Furthermore both the total tree height and

tree bole length (the height until the lowest big branch) were measured with the Nikon Forestry

Pro for approximatively 20 % of the trees in each plot. We aimed to cover all diameter classes.

Figure 8. The permanent sample plot setup. A, B, C and D indicate the four corners, E is the plot centre. Subplots are indicated with the roman numbers I to

IV. A was always located in the northwest of the plot.

Page 34: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

22

Lianas were tagged and measured if its diameters were greater than or equal to 0.1 m at any point

within 2.5 m of the ground, even if they had a diameter smaller than 0.1 m at 1.3 m from the

base. Each climbing liana stem that was in accordance with this criterion and that was separately

rooted, counted as one individual plant. Diameters were measured at three points:

at 1.3 m along the stem from the principal rooting point

at 1.3 m vertically above the ground

at the widest point on the stem within 2.5 m of the ground, including any deformity

The lianas were tagged 0.3 m above the POM, 1.3 m vertically above the ground. If lianas or trees

were clearly elliptic in cross-section at the POM, the stems were measured both conventionally

(with the measuring tape wrapped around the whole stem) and by measuring the linear distance

of the maximum dimension and the linear distance perpendicular to the latter after which the

geometric mean was calculated.

4.4 Sampling and processing of soil, litter, wood and leafs

At 5 places within each plot (1 near the centre and 4 randomly chosen in each subplot) the

mineral soil was sampled by drilling to a depth of 100 cm using a soil auger (Eijkelkamp). This

depth was sectioned into 6 segments: 0-5, 5-10, 10-20, 20-30, 30-50 and 50-100 cm. For each

depth the 5 samples were adequately mixed and the composite samples, just like all other samples

described hereafter, were sun-dried in the field as much as possible.

At 3 places within each plot, randomly chosen in 3 different subplots, litter, the O-horizon

(consisting of humus and roots) and the upper mineral soil - now for bulk density measurement -

were sampled. To collect a sample of the litter and the O-horizon, a square metal framework of

25x25 cm was used whereas the soil sample of the upper mineral layer was taken by means of a

100 cm³ steel Kopecky ring (height 5 cm, diameter 5.3 cm, Eijkelkamp).

Per stratum the most abundant tree species, covering 95% of the basal area, were determined and

used to sample leaves and wood cores. Per tree species at least 10 leaves of 3 randomly chosen

individuals in each stratum were collected by a climber. Preferably leaves receiving full sunlight

were collected but if this was not practically feasible, shade leaves were taken instead, while

carefully making notes of these light conditions. Wood samples were collected using an

increment borer (16”, Haglöf, Sweden) both on the uphill and on the downhill side of the trees,

approximately at DBH. Leafs were dried, stored and transported in paper envelopes whereas

paper straws were used for the wood core samples.

The soil samples of the upper mineral layer, needed to determine the soil bulk densities, were

dried and weighed in the lab of the Catholic University of Bukavu (Université Catholique de

Bukavu, UCB), Democratic Republic of Congo. All other samples were transported to Belgium

and dried for at least 48 hours in an oven at 60°C. After the fragmentation layer samples were

dried, they were sieved through a 1 cm screen, to roughly separate the extensive root system3

from the humus. To homogenize the samples, a planetary ball mill (PM 400, Retsch, Germany)

was used for the soil samples and an ultra centrifugal mill (ZM 200, Retsch, Germany) was used

for the litter, leaves, humus and roots with a diameter < 1 cm.

3 In all strata the fragmentation layer consisted of humus and an extensive system of fine roots that were highly connected

Page 35: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

23

Wood samples were used to determine the wood density with the water displacement method.

First all samples were accurately weighed and secondly they were submerged in a small water

reservoir on a precision balance. When fully submerged the weight of the displaced water was

calculated, being equal to the volume of the wood sample.

Of the soil, humus, roots, litter and leaves subsamples were taken and carefully weighed on a

precision balance. These subsamples where then analysed with an elemental analyser (ANCA-SL,

SerCon, Crewe, UK) coupled to an isotope ratio mass spectrometer (20-20, SerCon, Crewe, UK)

(EA-IRMS). Through the former carbon and nitrogen concentrations were obtained, through the

latter δ15N and δ13C were obtained. Additionally soil samples for a depth of 0-5, 5-10, 10-20 and

20-30 cm were weighed and mixed to obtain samples for a depth of 0-30 cm for each plot. These

composite samples were consequently analysed for P concentrations and exchangeable cations.

Total P and cation concentrations were measured at the Forest & Nature Lab (ForNaLab,

UGent) in Gontrode. Total P was measured colorimetrically with malachite green after

destruction in acid while atomic absorption spectroscopy after extraction with barium chloride

was used to determine K, Al, Na, Ca and Mg concentrations. Bio-available P was measured using

spectrophotometry at a wavelength of 650 nm after extraction with hydrochloric acid and anion

exchange membrane (AEM) strips. The lab protocol for this analysis is given in Appendix B.

4.5 Planimetric plot areas

Per subplot the distance of each side on a

horizontal plane was calculated by

multiplying the actual distance (20m) with

the cosine of the slope angle θ, which was

measured in the field (Formula 1). Since the

projection of a square on a horizontal plane

has to be a parallelogram, the opposite sides

need to equal. Consequently the averages of

the opposite sides on the horizontal plane

were taken. To calculate the surface area of

the parallelogram (Formula 2), i.e. the

estimated planimetric area of a subplot, the basis now has been determined but the height is still

lacking. Therefore the mean height, i.e. the altitudinal difference between two points, of the

diagonal in the square subplot was calculated using Formula 3 on all sides, after which the

projection of the diagonal could be calculated using the Pythagorean theorem (Formula 4). After

calculating the angle α between the two sides of the parallelogram using the law of cosines

(Formula 5), the height was easily obtained (Formula 6) and the surface area of the parallelogram

could be calculated. The planimetric areas of the four subplots (parallelograms) were summarized

to obtain a value for the whole plot. Slope values for the whole PSPs were calculated using the

following formula: slope angle=arccos(planimetric area/surface area).

Formula 1 Lp=cos(θ)*20 Formula 2 areap=bp*hp Formula 3 h=sin(θ)*20 Formula 4 Dp=√(20²+20²-hD²) Formula 5 α=arccos((bp²+zp²-Dp)/(2*bp*zp)) Formula 6 hp=sin(α)

Table 5. Formulas needed to calculate an estimation of the planimetric plot areas. Subscript p refers to the parallelogram; subscript D refers to the diagonal; b, z and h are respectively the basis, adjacent side and height.

Page 36: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

24

4.6 Estimation of carbon stocks

4.6.1 Aboveground carbon

As a first step in estimating aboveground carbon (AGC) in the plots, the improved pan-tropical

allometric relation for aboveground biomass (AGB) of individual trees by Chave et al. (2014) was

used on all measured trees. Subsequently these values were divided by 2, to obtain an estimation

of the AGC per tree, and summarized per plot. The formula by Chave et al. (2014) provides the

most accurate general estimation of AGB in tropical trees at hand. It relies on trunk diameter

(DBH), total tree height (H) and wood specific gravity (or wood density, ρ):

𝐴𝐺𝐵𝑡𝑟𝑒𝑒 = 0.0673 ∗ (𝜌 ∗ 𝐷𝐵𝐻2 ∗ 𝐻)0.976

where ρ is in g cm-3, DBH in cm and H in m. All DBH’s in the plots were measured in the field,

but measurements of all total tree heights and wood densities would have been too time-

consuming. Tree heights of the lacking trees were therefore determined using 16 published

diameter-height relationships (Table 6). First non-linear least-square estimates of the parameters

(a,b and c) were performed. Secondly the best relationship per plot was chosen according to the

Akaike Information Criterion (AIC). Consequently the lacking height values for the remaining

trees were estimated.

Table 6. Height-diameter models used for the performance test and total tree height estimates.

Formula Model type Source

𝐻 =1.3+𝑎∗(1−𝑒 – 𝑏 ∗ 𝐷𝐵𝐻 ^ 𝑐) Weibull Huang et al. (1992)

𝐻=1.3+𝑎∗(1−𝑒 − 𝑏 ∗ 𝐷𝐵H) 𝑐 Chapman-Richards Huang et al. (1992)

𝐻=1.3+𝑎/(1+𝑏 – 1 +𝐷𝐵𝐻 − 𝑐) Modified logistic Huang et al. (1992)

𝐻=1.3+𝑎∗𝑒 𝑏 / ( 𝐷𝐵𝐻 + 𝑐) Exponential Huang et al. (1992)

𝐻=𝑎∗(1−𝑒 – 𝑏 ∗ 𝐷𝐵𝐻 ^ 𝑐) Weibull without POM Scaranello and Alves (2012)

𝐻=1.3+𝑎/(1+𝑏∗𝑒 – 𝑐 ∗ 𝐷𝐵𝐻) Logistic Scaranello and Alves (2012)

𝐻=𝑎/(1+𝑏∗𝑒 – 𝑐 ∗ 𝐷𝐵𝐻) Logistic without POM Scaranello and Alves (2012)

𝐻=1.3+𝑏∗𝑒 𝑎 + 𝑏 / ( 𝐷𝐵𝐻 + 1) Exponential Scaranello and Alves (2012)

𝐻=𝑎∗𝑒 − 𝑏 ∗ 𝑒 ^ ( − 𝑐 ∗ 𝐷𝐵𝐻) Gompertz Scaranello and Alves (2012)

𝐻=1.3+𝑎∗𝐷𝐵𝐻/(𝑏+𝐷𝐵𝐻) Hyperbolic Scaranello and Alves (2012)

𝐻=1.3+𝐷𝐵𝐻²/(𝑎+𝑏∗𝐷𝐵𝐻) ² Hyperbolic Scaranello and Alves (2012)

𝐻=1.3+𝑎∗𝐷𝐵𝐻𝑏 Power Scaranello and Alves (2012)

𝐻=𝑎∗𝐷𝐵𝐻𝑏 Power without POM Scaranello and Alves (2012)

𝐻=𝑒 𝑎 + 𝑏 ∗ log (𝐷𝐵𝐻) Exponential Brown et al. (1989)

𝐻=𝑎−𝑏∗𝑒 – 𝑐 ∗ 𝐷𝐵𝐻 Non-linear exponential Feldpausch et al. (2012)

𝐻=𝑎∗(1-𝑒 – 𝑐 ∗ 𝐷𝐵𝐻) Non-linear exponential Banin et al. (2012)

Wood densities were measured if core samples were taken in the field (see section 4.4), if not, the

density was derived from the global wood density database DRYAD (Zanne et al., 2009). When

available the value at species-level was assigned, otherwise the value at genus-level was chosen. If

still no match was found, the plot average was assigned rather than averaging at family-level

because within-family wood densities show very high variability (Chave et al., 2006).

Page 37: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

25

4.6.2 Carbon in litter and belowground carbon

Soil carbon stocks to a depth of 30 cm and 100 cm were quantified by summarizing the stocks

from the aforementioned depth increments (0-5, 5-10, 10-20, 20-30, 30-50 and 50-100 cm). For

each depth increment, carbon stocks were determined as a product of bulk density (g cm-3),

carbon concentration (%) and thickness of the increment layer (cm). As bulk densities were only

measured in the first increment layer of the mineral soil (0-5 cm), those values were applied (as an

estimation) for the deeper increment layers as well. An estimation of the carbon stocks in litter,

roots and the O-horizon was calculated by multiplying the carbon concentration with the dry

weight, while accounting for the sampled area. Due to the small sample size, this only provided a

rough estimate.

4.7 Statistical analyses

Statistical analyses, as well as the carbon stock estimations in the previous section, were

performed using the open source programming software R (R Development Core Team, 2014).

Per stratum the average and standard deviation for all plot variables were calculated. To test if

there were significant differences between the 4 strata, first the Kruskal-Wallis test was carried

out and subsequently the Mann-Whitney U test (also called Wilcoxon two-sample test) was

performed. The latter detected significant differences (p<0.05) pairwise between strata and is

indicated in the tables with different letters. For instance if for a certain variable the values of

stratum 1 and 2 are significantly different, but those of stratum 2 and 3 on one hand and stratum

1 and 3 on the other hand are not significantly different, than the mean values would be indicated

with (a), (b) and (ab) respectively for stratum 1, 2 and 3. In addition simple linear regressions

were performed with altitude as explanatory variable.

Page 38: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

26

5 Results

5.1 Planimetric plot area

Since the distances of the plot sides were accurately measured in the field, real surface areas are

assumed to be 1600 m². Planimetric areas ranged from 1155.95 m² to 1590.79 m², accounting for

respectively 72.25 and 99.42 % of the surface area. 4 of the 19 plots displayed steep slopes (slope

angles > 27°). Note that the planimetric area and slope for plot 21 are missing. Due to logistical

problems the measurements of the slope angles on the sides of this plot could not be pursued.

Table 7. Planimetric plot areas, absolute and relative to the surface area, followed by the slope angles.

Stratum Plot Area (m²) Rel. area (%) Slope (°)

1 9 1155.94 72.25 43.74 10 1447.52 90.47 25.22 11 1433.04 89.56 26.41 12 1425.79 89.11 26.99 13 1498.24 93.64 20.54 2 6 1392.30 87.02 29.52 7 1543.25 96.45 15.31 17 1441.81 90.11 25.69 18 1435.85 89.74 26.18 21 / / /

3 19 1413.84 88.36 27.91 20 1485.27 92.83 21.83 22 1402.10 87.63 28.80 23 1557.43 97.34 13.25 24 1590.79 99.42 6.15 4 1 1461.95 91.37 23.98 2 1516.90 94.81 18.55 3 1532.40 95.78 16.71 4 1496.89 93.56 20.68 5 1508.28 94.27 19.49

5.2 Forest structure

6 forest structure variables are summarised per stratum in Table 8 and visualised in Figure 9.

Dominant tree height, here defined as the average of the three highest trees measured in each

plot, was taken as a proxy for canopy height and decreased 1.5 fold from stratum 1 to stratum 4.

The average diameter at breast height and the tree basal area also showed a decreasing trend,

however with lower significance (0.01<p<0.05 in contrast to p<0.001 for dominant tree height).

Remarkably, no significant differences were found for these variables between the two lower

strata and the two higher strata. Slenderness and stand basal area displayed no trends or

significant differences between any strata whereas the stem density showed an increase from an

average of 436 trees per ha in stratum 1 to 773 trees per ha in stratum 4. Values varied strongly

between plots of the same stratum and a significant difference was only found between stratum 2

and 3. Additionally Taveirne (2016) also reported an increase in tree dominance at the higher

strata. Dominant tree species and their share per stratum are given in Table 9.

Page 39: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

27

Figure 9. Average forest structure variables per PSP were plotted against the altitude. When significant (p<0.05) linear regressions were visualised and R² and p-values were given. (a) Dominant tree height in m; (b) Diameter at breast height in cm; (c) Slenderness, the ratio of tree height over DBH in m cm

-1; (d) Tree basal area in m²; (e) Stand basal area in m² ha

-1; (f)

Stem density, the number of trees per ha.

Table 8. Summary of the chosen variables that quantify the forest structure per stratum. DBH=diameter at breast height; slenderness=tree height/DBH; TBA=tree basal area; BA=(stand) basal area; KW gives the significance of the Kruskal-Wallis test: * = p<0.05, ** = p<0.01, *** = p<0.001; the letters between brackets indicate the results of the Mann-Whitney U test.

Table 9. Tree species dominance per stratum. Data retrieved from Taveirne (2016).

KW Stratum 1 Stratum 2 Stratum 3 Stratum 4

Dominant tree height (m) ** 29.22 ± 4.97 (a) 30.59 ± 2.09 (a) 20.36 ± 3.14 (b) 19.13 ± 0.84 (b) DBH (cm) ** 27.47 ± 2.77 (ab) 29.72 ± 1.06 (a) 20.99 ± 4.82 (b) 22.62 ± 3.52 (b) Slenderness (m cm-1) n 0.56 ± 0.05 0.57 ± 0.02 0.67 ± 0.07 0.60 ± 0.04 TBA (m²) ** 0.08 ± 0.02 (ab) 0.10 ± 0.01 (a) 0.05 ± 0.02 (b) 0.05 ± 0.02 (b)

BA (m² ha-1) n 34.02 ± 3.74 45.16 ± 9.27 31.39 ± 8.60 34.45 ± 4.32 Stem density (ha-1) * 436 ± 113 (ab) 443 ± 114 (a) 760 ± 234 (b) 773 ± 329 (ab)

Stratum 1 Stratum 2 Stratum 3 Stratum 4

Dominant tree species and its share

Cleistanthus polystachyus (15.5%) Grewia mildbraedii (13.8%) Strombosia scheffleri (6.3%)

Cleistanthus polystachyus (13.2%) Strombosia scheffleri (8.9%)

Syzygium guineense (30.7%) Psychotria mahonii (15.6%)

Podocarpus latifolius (33.0%) Psychotria mahonii (30.3%)

Page 40: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

28

5.3 Carbon stocks

The carbon stocks in ton ha-1 per PSP were estimated according to the methods described in

section 4.6. The Kruskal-Wallis test indicates significant differences for AGC and root carbon.

AGC decreases from stratum 2 to stratum 4. By contrast, there is a significant (p<0.05) increase

in C in the roots with rising altitude (Table 10). The estimates for AGC and BCG in ton ha-1 per

PSP were plotted relative to the altitude. Figure 10 shows the AGC estimations and demonstrates

a slight, although just not significant, declining trend in carbon. However, BGC estimations do

not show a correlation to altitude (Figure 12). Nevertheless Table 11 elaborates in more detail on

the soil C in the different increment layers and here a significant increase is seen in the 10-20cm

layer from stratum 2 to stratum 4. The table also demonstrates a decrease of C with depth. The

soil C values are also visualized in Figure 13 for three different depth increments (0-5cm, 10-

20cm and 50-100 cm). A strong correlation is shown with in the first depth with an R2 of 0.241.

The carbon stocks of the humus layer roots are visualized in Figure 11. A highly significant

regression shows that the ton carbon per hectare in the roots correlates to the altitude with an R2

of 0.444. The carbon stocks for the O-horizon, litter and BGC from 0-30 cm show no trends and

are given in Appendix C.

Table 10. Summary of the estimated carbon stocks per stratum in ton ha-1

(mean ± SD). KW gives the significance of the Kruskal-Wallis test: * = p<0.05, ** = p<0.01, *** = p<0.001; the letters between brackets indicate the results of the Mann-Whitney U test.

Table 11. Belowground carbon in g kg-1

for all depth increment layers (mean ± SD). The letters between brackets indicate the results of the Mann-Whitney U test.

KW Stratum 1 Stratum 2 Stratum 3 Stratum 4

AGC ** 120.88 ± 10.82 (a) 178.80 ± 26.74 (b) 98.50 ± 35.98 (ac) 89.49 ± 9.60 (c) C in Litter n 3.80 ± 0.68 3.87 ± 0.62 3.46 ± 0.71 3.24 ± 0.55 C in O-horizon n 28.81 ± 12.60 (a) 48.43 ± 10.50 (ab) 52.21 ± 19.08 (b) 44.24 ± 19.30 (ab)

C in Roots * 1.86 ± 0.58 (a) 2.55 ± 1.43 (ab) 4.64 ± 1.88 (b) 6.16 ± 3.20 (b) BGC (0-30 cm) n 89.02 ± 34.82 96.55 ± 5.67 98.64 ± 16.79 98.67 ± 17.90

BGC (0-100 cm) n 151.94 ± 49.88 160.60 ± 17.43 162.77 ± 34.26 161.09 ± 37.38

Stratum 1 Stratum 2 Stratum 3 Stratum 4

C (0-5 cm) 91.56 ± 33.12 99.66 ± 19.19 111.78 ± 22.82 135.00 ± 33.52 C (5-10 cm) 59.42 ± 28.36 58.70 ± 7.08 62.36 ± 12.62 83.76 ± 23.64 C (10-20 cm) 56.66 ± 38.84 (ab) 49.44 ± 6.55 (a) 60.12 ± 7.49 (ab) 64.36 ± 8.69 (b)

C (20-30 cm) 34.94 ± 13.89 36.88 ± 7.68 44.68 ± 6.47 42.14 ± 12.57

C (30-50 cm) 22.44 ± 9.34 21.66 ± 4.53 30.76 ± 13.66 23.68 ± 14.20

C (50-100 cm) 15.62 ± 7.68 13.34 ± 3.95 12.84 ± 3.52 20.60 ± 15.14

Page 41: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

29

Figure 10. Aboveground carbon estimations per PSP in ton ha-1

for each plot-altitude. A linear regression was tested and showed a decreasing trend with an R² of 0.148 but was just not significant at p=0.05.

Figure 11. Carbon stocks of the sampled roots in the humus layer in ton ha-1

per plot. The linear regression had an R² of 0.444 and was highly significant.

Page 42: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

30

Figure 12. Belowground carbon estimates per PSP in ton ha-1

plotted against altitude. No significant linear trend could be found.

Figure 13. Belowground carbon in g kg-1

for each PSP plotted against altitude for three different depth increment layers: 0-5 cm (red triangles), 10-20 cm (green circles) and 50-100 cm (blue squares). A linear regression was visualised (with the R² and p-value) as p < 0.05.

Page 43: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

31

5.4 Litter, fine roots and the O-horizon

The following paragraph summarizes the findings of Tables 12, 13 and 14. These tables show the

litter layer, the roots and the O-horizon per stratum regarding the succeeding four variables:

weight (g), C/N ratio, δ15N (‰) and δ13C (‰).

Only the mass of the roots differs significantly between the strata and more than triples with

altitude. The C/N ratio differs significantly for both the litter layer and O-horizon. Both δ15N

and δ13C show significant differences between strata. While δ15N demonstrates a decrease, δ13C

shows an increase with the altitude. Linear regressions are shown in Figures 14 and 15.

Figure 14 shows a significant linear increase of C/N ratio with altitude for the foliage and litter

layer, with a respective R2 of 0.825 and 0.209. However, no significant linear trend is found for

the O-horizon and roots. The C and N concentrations with changing altitude are plotted in

Appendix D.

Table 12. Summary of the chosen variables that quantify properties of the litter layer per stratum (mean ± SD). Weight (g); C/N ratio; δ

15N in ‰; δ

13C ‰; KW gives the significance of the Kruskal-Wallis test: * = p<0.05, ** = p<0.01, *** =

p<0.001; the letters between brackets indicate the results of the Mann-Whitney U test

KW Stratum 1 Stratum 2 Stratum 3 Stratum 4

Weight (g) n 49.48 ± 8.62 50.46 ± 8.35 43.44 ± 8.60 42.51 ± 6.67

C/N (-) * 17.34 ± 1.88 (a) 19.78 ± 0.89 (ab) 21.98 ± 2.39 (b) 19.91 ± 0.95 (ab)

δ15N (‰) * 5.09 ± 0.96 (a) 5.02 ± 1.38 (a) 1.55 ± 0.58 (b) -0.16 ± 1.01 (b)

δ13C (‰) ** -30.61 ± 0.33 (a) -29.46 ± 0.62 (b) -29.54 ± 0.32 (b) -28.35 ± 0.50 (c)

Table 13. Summary of the chosen variables that quantify properties of the O-horizon per stratum (mean ± SD). Weight (g); C/N ratio; δ

15N in ‰; δ

13C ‰; KW gives the significance of the Kruskal-Wallis test: * = p<0.05, ** = p<0.01, *** =

p<0.001; the letters between brackets indicate the results of the Mann-Whitney U test.

KW Stratum 1 Stratum 2 Stratum 3 Stratum 4

Weight (g) n 452.63 ± 173.28 (a) 893.93 ± 171.62 (b) 817.95 ± 313.48 (ab) 749.55 ± 289.38 (ab) C/N (-) * 15.03 ± 0.88 (a) 15.52 ± 1.19 (ab) 17.24 ± 0.98 (b) 14.71 ± 0.63 (a) δ15N (‰) ** 5.36 ± 1.05 (a) 5.15 ± 1.39 (a) 1.36 ± 0.50 (b) 0.82 ± 1.09 (b) δ13C (‰) * -28.54 ± 0.17 (a) -27.71 ± 0.66 (bc) -27.96 ± 0.76 (ab) -26.91 ± 0.34 (c)

KW Stratum 1 Stratum 2 Stratum 3 Stratum 4

Weight (g) * 24.52 ± 7.69 (a) 33.96 ± 18.41 (ab) 60.76 ± 23.91 (b) 78.46 ± 40.60 (b) C/N (-) n 24.02 ± 3.91 22.23 ± 4.05 36.16 ± 7.99 25.12 ± 2.78 δ15N (‰) * 2.34 ± 1.05 (a) 1.95 ± 0.70 (a) -0.53 ± 0.89 (b) 0.13 ± 2.06 (ab) δ13C (‰) ** -29.41 ± 0.24 (a) -28.04 ± 0.93 (b) -27.95 ± 0.42 (b) -26.58 ± 0.23 (c)

Table 14. Summary of the chosen variables that quantify properties of the roots per stratum (mean ± SD). Weight (g); C/N ratio; δ

15N in ‰; δ

13C ‰; KW gives the significance of the Kruskal-Wallis test: * = p<0.05, ** = p<0.01, *** =

p<0.001; the letters between brackets indicate the results of the Mann-Whitney U test.

Page 44: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

32

Figure 14. C/N ratios of the foliar, litter, O-horizon and root layer along an altitudinal gradient. (a) foliar C/N; (b) litter C/N; (c) O-horizon C/N; (d) root C/N. Linear regressions were visualized and R

2

values were given when p<0.05. (data for foliar CN was provided by (Taveirne, 2016))

Figure 15. δ15

N and δ13

C of the litter, O-horizon and roots in ‰ for different altitudes. (a) litter δ15

N in ‰ (b) litter δ

13C in ‰ (c) O-horizon δ

15N in ‰ (d) O-horizon δ

13C in ‰ (e) root δ

15N in ‰ (f) root δ

13C

in ‰. Linear regressions were visualized and R2

values were given when p<0.05.

Page 45: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

33

5.5 Mineral soil

Table 15 shows a significant difference between stratum 3 and stratum 4 for total P

concentration. No further significances are given. Soil N also does not significantly differ

between the different strata. However, a decrease in N concentration with soil depth can be

observed (Table 16).

Table 17 and Table 18 display results on the isotope distribution at different soil depths. Table 17

shows significant differences in δ15N (‰) for different depths per stratum, with a high

significance (p<0.01) for 20-30 cm. Except for the 30-50 cm layer, a decrease with altitude can be

observed starting from stratum 2. There are only significant differences in δ13C (‰) for the 30-50

cm and 50-100cm layers. The Mann-Whitney U test did show significant differences between

some strata except for the 20-30 cm layer (Table 18). However, the 0-5 cm layer does show a

significant linear regression, with slightly increasing δ13C for rising altitudes (Figure 18).

Figure 16 visualizes the belowground N (g kg-1) in three different soil layers, similar to

belowground carbon in Figure 13. It shows a significant linear increase of N in the first layer as

the altitude rises, with an R2 of 0.166. The bioavailable P and total amount of P are both plotted

against altitude in Figure 17, but both do not demonstrate any trends. Finally, Figure 18 shows

the δ13C and δ15N in ‰ for a soil depth of 0-5 cm. A slight linear increase of δ13C with altitude

can be seen, accompanied by an R2 of 0.172. δ15N on the other hand, displays a linear decrease

with rising altitude and an R2 of 0.635. Appendix E visualizes the δ13C and δ15N in ‰ for all soil

depths separately.

Table 15. P and micronutrients (mg kg-1

) per stratum (mean ± SD). Presin=Bioavailable phophorus; Ptot=Total phosphorus; Mg=Magnesium; Ca=Calcium; Na=Sodium; Al=Aluminum; K=Potassium; KW gives the significance of the Kruskal-Wallis test: * = p<0.05, ** = p<0.01, *** = p<0.001; the letters between brackets indicate the results of the Mann-Whitney U test.

KW Stratum 1 Stratum 2 Stratum 3 Stratum 4

Presin n 13.05 ± 5.63 8.30 ± 2.69 8.21 ± 4.06 9.59 ± 1.51

Ptot * 674.97 ± 258.03 (a) 998.30 ± 456.25 (a) 900.87 ± 586.09 (a) 288.14 ± 153.59 (b) Mg n 15.18 ± 5.96 13.08 ± 3.98 11.56 ± 4.75 14.70 ± 2.76 Ca n 27.27 ± 19.07 23.59 ± 19.14 10.34 ± 4.37 23.60 ± 16.54 Na n 4.21 ± 1.27 (a) 5.64 ± 1.29 (ab) 7.59 ± 1.99 (b) 5.36 ± 0.81 (ab)

Al n 16.04 ± 6.77 17.84 ± 8.65 19.96 ± 10.49 8.56 ± 9.83 K n 60.26 ± 14.26 67.80 ± 16.95 57.65 ± 12.13 56.60 ± 6.58

Table 16. Soil N (g kg-1

) depth profile per stratum (mean ± SD)

Stratum 1 Stratum 2 Stratum 3 Stratum 4

N (0-5 cm) 7.04 ± 2.50 6.86 ± 1.26 7.84 ± 1.42 9.26 ± 1.71 N (5-10 cm) 4.20 ± 1.62 4.10 ± 0.62 4.44 ± 1.05 5.50 ± 1.74 N (10-20 cm) 3.88 ± 2.24 3.50 ± 0.52 4.20 ± 0.59 3.94 ± 0.59 N (20-30 cm) 2.42 ± 0.83 2.62 ± 0.62 3.08 ± 0.70 2.32 ± 0.73 N (30-50 cm) 1.46 ± 0.61 1.56 ± 0.41 2.04 ± 1.06 1.40 ± 0.42 N (50-100 cm) 0.96 ± 0.49 0.98 ± 0.31 0.78 ± 0.28 0.94 ± 0.80

Page 46: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

34

Figure 16. Belowground N in g kg-1

for each PSP plotted against altitude for three different depth layers: 0-5 cm (red triangles), 10-20 cm (green circles) and 50-100 cm (blue squares). A linear regression was visualised (with the R² and p-value) as p < 0.05.

Table 17. Summary of the δ15

N in ‰ in the soil at different depths per stratum (mean ± SD). KW gives the significance of the Kruskal-Wallis test: * = p<0.05, ** = p<0.01, *** = p<0.001; the letters between brackets indicate the results of the Mann-Whitney U test.

KW Stratum 1 Stratum 2 Stratum 3 Stratum 4

δ15N (0-5 cm) ** 6.37 ± 1.23 (a) 6.42 ± 1.10 (a) 4.17 ± 0.44 (b) 0.97 ± 1.02 (c) δ15N (5-10 cm) ** 6.40 ± 1.25 (a) 7.61 ± 1.10 (a) 4.60 ± 0.49 (b) 2.18 ± 0.58 (c)

δ15N (10-20 cm) ** 6.31 ± 1.84 (ab) 7.54 ± 1.11 (a) 4.90 ± 0.33 (b) 2.27 ± 0.63 (c) δ15N (20-30 cm) *** 6.18 ± 0.80 (a) 7.43 ± 1.14 (a) 4.79 ± 0.65 (b) 2.21 ± 0.91 (c)

δ15N (30-50 cm) ** 6.33 ± 0.87 (a) 6.55 ± 1.91 (a) 5.04 ± 0.73 (a) 2.71 ± 0.59 (b) δ15N (50-100 cm) ** 5.90 ± 1.77 (ab) 7.63 ± 0.77 (a) 5.04 ± 0.78 (b) 2.87 ± 0.52 (c)

Table 18. Summary of the δ13

C in ‰ in the soil at different depths per stratum (mean ± SD). KW gives the significance of the Kruskal-Wallis test: * = p<0.05, ** = p<0.01, *** = p<0.001; the letters between brackets indicate the results of the Mann-Whitney U test.

KW Stratum 1 Stratum 2 Stratum 3 Stratum 4

δ13C (0-5 cm) n -27.04 ± 0.41 (a) -26.44 ± 0.43 (ab) -26.33 ± 0.33 (b) -26.32 ± 0.87 (ab) δ13C (5-10 cm) n -26.94 ± 0.23 (a) -26.17 ± 0.48 (b) -25.95 ± 0.67 (b) -26.24 ± 0.71 (ab) δ13C (10-20 cm) n -26.93 ± 0.35 (a) -26.04 ± 0.47 (b) -26.05 ± 0.52 (b) -26.03 ± 0.58 (ab) δ13C (20-30 cm) n -26.61 ± 0.56 -25.74 ± 0.42 -25.93 ± 0.55 -26.15 ± 0.62 δ13C (30-50 cm) * -26.12 ± 0.61 (a) -25.14 ± 0.34 (b) -25.74 ± 0.42 (ab) -26.82 ± 1.37 (a) δ13C (50-100 cm) * -25.65 ± 1.03 (ab) -24.58 ± 0.37 (a) -25.22 ± 0.54 (ab) -26.37 ± 0.77 (b)

Page 47: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

35

Figure 17. Bioavailable P (a) and total amount of P (b) in mg kg-1

relative to altitude. Presin= bioavailable P; Ptot= total amount of P. Linear regressions and R

2 values were not given as p>0.05.

Figure 18. δ13

C (a) and δ15

N (b) of the upper soil layer (0-5 cm) in ‰ relative to altitude. Linear regressions were visualized and R

2 values were given as p<0.05.

Page 48: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

36

6 Discussion

6.1 The correction for planimetric plot areas and steep slopes

Since PSPs were selected according to canopy closure and on relatively accessible terrain, the

prevalence of steep slopes was far lower than 75%, which was reported for the global planimetric

area of TMF occurring on steep slopes by Spracklen & Righelato (2014). However 4 of the 19

plots were characterised by a steep slope with a maximum slope angle of 43.74° in plot 9. Steep

slopes have an impact on forest structure through altering the incidence of landslides, light

conditions and the access to space (Dislich and Huth, 2012; Robert, 2003). Despite the impact of

the slopes on forest structure and carbon stocks it was not accounted for in this study because

data is not at hand and the effects are still not well known. Furthermore, the impact is not very

often taken into account in comparable studies. AGB in forest plot studies is typically (and

correctly) reported per unit area of the Earth’s surface (Spracklen and Righelato, 2014), therefor

our results were not adjusted for planimetric areas. Remote sensed data, on the other hand, do

report planimetric area and need to be corrected for slope to produce reliable estimates of

regional biomass storage in montane forests.

6.2 Changes in forest structure with altitude

The dominant tree height as a measure for canopy height showed the best trend of the forest

structure variables, decreasing with altitude. The decreasing trend of canopy height along a

transect with increasing altitude is widely confirmed in literature (e.g. Asner et al. (2014), Clark,

Hurtado & Saatchi (2015), Lieberman, Peralta & Hartshorn (1996) and Lovett, Marshall & Carr

(2006)), either for the whole transect or from a certain altitudinal level onwards. Furthermore a

decreasing trend of the DBH and tree basal area are found. Different reasons for these

physiognomic forest changes can be found in literature but in line with section 6.6 the most

convincible driver for a decrease in the three aforementioned structural parameters is a declining

net primary productivity, as observed for instance by (Asner et al., 2014). Slenderness did not

differ because both the diameter and tree height declined. Although tree basal area declined, total

basal area per ha did not show any trends. This can be explained because the stem density

increased. Stem densities in the lower strata were in the same order of magnitude as the average

stem density of 479 trees per ha based on 12 1-ha plots in four other Albertine Rift forests, as

reported by Eilu et al. (2004). However stem densities in the higher strata greatly exceeded this.

How can this be explained? By following the hypothesis of Banin et al. (2012) who stated that

dominant large-statured families create conditions in which only tall species can compete, thus

perpetuating a forest dominated by tall individuals from diverse families. This is in line with our

results, which showed an increasing dominance at the higher strata, e.g. of Podocarpus latifolius,

which effectively was the tallest tree species in the plots of the highest stratum.

6.3 Changing carbon stocks along an altitudinal gradient

After an increase from stratum 1 to stratum 2, the AGB declined as often demonstrated in other

TMFs. However exceptions exist. Culmsee et al. (2010) for instance found a slight increase in

AGB at higher altitude. This was due to an increase in wood specific gravity. Again the decline in

AGB is attributed to a decrease in NPP. Again in line with the prevailing literature (Dieleman et

al., 2013), indications of more belowground soil organic carbon at higher altitudes were found in

the topsoil and due to the estimates of the roots in the humus layer. Although we used a very

Page 49: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

37

rough methodology (we just wanted an general idea), a significant trend was found in the amount

of carbon in the sampled roots. This can be explained by a decreasing temperature which leads to

reduced microbial nutrient mineralisation rates, reduced abundance and activity mycorrhizal fungi

or substantial reductions in nutrient uptake rates of the roots because carrier activity in the

plasma membranes in temperature sensitive (Leuschner et al., 2007b).

6.3.1 Underestimated AGC in the lower strata

The AGB in the first stratum was significantly lower than in the second stratum. This could be an

effective representation of the reality or this could be an underestimation. If really lower, this

could be due to soil fertility, steeper slopes (the steepest slope was in the lowest stratum) or

because the lower stratum was near a village, so more accessible to illegally cut wood. However it

is also possible that an underestimation was made because too few large diameter trees were

measured. Often about 30 to 40 % of the AGB can be found in large-diameter trees (DBH>70

cm) (Brown and Lugo, 1992; Brown et al., 1997). We found 25 trees with a DBH>70 and of

those 4 were measured. The assigned values were always much lower than the measured values,

up to a difference of 24 m. The mean total tree heights are plotted in Figure 19 and the mean

values ± SD are given in the caption with the p of the kruskal wallis test and the letters indicating

the Mann whitney u test. It is assumed that the mean values of the first strata are actually higher.

Figure 19. Mean total tree heights per plot 13.03 ± 0.58 (ab), 13.68 ± 0.42 (a), 12.15 ± 0.88 (b), 12.27 ± 0.97 (b) (KW: p=0.02.

6.4 Mineral soil nutrient concentrations along an altitudinal gradient

N is a measure for organic material and this demonstrated an increase in the first 5 cm of the soil,

consistent with soil C. Total P is a measure for weathering and no significant trend was found.

Neither did the cations show any trends and neither did bio-available P show a trend, indicating

that P was not limiting.

Page 50: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

38

6.5 How do the isotopic compositions change with altitude and what do they tell us?

The most significant altitudinal trends to be found in this thesis resulted from the isotopic

analysis. δ15N decreased significantly with altitude in the litter layer, the O-horizon, the roots and

in all depth increment layers of the soil. In addition this trend was also detected in the canopy

foliage by Taveirne (2016). Different interpretations to explain variations in δ15N of plants and

soil exist and have been reviewed by Craine et al. (2015b). Since it is a single response variable

with multiple drivers no fully conclusive answer can be given. Yet several results point in the

same direction. A possible explanation for the observed trend could be that the higher values of

δ15N are due to a more ‘open’ nitrogen cycle at higher altitudes, with higher N availability and

greater losses via fractioning pathways, while the lower values indicate a more closed N limiting

cycle at higher altitudes. This is probably indirectly caused by the decreasing temperature.

Increasing N limitation from lowland to montane tropical forest has been stated by many (e.g.

Tanner et al. (1998); Townsend et al. (2008)). The hypothesis concerning ‘open or closed N

cycles’ was also confirmed by Martinelli et al. (1999) who compared δ15N between temperate and

tropical forests. Craine et al. (2015a) however, state that soils in warmer ecosystems possibly have

soil organic matter with higher δ15N values as a result of greater decomposition of organic matter

and/or greater concentrations of clay. The latter is then responsible for a greater share of their N

in mineral-associated organic matter instead of having a greater proportion of N being lost to

fractionating pathways. Next to this protection of mineral association, δ15N in soil organic matter

is possibly enriched by microbial processing. The higher N availability in a lot of low-altitude

ecosystems may also reduce the share of N lost in gaseous forms relative to leaching (Perakis et

al., 2011). Nevertheless in favour of the first hypothesis Rütting et al. (2015) recently found very

high N losses at the altitude of the lowest stratum in the same study site, Nyungwe Forest.

Furthermore the counter-arguments may be valuable in discussing the trend in δ15N in the soil

and O-horizon but a decreasing trend was also found in foliar δ15N which has been shown clearly

to reflect N availability (e.g. Choi et al. (2005); Garten and Van Miegroet, (1994)). At last, there

are still other results indicating a shift to a more closed N cycle and N limitation at higher

altitudes:

the increased C:N-ratio in leaves and litter (more C per unit N is assimulated), and

a shift towards more nutrient conservative forest communities with increasing altitude as

shown by Taveirne (2016).

δ13C shows a clear highly significant increasing trend in the litter, the O-horizon and the roots but

a weak trend in the first increment layer of the soil (0-5 cm) and no trends in deeper layers. A

general increase of δ13C with increasing altitude has been found in plants all over the globe

(Körner et al., 1988) and is probably caused by changing partial pressures of O2 and CO2 and

changing temperatures (Körner et al., 1991). The long term effect of the environment, i.e. the

altitudinal gradient, led to an adapted leaf size and structure which altered the mesophyll

conductance and explains the observed trend (Seibt et al., 2008). However δ13C in plants is also

effected by short-term environmental conditions dependent on the evaporative demand but this

was not relevant for the present study.

Page 51: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

39

7 General conclusion and future perspectives

All results pointed in one direction: with increasing altitude N limitation prevails and the N cycle

evolves from an open to a more closed cycle. With the limitation of N, productivity decreases,

resulting among others in a decrease of AGB. This is important in the light of future global

change. How will TMFs respond on a rising temperature, changing weather patterns and an

increased reactive N deposition? I recommend to recensus the PSPs in Nyungwe in 10 years to

evaluate the evolutions.

Page 52: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

40

8 List of references

Aldrich, M., Billington, C., Edwards, M. and Laidlaw, R.: Tropical montane cloud forests: an urgent priority for conservation, WCMC Biodivers. Bull., (2), 1–17, 1997.

Alves, L. F., Vieira, S. A., Scaranello, M. A., Camargo, P. B., Santos, F. A. M., Joly, C. A. and Martinelli, L. A.: Forest structure and live aboveground biomass variation along an elevational gradient of tropical Atlantic moist forest (Brazil), For. Ecol. Manage., 260(5), 679–691, 2010.

Anderson-Teixeira, K. J., Wang, M. M. H., McGarvey, J. C. and LeBauer, D. S.: Carbon dynamics of mature and regrowth tropical forests derived from a pantropical database (TropForC-db)., Glob. Chang. Biol., 2016.

Ashton, L. A., Odell, E. H., Burwell, C. J., Maunsell, S. C., Nakamura, A., Mcdonald, W. J. F. and Kitching, R. L.: Altitudinal patterns of moth diversity in tropical and subtropical Australian rainforests, Austral Ecol., 41(2), 197–208, 2016.

Ashton, P. S.: Floristic zonation of tree communities on wet tropical mountains revisited, Perspect. Plant Ecol. Evol. Syst., 6(1–2), 87–104, 2003.

Asner, G. P., Anderson, C. B., Martin, R. E., Knapp, D. E., Tupayachi, R., Sinca, F. and Malhi, Y.: Landscape-scale changes in forest structure and functional traits along an Andes-to-Amazon elevation gradient, Biogeosciences, 11(3), 843–856, 2014.

Attiwill, P. M. and Adams, M. a: Nutrient cycling in forests, New Phytol., 124(4), 561–582, 1993.

Bach, K. and Gradstein, S. R.: A comparison of six methods to detect altitudinal belts of vegetation in tropical mountains, Ecotropica, 17(1), 1–13, 2011.

Banin, L., Feldpausch, T. R., Phillips, O. L., Baker, T. R., Lloyd, J., Affum-Baffoe, K., Arets, E. J. M. M., Berry, N. J., Bradford, M., Brienen, R. J. W., Davies, S., Drescher, M., Higuchi, N., Hilbert, D. W., Hladik, A., Iida, Y., Salim, K. A., Kassim, A. R., King, D. A., Lopez-Gonzalez, G., Metcalfe, D., Nilus, R., Peh, K. S. H., Reitsma, J. M., Sonké, B., Taedoumg, H., Tan, S., White, L., Wöll, H. and Lewis, S. L.: What controls tropical forest architecture? Testing environmental, structural and floristic drivers, Glob. Ecol. Biogeogr., 21(12), 1179–1190, doi:10.1111/j.1466-8238.2012.00778.x, 2012.

Batjes, N. H.: Mapping soil carbon stocks of Central Africa using SOTER, Geoderma, 146(1-2), 58–65, 2008.

Battipaglia, G., Zalloni, E., Castaldi, S., Marzaioli, F., Cazzolla-Gatti, R., Lasserre, B., Tognetti, R., Marchetti, M. and Valentini, R.: Long tree-ring chronologies provide evidence of recent tree growth decrease in a central african tropical forest, PLoS One, 10(3), 1–21, 2015.

Boisvenue, C. and Running, S. W.: Impacts of climate change on natural forest productivity - Evidence since the middle of the 20th century, Glob. Chang. Biol., 12(5), 862–882, 2006.

Braun-Blanquet, J.: Pflanzensoziologie, Springer, Wien., 1964.

Brown, S. and Lugo, A. E.: Aboveground biomass estimates for tropical moist forests of Brazilian Amazon., Interciencia (Venezuela), 17(1), 8–18, 1992.

Page 53: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

41

Brown, S., Gillespie, A. J. R. . and Lugo, A. E.: Biomass Estimation Methods for Tropical Forests with Applications to Forest Inventory Data, For. Sci., 35(4), 881–902, 1989.

Brown, S., Schroeder, P. and Birdsey, R.: Aboveground biomass distribution of US eastern hardwood forests and the use of large trees as an indicator of forest development, For. Ecol. Manage., 96(1-2), 37–47, doi:10.1016/S0378-1127(97)00044-3, 1997.

Bruijnzeel, L.: Hydrology of tropical montane cloud forests: a reassessment, L. use water Resour. Res., 1, 1–18, 2001.

Bruijnzeel, L. A. and Hamilton, L. S.: Decision time for cloud forests, IHP humid Trop. Program., (13), 2000.

Bruijnzeel, L. A. and Veneklaas, E. J.: Climatic Conditions and Tropical Montane Forest Productivity : The Fog Has Not Lifted Yet, Ecology, 79(1), 3–9, 1998.

Bruijnzeel, L. A., Kappelle, M., Mulligan, M. and Scatena, F. N.: Tropical Montane Cloud Forests: state of knowledge and sustainability perspectives in a changing world, in Tropical Montane Cloud Forest: Science for Conservation and Management, pp. 691–740, Cambridge University Press., 2011.

Bubb, P., May, I., Miles, L. and Sayer, J.: Cloud forest agenda., 2004.

Bustamante, M. M. C., Roitman, I., Aide, T. M., Alencar, A., Anderson, L. O., Aragão, L., Asner, G. P., Barlow, J., Berenguer, E., Chambers, J., Costa, M. H., Fanin, T., Ferreira, L. G., Ferreira, J., Keller, M., Magnusson, W. E., Morales-Barquero, L., Morton, D., Ometto, J. P. H. B., Palace, M., Peres, C. A., Silvério, D., Trumbore, S. and Vieira, I. C. G.: Toward an integrated monitoring framework to assess the effects of tropical forest degradation and recovery on carbon stocks and biodiversity., Glob. Chang. Biol., 22(1), 92–109, 2016.

Callicott, J. B.: Environmental Ethics: I. Overview, in Encyclopedia of Bioethics, vol. 2, edited by S. Post, pp. 1–10, Macmillan Publishers, New York., 2004.

Chave, J., Muller-landau, H. C., Baker, T. R., Easdale, T. A. and Webb, C. O.: Regional and Phylogenetic Variation of Wood Density across 2456 Neotropical Tree Species Hans ter Steege and Campbell O . Webb Published by : Wiley Stable URL : http://www.jstor.org/stable/40061964 Accessed : 19-04-2016 18 : 42 UTC Your use of the JSTOR a, Ecol. Appl., 16(6), 2356–2367, 2006.

Chave, J., Réjou-Méchain, M., Búrquez, A., Chidumayo, E., Colgan, M. S., Delitti, W. B. C., Duque, A., Eid, T., Fearnside, P. M., Goodman, R. C., Henry, M., Martínez-Yrízar, A., Mugasha, W. A., Muller-Landau, H. C., Mencuccini, M., Nelson, B. W., Ngomanda, A., Nogueira, E. M., Ortiz-Malavassi, E., Pélissier, R., Ploton, P., Ryan, C. M., Saldarriaga, J. G. and Vieilledent, G.: Improved allometric models to estimate the aboveground biomass of tropical trees, Glob. Chang. Biol., 20(10), 3177–3190, 2014.

Choi, W., Chang, S. and Allen, H.: Irrigation and fertilization effects on foliar and soil carbon and nitrogen isotope ratios in a loblolly pine stand, For. Ecol. …, doi:10.1016/j.foreco.2005.03.0I6, 2005.

Ciais, P., Piao, S.-L., Cadule, P., Friedlingstein, P. and Chédin, A.: Variability and recent trends in the African terrestrial carbon balance, Biogeosciences, 6(9), 1935–1948, 2009.

Page 54: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

42

Cizungu, L. N.: Dynamics and loss of reactive nitrogen in a central African tropical mountain forest and Eucalyptus plantation., 2014.

Clark, D. B., Hurtado, J. and Saatchi, S. S.: Tropical Rain Forest Structure, Tree Growth and Dynamics along a 2700-m Elevational Transect in Costa Rica., PLoS One, 10(4), 1–18, 2015.

Costanza, R., D’Arge, R., de Groot, R., Farber, S., Grasso, M., Hannon, B., Limburg, K., Naeem, S., O’Neill, R. V., Paruelo, J., Raskin, R. G., Sutton, P. and van den Belt, M.: The value of the world’s ecosystem services and natural capital, Nature, 387(6630), 253–260, 1997.

Cox, P. M., Pearson, D., Booth, B. B., Friedlingstein, P., Huntingford, C., Jones, C. D. and Luke, C. M.: Sensitivity of tropical carbon to climate change constrained by carbon dioxide variability., Nature, 494(7437), 341–4, 2013.

Crafford, J., Strohmaier, R., Muñoz, P., Oliveira, T. De, Lambrechts, C., Wilkinson, M., Burger, A. and Bosch, J.: The Role and Contribution of Montane Forests and Related Ecosystem Services to the Kenyan Economy, UNEP., 2012.

Craine, J. M., Elmore, A. J., Wang, L., Augusto, L., Baisden, W. T., Brookshire, E. N. J., Cramer, M. D., Hasselquist, N. J., Hobbie, E. A., Kahmen, A., Koba, K., Kranabetter, J. M., Mack, M. C., Marin-Spiotta, E., Mayor, J. R., McLauchlan, K. K., Michelsen, A., Nardoto, G. B., Oliveira, R. S., Perakis, S. S., Peri, P. L., Quesada, C. A., Richter, A., Schipper, L. A., Stevenson, B. A., Turner, B. L., Viani, R. A. G., Wanek, W. and Zeller, B.: Convergence of soil nitrogen isotopes across global climate gradients., Sci. Rep., 5, 8280, doi:10.1038/srep08280, 2015a.

Craine, J. M., Brookshire, E. N. J., Cramer, M. D., Hasselquist, N. J., Koba, K., Marin-Spiotta, E. and Wang, L.: Ecological interpretations of nitrogen isotope ratios of terrestrial plants and soils, Plant Soil, 396(1-2), 1–26, doi:10.1007/s11104-015-2542-1, 2015b.

Cramer, W., Bondeau, A., Schaphoff, S., Lucht, W., Smith, B. and Sitch, S.: Tropical forests and the global carbon cycle: impacts of atmospheric carbon dioxide, climate change and rate of deforestation, Philos. Trans. R. Soc. London B Biol. Sci., 359(1443), 331–343, 2004.

Culmsee, H., Leuschner, C., Moser, G. and Pitopang, R.: Forest aboveground biomass along an elevational transect in Sulawesi, Indonesia, and the role of Fagaceae in tropical montane rain forests, J. Biogeogr., 37(5), 960–974, 2010.

Dickson, B., Blaney, R., Miles, L., Regan, E., Soesbergen, A. Van, Väänänen, E., Blyth, S., Harfoot, M., Martin, C., Mcowen, C. and Newbold, T.: Towards a global map of natural capital : key ecosystem assets, UNEP., 2014.

Dieleman, W. I. J., Venter, M., Ramachandra, A., Krockenberger, A. K. and Bird, M. I.: Soil carbon stocks vary predictably with altitude in tropical forests: Implications for soil carbon storage, Geoderma, 204-205, 59–67, 2013.

Dislich, C. and Huth, A.: Modelling the impact of shallow landslides on forest structure in tropical montane forests, Ecol. Modell., 239, 40–53, 2012.

Don, A., Schumacher, J. and Freibauer, A.: Impact of tropical land-use change on soil organic carbon stocks - a meta-analysis, Glob. Chang. Biol., 17(4), 1658–1670, 2011.

Ehleringer, J. R., Buchmann, N. and Flanagan, L. B.: Carbon Isotope Ratios in Belowground Carbon Cycle Processes, Ecol. Appl., 10(2), 412–422, 2000.

Page 55: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

43

Eilu, G., Hafashimana, D. L. N. and Kasenene, J. M.: Density and species diversity of trees in four tropical forest of the Albertine Rift , Western Uganda four tropical forests of the Albertine rift , western Uganda, Divers. Distrib., 10(February 2016), 303–312, 2004.

Falkowski, P. G., Scholes, R. J., Boyle, E., Canadell, J., Canfield, D., Elser, J., Gruber, N., Hibbard, K., Högberg, P., Linder, S., Mackenzie, F. T., Moore, B., Pedersen, T., Rosenthal, Y., Seitzinger, S., Smetacek, V. and Steffen, W.: The global carbon cycle: a test of our knowledge of earth as a system., Science (80-. )., 290(5490), 291–296, 2000.

FAO: World reference base for soil resources 2014. International soil classification system for naming soils and creating legends for soil maps., 2014.

Feldpausch, T. R., Lloyd, J., Lewis, S. L., Brienen, R. J. W., Gloor, M., Monteagudo Mendoza, A., Lopez-Gonzalez, G., Banin, L., Abu Salim, K., Affum-Baffoe, K., Alexiades, M., Almeida, S., Amaral, I., Andrade, A., Aragão, L. E. O. C., Araujo Murakami, A., Arets, E. J. M., Arroyo, L., Aymard C., G. A., Baker, T. R., Bánki, O. S., Berry, N. J., Cardozo, N., Chave, J., Comiskey, J. A., Alvarez, E., De Oliveira, A., Di Fiore, A., Djagbletey, G., Domingues, T. F., Erwin, T. L., Fearnside, P. M., França, M. B., Freitas, M. A., Higuchi, N., Honorio C., E., Iida, Y., Jiménez, E., Kassim, A. R., Killeen, T. J., Laurance, W. F., Lovett, J. C., Malhi, Y., Marimon, B. S., Marimon-Junior, B. H., Lenza, E., Marshall, A. R., Mendoza, C., Metcalfe, D. J., Mitchard, E. T. A., Neill, D. A., Nelson, B. W., Nilus, R., Nogueira, E. M., Parada, A., S.-H. Peh, K., Pena Cruz, A., Peñuela, M. C., Pitman, N. C. A., Prieto, A., Quesada, C. A., Ramírez, F., Ramírez-Angulo, H., Reitsma, J. M., Rudas, A., Saiz, G., Salomão, R. P., Schwarz, M., Silva, N., Silva-Espejo, J. E., Silveira, M., Sonké, B., Stropp, J., Taedoumg, H. E., Tan, S., Ter Steege, H., Terborgh, J., Torello-Raventos, M., Van Der Heijden, G. M. F., Vásquez, R., Vilanova, E., Vos, V. A., White, L., Willcock, S., Woell, H. and Phillips, O. L.: Tree height integrated into pantropical forest biomass estimates, Biogeosciences, 9(8), 3381–3403, doi:10.5194/bg-9-3381-2012, 2012.

Fernández-Martínez, M., Vicca, S., Janssens, I. a., Sardans, J., Luyssaert, S., Campioli, M., Chapin III, F. S., Ciais, P., Malhi, Y., Obersteiner, M., Papale, D., Piao, S. L., Reichstein, M., Rodà, F. and Peñuelas, J.: Nutrient availability as the key regulator of global forest carbon balance, Nat. Clim. Chang., 4(June), 471–476, doi:doi:10.1038/nclimate2177, 2014.

Fisher, J. B., Malhi, Y., Torres, I. C., Metcalfe, D. B., van de Weg, M. J., Meir, P., Silva-Espejo, J. E. and Huasco, W. H.: Nutrient limitation in rainforests and cloud forests along a 3,000-m elevation gradient in the Peruvian Andes, Oecologia, 172(3), 889–902, 2013.

Flenley, J. R.: Cloud Forest, the Massenerhebung Effect, and Ultraviolet Insolation, in The Puerto Rico tropical cloud forest symposium, edited by L. S. Hamilton, J. O. Juvik, and F. N. Scatena, pp. 150–155, Springer-Verlag., 1995.

Gardner, T. A., Barlow, J., Sodhi, N. S. and Peres, C. A.: A multi-region assessment of tropical forest biodiversity in a human-modified world, Biol. Conserv., 143(10), 2293–2300, 2010.

Garten, C. T. and Van Miegroet, H.: Relationships between soil nitrogen dynamics and natural 15N abundance in plant foliage from Great Smoky Mountains National Park, Can. J. For. Res., 1636–1645, doi:10.1017/CBO9781107415324.004, 1994.

Geist, H. J. and Lambin, E. F.: Proximate Causes and Underlying Driving Forces of Tropical Deforestation, Bioscience, 52(2), 143, 2002.

Gharahi Ghehi, N., Werner, C., Cizungu Ntaboba, L., Mbonigaba Muhinda, J. J., Van Ranst, E.,

Page 56: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

44

Butterbach-Bahl, K., Kiese, R. and Boeckx, P.: Spatial variations of nitrogen trace gas emissions from tropical mountain forests in Nyungwe, Rwanda, Biogeosciences, 9(4), 1451–1463, 2012.

Gibson, L., Lee, T. M., Koh, L. P., Brook, B. W., Gardner, T. a., Barlow, J., Peres, C. a., Bradshaw, C. J. a., Laurance, W. F., Lovejoy, T. E. and Sodhi, N. S.: Primary forests are irreplaceable for sustaining tropical biodiversity, Nature, 478(7369), 378–381, 2011.

González, G., Willig, M. R. and Waide, R. B.: Ecological Gradient Analyses in a Tropical Landscape, edited by G. González, M. R. Willig, and R. B. Waide, Wiley-Blackwell., 2013.

Grace, J., Mitchard, E. and Gloor, E.: Perturbations in the carbon budget of the tropics, Glob. Chang. Biol., 20(10), 3238–3255, 2014.

Gregory, P. J. and Nortcliff, S., Eds.: Soil Conditions and Plant Growth, Blackwell Publishing Ltd., 2013.

Grubb, P. J., Lloyd, J. R., Pennington, T. D. and Whitmore, T. C.: A Comparison of Montane and Lowland Rain Forest in Ecuador I. The Forest Structure, Physiognomy, and Floristics, J. Ecol., 51(3), 567–601, 1963.

Grytnes, J. A. and Beaman, J. H.: Elevational species richness patterns for vascular plants on Mount Kinabalu, Borneo, J. Biogeogr., 33(10), 1838–1849, 2006.

Hamilton, A. C.: A quantitative analysis of altitudinal zonation in Uganda forests, Vegetatio, 30(2), 99–106, 1975.

Hamilton, L. S.: Mountain Cloud Forest Conservation and Research : A Synopsis, Mt. Res. Dev., 15(3), 259–266, 1995.

Hemp, A.: Continuum or zonation? Altitudinal gradients in the forest vegetation of Mt. Kilimanjaro, Plant Ecol., 184(1), 27–42, 2006.

Herbert, D. a. and Fownes, J. H.: Phosphorus limitation of forest leaf area and net primary production on a highly weathered soil, Biogeochemistry, 29(3), 223–235, doi:10.1007/BF02186049, 1995.

Hodkinson, I. D.: Terrestrial insects along elevation gradients: species and community responses to altitude, Biol. Rev., 80, 489–513, 2005.

Hopkins, W. G. and Hüner, N. P. A.: Introduction to Plant Physiology, edited by 3rd, Wiley, J., 2004.

Hosonuma, N., Herold, M., De Sy, V., De Fries, R. S., Brockhaus, M., Verchot, L., Angelsen, A. and Romijn, E.: An assessment of deforestation and forest degradation drivers in developing countries, Environ. Res. Lett., 7(4), 044009, 2012.

Huang, S., Titus, S. J. and Wiens, D. P.: Comparison of nonlinear height-diameter functions for major Alberta tree species, Can. J. For. Res., (22), 1297–1304, 1992.

IPCC: IPCC meeting on current scientific understanding of the processes affecting terrestrial carbon stocks and human influences upon them., 2003.

Jobbágy, E. G. and Jackson, R. B.: the Vertical Distribution of Soil Organic Carbon and Its, Ecol. Appl., 10(2), 423–436, doi:10.1890/1051-0761(2000)010[0423:TVDOSO]2.0.CO;2, 2000.

Page 57: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

45

Jolly, W. M., Nemani, R. and Running, S. W.: A generalized, bioclimatic index to predict foliar phenology in response to climate, Glob. Chang. Biol., 11(4), 619–632, doi:10.1111/j.1365-2486.2005.00930.x, 2005.

Jones, A., Breuning-Madsen, H. Brossard, M., Dampha, A., Deckers, J., Dewitte, O., Gallali, T., Hallett, S., Jones, R., Kilasara, M., Le Roux, P., Micheli, E., Montanarella, L., Spaargaren, O., Thiombiano, L., Van Ranst, E., Yemefack, M., Zougmoré, R. and (eds.): Soil Atlas of Africa, edited by European Commission, Publications Office of the European Union, Luxembourg., 2013.

Joseph Wright, S.: The carbon sink in intact tropical forests, Glob. Chang. Biol., 19(2), 337–339, 2013.

Kapos, V., Rhind, J., Edwards, M. and Price, M. F.: Developing a map of the world’s mountain forests., 2000.

Kaspari, M., Garcia, M. N., Harms, K. E., Santana, M., Wright, S. J. and Yavitt, J. B.: Multiple nutrients limit litterfall and decomposition in a tropical forest, Ecol. Lett., 11(1), 35–43, doi:10.1111/j.1461-0248.2007.01124.x, 2008.

Keenan, R. J., Reams, G. A., Achard, F., de Freitas, J. V., Grainger, A. and Lindquist, E.: Dynamics of global forest area: Results from the FAO Global Forest Resources Assessment 2015, For. Ecol. Manage., 352, 9–20, 2015.

Kitayama, K. and Aiba, S. I.: Ecosystem structure and productivity of tropical rain forests along altitudinal gradients with contrasting soil phosphorus pools on Mount Kinabalu,Borneo, J. Ecol., 90(1), 37–51, 2002.

Körner, C.: The use of “altitude” in ecological research, Trends Ecol. Evol., 22(11), 569–574, 2007.

Körner, C., Farquhar, G. D. and Roksandic, Z.: A Global Survey of Carbon Isotope Discrimination in Plants from High Altitude, , 74(4), 623–632, 1988.

Körner, C., Farquhar, G. D. and Wong, S. C.: Carbon isotope discrimination by plants follows latitudinal and altitudinal trends, Oecologia, 88(1), 30–40, doi:10.1007/BF00328400, 1991.

Körner, C., Ohsawa, M., Spehn, E., Berge, E., Bugmann, H., Groombridge, B., Hamilton, L., Hofer, T., Ives, J., Jodha, N., Messerli, B., Pratt, J., Price, M., Reasoner, M., Rodgers, A., Thonell, J. and Yoshino, M.: Mountain systems, Ecosyst. Hum. well-being Curr. State …, 683–716, 2005.

Krashevska, V., Bonkowski, M., Maraun, M., Ruess, L., Kandeler, E. and Scheu, S.: Microorganisms as driving factors for the community structure of testate amoebae along an altitudinal transect in tropical mountain rain forests, Soil Biol. Biochem., 40(9), 2427–2433, 2008.

Latham, R. E. and Ricklefs, R. E.: Continental Comparisons of Temperate-Zone Tree Species Diversity, in Species Diversity in Ecological Communities: Historical and Geographical Perspectives, pp. 294–314, University of Chicago Press., 1993.

Leuschner, C., Moser, G., Bertsch, C., Röderstein, M. and Hertel, D.: Large altitudinal increase in tree root/shoot ratio in tropical mountain forests of Ecuador, Basic Appl. Ecol., 8(3), 219–

Page 58: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

46

230, 2007a.

Leuschner, C., Moser, G., Bertsch, C., R??derstein, M. and Hertel, D.: Large altitudinal increase in tree root/shoot ratio in tropical mountain forests of Ecuador, Basic Appl. Ecol., 8(3), 219–230, doi:10.1016/j.baae.2006.02.004, 2007b.

Leuschner, C., Zach, A., Moser, G., Homeier, J., Graefe, S., Hertel, D., Wittich, B., Soethe, N., Iost, S., Röderstein, M., Horna, V. and Wolf, K.: The Carbon Balance of Tropical Mountain Forests Along an Altitudinal Transect, in Ecosystem Services, Biodiversity and Environmental Change in a Tropical Mountain Ecosystem of South Ecuador, Ecological Studies, vol. 221, pp. 117–139., 2013.

Lewis, S. L., Lopez-Gonzalez, G., Sonké, B., Affum-Baffoe, K., Baker, T. R., Ojo, L. O., Phillips, O. L., Reitsma, J. M., White, L., Comiskey, J. A., K, M.-N. D., Ewango, C. E. N., Feldpausch, T. R., Hamilton, A. C., Gloor, M., Hart, T., Hladik, A., Lloyd, J., Lovett, J. C., Makana, J.-R., Malhi, Y., Mbago, F. M., Ndangalasi, H. J., Peacock, J., Peh, K. S.-H., Sheil, D., Sunderland, T., Swaine, M. D., Taplin, J., Taylor, D., Thomas, S. C., Votere, R. and Wöll, H.: Increasing carbon storage in intact African tropical forests, Nature, 457(7232), 1003–1006, 2009.

Lewis, S. L., Sonké, B., Sunderland, T., Begne, S. K., Lopez-Gonzalez, G., van der Heijden, G. M. F., Phillips, O. L., Affum-Baffoe, K., Baker, T. R., Banin, L., Bastin, J.-F., Beeckman, H., Boeckx, P., Bogaert, J., De Cannière, C., Chezeaux, E., Clark, C. J., Collins, M., Djagbletey, G., Djuikouo, M. N. K., Droissart, V., Doucet, J.-L., Ewango, C. E. N., Fauset, S., Feldpausch, T. R., Foli, E. G., Gillet, J.-F., Hamilton, A. C., Harris, D. J., Hart, T. B., de Haulleville, T., Hladik, A., Hufkens, K., Huygens, D., Jeanmart, P., Jeffery, K. J., Kearsley, E., Leal, M. E., Lloyd, J., Lovett, J. C., Makana, J.-R., Malhi, Y., Marshall, A. R., Ojo, L., Peh, K. S.-H., Pickavance, G., Poulsen, J. R., Reitsma, J. M., Sheil, D., Simo, M., Steppe, K., Taedoumg, H. E., Talbot, J., Taplin, J. R. D., Taylor, D., Thomas, S. C., Toirambe, B., Verbeeck, H., Vleminckx, J., White, L. J. T., Willcock, S., Woell, H. and Zemagho, L.: Above-ground biomass and structure of 260 African tropical forests., Philos. Trans. R. Soc. Lond. B. Biol. Sci., 368, 1–12, 2013.

Lieberman, D., Lieberman, M., Peralta, R. and Hartshorn, G.: Tropical forest structure and composition on a large scale altitudinal gradient in Costa Rica, J. Ecol., 84(2), 137–152, 1996.

Lovett, J. C.: Continuous Change in Tanzanian Moist Forest Tree Communities with Elevation, J. Trop. Ecol., 14(5), 719–722, 1998.

Lovett, J. C., Marshall, A. R. and Carr, J.: Changes in tropical forest vegetation along an altitudinal gradient in the Udzungwa Mountains National Park, Tanzania, Afr. J. Ecol., 44, 478–490, 2006.

Luyssaert, S., Schulze, E.-D. D., Börner, A., Knohl, A., Hessenmöller, D., Law, B. E., Ciais, P., Grace, J., Boerner, A., Hessenmoeller, D., Borner, A. and Hessenmoller, D.: Old-growth forests as global carbon sinks, Nature, 455(7210), 213–215, 2008.

Malhi, Y., Phillips, O. L., Lloyd, J., Baker, T., Wright, J., Almeida, S., Arroyo, L., Frederiksen, T., Grace, J., Higuchi, N., Killeen, T., Laurance, W. F., Leano, C., Lewis, S., Meir, P., Monteagudo, A., Neill, D., Nunez Vargas, P., Panfil, S. N., Patino, S., Pitman, N., Quesada, C. a., Rudas-Ll, A., Salomao, R., Saleska, S., Silva, N., Silveira, M., Sombroek, W. G., Valencia, R., Vasquez Martinez, R., Vieira, I. C. G. and Vinceti, B.: An international network to monitor the structure, composition and dynamics of Amazonian forests (RAINFOR), J. Veg. Sci., 13, 439–

Page 59: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

47

450, 2002.

Malhi, Y., Roberts, J. T., Betts, R. a, Killeen, T. J., Li, W. and Nobre, C. a: Climate Change, Deforastation, and the Fate of the Amazon, Science (80-. )., (January), 169–172, 2008.

Malhi, Y., Silman, M., Salinas, N., Bush, M., Meir, P. and Saatchi, S.: Introduction: Elevation gradients in the tropics: Laboratories for ecosystem ecology and global change research, Glob. Chang. Biol., 16(12), 3171–3175, 2010.

Martin, P. H., Fahey, T. J. and Sherman, R. E.: Vegetation zonation in a neotropical montane forest: Environment, disturbance and ecotones, Biotropica, 43(5), 533–543, 2011.

Martinelli, L. a., Piccolo, M. C., Townsend, a. R., Vitousek, P. M., Cuevas, E., McDowell, W., Robertson, G. P., Santos, O. C. and Treseder, K.: Nitrogen stable isotopic composition of leaves and soil: Tropical versus temperate forests, Biogeochemistry, 46(1-3), 45–65, doi:10.1007/BF01007573, 1999.

Millennium Ecosystem Assessment: Ecosystems and human well-being: synthesis., 2005.

Mittermeier, R. a, Mittermeier, C. G., Brooks, T. M., Pilgrim, J. D., Konstant, W. R., da Fonseca, G. a B. and Kormos, C.: Wilderness and biodiversity conservation., Proc. Natl. Acad. Sci. U. S. A., 100(18), 10309–13, 2003.

Moser, G., Hertel, D. and Leuschner, C.: Altitudinal change in LAI and stand leaf biomass in tropical montane forests: A transect study in ecuador and a pan-tropical meta-analysis, Ecosystems, 10(6), 924–935, 2007.

Myers, N., Mittermeier, R. a, Mittermeier, C. G., da Fonseca, G. a and Kent, J.: Biodiversity hotspots for conservation priorities., Nature, 403(6772), 853–8, 2000.

Myers, R. J. K.: Temperature effects on ammonification and nitrification in a tropical soil, Soil Biol. Biochem., 7(2), 83–86, doi:10.1016/0038-0717(75)90003-6, 1975.

Pachauri, R. K. and Meyer, L. A.: IPCC, 2014: Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change., 2014.

Pansu, M. and Gautheyrou, J.: Handbook of Soil Analysis. Mineralogical, Organic and Inorganic Methods, Springer., 2006.

Perakis, S. S., Sinkhorn, E. R. and Compton, J. E.: δ15N constraints on long-term nitrogen balances in temperate forests, Oecologia, 167(3), 793–807, doi:10.1007/s00442-011-2016-y, 2011.

Phillips, O., Baker, T., Feldpausch, T. and Brienen, R.: Field manual for plot establishment and remeasurement - RAINFOR., 2016.

Plumptre, A. J., Michel, M., Fashing, P. J., McNeilage, A., Ewango, C., Kaplin, B. A. and Liengola, I.: Biodiversity Surveys of the Nyungwe Forest Reserve In S.W. Rwanda., 2002.

Plumptre, A. J., Davenport, T. R. B., Behangana, M., Kityo, R., Eilu, G., Ssegawa, P., Ewango, C., Meirte, D., Kahindo, C., Herremans, M., Peterhans, J. K., Pilgrim, J. D., Wilson, M., Languy, M. and Moyer, D.: The biodiversity of the Albertine Rift, Biol. Conserv., 134(2), 178–194, 2007.

Page 60: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

48

R Development Core Team: A language and environment for statistical computing. R Foundation for Statistical Computing, 2014.

Raich, J. W., Russell, A. E., Kitayama, K., Parton, W. J. and Vitousek, P. M.: Temperature influences carbon accumulation in moist tropical forests, Ecology, 87(1), 76–87, 2006.

Van Ranst, E., Delvaux, B., Baert, G., Imerzoukène, S., Jamagne, P., Ndayiragije, S. and Pineros Garcet, J. D.: Carte Pédologique du Rwanda. Feuille 29 : Carte Pédologique de Rwesero et Notice explicative, UGent/MINAGRI., 2000a.

Van Ranst, E., Delvaux, B., Baert, G., Imerzoukène, S., Jamagne, P., Ndayiragije, S. and Pineros Garcet, J. D.: Carte pédologique du Rwanda. Feuille 30 : Carte Pédologique de Mushubi et Notice explicative, UGent/MINAGRI., 2000b.

Van Ranst, E., Delvaux, B., Baert, G., Imerzoukène, S., Jamagne, P., Ndayiragije, S. and Pineros Garcet, J. D.: Carte Pédologique du Rwanda. Feuille 38 : Carte Pédologique de Bugumya et Notice explicative, UGent/MINAGRI., 2000c.

Van Ranst, E., Delvaux, B., Baert, G., Imerzoukène, S., Jamagne, P., Ndayiragije, S. and Pineros Garcet, J. D.: Carte Pédologique du Rwanda. Feuille 39 : Carte Pédologique de Nshili et Notice explicative, UGent/MINAGRI., 2000d.

Richards, P. W.: The tropical rain forest: an ecological study., 1952.

Ricketts, T. H., Daily, G. C., Ehrlich, P. R. and Michener, C. D.: Economic value of tropical forest to coffee production, Proc. Natl. Acad. Sci. U. S. A., 101(34), 12579–12582, 2004.

Robert, A.: Simulation of the effect of topography and tree falls on stand dynamics and stand structure of tropical forests, Ecol. Modell., 167(3), 287–303, 2003.

Rütting, T., Cizungu Ntaboba, L., Roobroeck, D., Bauters, M., Huygens, D. and Boeckx, P.: Leaky nitrogen cycle in pristine African montane rainforest soil, Global Biogeochem. Cycles, 29(10), 1754–1762, doi:10.1002/2015GB005144, 2015.

Rwanda Travel Guide: About Rwanda: Nyungwe National Park, 2016.

Sandbrook, C. G. and Burgess, N. D.: Biodiversity and ecosystem services: Not all positive, Ecosyst. Serv., 12(January), 29, 2015.

Scaranello, M. A. da S. and Alves, L. F.: Height-diameter relationships of tropical Atlantic moist forest trees in southeastern, Sci. Agric, 69(February), 26–37, doi:10.1590/S0103-90162012000100005, 2012.

Schulze, E. D., Beck, E. and Muller-Hohenstein, K.: Plant ecology, Springer-Verlag., 2002.

Schulze, E.-D., Heimann, M., Harrison, S., Holland, E., Lloyd, J., Prentice, C. I. and Schimel, D.: Global Biogeochemical Cycles in the Climate System, Academic Press, Jena, Germany., 2001.

Schwartz, D. and Namri, M.: Mapping the total organic carbon in the soils of the Congo, Glob. Planet. Change, 33(1-2), 77–93, 2002.

Seibt, U., Rajabi, A., Griffiths, H. and Berry, J. A.: Carbon isotopes and water use efficiency: Sense and sensitivity, Oecologia, 155(3), 441–454, doi:10.1007/s00442-007-0932-7, 2008.

van der Sleen, P., Groenendijk, P., Vlam, M., Anten, N. P. R., Boom, A., Bongers, F., Pons, T.

Page 61: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

49

L., Terburg, G. and Zuidema, P. a.: No growth stimulation of tropical trees by 150 years of CO2 fertilization but water-use efficiency increased, Nat. Geosci., 8(January), 24–28, 2014.

Slik, J. W. F., Arroyo-Rodriguez, V., Aiba, S.-I., Alvarez-Loayza, P., Alves, L. F., Ashton, P., Balvanera, P., Bastian, M. L., Bellingham, P. J., van den Berg, E., Bernacci, L., da Conceicao Bispo, P., Blanc, L., Bohning-Gaese, K., Boeckx, P., Bongers, F., Boyle, B., Bradford, M., Brearley, F. Q., Breuer-Ndoundou Hockemba, M., Bunyavejchewin, S., Calderado Leal Matos, D., Castillo-Santiago, M., Catharino, E. L. M., Chai, S.-L., Chen, Y., Colwell, R. K., Robin, C. L., Clark, C., Clark, D. B., Clark, D. A., Culmsee, H., Damas, K., Dattaraja, H. S., Dauby, G., Davidar, P., DeWalt, S. J., Doucet, J.-L., Duque, A., Durigan, G., Eichhorn, K. A. O., Eisenlohr, P. V., Eler, E., Ewango, C., Farwig, N., Feeley, K. J., Ferreira, L., Field, R., de Oliveira Filho, A. T., Fletcher, C., Forshed, O., Franco, G., Fredriksson, G., Gillespie, T., Gillet, J.-F., Amarnath, G., Griffith, D. M., Grogan, J., Gunatilleke, N., Harris, D., Harrison, R., Hector, A., Homeier, J., Imai, N., Itoh, A., Jansen, P. A., Joly, C. A., de Jong, B. H. J., Kartawinata, K., Kearsley, E., Kelly, D. L., Kenfack, D., Kessler, M., Kitayama, K., Kooyman, R., Larney, E., Laumonier, Y., Laurance, S., Laurance, W. F., Lawes, M. J., Amaral, I. L. do, Letcher, S. G., Lindsell, J., Lu, X., Mansor, A., Marjokorpi, A., Martin, E. H., Meilby, H., Melo, F. P. L., Metcalfe, D. J., Medjibe, V. P., Metzger, J. P., Millet, J., Mohandass, D., Montero, J. C., de Morisson Valeriano, M., Mugerwa, B., Nagamasu, H., Nilus, R., et al.: An estimate of the number of tropical tree species, PNAS, 112(24), 7472–7477, 2015.

Soethe, N., Lehmann, J. and Engels, C.: Nutrient availability at different altitudes in a tropical montane forest in Ecuador, J. Trop. Ecol., 24(04), 397–406, doi:10.1017/S026646740800504X, 2008.

Spracklen, D. V. and Righelato, R.: Tropical montane forests are a larger than expected global carbon store, Biogeosciences, 11(10), 2741–2754, 2014.

Tanner, E. V. J., Vitousek, P. M., Cuevas, E. and Jan, N.: Experimental Investigation of Nutrient Limitation of Forest Growth on Wet Tropical Mountains, Ecology, 79(1), 10–22, 1998.

Taveirne, C.: Functional diversity study on an altitudinal forest transect in Central Africa, Nyungwe National park, Rwanda, UGent., 2016.

Townsend, A. R., Asner, G. P. and Cleveland, C. C.: The biogeochemical heterogeneity of tropical forests, Trends Ecol. Evol., 23(8), 424–431, doi:10.1016/j.tree.2008.04.009, 2008.

Townsend, A. R., Cleveland, C. C., Houlton, B. Z., Alden, C. B. and White, J. W. C.: Multi-element regulation of the tropical forest carbon cycle, Front. Ecol. Environ., 9(1), 9–17, doi:10.1890/100047, 2011.

Troll, C.: Die tropische Gebirge: Ihre dreidimensionale klimatische und pflanzengeographische Zonierung., 1959.

USAID and Wildlife Conservation Society: Sustaining biodiversity conservation in and around Nyungwe National Park (NNP)., 2013.

Velescu, A., Valarezo, C. and Wilcke, W.: Response of Dissolved Carbon and Nitrogen Concentrations to Moderate Nutrient Additions in a Tropical Montane Forest of South Ecuador, Front. Earth Sci., 4(58), 1–18, doi:10.3389/feart.2016.00058, 2016.

Page 62: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

50

Vitousek, P. M.: Litterfall , Nutrient Cycling , and Nutrient Limitation in Tropical Forests, Ecology, 65(1), 285–298, 1984.

Vitousek, P. M. and Farrington, H.: Nutrient limitation and soil development: Experimental test of a biogeochemical theory, Biogeochemistry, 37(1), 63–75, doi:10.1023/A:1005757218475, 1997.

Vitousek, P. M., Aplet, G., Turner, D. and Lockwood, J. J.: The Mauna Loa environmental matrix: foliar and soil nutrients, Oecologia, 89(3), 372–382, 1992.

Waide, R. B., Zimmerman, J. K. and Scatena, F. N.: Controls of Primary Productivity : Lessons from the Luquillo Mountains in Puerto Rico, Ecology, 79(1), 31–37, 1998.

Watanabe, Y., Martini, J. E. and Ohmoto, H.: Geochemical evidence for terrestrial ecosystems 2.6 billion years ago., Nature, 408(6812), 574–8, doi:10.1038/35046052, 2000.

van de Weg, M. J., Meir, P., Grace, J. and Atkin, O. K.: Altitudinal variation in leaf mass per unit area, leaf tissue density and foliar nitrogen and phosphorus content along an Amazon-Andes gradient in Peru, Plant Ecol. Divers., 2(3), 243–254, 2009.

Williams, C. a, Hanan, N. P., Neff, J. C., Scholes, R. J., Berry, J. a, Denning, a S. and Baker, D. F.: Africa and the global carbon cycle., Carbon Balance Manag., 2, 3, 2007.

WWF: Albertine Rift Montane Forests - A Global Ecoregion, [online] Available from: http://wwf.panda.org/about_our_earth/ecoregions/albertine_montane_forests.cfm (Accessed 16 May 2016), 2011.

Zanne, A. E., Lopez-Gonzalez, G., Coomes, D. A., Ilic, J., Jansen, S., Lewis, S. L., Miller, R. B., Swenson, N. G., Wiemann, M. C. and Chave, J.: Global wood density database. Dryad, 2009.

Zhou, X., Fu, Y., Zhou, L., Li, B. and Luo, Y.: An imperative need for global change research in tropical forests, Tree Physiol., 33(9), 903–912, 2013.

Page 63: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

51

9 Appendices

A. RAINFOR field work database codes for living trees

a = alive normal

b = alive, broken stem/top &resprouting, or at least phloem/xylem. Note at what height stem is broken.

c = alive, leaning by ≥ 10%

d = alive, fallen (e.g. on ground)

e = alive, tree fluted and/or fenestrated

f = alive, hollow

g = alive, rotten

h = multiple stemmed individual (each stem > 99 mm gets a number), always use with another code – e.g. if a tree is normal and with multiple stems, use ‘ah’, etc.

i = alive, no leaves/few leaves

j = alive, burnt

k = alive snapped < 1.3 m

l= alive, has liana >=10cm d on stem or in canopy

m=covered by lianas (note only in the case where canopy is at least 50% covered by lianas, even where no individual liana reaches 10 cm d)

n = new (recruit), always use with another code – e.g. if a tree is normal and new the code = ‘an’, if a tree is broken and new the code is ‘bn’, etc.

o= lightning damage

p= cut

q= bark loose or flaking off

s= has strangler

t = is a strangler

z = alive, declining productivity (nearing death, diseased etc.)

Note, tree status codes can be used together in whatever combination is necessary! Thus, for example, a multiple stemmed, leaning and broken tree would be coded bch.

Page 64: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

52

B. Lab protocol for soil bioavailable P

The resin P method is an adequate method that enables the extraction of biological available

forms of inorganic P (Pansu and Gautheyrou, 2006). It makes use of anion exchange membranes

(AEM). Before starting the method, it is very important to clean the AEM strips. Therefore, the

strips were washed 3 times with milli-Q® water (MerckMillipore, Darmstadt, Germany), making

sure all residual P, from possible previous uses, has been removed. Subsequently, the strips were

‘activated’, i.e. regenerated into the bicarbonate form by placing the strips into 0.5 M NaHCO3

(pH 8.5).

To measure the available P, 1 g of soil was weighed into a plastic jar containing 2 activated AEM

strips and 30 ml of milli-Q® water. This jar was shaken end-over-end for 16 hours at room

temperature. Consequently, the membranes were removed, cleaned with as little milli-Q® water

as possible to remove soil particles, placed into another plastic jar with 20 ml of 0.5 M HCL, and

shaken for another 16 hours. Subsequently, the strips were removed and the HCl desorption

solution was ready for P analysis.

As the desorption solution was too acid to measure with the auto-analyser, a phosphate

colorimetric kit (Sigma-Aldrich) was used to measure absorbances at a wavelength of 650 nm. In

the end bio-available P concentrations were determined via a calibration curve.

Page 65: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

53

C. Graphs of the additional carbon stocks

Figure 20. Carbon estimations per PSP n ton ha-1

for each plot-altitude. (a) BGC* is the belowground carbon in the top 30 cm (b) is the carbon in the O-horizon and (c) is the carbon in the Litter. No linear regressions and R

2 values are shown as

all p>0.05.

Page 66: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

54

D. Graphs of N & C in litter, O-horizon and roots

Figure 13. Concentrations of nitrogen and carbon in g g-1

. (a) and (b) are for litter, (c) and (d) are for the O-horizon and € and (f) are for the roots. Linear regressions were visualized and R

2 values were given when p<0.05.

Page 67: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

55

E. Graphs of δ13C and δ 15N in the soil at different depths

Figure 22. δ15

N and δ13

C of the soil at different depths in ‰. (a) δ15

N in ‰ for 0-5 cm (b) δ15

N in ‰ for 5-10 cm (c) δ15

N in ‰ for 10-20 cm (d) δ

15N in ‰ for 20-30 cm (e) δ

15N in ‰ for 30-50 cm and (f) δ

15N in ‰ for 50-100 cm. Linear regressions

were visualized and R2

values were given when p<0.05.

Page 68: Carbon storage and nutrient shifts along an altitudinal ...lib.ugent.be/fulltxt/RUG01/002/305/127/RUG01... · montane forests (TMFs) in Rwanda and Ecuador. Two students studied functional

56

Figure 23. δ13

C of the soil at different depths in ‰. (a) δ13

C in ‰ for 0-5 cm (b) δ13

C in ‰ for 5-10 cm (c) δ13

C in ‰ for 10-20 cm (d) δ

13C in ‰ for 20-30 cm (e) δ

13C in ‰ for 30-50 cm (f) δ

13C in ‰ for 50-100 cm. Linear regressions were

visualized and R2

values were given when p<0.05.


Recommended