Anda di halaman 1dari 245

J. S.

Singh
R. K. Chaturvedi

Tropical Dry
Deciduous Forest:
Research Trends
and Emerging
Features
Tropical Dry Deciduous Forest: Research Trends
and Emerging Features
J. S. Singh • R. K. Chaturvedi

Tropical Dry Deciduous


Forest: Research Trends
and Emerging Features
J. S. Singh R. K. Chaturvedi
Department of Botany Center for Integrative Conservation
Banaras Hindu University Xishuangbanna Tropical Botanical Garden
Varanasi, Uttar Pradesh, India Chinese Academy of Sciences
Mengla, Yunnan, China

ISBN 978-981-10-7259-8    ISBN 978-981-10-7260-4 (eBook)


https://doi.org/10.1007/978-981-10-7260-4

Library of Congress Control Number: 2018934838

© Springer Nature Singapore Pte Ltd. 2017


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, express or implied, with respect to the material contained herein or for any errors
or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims
in published maps and institutional affiliations.

Printed on acid-free paper

This Springer imprint is published by the registered company Springer Nature Singapore Pte Ltd. part of
Springer Nature.
The registered company address is: 152 Beach Road, #21-01/04 Gateway East, Singapore 189721,
Singapore
Preface

Tropical dry deciduous forests (TDFs) occur in severe and extremely variable cli-
mate characterised by low annual rainfall and 5–6 months of dry period within the
annual cycle. The original global extent of TDFs is not clearly documented; how-
ever, these forests exhibit worldwide distribution. The canopy of trees in these for-
ests ranges from multi-layered (forest) to highly discontinuous (savannas). Although
TDFs exhibit lower species richness compared to rain forests, the structural and
physiological diversity in various life forms of TDFs is very high compared to those
in rainforests. The majority of woody species in the TDF exhibit drought deciduous-
ness as a response to the long dry period in the annual cycle.
Soils of about half of all tropical forests have been reported to be highly degraded,
leached and impoverished. Plants growing in such habitats have developed mecha-
nisms to conserve nutrients. In TDFs, the influence of the soil microbial community
is the immediate controlling factor for carbon (C) mineralization rates in soils; it is
generally assumed that the Birch effect (release of large CO2 pulses from the dry
soils after rewetting) occurs when the microbial osmolytes are released upon rewet-
ting and/or when organic substrates released by alteration in soil structure are rap-
idly metabolised. The drying-rewetting cycles could have strong impact over the
biogeochemical cycles in the TDF ecosystems, as the nutrients which are immobi-
lized in the microbial biomass are mineralized quickly at the start of the wet season,
meeting the nutrient demand of plants.
Occurring in such severe and extremely variable environment, the vegetation of
TDF is a mosaic of communities which are dynamic in space and time and whose
distribution in non-contiguous patches is determined by soil texture. The TDF is a
productive and aggrading ecosystem with three levels of superimposed mineral
cycling, viz. internal, short-term through short-lived components and long-term
through long-lived components. Disproportionate to their structural role, short-lived
components play a dominating role in ecosystem function. The phenological clock
of the forest is set during the inter-phase of winter and summer, ensuring full advan-
tage of the short rainy season that follows. Plants withdraw substantial amounts of
nitrogen (N) and phosphorus (P) from senescing leaves so that each unit of nutrient
absorbed from the soil is used for the development of successive generations of

v
vi Preface

f­ oliage. Habitat heterogeneity results in nutrient hot spots, which attract fine roots
to support tree growth. The forest has a patchy distribution of C stock. The species
vary markedly in terms of functional traits, plasticity and C accumulation rate. The
growth response of trees of different species is modulated by alterations in key
functional traits.
Analysis of the relationships among seed size, plant distribution and abundance
in a TDF of northern India showed that the mean seed mass was significantly associ-
ated with the mode of dispersal as well as shade tolerance, with the major propor-
tion of small-seeded species being wind-dispersed, whereas all the large-seeded
species were found to be mammal-dispersed. Mechanism of regeneration in TDFs
also varies substantially from the tropical humid forests; for example, sprouting in
TDF is considered to be relatively more significant due to germination constraints
and high mortality of seedlings. Most TDF species have been observed to re-sprout
after disturbances, start with a vigorous shoots and do not carry the burden of vul-
nerable life stages.
The TDFs experience intensive anthropogenic disturbance and are among the
most threatened ecosystems in the world. The biotic disturbances have resulted in
fragmentation and ecosystem conversion, and therefore, the TDFs exhibit changes
in biomass, productivity, soil microbial biomass, etc. Yet, these forests provide a
suite of ecosystem services which are the basis for the livelihoods of millions of
people.
Several studies have suggested that the future drought severity might have sev-
eral possibilities, for example, decline in the annual rainfall, alteration in the timing
and duration of rainfall, alteration in the extent of dry season, or multi-year drought
combined with sequential low rainfall; therefore, concerted research effort is needed
to understand the response of TDFs to such drought conditions. Although the TDF
has shown some resistance to the invasion of exotic species, large areas of the TDF
are still being invaded by invasives, affecting the regeneration of indigenous tree
species and interfering with forestry operations. This aspect has been little studied
and needs extensive investigation.
In order to assess conservation status of this forest type, and to develop appro-
priate management strategies, information on its distribution pattern, climate,
structure and functional traits of the vegetation, phenology, strategies against
drought and nutrient poverty, and disturbance effects needs to be synthesized.
Intensive exploitation has fragmented and converted significant portions of the
TDF. The converted ecosystems have reduced amounts of microbial C, N and P,
and reduced proportion of macro-aggregates in the soil. Macro-aggregates are
dominated by fungal-based food web organisms conducive to C retention as
opposed to micro-aggregates which have bacteria-dominated food web with rapid
C turnover. These ecosystem conversions contribute to atmospheric loading of
CO2. Systems with small structure (low standing crop of C) maintain C flux rates
similar to systems with large structure (high standing crop of C) at the cost of C
conservation.
Preface vii

In this monograph, we review important studies on TDFs across the globe, par-
ticularly those conducted at the Banaras Hindu University on the northern dry
deciduous forests. This report suggests that the TDFs, which experience drought
and occur on nutrient poor sites, have adaptations such as varying degree of decidu-
ousness and a variety of nutrient conservation strategies.

J. S. Singh
R. K. Chaturvedi
Contents

1 Introduction.............................................................................................. 1
2 Global Coverage, Climate and Soil........................................................ 13
2.1 Global Coverage............................................................................... 13
2.2 Climate Variability............................................................................ 17
2.3 Edaphic Features............................................................................... 19
2.3.1 Spatial Heterogeneity and Availability of Nutrients............. 20
2.3.2 Soil Microbial Biomass and Ecosystem Functioning........... 24
2.3.3 Carbon Balance and CO2 Flux.............................................. 27
2.3.4 Carbon Balance and Global Change..................................... 28
3 Vegetation Attributes...............................................................................  31
3.1 The Tropical Dry Forest.................................................................... 31
3.1.1 Tropical Dry Forest Vegetation of India................................ 32
3.1.2 Small-Scale Variations in Environmental Factors
and the Distribution of Woody Species................................. 43
3.1.3 Composition and Dynamics of Tropical Dry Forest
in Relation to Soil Texture.................................................... 43
3.1.4 Effect of Woody Plant Canopies on Species Composition
and Diversity of Ground Vegetation...................................... 45
3.2 Seasonality, Phenology and Deciduousness..................................... 47
3.2.1 Phenology of Seasonally Tropical Dry Forest: Two
Case Studies.......................................................................... 53
3.2.2 Reproductive Phenology....................................................... 57
3.2.3 Seasonal Variation in Tree Water Status............................... 58
3.3 Seasonality, Leaf Attributes and Growth Rates................................. 60
3.4 Plant Strategies Against Drought...................................................... 63
3.5 Relations Between Water Balance, Wood Traits and Phenological
Behaviour.......................................................................................... 65

ix
x Contents

4 Plant Traits and Regeneration................................................................  69


4.1 Plant Functional Traits...................................................................... 69
4.1.1 Leaf Traits............................................................................. 71
4.1.2 Stem and Root Traits............................................................. 75
4.1.3 Reproductive Traits............................................................... 75
4.1.4 Functional Trait Syndrome for Maintaining Growth
in Seasonally Dry Environments........................................... 77
4.2 Plant Functional Types...................................................................... 78
4.3 Seed Size: A Key Trait Determining Species Distribution............... 79
4.4 Seed and Seedling Ecology............................................................... 84
4.4.1 Seed Viability and Dormancy............................................... 84
4.4.2 Seed Size, Germination, Water Stress and Seedling
Growth.................................................................................. 85
4.4.3 Major Environmental Factors Influencing Seedling
Recruitment........................................................................... 88
4.5 Regeneration..................................................................................... 107
5 Productivity and Nutrient Cycling......................................................... 111
5.1 Carbon Stock and Biomass Accumulation Pattern........................... 111
5.1.1 Dry Matter Dynamics, Storage and Flux
of Nutrients in TDF of India: A Case Study......................... 115
5.1.2 Carbon Stock and Accumulation in Woody Species
of TDF in India..................................................................... 126
5.2 Nutrient Cycling and Nutrient Conservation Strategies................... 130
5.2.1 Effect of Precipitation Regime on N and P Cycles............... 130
5.2.2 Foliar Nutrient Concentration and Temporal Variation........ 132
5.2.3 Litter Decomposition............................................................ 140
6 Influence of Biotic Pressure and Land-Use Changes............................ 149
6.1 Biotic Disturbance in Tropical Dry Forests...................................... 149
6.2 Tree Species Composition, Dispersion and Diversity: A Case
Study from a Tropical Dry Forest Region of India........................... 154
6.3 Impact of Forest Fragmentation and Edge Effects on Species
Diversity............................................................................................ 155
6.4 Diversity, Recruitment and Future Composition of Juveniles
and Saplings ..................................................................................... 159
6.5 Fodder and Fuel Extraction and Functioning of Village
Ecosystems........................................................................................ 164
6.6 Effect of Cultivation on Soil Subsystem........................................... 165
6.6.1 Viable Nitrifier Community and Nutrient Availability
in TDF and Derived Cropland Sites...................................... 165
6.6.2 Comparative Analysis of Microbial C, N and P in TDF
and the Alternate Land Uses................................................. 168
6.7 Invasive Alien Plants and Their Effect on TDFs in India................. 173
Contents xi

6.8 Degradation and Savannization......................................................... 175


6.9 Water-Stable Aggregates and Microbial Biomass in TDF
and Derived Ecosystems................................................................... 178
7 Research Perspectives.............................................................................. 191

References......................................................................................................... 197
About the Authors

Dr. J. S. Singh is an Indian ecologist, academic and a former Professor of Botany


at Banaras Hindu University.
The Council of Scientific and Industrial Research awarded Singh the Shanti
Swarup Bhatnagar Prize, one of India’s most prestigious science awards, in 1980.
He was selected for the Pitamber Pant National Environment Fellowship by the
Ministry of Environment, Forest and Climate Change in 1984, and for the
Pranavanand Saraswati Award by the University Grants Commission of India in
1985. The Indian Botanical Society awarded him the Birbal Sahni Gold Medal in
1999 and he received the Prof. S. B. Saksena Memorial Medal of the Indian National
Science Academy the same year. He has won the Honor of Distinction of the Society
for Protection of Environment and Sustainable Development (2003) and the Lifetime
Achievement Award of the AWA (2005).
Dr. R. K. Chaturvedi is Postdoctoral Fellow at Xishuangbanna Tropical Botanical
Garden, Chinese Academy of Sciences, Yunnan, China. He has received the National
Natural Science Foundation of China (NSFC) Research Fund for International
Young Scientists. His publications include several papers in high impact journals
for his work done in the northern dry deciduous forest of India.

xiii
Chapter 1
Introduction

Tropical dry forest (TDF) is found in regions characterized by several months of


severe drought within the annual cycle, with most of the rain falling during a brief
wet season (Murphy and Lugo 1986; Mooney et al. 1995). Several terms have been
used for this vegetation type such as seasonally dry tropical forest (SDTF), tropical
dry deciduous forest, monsoon forest, caatinga, cuabal, etc. Description of savannas
of Asia by Ratnam et al. (2016) and of ‘tropical grassy biomes’ by Lehmann and
Parr (2016) confuses the meaning of TDF. Lehmann and Parr (2016) state that “In
many regions, including Madagascar, Southeast Asia and South America, grassy
biomes have historically been considered either a degraded form of forest of anthro-
pogenic origin” having been created by tree clearing, burning and grazing or a sub-
climax or secondary successional stage. Murphy et al. (2016) have stated that “The
Earth’s tropical landscapes are dominated by two strongly contrasting biomes:
savannahs and grasslands on the one hand and closed-canopy forests on the other”
and “There has been a widespread misconception that TDFs are anthropogenically
degraded forests”.
The ecotones between TDFs and other vegetation types may vary from broad to
remarkably narrow, for example, in some inter-Andean valleys which exhibit change
in precipitation conditions over short distances (Pennington et al. 2009). The TDF
biome may be considered to include many communities which reciprocally and
preferentially act as immigrant sources in the process of evolution (Pennington et al.
2009). Such kind of evolutionary definition of TDF biomes corresponds with the
definition of metacommunity given by Hubbell (2001). In the neotropics, the TDFs
are found in frost-free regions experiencing rainfall less than c. 1800 mm year−1 and
a period of almost 5–6 months with less than 100 mm rainfall (Gentry 1995; Murphy
and Lugo 1986). Compared to tropical rain forests, TDFs exhibit lower canopy
cover as well as basal area (Murphy and Lugo 1986), and often thorny and succulent
species are commonly seen particularly in the drier formations. The TDF metacom-
munity also shows high levels of dispersal limitation, which confines the endemic
species to a single TDF area. These endemic species are often monophyletic and
relatively old. Moreover, these highly localized endemic species are often rare

© Springer Nature Singapore Pte Ltd. 2017 1


J. S. Singh, R. K. Chaturvedi, Tropical Dry Deciduous Forest: Research Trends
and Emerging Features, https://doi.org/10.1007/978-981-10-7260-4_1
2 1 Introduction

g­ lobally but common locally. The global TDF metacommunity could also contain
the bush thickets and other seasonally dry vegetation formations from the Old World
(Schrire et al. 2005a, b).
Typically, TDFs are described as the tropical forests which experience 250–
2000 mm of annual rainfall and a drought period of at least 3–4 months (Murphy and
Lugo 1986). However, Pennington et al. (2006a) introduced a broader interpretation,
according to which TDF includes vegetation which experiences a minimum drought
period of 5–6 months, leading to strong seasonality in ecological processes and func-
tions. Therefore, the lack of rainfall for a prolonged part of the year is the most impor-
tant factor producing the true dry forest, in which plants and animals are characterized
by having specific mechanisms for their survival in the dry season. The plant and
animal diversity in TDF is remarkably high. These forests exhibit heterogeneous veg-
etation with a range of formations varying from tall forests in moist forests to cactus
scrub in very dry habitats; however the dominance of semi-­deciduous to deciduous
trees is the most common feature (Murphy and Lugo 1986; Pennington et al. 2006a).
Depending upon the distance from the sea, altitude and latitude, the variations in
temperature between the seasons in TDFs can be broad or narrow. For instance, in
North Australia, savannas which are located near the tropics (12–20°S) and also
close to the sea experience uniformly high temperature (the maximum ranges from
30.5 to 31.8 °C and minimum from 20.0 to 24.8 °C) (Eamus 1999). In contrast, at
the tropical margins or at higher altitudes, temperatures can reach freezing point
with seasonal variations of 20 °C or more (Eamus 1999). However, the TDF ecosys-
tems are generally present in the areas where the annual mean air temperature is
greater than 20 °C and the annual mean temperature in coldest month is greater than
13 °C (Archibald 1995).
Seasonally dry ecosystems are characterized by having distinct wet and dry sea-
sons. The duration of dry season may vary from 1 to 7 months (Johnson and Tothill
1985). Rainfall conditions in TDFs can be bimodal (central and southern Africa) or
unimodal (North Africa, Central America and India) (Archibald 1995). The distin-
guishing feature of the seasonality in the TDFs and woodlands compared to the
temperate seasonal climates is the phenomenon of reversal in association between
temperature and rainfall in the annual cycle (Johnson and Tothill 1985). The length
of dry season in TDFs can be shorter ranging from 1 to 3 months as observed in
humid regions, or it may be longer ranging from 5 to 8  months as found in dry
regions (Johnson and Tothill 1985).
Among plants in the TDFs, deciduousness is the single most important adapta-
tion to the extended droughts and has also been reported to be a good predictor of
drought survival (Poorter and Markesteijn 2008). Most of the trees drop their leaves
after the end of rainy season, and essentially halt photosynthesis, as they otherwise
cannot sustain water loss during the dry season (Murphy and Lugo 1986). Although
when the rainfall ends the upper soil layers soon become dry, the availability of
subsoil moisture shows wide variations at different sites and influences the degree
of desiccation particularly in trees with low stem water storage capacity. The dense,
high hill forests experiencing high root competition exhibit higher degree of
­desiccation compared to the savanna where trees are widely spaced (Borchert
1 Introduction 3

1994a). Before the leaves are shed, these species efficiently reabsorb nutrients which
are utilized during the drought period and in the early growing season (Aerts 1996;
Givnish 2002). Earlier studies have reported that during the dry period, deciduous
species exhibit minimum water loss and therefore become dominant in drier forests
along a gradient of precipitation (Murphy and Lugo 1986; Santiago et al. 2004).
Most of the tree species exhibit flowering or leaf flushing soon after the shedding
of leaves (Borchert 1994c). Reich and Borchert (1982, 1984) studied the season-­wise
changes in the tree stem diameter and associated the diameter  change with the
changes in the tree water status, and reported that the increasing water stress during
early dry season generally initiated leaf shedding. After leaf shedding most of the
tree species rehydrated, and this rehydration was usually before flowering or flushing
(Borchert 1980, 1991; Reich and Borchert 1982, 1984). It has been reported that the
trees showing dry-season leaf flushing might overcome the soil water stress either by
their deep tap root system or by the use of stored water in their vascular conductive
tissues, as the transpiration is absent during the leafless period (Bullock and Solís-
Magallanes 1990; Murali and Sukumar 1993). Moreover, it has been suggested that
the leaf flushing time is affected by the need of the plant for harvesting more light and
to minimize competition among the sites which are physiologically active for growth
and reproductive functions (Alvim and De 1964; Wareing and Patrick 1975).
In the tropical dry region experiencing low water availability, plants adapt to
drought by allocating more biomass to roots (Mokany et  al. 2006; Roa-Fuentes
et al. 2012). The drought-adapted plant roots have also been observed to contain
certain biopolymers, for example, cutin and suberins, which make the cell walls
resistant to water and air and provide protection against the attack of harmful
microbes (Boom et  al. 2005). The carbon derived from plant roots is more effi-
ciently retained in soils compared to the above-ground carbon inputs by leaves
because of the presence of higher quantities of resistant biopolymers in roots
(Lorenz et  al. 2007; Feng et  al. 2008; Pisani et  al. 2014) and also due to higher
chances for interacting physically and chemically with soil particles (Rasse et al.
2005). Therefore, the carbon allocated by plant roots affects the composition of the
soil organic matter (Schmidt et al. 2011) and contributes efficiently to the soil car-
bon dynamics and sequestration.
An important feature of TDFs is the seasonal phenology which is controlled by
a long and extreme dry season. Growth and reproduction occur mostly in a short wet
season. The occurrence of a short wet season also limits the possibilities of seed
germination, seedling establishment, regeneration and the process of natural dry
forest succession (Opler et  al. 1976; Quesada et  al. 2009). Many key ecological
processes take their cue from the arrival of the rains. In brief, as the new leaves
begin catching sunlight, growth of plants which was stalled during the dry months
shifts into high gear. On the forest floor, the accumulated leaf litter is moistened by
precipitation, and insects and bacteria initiate the decomposition process leading to
release of nutrients which is reused by the growing plants. As the rainy season is
about to end, trees drop their leaves and profusion of flowers is produced in a
firework-­like display. This display is actually for attracting the pollinators. Plants
remain mostly dormant during the dry season, using their stored water, and wait for
4 1 Introduction

the rains. Leaf litter on the ground dries, and animals such as insects and frogs
return to their chambers to wait for the return of the favourable season. As the dry
season approaches its conclusion, fruits start forming in the pollinated flowers.
From this period onwards, fruits and leaves are in abundance, and all animals, from
birds to mammals to insects and lizards, start foraging voraciously to regain the
weight which they had lost during the dry season and begin looking for mates. In
order to take advantage of the surplus fruit and other foods, young are produced
quickly. Soon the forest is teeming with new life, bright green and full of the sounds
of animals stocking up in preparation for yet another long, hot dry season. Studies
on these phenological patterns of TDF also play an important role in remote sensing
since they function as a key element for understanding large-scale monitoring pro-
cesses aimed at recording local, regional and global ecosystem processes (e.g. net
primary productivity) (Quesada et al. 2009).
For the integrated predictive interpretation of phenology, it is necessary to quan-
tify links between species-wise variation in life-history traits and physiological pro-
cesses across temporal scales (Visser et al. 2010). The large variation in phenology
observed among the trees of dry forest biome experiencing same type of seasonality
in climatic drought has been expected to be caused due to differences among vari-
ous components constituting soil-plant-atmosphere continuum that governs tree
water status (Hinckley et  al. 1991; Borchert 1994a). For example, in a TDF of
Puerto Rico, Lasky et  al. (2016) reported significant evidence for flowering syn-
chrony across the plant communities at multiple timescales. Evidently, their find-
ings indicated the influence of rainfall or moisture in driving the synchrony (Borchert
1983). However, apart from rainfall, biotic interactions, such as flower and pollina-
tor interaction, can also influence flowering synchrony (Janzen 1967; Elzinga et al.
2007). Further, Lasky et al. (2016) reported rainy season as the community peak
flowering period, while in other TDFs the dry season is often observed as the peak
period of flowering (e.g. Frankie et al. 1974; McLaren and McDonald 2005; Singh
and Kushwaha 2006; Selwyn and Parthasarathy 2006).
In the study of Prasad and Hegde (1986) in the TDF at Bandipur (India, 11°39′N,
rainfall 1348  mm), among the 13 most commonly found tree species, only four
exhibited flowering within the season when leaves attained maturity, while none of
the species showed flowering near the end of the rainy season. Monasterio and
Sarmiento (1976) at Los Llanos reported dry season as the peak period of leaf flush-
ing, flowering and fruiting in evergreen or brevi-deciduous trees growing in savanna
as well as forest. In TDF, particularly for the deciduous trees, leaf flushing occurs
either before or just after the start of rains, but the trees exhibit diverse patterns for
reproductive period. Studies have reported alleviation of drought as the principal
external governing factor commonly responsible for the start of leaf flushing for
majority of the species and for initiation of reproduction for some species (Alvim
and De 1960; Opler et al. 1976; Augspurger 1980, 1983; see also Idso et al. 1978).
Effect of soil moisture becomes even more important for the dry forests ­experiencing
two annual rainy seasons where the deciduous trees become strictly opportunistic
regarding the moisture availability (Lieberman 1982; Tyrell and Coe 1974).
However, the significance of soil temperature as an important limiting factor has not
1 Introduction 5

been indicated in the phenological studies (Malaisse 1974). It seems true that phe-
nology of the TDF has not been properly studied, although moisture stress is more
commonly reported as a primary factor governing the processes of phenological
events (Murphy and Lugo 1986). In several other studies, significance of biotic fac-
tors, such as seasonality in pollinator populations, predators, competitors and seed
dispersal agents in influencing the evolution of phenological patterns in tropics, has
been reported (Lieberman 1982). According to Singh and Singh (1992a), the pheno-
logical cycle of the TDF is expected to be set during the transition period of winter
and summer seasons which enables the plant community to take sufficient advan-
tage of the available soil moisture during the rainy season for germination, recruit-
ment and productivity leading to increase in the period of food availability to
animals which are also involved in transfer of pollen and seed dispersal services.
Wright (1991) emphasized that the factors that may trigger the phenological events
could be water stress, humidity, temperature and timing of nutrient release.
TDFs and savanna can potentially grow at high rates; however, their capacity
to grow is strictly determined by climate and nutrients. Soils of about half of all
tropical forests have been reported to be highly degraded, leached and impover-
ished; therefore the ecosystem needs to develop certain mechanisms for nutrient
conservation (Sanchez 1976; Jordan 1985). The nutrient mineralization is of
pulsed nature, where sudden additions of organic matter through leaching of
nutrients from epiphytes or lysis of microbial biomass are commonly observed.
The pulsed drying and wetting cycles can significantly affect detrital food
chains resulting into pulses of nutrient mineralization; however the variations
between pulsed and steady-state nutrient mineralization have not always been
appreciated (Hunt et al. 1989). Moreover, changes in the quantity and timing of
precipitation indirectly affect soil respiration through their effects on the sub-
strate availability for microbial growth. Recognition of the contribution of soil
microorganisms to nutrient mineralization has increased the interest of research-
ers in measurement of the nutrients accumulated in the microbial biomass
(Jenkinson and Powlson 1976; Srivastava and Singh 1988; Singh et al. 1989).
The ecology and patchy global distribution of seasonally dry tropical forest
(SDTF) have distinctively structured the evolutionary history and biogeography of
woody plant groups that are confined to it. It has also been anticipated that the dis-
tribution of important woody plants in the dry tropics has been modified by humans
and therefore humans are supposed to be partly responsible for their wide geo-
graphic distribution (Levis et al. 2017). SDTFs have few widespread woody plant
species causing high β-diversity between separate areas of forests. These separate
areas contain geologically old, monophyletic clades of endemic plant species that
often have geographically structured intraspecific genetic variation. According to
Gentry (1995), the most frequent species in these forests generally belong to a few
important families, and the dominance of species is commonly affected by ecologi-
cal features of the habitat (De Souza et  al. 2007). These patterns of diversity,
­endemism and phylogeny indicate a stable, dispersal-limited SDTF system. SDTF
species tend to belong to larger clades confined to this vegetation, exemplifying
phylogenetic niche conservatism, and it has been argued that this is evidence that
6 1 Introduction

the SDTF is a metacommunity (biome) for woody plant clades. That phylogenetic,
population genetic, biogeographic and community ecological patterns differ in
woody plants from tropical rain forests and savannas suggests a hypothesis that
broad ecological settings strongly influence plant diversification in the tropics.
Forman (1995) and Lomolino (2001) have argued that large-sized patches con-
tain higher number of species compared to small-sized patches and also patch size
is more effective than isolation, patch age and several other factors in predicting
species richness. Species-area curves are useful for understanding the associations
between the species richness and the patch size, and consequently the probable
changes in species richness can be analysed in the forest areas with variable patch
sizes. By using species-area curve methods, a thumb rule calculation has indicated
that if 90% of the habitat is lost, it results into the loss of 50% of species (Heywood
and Stuart 1992). With the help of species-area curve, there have been many esti-
mates of the rate of extinction in tropical forests (Lovejoy 1980; Simberloff 1986;
Raven 1987; Myers 1988; Reid and Miller 1989; Reid 1992). The area-based extinc-
tion model applies the widely known species-area relationship (Mac Arthur and
Wilson 1967) for prediction of the loss of species in the areas which are fragmented
(Boeklen and Simberloff 1987). It has been anticipated that if the size of a forest is
reduced, it will result into the species loss according to certain gradients (Hill and
Curran 2003). The consequence of habitat fragmentation is the loss of forest, par-
ticularly for the sites which are important for endemism and species diversity.
A direct impact of fragmentation in forests is the formation of forest edges and
the consequent edge effects, which are created by the interaction among the adja-
cent ecosystems abruptly separated by a transition zone (Murcia 1995; Ries et al.
2004). The influence of multiple nearby edges, exhibiting complicated fragment
shapes, could result into complex edge effects. While applying the additive edge
model of Malcolm (1994) to understand the effects of variety of fragment shapes,
Fernandez et  al. (2002) found that it was almost impossible to detect identical
nearest-­edge gradient (i.e. gradients which extend from a particular location in a
fragment to the nearest edge) in the fragments which are irregularly shaped. In spite
of the significant potential importance of the contribution of edge additivity in frag-
mented forest patches, very little research has been conducted in this field (e.g. Ries
et al. 2004; Ewers et al. 2007).
In India, the species composition, distribution and diversity in TDF are influ-
enced by small-scale differences in environmental parameters leading to patchiness
in communities (Chaturvedi et al. 2011a; Chaturvedi and Raghubanshi 2014, 2015).
The environmental heterogeneity and disturbance are the primary factors of patch
formation. This patchy distribution of tree communities in TDFs results in uneven
distribution of the above-ground tree biomass as well as carbon accumulation
capacity (Chaturvedi et al. 2011b, c, 2012, 2017a). The deciduous forest in India has
been reported to be a mosaic of plant communities exhibiting distinct species com-
position; the distribution of these communities in noncontiguous patches results
into immense diversity (Jha and Singh 1990). It has also been reported that the
texture determines most of the characteristics of the soil, including permeability,
capacity to retain water, degree of aeration, capacity of storing plant-available nutri-
1 Introduction 7

ents in the clay-humus complex, ability to withstand mechanical damage of the top
soil and lastly the capacity to support a permanent plant cover (Jha and Singh 1990).
The Indian TDF contains a wide variation in the gradients of soil resources and
availability of light. Moreover, seeds and seedlings are consumed and dispersed by
a high diversity of vertebrate and invertebrate animals playing an important role in
the development of community structure. Life-history traits, for example, shade
tolerance, also affect the optimization of seed size and therefore indirectly affect
community structure. The dry forest communities produce patchy canopies leading
to formation of forest floor microenvironments having variable resource availabil-
ity, which provide specific niches suitable for the development of small as well as
large-seeded species (Khurana et al. 2006). Seedlings produced from large-seeded
species might have better capacity to cope with the carbon deficit temporarily result-
ing from the moderate disturbances such as grazing and browsing (see also
Armstrong and Westoby 1993). Larger seeds are also rich in various secondary
compounds, which might be translocated to seedlings from the seeds, leading to
reduction of the load of herbivory (Foster 1986). When the disturbance pressure is
severe, species having small- and medium-sized seeds are benefitted by their greater
ability of dispersal. According to Baker (1972), Foster (1986) and Hammond and
Brown (1995), various studies have reported that changes in environmental condi-
tions such as fluctuations in temperature, light intensity, rainfall and predation and
herbivory could lead to variations in seed size within the plant population and main-
tain the population with different seed sizes. Therefore, the biotic and abiotic pro-
cesses together determine the properties of species pools which lead to the formation
of local community structure. Besides this, the abundance of large-seeded species
with fleshy fruits attracts frugivorous animals, which indicates that in the least dis-
turbed sites, species in other trophic levels (i.e. consumers) could also be suffi-
ciently supported, resulting into further enrichment of the biodiversity (Khurana
et al. 2006).
Besides having unique biodiversity and a high percentage of endemic species
(Trejo and Dirzo 2002), TDFs also contain a large functional diversity of plants,
particularly in terms of phenological strategies of leaf (Eamus 1999). For example,
in trees the leaf habit strategies range from evergreen to drought-deciduous, while
an intermediate strategy is adapted by the brevi-deciduous species which shed their
leaves at the start of the dry season and then flush new leaves (Quesada et al. 2009).
TDFs are seasonally deciduous and have low leaf area index, leading to low shade
that almost disappears in the dry season. This low shade allows the sub-canopy trees
to positively contribute to the forest carbon balance during the wet season. Moreover,
due to the high light pulse in the dry season, the evergreens which remain physio-
logically active accumulate large amount of carbon. It may be expected that plants
growing in TDFs will exhibit low adaptation to shade, while the severe dry season
could function as a strong environmental filter.
Modes of regeneration in TDF determine survival after a disturbance and also
influence growth and survival when the disturbance is removed. Compared to wet
forests, our understanding of regeneration processes in TDFs is very poor (Meli
2003; Vieira and Scariot 2006). The regeneration processes in these TDFs are
8 1 Introduction

likely to be influenced by a complex interplay of biotic and abiotic factors (Powers


et al. 2009a). Biotic or management-related factors comprise the land-use pattern,
factors affecting seed arrival (e.g. seed sources and dispersing agents) and factors
that influence germination and establishment (e.g. competition from remaining
pasture grasses, depth of litter layer, scarcity of mycorrhizal symbionts and seed
predators) (Bazzaz and Pickett 1980; Ewel 1980; Wijdeven and Kuzee 2000;
Khurana and Singh 2001a; Hooper et al. 2005). Abiotic factors influencing regen-
eration comprise regular fire disturbance, temperature, precipitation, light and
nutrient availability, and drought conditions (Janzen 1988a; Gerhardt 1993; Campo
and Vazquez-Yanes 2004; Ceccon et al. 2004; Vargas-Rodriguez et al. 2005).
Primary productivity in TDFs is largely controlled by the timing and amount of
rainfall. Production of leaves as well as photosynthesis is affected by availability of
water, and the nutrient dynamics and ecosystem productivity are constrained and
controlled by interannual as well as rainy season precipitation. According to Singh
et al. (1992), the biological cycle of nutrients is considered to be one of the most
important processes supporting organic productivity. The biogeochemical cycling
in relatively dry ecosystems is particularly influenced by both quality and quantity
of organic matter in association with soil microclimate (Burke et  al. 1989). The
microbial dynamics, fine root production and organic matter decomposition are
regulated by the availability of soil water. Therefore, the seasonal and annual varia-
tion in rainfall is strongly associated with the nutrient availability in TDFs. Besides
water availability, primary productivity and nutrient dynamics in TDFs are also
influenced by biotic pressure and land-use change.
Length of monsoon season and the distribution of precipitation (i.e. timing,
duration and magnitude) in TDF primarily determine whether the ecosystem will
function as a net sink or a source of carbon at the annual timescale. In northwest
Mexico, where TDF experiences bimodal precipitation, the CO2 produced mainly
from heterotrophic respiration (net C source) before monsoon period is captured by
the TDF vegetation (net C sink) exhibiting high strength of primary productivity
during the North American monsoon (NAM) (Verduzco et  al. 2015). When the
monsoon season rainfall exceeds 350–400 mm, the available soil moisture becomes
sufficient for promoting the growth of shallow as well as deep-rooted plants
(Huxman et al. 2004; Méndez-Barroso et al. 2009), and contributes to the net car-
bon uptake. Such kind of carbon flux dynamics exhibiting a shift from the net car-
bon source to net carbon sink in an annual cycle could also be expected in other
seasonally dry ecosystems experiencing bimodal precipitation regime (Bullock
et  al. 1995; Verduzco et  al. 2015). Campo and Merino (2016) highlighted the
importance of direct and indirect influence of rainfall regime on carbon distribution
and turnover in SDTF ecosystems. The investigation on litterfall, and its decompo-
sition in the forest floor, and the carbon storage in the mineral soil by Campo and
Merino (2016) showed that in SDTF ecosystems, the decrease in precipitation for a
long time period would result in an increase of organic layer and mineral soil car-
bon storage, particularly due to reduced decomposition and increased chemical
recalcitrance of organic matter occasioned by changes in the composition of litter
and most probably by alteration in wildfire patterns. This might convert these
1 Introduction 9

SDTFs into significant soil carbon sinks under the predicted increase in drought
periods if the strength of primary productivity is maintained. The functioning of
majority of ecosystems is evidently associated with primary production, which is
commonly affected by the nutrients availability, and this availability of nutrients in
turn is influenced by their distribution and the cycling rates at the ecosystem level
(Chaturvedi and Singh 1987). A literature review on nutrient cycling and productiv-
ity in tropical and subtropical dry forests (Lugo et al. 1978; Arnason and Lambert
1982; Brown and Lugo 1982; Lugo and Murphy 1986; Murphy and Lugo 1986)
exhibited scanty information regarding the role of short-lived components such as
tree foliage, fine roots and herbaceous plants. Alvim and De (1964) and Wareing
and Patrick (1975) suggested that the different physiologically active components
of a plant might compete for acquiring water, nutrients and metabolites. The result
of such internal competition could be the differential time of the optima for the
plant’s functioning (Lieberman 1982). Edwards (1982) identified two types of min-
eral turnover occurring in forests: a rather significant rapid cycle in leaf litter, twig
litter and throughfall and a slower mineral cycling constituting larger woody com-
ponent. Edwards (1982) further suggested that short-term mineral cycling is rela-
tively rapid in comparison to the large amount of minerals stored in the vegetation.
In forest ecosystems a significant proportion of net primary production is accounted
by the fine roots (Santantonio et  al. 1977; Singh and Singh 1991a). Also, plants
allocate a large quantity of photosynthate to roots (Perry et  al. 1989); therefore,
production of fine roots is remarkably important for the cycling of nutrient as well
as carbon in forest ecosystems (Vogt et al. 1986; McClaugherty et al. 1982).
Leaf growth for the redevelopment of the canopy needs substantial quantity of
nutrients and water. In many TDF trees, elimination of the water loss in transpira-
tion by leaf shedding together with the effective use of residual water enables the
stem tissues to rehydrate for supporting the subsequent flushing of leaves during the
dry season (Borchert 1994a). The ability of soil to supply nutrients (Singh et  al.
1989; Raghubanshi 1992; Roy and Singh 1995) and the capacity for the uptake of
nutrients by roots are, however, minimum in the dry season due to a remarkably low
availability of soil water (Pandey and Singh 1992a).
Before the leaf abscission, deciduous woody plants retranslocate or resorb the
nutrients, and if they are able to reduce the concentration of N and P in senesced
leaves below 0.7% and 0.05%, respectively, nutrient resorption is considered to be
highly profitable and, subsequently, an effective mechanism for nutrient conserva-
tion (Killingbeck 1996). Resorption proficiency is an attribute that exhibits intrinsic
variations among species present in an ecosystem and also among individuals within
the same species (Killingbeck 1996; Richardson et al. 2005). Recent studies indi-
cated that this plant attribute better describes resorption in association with changes
in nutrient availability, status of plant nutrient and foliar N:P ratios (Lal et al. 2001a;
Wright and Westoby 2003; Rentería et al. 2005; Ratnam et al. 2008). McGroddy
et al. (2004) reported that the nutrient resorption, measured as the concentration of
nutrients in litterfall, is considered as a globally important mechanism, particularly
for P.  Yuan and Chen (2009a) suggested that the greater N:P ratios in senesced
leaves of trees in tropical forests compared to those of boreal forests indicates higher
10 1 Introduction

limitation of P and a greater tendency for P resorption compared to N in the leaves


of tropical trees.
The important role of water availability in the transformations of soil N and P,
and N and P availability in ecosystems experiencing low or markedly seasonal
quantities of rainfall (Davidson et al. 1993; Campo et al. 1998; Austin et al. 2004)
indicates that plant traits such as leaf N and P and resorption efficiency should
exhibit significant response to this key control on the dynamics of ecosystem. The
observations of Rentería et al. (2005) with six deciduous tree species of TDF indi-
cate that the concentrations of P but not N in green and senesced leaf show changes
according to the topographic-related changes in soil nutrients and availability of
water and also according to changes in annual rainfall. The findings of these authors
suggest that the resorption in TDF ecosystems is mostly controlled by water avail-
ability than the soil nutrient availability. Rentería and Jaramillo (2011) analysed the
core leaf traits, for example, leaf mass per area (LMA) and concentrations of leaf
nitrogen (N) and phosphorus (P) and resorption of nutrients in 21 woody species of
a TDF ecosystem located in western Mexico, and observed that the efficiency of
resorption and proficiency of P, but not N, enhanced in years experiencing low rain-
fall, which indicates that the nutrient resorption costs relative to acquisition costs
from soil vary between N and P, and also P conservation becomes higher when
rainfall becomes lower in the Chamela TDF.
Generally, it has been observed that species growing in low-nutrient habitats
exhibit higher efficiency for nutrient resorption compared to the species belonging
to the high-nutrient habitats (Aerts 1996). However, it is also evident that nutrient
resorption efficiency is apparently not very responsive to variations in nutrient sup-
ply and might not determine the composition and distribution of species across the
habitats differing in soil nutrient content (Aerts and Chapin 2000). Soil moisture
availability is among the several possible proposed controlling factors involved in
nutrient absorption (Boerner 1985; Del Arco et  al. 1991; Escudero et  al. 1992;
Pugnaire and Chapin 1993; Demars and Boerner 1997).
Although the site fertility is commonly reflected by leaf nutrient concentration,
the high interspecific differences observed in growth rates make tissue concentra-
tions of plants growing in wild habitats to be less sensitive indicators for the avail-
ability of soil nutrients (Chapin III 1980). Studies on the association between the
concentration of leaf nutrients and the availability of soil nutrients in the tropics
have reported conflicting results. For instance, in Hawaii, Harrington et al. (2001)
observed that foliar nutrients in trees were largely controlled by the supply of nutri-
ents. Austin and Vitousek (1998) also found that a decreasing availability of soil P
was related to the decrease in the concentration of foliar P; however this relationship
was not observed with N. In contrast, Lal et al. (2001b) reported no significant dif-
ference in average foliar P between the two sites in a TDF in India, exhibiting sig-
nificantly high variation in fertility. Rentería et al. (2005) measured the efficiency of
resorption in tree species located at the contrasting topographic positions (i.e. top
vs. bottom and north vs. south aspect) in a TDF in Mexico and observed that water
is more important compared to soil nutrient availability in controlling the resorption
1 Introduction 11

efficiency in Chamela TDF.  However, it may be possible that the interaction of


water availability and soil nutrients controls the resorption proficiency.
The mechanisms of nutrient retention and withdrawal are highly effective in sys-
tems poor in soil nutrients (Chapin III 1980; Singh et al. 1984). The study of Singh
and Singh (1991b) reported about 67% N, 70% P and 79% K retranslocated before
the abscission of leaves. In a Puerto Rican subtropical dry forest, Lugo and Murphy
(1986) reported a conservative recycling of P relative to N, with 65% of P showing
retranslocation prior to leaf fall. Staaf (1982) and Staaf and Berg (1981) suggested
that an efficient retranslocation capacity, especially of essential elements which are
in short supply, is a characteristic feature of tree species growing in a climax forest.
Such kind of conservative mechanism is useful because it leads to a certain degree
of independence from soil; however, it also results in a reduction in the transfer of
nutrients through litter, a factor which supports an even availability of nutrients and
continuous cycling in the ecosystem (Staaf and Berg 1981; Singh et al. 1984).
During the past few decades, in the studies examining the maintenance of forest
soil fertility, considerable attention is being given on decomposition of leaf litter,
which is an important process of returning nutrients and organic matter to the forest
soils (Swift et al. 1979; Hobbie 1992; Moretto et al. 2001; Xuluc-Tolosa et al. 2003;
Moore et  al. 2006; Sayer 2006). The complex process of litter decomposition is
affected by both biotic (Hättenschwiler et al. 2005) and abiotic factors (Austin and
Vivanco 2006; Weider et al. 2009). Studies have highlighted a close relationship of
decomposition with the carbon and nitrogen contents in the litter (Meentemeyer
1978; Swift et al. 1979; Fog 1988; Kemp et al. 2003). The quality of litter is also
expected to change according to environmental conditions at various spatial scales.
The effect of the quality of litter on decomposition rate is evident from variable rates
of decay of different types of plant tissues. Different species exhibit variations
regarding the release pattern of nutrients, which is particularly associated with qual-
ity of litter, seasonal conditions and ecological factors (Arunachalam et al. 2003;
Abiven et al. 2005). Although the decomposition rate of litter mostly depends on the
quality of litter, biota and microclimate (Seastedt et al. 1983), there is also a sub-
stantial effect of edaphic properties of the habitat (Heneghan et al. 1998).
Tropical and subtropical dry forest life zones (sensu lato Holdridge 1967; the dry
life zone(s)) provide favourable climate for living, agricultural activities and abun-
dance of fuelwood; therefore the human population has been most attracted by
TDFs (Murphy and Lugo 1986, 1995). The human activities have converted these
forests into pastures and agricultural fields and have removed the above-ground
woody biomass unsustainably to satisfy fuelwood demand (Murphy and Lugo 1986;
Maass 1995). The intensive human activities have degraded soils of TDF making it
compact, nutrient poor and eroded (Maass 1995). Further, these human activities
have fragmented the landscapes of dry life zones (Ramjohn et al. 2012).
Biodiversity is extremely important for the survival and economic develop-
ment of human beings and for the stability and functioning of ecosystems (Singh
2002). Political and scientific community is much concerned for the protection
of biodiversity since the species extinction rates are increasing at an alarming
rate mostly due to anthropogenic activities (Ehrlich and Wilson 1991). Many
12 1 Introduction

types of environmental changes determine or affect the processes that can both
enrich and erode diversity (Sheil 1999). According to UNEP (2001), the impor-
tant factors responsible for the biodiversity loss in India are destruction of habi-
tat, over exploitation, pollution and introduction of alien species. The disturbances
mostly due to impact of these factors influence the dynamics of forest, diversity
of tree species and structure of communities at the local and regional scales
(Sumina 1994; Burslem and Whitmore 1999; Hubbell et al. 1999).
Due to biotic pressure and the effect of changing climate, TDFs are degrading at
an alarming rate. Natural resource-based economy in developing countries and the
high rates of human population growth subject tropical landscapes to rapid rates of
land conversion. Moreover, since these forests are easily converted to cattle pasture
by logging and burning, little cover of TDF is now remaining. In many areas, the
remaining mature TDFs are less than 10% of the original extent (Murphy and Lugo
1986; Janzen 1988b; Bullock et al. 1995). Recently, the results from the analyses of
Landsat data reported that the most rapid deforestation rates occurred from the year
2000 to 2012 and, particularly in the TDF regions of Paraguay and Argentina, where
during the past few decades, large-scale expansion of extensive industrialized agri-
cultural has been observed (Gasparri and Grau 2009; Hansen et al. 2013). Primarily,
TDFs are gradually shifting towards agricultural lands, tree plantations or grass-
lands for grazing (Murphy and Lugo 1986). Thus, due to these tremendous rates of
loss, organisms that were once common in these forests now face extinction, merely
for lack of habitat. Newmark et al. (2017) have argued that habitat loss is the most
important cause of species extinction, but such extinctions are often delayed, pro-
viding an opportunity to conserve species through habitat restoration. Further, rapid
regeneration of forest among fragments is important because such regenerating for-
ests will connect fragments and allow immigration of species from source popula-
tions in order to rescue the endangered populations.
Furthermore, because few functioning dry forest ecosystems remain (the forest
is reduced to small, isolated patches in most parts of the world due to fragmenta-
tion), their ecology is poorly studied, and their fauna and flora are far less well
understood compared to the much better-studied rain forests.
In this communication we review important studies across the globe, with a
focus on the northern dry deciduous forests of India, to elucidate the research trends
and the salient features of the structure and functioning of TDFs and the ecosystems
derived from the same.
Chapter 2
Global Coverage, Climate and Soil

Although tropical dry forests (TDFs) are widely distributed, they are considered
among the most threatened ecosystems, the rate of disturbance and deforestation
being much higher in these forests as compared to other tropical biomes.
Physiognomically, these forests range from low scrub on the conspicuously dry
sites to tall forests on mesic sites, the tree canopy ranging from multilayered to
highly discontinuous. Most of the trees remain deciduous during the dry period of
the year, and the deciduousness increases with the declining rainfall. The Indian dry
deciduous forests or monsoon forests, sharing 38.2% of the total forest cover, occur
in both northern India and in South Deccan Plateau, mainly in areas which are warm
year-round, and where annual rainfall ranges from 500 to 1500 mm. These forests
occur on highly weathered, leached and impoverished soils and have developed
mechanisms to conserve nutrients. Soil microbial biomass, in these forests, acts as
a sink as well as source of plant-available nutrients. In this chapter we summarize
the global distribution of TDFs and their environmental setting, both climatic and
edaphic.

2.1  Global Coverage

Tropical dry forests (TDFs) comprise about 46% of tropical forests (Olson et  al.
2001) and are considered among the least protected and most disturbed ecosystems
on the earth (Hoekstra 2005; Janzen 1988a, b). According to FAO (2000, 2008) and
FAO and JRC (2012), tropical forests cover about 30% of the world’s land area and
50% of the world’s forested area which is around 4 billion ha. According to Galicia
et al. (2008), the TDF occupies 42% of the tropical forest area. Broadly, TDF vegeta-
tion is dominated by deciduous trees and occurs in regions with marked seasonality
in precipitation (Murphy and Lugo 1986; Sánchez-Azofeifa et al. 2005; Miles et al.
2006; Pennington et  al. 2006a, b). On the basis of forest type and location along
precipitation gradients, the percentage of deciduous trees in TDFs varies from 50%

© Springer Nature Singapore Pte Ltd. 2017 13


J. S. Singh, R. K. Chaturvedi, Tropical Dry Deciduous Forest: Research Trends
and Emerging Features, https://doi.org/10.1007/978-981-10-7260-4_2
14 2  Global Coverage, Climate and Soil

Fig. 2.1  Global distribution of tropical dry forests (FAO 2000)

to 100% (Medina 2005). The original extent of TDFs around the globe is not known
precisely (Murphy and Lugo 1986; Sánchez-Azofeifa et al. 2005; Pennington et al.
2006a, b); however, these forests have been observed worldwide and are mostly asso-
ciated with savannas (Furley et al. 1992; Pennington et al. 2000). The global extent
of TDFs varies depending on the method of estimation and definition and ranges
from 1 to 7 million km2 (Mayaux et  al. 2005; Miles et  al. 2006; Grainger 1996).
According to the FAO map of the TDF global ecological zone (Fig. 2.1), the largest
areas of TDF are in South America, sub-Saharan Africa and northeast India. Presence
of these forests is also observed throughout Southeast Asia, northern Australia and
parts of the Pacific, Central America and the Caribbean. In India, the share of TDF is
38.2% of the total forest cover (MoEF 1999).
Compared to the rain forest, the physiognomy of TDF is much variable and
ranges from low scrub on the driest sites to tall forests on the most mesic sites
(Pennington et al. 2009). Major biomes that the TDF contact include savanna wood-
land (e.g. the cerrado of Brazil), lowland tropical rain forest (e.g. in the Chiquitano
region of Bolivia) and montane forest (e.g. in inter-Andean valleys) (Pennington
et al. 2009). The TDF vegetation is mostly deciduous during the dry season, and
with the declining rainfall, the degree of deciduousness increases, although more
evergreen and succulent species have been observed in the driest forests compared
to forests occurring on mesic sites (Mooney et al. 1995).
According to Eamus (1999), the tree canopy of seasonally dry forests and wood-
lands ranges from multilayered (forest) to highly discontinuous (savanna). Further,
2.1  Global Coverage 15

with reference to the proportion of deciduous species or stems in the canopy, the
TDFs are sometimes subdivided into deciduous, semi-deciduous, or semi-­evergreen,
but the definitions of these terms are still not clear (Griscom and Ashton 2011). On
a continuum, rain forest and savanna or the other seasonally dry ecosystems repre-
sent points as the two extremes. Savannas and thorn woodlands could be thought of
as TDF variants and by historical fire and disturbance regimes, they have been mod-
ified and assumed new structures and states (Denevan 1992; Cooke and Ranere
1992). Finally, as Ratnam et al. (2016) have argued, the forests and savannahs occur
as a mosaic of alternate states within a landscape (Hirota et al. 2011; Staver et al.
2011; Bond and Parr 2010). Savannas are found under similar or slightly wetter
climates than TDFs, and these two biomes can coexist in close proximity (Pennington
et al. 2009). Although savannas are usually included in the distribution maps of TDF
(Murphy and Lugo 1986; Miles et al. 2006), there are certain ecophysiological dif-
ferences between the two ecosystems (Pennington et al. 2009). While grasses are a
minor element in the ground layer of TDFs (Mooney et  al. 1995), savannas are
characterized by the abundance of xeromorphic, fire-tolerant grasses. Leaves of
savanna trees are frequently sclerophyllous, evergreen owing to nutrient-limited
soils (Ratter et al. 1997). Succulent species lack adaptation to fire; therefore, they
are almost entirely absent from savannas. Savanna plants show clear adaptation to
fire and possess thick, corky bark. They have the ability to root sprout from under-
ground storage and protected buds. Such characteristics demonstrate that in the
ecology and evolution of savannas, fire has been a key factor. However, such adapta-
tions are lacking in TDF plants (although most TDF plants are good sprouters),
which suggests that despite the current anthropogenic fire exposure by humans, fire
could not be considered as an evolutionary force in this biome.
Compared to the land-use/cover change processes in other tropical biomes, the
rate of disturbance and deforestation in TDFs is much higher (Hoekstra et al. 2005).
These forests are generally seen as small fragments, the fragmentation being mostly
natural or by human activities, and therefore, biologists ignore these small remnants
(Pennington et al. 2009). As a result of different socio-economic forces, and being
subjected to different agricultural land uses, the TDF landscape is a mixture of for-
ests undergoing different stages of ecological succession (Cao et al. 2015). Large
areas of TDFs have been cleared for agriculture because both the climate and soils
that support TDFs are also suitable for many types of farming and ranching (Trejo
and Dirzo 2000). The original global extent of TDF is not known; however, it has
been estimated that 48.5% of total TDF area has been converted to other land uses
(Hoekstra et al. 2005).
Pennington et al. (2009) suggested that the Neotropical SDTFs are characterized
by large changes in taxonomic composition with distance, i.e. high β-diversity from
the species to family level (e.g. Castillo-Campos et al. 2008; Gillespie et al. 2000;
Linares-Palomino et al. 2010; Trejo and Dirzo 2002). Although an influential review
by Gentry (1995) emphasized a floristic similarity at the family level, a more hetero-
geneous pattern of diversity is detected even at this high taxonomic level for
Neotropical SDTF (e.g. Linares-Palomino 2006; Lott and Atkinson 2006; De
Queiroz 2006). Despite high levels of β-diversity, there is perhaps a phylogenetic
16 2  Global Coverage, Climate and Soil

integrity to the SDTF floristic elements, e.g. Leguminosae is the most species-rich
family everywhere with the exception of the Caribbean (Lugo et  al. 2006) and
Florida (Gillespie 2006), where Myrtaceae dominates.
Pennington et al. (2009) suggested that Neotropical SDTFs are part of a global
metacommunity of tropical vegetation that experiences erratic water availability.
The tropical Asian forests growing in monsoon climates in Thailand, Burma,
­Indo-­China, the Indian subcontinent and Sri Lanka (see overview of Whitmore
1975), or the TDFs of Australia and the Pacific Islands, could be part of the same
global SDTF metacommunity, as suggested by the climate where they grow, which
matches that of areas where Neotropical SDTF occurs. These Asian dry forests,
however, lack the succulent flora element so characteristic of Neotropical, Afro-
Madagascan and Arabian SDTFs, and many are less rich in Leguminosae (e.g. in
Thailand; Rundel and Boonpragob 1995) than African, Arabian, or Neotropical
forests.
The Indian dry deciduous forests or monsoon forests that occur in both northern
India and in South Deccan Plateau, mainly in areas which are warm year-round, and
where annual rainfall ranges from 500 to 1500 mm (Singh and Chaturvedi 2017)
have species which have their own specific time of leaf flushing. In spite of being
less biologically diverse than rain forests, the Indian dry deciduous forests are home
to a wide variety of wildlife. Some of the most notable animals include monkeys,
large cats, parrots, various rodents and ground-dwelling birds. The mammalian bio-
mass also tends to live in higher numbers in dry forests than in rain forests. Sal is
the most economically significant tree found in the dry deciduous forests. Several
studies have revealed that the dry deciduous forests have replaced and are replacing
the erstwhile moist deciduous forests.
Although the northern dry deciduous forests of India are not remarkably rich in
number of plant species, they do harbour several large vertebrates, which include
the largest and most charismatic carnivore of Asia, the tiger (Panthera tigris). The
known mammal fauna in these forests consists of 68 species, but there are no eco-­
regional endemic species. The threatened species include the tiger, wild dog, sloth
bear and chousingha. These forests are mainly located across the Indian states of
Bihar, Orissa and Madhya Pradesh (Singh and Chaturvedi 2017). They can also be
found in the rain shadow of the Eastern Ghats Mountain Range. The Indian dry
deciduous forests have a three-storied structure, with an upper canopy at 15–25 m,
an understory at 10–15 m and undergrowth at 3–5 m (Singh and Chaturvedi 2017).
The vegetation in these forests is made up of associations of Boswellia serrata,
Anogeissus latifolia, Acacia catechu, Terminalia tomentosa, Terminalia paniculata,
Terminalia belirica, Chloroxylon swietenia, Albizzia amara, Cassia fistula,
Hardwickia binata, Dalbergia latifolia, Pterocarpus marsupium, Stereospermum
suaveolens, Spondias pinnata, Cleistanthus collinus, Acacia lenticularis, Flacourtia
indica, Butea monosperma, etc.
The South Deccan Plateau dry deciduous forests are located on the lee side of the
Western Ghats Mountain Range. The forests in this region extend across the ­southern
Indian states of Karnataka and Tamil Nadu, and their vegetation is highly influenced
by climate (Singh and Chaturvedi 2017). The annual rainfall in these forests ranges
2.2  Climate Variability 17

from 900 to 1500  millimetres (mm). As the tall Western Ghats Mountain Range
intercepts the moisture from the southwest monsoon, the eastern slopes and the
Deccan Plateau receive very little rainfall. The forests are also home to several
important populations of India’s large threatened vertebrates (Singh and Chaturvedi
2017). One of the most important among them is the elephant population that ranges
from the Nilgiri Hills to the Eastern Ghats. Some of the other important species
include wild dog, sloth bear, chousingha, gaur and grizzled giant squirrel.

2.2  Climate Variability

The characteristic biotemperature (average of the Celsius temperatures where veg-


etation growth takes place relative to the annual period) of TDF is greater than
17 °C, and the potential evapotranspiration to precipitation ratio is 1–2 (Holdridge
1967). The precipitation ranges from 500 to 2000 mm per year with little or no pre-
cipitation for 4–6 months (Janzen 1983; Murphy and Lugo 1995; Luttge 1997). A
characteristic feature of these forests is the pronounced seasonality in rainfall with
several months of drought each year (Mooney et al. 1995). According to Sánchez-­
Azofeifa et al. (2005), the dominant trees in TDFs are typically deciduous where the
drought deciduous trees comprise at least 50% of the tree population. The mean
annual temperature is >20 °C (and the mean temperature of the coldest month is
higher than 13 °C, Archibald 1995), total annual precipitation ranges between 700
and 2000 mm, and there are 3 or more dry months (with precipitation <100 mm)
every year, resulting in strong seasonality in ecological processes and functions (see
also Allen et  al. 2017). Rainfall can be bimodal (central and southern Africa) or
unimodal (North Africa, Central America and India) (Archibald 1995). The dry sea-
son can be short (e.g. 1–3 months in humid regions) or long (e.g. 5–8 months in dry
regions) (Johnson and Tothill 1985).
Rainfall seasonality means that soil moisture content peaks in the wet season and
is at its lowest by the end of the dry season (Fig. 2.2a). During the dry season, there
is decline in the tree canopy cover with no understorey grass. The evaporation at the
soil surface tends towards zero and consequently there is increase in atmospheric
water vapour pressure deficit (VPD) through the dry season (i.e. relative humidity,
RH, declines and evaporative demand increases). Between morning and afternoon,
the VPD increases in both the wet and the dry season (Fig. 2.2b). Figures 2.3 and
2.4 show the marked variability in monthly rainfall in the northern tropical decidu-
ous forest of India.
18 2  Global Coverage, Climate and Soil

a 400 20 b 5

350 18

Volumetric soil moisture (%)


4
300 16

14
Rainfall (mm)

LAVPD (kPa)
250 3
200 12

150 10 2

100 8
1
50 6

0 4 0
D J F M A M J J A S O N D D J F M A M J J A S O N D J F
Month Month

Fig. 2.2 (a) Rainfall (bars), with a distinct dry season between the start and end of the wet season.
Once rains have ceased, percentage soil moisture (filled circles), especially in the upper 1  m,
declines substantially because of transpiration by trees and percolation to deeper layers. (b) Leaf-­
to-­air vapour pressure difference (LAVPD) increases from morning (open circles) through to the
afternoon (filled circles) in both wet and dry seasons, but the increase is largest in the dry season
(April–October) (Source: Eamus 1999)

600

500
Average rainfall (mm)

400

300

200

100

0
1 2 3 4 5 6 7 8 9 10 11 12
Months
Fig. 2.3  Average monthly rainfall (with SD) variability in the northern tropical deciduous forest
of India. Data from 1990 to 1998 (J.S. Singh, unpublished)
2.3  Edaphic Features 19

800

700

600
Rainfall (mm)

500

400

300

200

100

0
1 7 13 19 25 31 37 43 49 55 61 67 73 79 85 91 97 103 109
Months

Fig. 2.4  Monthly rainfall variability from January 1990 to December 1998, in the northern tropi-
cal deciduous forests of India (J.S. Singh, unpublished)

2.3  Edaphic Features

Since more than half of the dry forests in the world occur on highly weathered,
leached and impoverished soils, they are required to develop mechanisms to con-
serve nutrients (Sánchez 1976; Jordan 1985; Singh et al. 1989). Further, due to low
humidity, the availability of nutrients is restricted; and for the regeneration, growth
and survival of seedling communities in TDF, availability of soil water is considered
a key factor (Lieberman 1982; Reich and Borchert 1984; Lieberman and Lieberman
1984; Ceccon et al. 2004). Compared to the rain forests, much higher root/shoot
ratio has been observed in TDF plants, which is considered as an adaptation to the
low soil water availability (Bullock 1990). In some TDF species, consistent correla-
tions between certain compositional and structural characteristics of seedlings and
soil properties have been reported (González and Zak 1996; Johnston 1992; Ceccon
et al. 2002). Ceccon et al. (2002) found positive correlations between the number of
individuals and species richness of seedlings with soil phosphorus content. In a TDF
in Chamela, Mexico, studies reported that the fast-growing species required high
availability of light, nutrients and phosphorus and received less or no benefit from
mycorrhizae fungi (Rincon and Huante 1993; Huante et al. 1993, 1995). It has been
reported that these species possessed a forked root pattern (Huante et  al. 1988)
which helped to explore and use areas with high resource availability (Fitter 1985).
Richness and survival of seedling species in a TDF in Yucatán, Mexico, have been
reported to increase after fertilizer application (N and P) (Ceccon et al. 2003), which
20 2  Global Coverage, Climate and Soil

determined the level of success in the establishment of TDF species to a great extent,
conforming to the findings of Vitousek (1984) regarding phosphorus deficiency in
tropical forests.

2.3.1  Spatial Heterogeneity and Availability of Nutrients

Habitat heterogeneity leads to patchy accumulation of nutrients and organic matter


for locally intensified nutrient supply in TDFs occurring on nutrient-poor soils. Roy
and Singh (1994, 1995) quantified the effect of habitat heterogeneity on available
nutrients, active organic matter pool and nutrient supply rates in a northern tropical
dry deciduous forest of India. In this study they demarcated permanent plots on the
hillslope and hilltop and identified microsites in the form of troughs and flats in each
plot. Roy and Singh (1994) observed quite distinct physicochemical and biological
features for the two kinds of microsites (troughs and flats) on the forest floor.
The troughs were richer in soil nutrients and organic matter (Tables 2.1 and 2.2)
and showed greater microbial biomass and microbial nutrients (Table 2.3), higher
rates of nitrogen mineralization (Figs. 2.5a and 2.6), greater accumulation of leaf
litter (Fig. 2.5b), greater concentration of herbaceous shoot and fine roots (Table 2.4,
Figs. 2.5c, d and 2.7) and greater litter decomposition (Table 2.5), compared to the
flats. These microsites richer in nutrients furnish pockets of nutrient reserves in the
otherwise nutrient-poor soil to be exploited by the plants. Microsites richer in avail-
able plant nutrients have been found to be an important source of mineral nutrients
for plants in a variety of environments (Chapin 1980; Bloom et al. 1985; Robinson
and Rorison 1983; Grime et al. 1986). Burke et al. (1989) found highest total and
available N and P in topographic depressions. Roy and Singh (1994) also reported

Table 2.1  Physicochemical properties of the soil from trough and flat microsites on the forest
floor (values are mean ± 1 SE)
Troughs Flats LSD at 5%
Fine particles (silt + clay) (%) 32.0 ± 1.8 22.7 ± 1.4 5.14
Bulk density (g cm−3) 1.27 ± 0.02 1.34 ± 0.02 0.06
Water holding capacity (%) 48.2 ± 4.1 35.4 ± 1.9 10.00
Organic C (%) 2.45 ± 0.28 1.14 ± 0.23 8.81
Total N (%) 0.173 ± 0.02 0.076 ± 0.01 0.055
Total P (%) 0.033 ± 0.002 0.024 ± 0.002 0.007
Total Ca (%) 0.362 ± 0.04 0.332 ± 0.03 0.016
Total K (%) 0.656 ± 0.02 0.497 ± 0.02 0.09
Total Na (%) 0.070 ± 0.01 0.055+0.002 0.016
C:N 14.2 15.0
C:P 74.2 47.3
Source: Roy and Singh (1994)
2.3  Edaphic Features 21

Table 2.2  Seasonal averages of concentrations of available nutrients in soil from troughs and flats
(μg g−1 dry soil ±1 SE)
Hilltop Hillslope
NO3-N Troughs Flats Troughs Flats
Winter 0.94 ± 0.06 0.72 ± 0.05 1.02 ± 0.26 0.77 ± 0.06
Summer 6.00 ± 0.11 3.61 ± 0.10 7.78 ± 0.92 6.01 ± 0.11
Rainy 1.71 ± 0.12 1.25 ± 0.08 1.91 ± 0.66 1.33 ± 0.12
NH4-N
Winter 3.66 ± 0.4 2.42 ± 0.30 4.51 ± 0.50 3.38 ± 0.32
Summer 6.51 ± 0.36 3.86 ± 0.33 10.00 ± 0.59 6.15 ± 0.33
Rainy 2.62 ± 0.44 0.83 ± 0.80 4.90 ± 0.36 3.26 ± 0.55
PO4-P
Winter 2.48 ± 0.28 1.43 ± 0.12 4.70 ± 0.51 3.40 ± 0.28
Summer 3.84 ± 0.35 2.55 ± 0.12 6.00 ± 0.23 4.05 ± 0.40
Rainy 1.80 ± 0.17 1.18 ± 0.10 3.14 ± 0.19 1.96 ± 0.21
Source: Roy and Singh (1995)

Table 2.3  Microbial C, N and P in soil as amounts measured in post-fumigation flush (μg g−1 soil
±1 SE)
Microsites Winter Summer Rainy
Microbial C
Flats 248 ± 21 417 ± 23 103 ± 10
Troughs 352 ± 22 548 ± 39 142 ± 7
Microbial N
Flats 32 ± 3 50 ± 3 10 ± 1
Troughs 48 ± 4 69 ± 4 15 ± 1
Microbial P
Flats 11 ± 1 19 ± 2 4±1
Troughs 19 ± 2 28 ± 2 5±1
Source: Roy and Singh (1994)

that the fine lateral roots of trees, which constituted a majority of the root biomass,
are attracted and proliferate to a greater extent in the troughs which exhibit greater
N-mineralization rate than flats. A suite of potential mechanisms possessed by
plants has been known for exploiting the heterogeneity of the soil, which includes
proliferation of roots and change in nutrient uptake kinetics (Drew and Saker 1975;
Jackson and Caldwell 1993). At a rich microsite, the greater rooting density short-
ens diffusion pathways leading to more rapid uptake of nutrients compared to situ-
ations of widely separated roots (Barley 1970). Jackson et al. (1990) reported large
and rapid increase in the uptake kinetics of plant roots after creating nutrient-rich
soil patches in the field condition. Gibson (1988) argued that the presence of active
lateral roots in the hollows suggests a mechanism of nutrient redistribution whereby
the plants on the hummocks could be exploiting nutrients from the hollows.
22 2  Global Coverage, Climate and Soil

50 a

40 1200 b
NITROGEN MINERALIZATION

1000

LITTER ACCUMULATION
30
800
µg g–1 mo–1

20 600

gm–2
400
10
200

0 0

300 c

600
d
HERBACEOUS SHOOT

200 500
FINE ROOT BIOMASS

400
gm–2

gm–2

300
100
200

100

0 0
S O N D J F M A M J J A S O N D J F M A M J J A
MONTH MONTH

Fig. 2.5  Mean monthly (a) N-mineralization rate, (b) standing crop of litter, (c) herbaceous shoot
biomass and (d) fine root biomass (live + dead) in the troughs (○) and flats (∆) on the forest floor.
Bars represent ±1 SE (Source: Roy and Singh 1994)

Existence of such spatial heterogeneity in the nutrient-poor dry forest has been
claimed as a nutrient conserving mechanism.
Roy and Singh (1995) reported that the troughs where the proportion of ammo-
nium was greater than the nitrate showed greater fine root production as well as
turnover than the flats. Roy and Singh (1995) suggested that the two forms of nitro-
gen dominate in two different parts of the year (wet season, NO3− domination; dry
season, NH4+ domination), and irrespective of the form of N present, fine root
­development is facilitated by the greater availability of N. As a result of enzymatic
­digestion, the more complex proteins and allied compounds of the organic matter
are simplified and hydrolysed to NH4+. These authors observed a sharp increase in
the rate of N-mineralization at the start of the rainy season in all the microsites with
simultaneous reduction of litter mass. Sorensen (1974) and Birch (1958, 1960)
2.3  Edaphic Features 23

Fig. 2.6 N-mineralization 40 HILL SLOPE


rate in the troughs and flats
of the hilltop and the
hillslope sites in different 30
months. (Bars represent ±1
SE). (○) trough; (∆) flat
(Source: Roy and Singh 20
1995)

N mineralization rate (µg g–1 month–1)


10

HILL TOP
50

40

30

20

10

0
S O N D J F M A M J J A

Month

Table 2.4  Seasonal averages of total fine root biomass in patchy and adjacent microsites (g m−2 ±
1SE)
Hilltop Hillslope
Troughs Flats Troughs Flats
Winter 249 ± 25 153 ± 17 436 ± 43 165 ± 24
Summer 174 ± 22 113 ± 16 273 ± 49 214 ± 33
Rainy 374 ± 49 147 ± 19 572 ± 38 291 ± 22
Source: Roy and Singh (1995)
24 2  Global Coverage, Climate and Soil

Fig. 2.7  Fine root biomass 600 HILL TOP


(live + dead) in the troughs
and flats of the hilltop and
hillslope sites in different 400
months. (Bars represent ±1

Fine root biomass (g m–2)


SE). (○) trough; (∆) flat
200
(Source: Roy and Singh
1995)
0
700 HILL SLOPE
600

400

200

0
S O N D J F M A M J J A
Month

Table 2.5  Dry matter loss of Troughs Flats


enclosed leaf litter placed in
Mean relative decomposition (mg 9.69 ± 0.08 9.38 ± 0.15
flats and troughs on the forest
g−1 day−1)
floor
Instantaneous decay rate (annual) 2.26 ± 0.11 1.93 ± 0.13
Time for 50% decomposition (days) 113 ± 6.35 133 ± 9.8
Time for 95% decomposition (days) 488 ± 28.3 576 ± 43
Source: Roy and Singh (1994)

reported that N-mineralization increases when dry soil is rewetted with water and
that the extent of flush of N-mineralization in East African soils once rewetted was
correlated with the extent of dry season and predicted that maximum mineralization
occurred during the rainy season. The period of increased N-mineralization coin-
cides with the active plant growth period (Roy and Singh 1995).

2.3.2  Soil Microbial Biomass and Ecosystem Functioning

Singh et al. (1989) investigated the factors affecting plant growth in nutrient-poor
ecosystems of TDFs of India. Their study reported that the contents of extractable
ammonium-N and phosphate-P were low and varied considerably within the annual
cycle. In forest soils, nitrate-N varied from 0.4 to 5.6 μg g−1 dry soil and in savanna
soils from 0.4 to 0.8 μg g−1. Values for ammonium-N were in the ranges 1.2–4.0 and
2.0–5.0  μg g−1 for forest and savanna sites, respectively. The contents of
2.3  Edaphic Features 25

Fig. 2.8  Nitrification in 22


tropical dry forest and
20
savanna sites. ●-·-●.
forest; ○─○. grass 18

Nitrification (µg g–1 dry soil 30d–1)


savanna; and ○---○. 16
bamboo savanna (bars.
1s.e.) (Source: Singh et al. 14
1989) 12
10
8
6
4
2
0
–2
–4
Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr May
Months

Table 2.6  Microbial biomass C, N and P in dry tropical ecosystems


Season C N P
Forest
Rainy 487 ± 9.2 51 ± 0.3 20 ± 1.7
Winter 662 ± 14.8 70 ± 1.2 29 ± 3.3
Summer 744 ± 10.3 88 ± 4.1 31 ± 1.6
Savanna
Rainy 262 ± 7.5 31 ± 0.2 11 ± 1.5
Winter 400 ± 7.5 32 ± 0.7 19 ± 1.6
Summer 520 ± 9.4 46 ± 1.7 23 ± 1.8
Biomass C, N and P contents given as μg g−1 dry soil. Differences due to seasons were significant
at P < 0.001 for both the forest and savanna ecosystems
Source: Singh et al. (1989)

phosphate-­P extracted using NaHCO3 varied from 1.4 to 4.8 μg g−1 for forest soils
and from 2.0 to 4.0 μg g−1 for savanna soils; the highest levels of phosphate-P as
well as inorganic-­N occurred in the summer. These authors reported a marked sea-
sonality in the rate of nitrification which peaked during the rainy season (Fig. 2.8);
all soils showed little or no net nitrification during the dry period. The peak nitrate
production rates in these ecosystems were lower than the minimum rates reported
for tropical rain forest sites in Brazil, Costa Rica and Panama (Vitousek and Matson
1988). Low nutrient pools and low rates of nitrification observed by Singh et  al.
(1989) show that the TDF and savanna systems are extremely nutrient poor. The
highest and the lowest levels of microbial biomass N occurred in the dry and rainy
season, respectively (Table 2.6), and the range of values measured (31–88 μg g−1)
26 2  Global Coverage, Climate and Soil

fell within that reported for the Amazonian rain forest sites (Livingston et al. 1988).
Microbial immobilization of 15N is reported to be greater at sites where the rates of
nitrification are low than at sites where they are high (Vitousek and Matson 1988).
Furthermore, the activity of soil microbes is less sensitive to soil water potential
than is water uptake by the plants (Calder 1957; Semb and Robinson 1969), and the
substantial amount of water which is possibly present during summer at high ten-
sion becomes available to microbes but remains unavailable to plants (Sanchez
1976). It seems, therefore, that microbial activity in these systems continues during
dry months but that plant growth is drastically reduced, curtailing the demand for
nutrients.
Singh et al. (1989) argued that mineralization products and nutrients that move
upward through capillary action in the dry season are immobilized in the microbial
biomass. In the TDF and savanna ecosystems, the microbial biomass and nutrient
pools declined as N-mineralization increased in the rainy season when plant growth
was most rapid. Evidently the N immobilized in the microbial biomass (dead micro-
bial cells, microbial metabolites) is an important source of N for plants (Cameron and
Posner 1979). Grazing of microbes by soil fauna such as protozoa and nematodes
results in the enhanced turnover of microbes in the rainy season. It is known that there
is a tremendous increase in the populations of fungivorous fauna and protozoa in the
rainy season (Dash and Guru 1980; Singh and Mukherjee 1986). Coleman et  al.
(1977) and Clarholm (1985) have argued that grazing of bacteria is necessary to
release nutrients for plant uptake. The turnover of N from dead microbial cells is
about five times that from native soil organic N (Marumoto et al. 1977; Stevenson
1986). It has been reported that in low-N soils, microbial biomass is as effective as
nitrate-N in supplying N to wheat (Lathbridge and Davidson 1983). Singh et  al.
(1989) have concluded that microbial biomass in these nutrient-poor systems acts as
a sink and a source of nutrients. Thus, the principal function of the microbial biomass
during the dry period (high biomass, low turnover) is to conserve and accumulate
nutrients in a biologically active form, when the plant activity is low and their effi-
ciency of nutrient extraction from soil is low, and then during the monsoon period
(low biomass, high turnover) to release them rapidly to initiate plant growth.
The first wet season rains initiate the decomposition of litter accumulated on the
soil surface during the dry period (Cornejo et al. 1994), and according to Van veen
et al. (1984) and Cabrera (1993), drying-rewetting cycles in the seasonally dry habi-
tats enhance the replenishment of available soil N pools from microbial, recalcitrant
or physically protected N pools. In seasonally dry Indian (Singh et al. 1989) and
Mexican forests (Davidson et al. 1993), rewetting of dry soils resulted in a large
pulse of N-mineralization attributed primarily to the lysis of soil microbial biomass.
Thus, plants provide carbon to the microbial community, and the latter mineralizes
the plant and animal residues and soil organic matter to make nutrients available for
plant growth.
In the TDF ecosystems, immobilization and release of N are governed by micro-
relief and seasonality (Roy and Singh 1994, 1995; Singh et al. 1989). During nitri-
fication, the first step, conversion of ammonium to nitrite, is mediated by
2.3  Edaphic Features 27

ammonia-oxidizing bacteria, and the conversion of nitrite to nitrate is brought about


by nitrite-oxidizing bacteria; both groups are gram-negative chemoautotrophic bac-
teria belonging to Nitrobacteraceae (Schmidt 1982; Watson et al. 1981). Sabey et al.
(1959) argue that the size of the nitrifier population could be an important factor
controlling nitrification, and the quantification of nitrifier population could yield
important information on soil nitrification process and rate (Belser 1979). Jha et al.
(1996a) compared the N-mineralization rates and the size of the viable community
of nitrifying bacteria for a dry forest site and an adjoining cropland site. These
authors observed a strong positive relationship between N-mineralization and soil
moisture, which indicated that the size of the mineralization process was moisture
limited. The activity of nitrifiers was inhibited more than that of ammonifiers at soil
moisture potentials below −1.5 MPa (Dommergues 1966). At soil water potentials
greater than −0.6 MPa, the activity of nitrifying bacteria is limited by the substrate
availability, whereas at water potentials of less than −0.6 MPa, cell dehydration is
the more important inhibiting factor (Stark and Firestone 1995). Soil moisture
affects the activity of nitrifying bacteria through both dehydration and substrate
limitation, and this sensitivity of nitrifiers to soil moisture is related to their high
energy requirement (Sprent 1987). In the study of Jha et  al. (1996a), the MPN
counts of ammonia- and nitrite-oxidizing bacteria were largest in rainy season and
smallest in the dry season and were significantly and positively related to soil mois-
ture. Vitousek and Matson (1982) found that the application of actively nitrifying
soil as an inoculum to soils before incubation increased the production of nitrate
and, hence, they suggested that nitrification is regulated by nitrifier populations.
Moreover, size of the viable community of nitrifiers reported by Jha et al. (1996a)
was positively related with nitrification as well as N-mineralization rates.
Campo and Merino (2016) found significant variation in organic C and microbial
biomass and N concentration in the soils of seasonally dry forest of Yucatan
Peninsula. Black soils were characterized by greater C and N concentrations than
thick clayey red soils and there occurred a decreasing gradient of differences
between soils in microbial biomass C concentrations with increased mean annual
precipitation. Nevertheless, microbial biomass C, soil organic C and soil N all were
highest in the drier site compared to intermediate or subhumid site. Campo and
Merino (2016) concluded that in the long run, reduced rainfall in the seasonally dry
forest of Yucatan Peninsula would result in increased organic layer and mineral soil
C storage, due to lower decomposition and higher chemical recalcitrance of organic
matter. Thus, under the predicted longer drought periods, these seasonally TDFs
could be significant soil C sinks.

2.3.3  Carbon Balance and CO2 Flux

In the phenomenon known as “Birch effect”, large CO2 pulses are released when the
dry soils are rewetted (Birch 1958). In arid or seasonally dry ecosystems, these CO2
pulses may contribute substantially to the annual soil CO2 flux (Schimel et al. 2007;
28 2  Global Coverage, Climate and Soil

Wang et al. 2015). Therefore, relatively small alteration in the quantity or timing of
precipitation regime may have a large influence on ecosystem carbon balance in
seasonal forests. The size of rewetting CO2 pulses is apparently determined by the
magnitude of the rewetting event (i.e. soil water potential change; Lado-Monserrat
et al. 2014), the frequency of rewetting events (Fierer and Schimel 2003) and the
time gap between each precipitation pulse (Xiang et al. 2008). According to Moyano
et al. (2012), these climatic influences may be partly affected by soil texture, as the
clay-rich soils have been observed to attenuate the responses to rewetting events
(Harrison-Kirk et al. 2014). Ultimately, impact of the soil microbial community is
the most proximate control for the C mineralization rates in soils; it is supposed that
the Birch effect starts when the microbial osmolytes are released immediately after
rewetting and/or when organic substrates released by alteration in soil structure are
rapidly metabolized (Unger et  al. 2010; Jenerette and Chatterjee 2012; Moyano
et al. 2013). These drying-rewetting cycles may have strong impact over the biogeo-
chemical cycles in the TDF ecosystems, as nutrients which are immobilized during
the dry season in microbial biomass are mineralized rapidly at the start of the wet
season, meeting the plant nutrient demand (Singh et al. 1989). The complexity of
the interactions among the seasonality of precipitation, microbial biomass dynamics
and plant growth is complex and yields a nonlinear relationship between rainfall
and C storage (Rohr et al. 2013). Therefore, under the future climate change sce-
nario, prediction of the C cycle dynamics in TDF is difficult by the ecosystem mod-
elling methods. Waring et  al. (2016) argued that at the fine spatial scale, TDF
exhibits extreme heterogeneity in the biogeochemical pools and fluxes; therefore,
the extrapolation of the pattern from local to regional scales could become more
complicated. To understand the relationships between precipitation and soil C
cycling in TDF, Waring and Powers (2016) investigated seasonal transitions of soil
respiration in a regenerating TDF in Guanacaste, Northwest Costa Rica. In this
experiment they tried to determine the factor controlling the magnitude of CO2
fluxes with changes in rainfall pattern, and they also tried to represent these dynam-
ics in predictive ecosystem models. Other studies in TDF have shown that microbial
N utilization is strongly dependent on the availability of dissolved organic carbon
(DOC) (Montaño et al. 2007).

2.3.4  Carbon Balance and Global Change

The quantity and seasonality of precipitation are already changing over many areas
of the dry tropics (Feng et al. 2013), which highlights the significance of a mecha-
nistic framework for the prediction of soil responses to changes in precipitation
regime. In both the enzyme-catalysed and modified conventional models, when all
other factors are considered constant, maximum changes in soil water potential
(ΔWP) resulted in larger CO2 pulses leading to greater total losses of CO2, as has
been reported in empirical studies (Lado-Monserrat et al. 2014). This suggests that
when the soil becomes very dry in more severe droughts, Birch effect may become
2.3  Edaphic Features 29

intense, leading to net ecosystem C losses. The models also predict the attenuation
of the size of CO2 pulse due to increased number of wetting events. Therefore,
belowground C storage may be affected by the timing of precipitation, even if the
total amount of rainfall remains the same. It has also been found that larger but more
sporadic precipitation events lead to more CO2 loss compared to the smaller but
more frequent ones. It has been anticipated that if trees alter the belowground allo-
cation in response to major rewetting events (Doughty et al. 2014), the change may
be observed in the relative sizes of DOC and microbial biomass pools. Experimental
observations and model simulations help to highlight the complexity of the pro-
cesses controlling belowground carbon cycling in TDF (Waring and Powers 2016).
In order to predict ecosystem-scale responses to climate change, more experiments
are needed to explore relationships among microbial biomass dynamics, DOC
availability and belowground C cycling.
In conclusion, TDF is widely distributed and experiences rather harsh climate
with several dry months within the annual cycle and has to cope up with also
nutrient-­poor soil. These conditions have prompted the vegetation to develop adap-
tational features to allow them to survive and grow. As an adaptation to the low soil
water availability, TDF species possess high root/shoot ratio. The study in Indian
TDF revealed that habitat heterogeneity in the TDF results into the patchy distribu-
tion of nutrients. The “trough” microsites at the ground surface accumulate litter
and consequently trap the nutrients. These troughs are the patches of intense bio-
logical activity, and possess higher amounts of organic C, microbial biomass, total
and mineral-N and available P, and consequently have greater nutrient supply poten-
tial than “flat” microsites. The trough microsites also support higher amounts of fine
roots which mine the nutrients for tree growth, in the otherwise nutrient-poor
­habitat. It has also been reported that during the dry period, soil nutrients are con-
served and accumulated in microbial biomass in a biologically active form, and
during the monsoon period, these conserved nutrients are released rapidly to sup-
port the initial plant growth.
Chapter 3
Vegetation Attributes

Although the tropical dry forests (TDFs) are substantially species poor compared to
the rain forests, they contain a large number of endemic and rare species and are
characterized by an immense phenological diversity, as they harbour deciduous,
brevi-deciduous and evergreen species. Leaf flushing in the TDF begins with the
rise of temperature and peaks in the hottest and driest month of the year. The dry
season leaf flushing permits the renovation of the canopy before the monsoon rain-
fall begins, so that the plants are able to take full advantage of the short rainy season
for primary production and growth. Distribution of plant communities and their
species and structural diversity is highly affected by soil water and soil nutrient
status. The tropical dry deciduous forests of India comprise two subgroups, south-
ern and northern. The southern subgroup comprises climax types, (1) teak forests,
(2) red sander forests and mixed forests, while the northern subgroup comprises two
types, (1) sal-bearing forests and (2) mixed forests without sal. In these forests there
is a patchy distribution of communities, and soil is an important determinant of the
distribution of the patches. In this chapter, we provide an overview of the structure,
and the distribution and diversity of TDF vegetation. We emphasize the seasonality,
phenology, deciduousness, leaf attributes and growth rates, and discuss features
such as strategies against drought.

3.1  The Tropical Dry Forest

The TDFs contain lesser number of species than rain forests, but the structural
and physiological diversity in life forms is conspicuously greater compared to
the rain forests (Mooney et  al. 1995). The majority of woody species in the
TDF exhibit drought deciduousness as a response to the long dry period in the
annual cycle (Frankie et  al. 1974; Bullock and Solís-Magallanes 1990; Lobo
et  al. 2003). However, both the number of species and the number of

© Springer Nature Singapore Pte Ltd. 2017 31


J. S. Singh, R. K. Chaturvedi, Tropical Dry Deciduous Forest: Research Trends
and Emerging Features, https://doi.org/10.1007/978-981-10-7260-4_3
32 3  Vegetation Attributes

individuals showing drought deciduousness vary, and a mix of deciduous and


evergreen species results in a unique phenological complexity in the TDF
(Burnham 1997).
In contrast to the general trend of increasing species richness towards equa-
tor (Fischer 1960; Gentry 1988; Chazdon and Denslow 2002), the most diverse
Neotropical TDFs are the farthest from the equator in Mexico and South
America (Argentina, Brazil, Bolivia and Paraguay) (Gentry 1995). Further, the
driest areas in TDF are the most species rich, for example, a greater diversity
has been found in samples from Chamela, Mexico and Quiapaca, Bolivia, than
in samples from Guanacaste, Costa Rica, which receives more rainfall (Gentry
1995; Lobo et al. 2003). The species richness of Mexican TDFs is associated
with an outstanding rate of species turnover among sites along with elevated
number of endemic and rare species (Trejo and Dirzo 2002; Gordon et al. 2004;
Trejo 2005).
A characteristic feature of TDFs is high floristic endemism, 43% in Mexico
(Rzedowski 1991) to 73% regional endemism in South American TDFs (Gentry
1982). Joly (1970) has estimated that in the cerrado region of the Brazilian dry
forest, up to 70% of all species of certain families may be endemic. The num-
ber of species in semi-deciduous forests (50–70) is greater than in tropical
deciduous forests (Gentry 1995) where species richness is significantly lower
due to human activity, such as burning, grazing and firewood collection
(Chaturvedi 2010). Forest structure and species diversity have been widely
studied in various regions of the tropics (e.g. Aiba and Kitayama 1999;
Ayyappan and Parthasarathy 1999; Gillespie et  al. 2000; Swamy et  al. 2000;
Huang et al. 2003; Brearley et al. 2004). Gentry (1988) recorded 275–283 tree
species ha−1 for trees ≥10 cm in diameter at breast height (DBH) at Yanamono
and Mishana near Iquitos, Peru. Valencia et al. (1994) enumerated 307 species
ha−1 for the same DBH in Amazonian Ecuador. These figures formed the
world’s highest record of tree species on a hectare basis for stems ≥10 cm DBH
(Ayyappan and Parthasarathy 1999).

3.1.1  Tropical Dry Forest Vegetation of India

The greater part of the rain in the dry deciduous region of India (in Uttar Pradesh,
Bihar, Orissa, Punjab, Haryana, Madhya Pradesh, Gujarat, Maharashtra, Madhya
Pradesh, Karnataka and Tamil Nadu (Fig. 3.1)) falls during July to September. The
maximum summer temperature may be as high as 48 °C and the minimum −2.2 to
6.1 °C. The dry forest once covered more than half of the area (about 170 million
ha) of the country. However, on account of population pressure, the dry forests now
cover only around 26–27 million ha.
3.1  The Tropical Dry Forest 33

Fig. 3.1  Potential distribution of tropical dry deciduous forest in India. The contour lines repre-
sent annual rainfall in centimetre (Source: Singh and Singh 2011)

The TDFs in India are classified into (i) dry deciduous forests, characterized by
abundance of shrubs and open canopy of small trees, which experience about
6 months of dry period in the annual cycle; (ii) thorn forests which experience more
than 6 months of dry period each year and are characterized by the sparse distribu-
tion of small, mostly thorny trees with abundance of shrubs; and (iii) dry evergreen
forests experiencing high temperature and small rainfall during the summer season
(Table 3.1). Bamboos are absent from these forests, but grasses and small trees are
abundant.
Table 3.1  Diversity of tropical dry forest ecosystems in India: forest types, their dominant species and distribution (Based on Champion and Seth 1968)
34

Major Group (characteristic Area %


group features) (mha) area Subgroups and their categories Occurrence
Tropical 5—tropical dry deciduous 29.4 38.2 5A—Southern tropical dry deciduous forests [C1, dry Maharashtra, Madhya Pradesh, Andhra
dry forests (entirely deciduous teak-bearing forest: (1a) very dry teak forest, (1b) dry teak Pradesh, Madras, Mysore
forests or nearly so; rainfall forest; C2, Dry red sanders-bearing forest (Pterocarpus
800–1200 mm; santalinus); C3, Southern dry mixed deciduous forest
temperature 25–32 °C; (Terminalia tomentosa)]
some forests bear Tectona 5B—Northern tropical dry deciduous forests [C1, dry Punjab, Uttar Pradesh, Bihar, Orissa,
grandis or Shorea robusta; sal-bearing forest: (1a) dry Siwalik sal forest, (1b) dry plain West Bengal
others are devoid of these sal forest, (1c) dry peninsular sal forest; C2, Northern dry
species and have species of mixed deciduous forest (Acacia catechu)]
Anogeissus and Degradation stages of tropical dry deciduous forests [DS1, Maharashtra, Madhya Pradesh, Andhra
Terminalia, etc.) dry deciduous scrub (Nyctanthes arbortristis); DS2, dry Pradesh, Madras, Mysore, Punjab, Uttar
savanna forest (Gardenia turgida); DS3, (Euphorbia scrub) Pradesh, Bihar, Orissa, West Bengal
(Euphorbia neriifolia); DS4 (dry grassland) (Sehima
nervosum)]
General edaphic types of dry deciduous forests [E1, Uttar Pradesh, Rajasthan, Madhya
Anogeissus pendula forest; DS1, Anogeissus pendula scrub; Pradesh, Maharashtra, Andhra Pradesh,
E2, Boswellia forest (Boswellia serrata); E3, Babul forest Punjab, Bihar, Kerala, Gujarat, Mysore,
(Acacia arabica); E4, Hardwickia forest (Hardwickia Madras
binata); E5, Butea forest (Butea monosperma); E6, Aegle
forest (Aegle marmelos); E7, laterite thorn forest (Gardenia
spp.); E8, saline/alkaline scrub savanna: (8a) Phoenix
savannah (Phoenix sylvestris), (8b) Babul savannah (Acacia
arabica), (8c) Salvadora-Tamarix scrub; E9, dry bamboo
brake (Dendrocalamus strictus)]
General seral types of dry deciduous forests [1S1, dry Rajasthan, Maharashtra, Punjab, Uttar
tropical riverine forest (Terminalia arjuna); 1S2, Khair- Pradesh, Bihar, West Bengal, Assam
sissoo forest (Acacia catechu); 1S3, Inundation babul forest
(Acacia arabica); 2S1, secondary dry deciduous forest
3  Vegetation Attributes

(Prosopis spicigera)]
Major Group (characteristic Area %
group features) (mha) area Subgroups and their categories Occurrence
6—tropical thorn forests 5.2 6.7 6A—Southern tropical thorn forests [C1, Southern thorn Madhya Pradesh, Maharashtra, Madras,
(deciduous with low forest (Acacia leucophloea); C2, Karnatak umbrella thorn Mysore
thorny trees, rainfall forest (Acacia planifrons); DS1, Southern thorn scrub
200–800 mm, temperature (Ziziphus xylopyrus); DS2, Southern Euphorbia scrub
27–30 °C, Acacia, (Euphorbia tirucalli)]
Balanites, Prosopis, 6B—Northern tropical thorn forests [C1, desert thorn Punjab, Uttar Pradesh, Madhya Pradesh,
Salvadora, etc., forest (Prosopis spicigera); C2, ravine thorn forest (Acacia Gujarat, Rajasthan
predominate) leucophloea); DS1, Ziziphus scrub (Acacia leucophloea);
DS2, tropical Euphorbia scrub (Euphorbia neriifolia)]
3.1  The Tropical Dry Forest

General edaphic, degraded and seral types of moist Rajasthan, Gujarat, Punjab
deciduous forests [E1 (Euphorbia scrub), Euphorbia
neriifolia; E2, Acacia senegal forest; E3, Rann saline thorn
forest (Prosopis juliflora); E4, Salvadora scrub (Salvadora
oleoides); DS1, Cassia auriculata scrub; 1S1, desert dune
scrub (Prosopis spicigera)]
7—tropical dry evergreen 0.1 0.1 [C1, Tropical dry evergreen forest (Manilkara hexandra); Andhra Pradesh, Madras
forests (hard-leaved DS1, tropical dry evergreen scrub (Memecylon edule)]
evergreen trees, rainfall
870–1200 mm, dominated
by Manilkara, Memecylon,
etc.)
35
36 3  Vegetation Attributes

Under the tropical dry deciduous forest formation of India, two subgroups,
southern and northern, are clearly distinguished (Champion and Seth 1968)
(Table 3.2). The southern subgroup comprises climax types: (1) teak (Tectona gran-
dis) forests, (2) red sander (Pterocarpus santalinus) forests and (3) mixed forests.
Tectona grandis, Anogeissus latifolia, Diospyros melanoxylon, Boswellia serrata,
Emblica officinalis, Acacia leucophloea, Bridelia retusa, Wrightia tinctoria,
Pterocarpus marsupium, etc. are the most common species in the southern types.
Besides, Tectona grandis, Santalum album and Pterocarpus santalinus are the other
economically important species in these forests. Dendrocalamus strictus is the most
common bamboo. The northern subgroup comprises two types (1) Shorea robusta-
bearing forests and (2) mixed forests without Shorea robusta. Anogeissus latifolia,
Buchanania lanzan, Terminalia tomentosa, Emblica officinalis and Lannea coro-
mandelica are the main associates of Shorea robusta. In the mixed type, species
composition is essentially the same, except that Shorea robusta is absent. Several
edaphic climax types and seral stages have been recognized within the two sub-
groups (Table 3.2), and these are described in detail by Champion and Seth (1968).

Table 3.2  Tropical dry deciduous forest types of India (Based on Champion and Seth 1968)
Southern tropical dry deciduous Northern tropical dry deciduous forests
forests
Dry teak-bearing forest Dry sal-bearing forest
Dry teak forest Dry Siwalik sal forest
Very dry teak forest Dry plain sal forest
Dry red sanders-bearing forest Dry peninsular sal forest
Southern dry mixed deciduous Northern dry mixed deciduous forest
forest
Degradation stages Edaphic types Seral types
Dry deciduous scrub Anogeissus pendula forest Dry tropical riverain forest
Dry savanna forest Anogeissus pendula scrub Khair-sissoo (Acacia catechu)
(Dalbergia sissoo) forest
Dry grassland Boswellia forest Inundation babul (Acacia
nilotica) forest
Euphorbia scrub Babul (Acacia nilotica) Secondary dry deciduous forest
forest
Hardwickia forest
Butea forest
Aegle forest
Laterite thorn forest
Saline/alkaline scrub
savanna
Phoenix savanna
Babul (Acacia nilotica)
savanna
Salvadora-Tamarix scrub
Dry bamboo brake
Degradation stages and edaphic and seral types occur in both southern and northern areas
3.1  The Tropical Dry Forest 37

Sagar and Singh (2005) examined the pattern of tree species diversity and forest
structure in the Vindhyan region of the northern tropical dry forest of India. They
collected data from a total of 1500 quadrats distributed over five 3 ha permanent
plots in five sites, namely, Hathinala, Khatabaran, Majhauli, Bhawani Katariya and
Kota, in the years 1998–2000 (Fig. 3.2). Later Sagar and Singh (2006) examined the
relationships between forest basal area and diversity components (number of spe-
cies and evenness). These studies are summarized below.
Sagar and Singh (2005) reported a total of 65 species with 136,983 stems
(Table 3.3). The number of species and number of individuals varied from 12 to 50
and 8063–65,332 per site. Table 3.4 shows the total number of species in different
categories (established seedlings, saplings and adults) at the five study sites (Sagar
and Singh 2005). In the same study region, the average number of species ha−1
reported by Sagar and Singh (2006) varied from 4 to 23 (Table 3.5), which was rela-
tively low compared to 30–50 tree species ha−1 in dry insular forests in south-­western

Fig. 3.2  Location of the study area. Numbers 1, 2, 3, 4 and 5, respectively, indicate the approxi-
mate locations of the Hathinala, Khatabaran, Majhauli, Bhawani Katariya and Kota sites in the
Vindhyan tropical dry forest of India (Source: Sagar and Singh 2005)
38 3  Vegetation Attributes

Table 3.3  Summary of stem inventory in different stages from five 3 ha permanent plots in the
tropical dry forest of India
Site Seedlings Saplings Adults Total
Hathinala 60,900 3175 1257 65,332
Khatabaran 5975 1250 838 8063
Majhauli 19,650 6525 1187 27,362
Bhawani Katariya 8500 2125 646 11,271
Kota 21,125 3725 105 24,955
Total 116,150 16,800 4033 136,983
Source: Sagar and Singh (2005)

Table 3.4  Total number of species in different life cycle stages on the five sites
Bhawani Total for five
Category Hathinala Khatabaran Majhauli Katariya Kota sites
Adult 31 30 23 22 7 49
(≥9.6 cm)
Sapling 22 8 18 17 6 36
Seedling 47 17 27 19 7 57
Only as adult 3 15 5 4 3 7
Only as sapling 0 0 0 0 1 0
Only as 16 2 9 1 2 14
seedling
Source: Sagar and Singh (2005)

Table 3.5  Vegetation structure of the five sites in a tropical dry forest of northern India
Vegetation Bhawani
parameters Hathinala Khatabaran Majhauli Katariya Kota
Number of species 23 (2.89) 22 (0.00) 18 (0.67) 17 (1.73) 4 (0.33)
(ha−1)
Basal area (m2 ha−1) 8.50 (1.54) 13.78 (2.10) 8.72 (0.66) 6.17 (0.72) 1.31 (0.56)
Stem density (ha−1) 419.00 279.33 395.67 215.33 35.00
(64.01) (50.61) (17.68) (23.13) (10.69)
CV of basal area 60.01 (4.75) 83.97 (8.34) 131.92 98.55 (4.86) 244.18
(66.40) (14.10)
Fisher’s alpha 5.25 (0.66) 5.70 (0.33) 3.98 (0.17) 4.23 (0.13) 1.08 (0.04)
diversity
Whittaker’s index of 5.20 (0.66) 4.93 (1.41) 3.93 (0.29) 3.80 (0.265) 1.16 (0.06)
evenness
Values are mean of three 1 ha plots at each site, and values in parentheses are ±1SE
Source: Sagar and Singh (2006)

Puerto Rico (Murphy and Lugo 1986). The mean stem density ha−1 in the study of
Sagar and Singh (2006) ranged from 35 to 419 as compared to 294–559 trees ha−1
(Jha and Singh 1990) and 315–494 trees ha−1 (Pandey and Shukla 2001) in other
seasonally dry forests of India. Thus, the TDF studied by Sagar and Singh (2006)
was relatively poor in species richness as well as in tree stem density as compared
to other forests.
3.1  The Tropical Dry Forest 39

The adults (individuals ≥9.6 cm DBH) were enumerated in 300, 10 × 10 m quad-
rats, and established seedlings (≥30 cm height but <3.2 cm diameter) and saplings
(≥3.2 to <9.6 cm DBH) were enumerated in 300, 2 × 2 m quadrats at each of the five
TDF sites. The established seedling and sapling stems were scaled up in same unit
as adults.
Sagar and Singh (2006) recorded the average basal area per site in the range of
1.31–13.78 m2 ha−1 which was lower than the range (17–40 m2 ha−1) reported for a
Mexican dry forest by Murphy and Lugo (1986). The reported basal areas from
other studies for dry forest communities are 7–23 m2 ha−1 (Jha and Singh 1990),
10.79–20.44 m2 ha−1 (Parthasarathy and Sethi 1997) and 3.9–16.7 m2 ha−1 (Backeus
et al. 2006). For tree species having stems ≥10 cm DBH, Top et al. (2009) found a
basal area of 23 m2 ha−1 in Kampong Thom Province, Cambodia. Chittibabu and
Parthasarathy (2000) recorded 23 m2 ha−1 in a highly disturbed site of Mottukkadu
shola and 54 m2 ha−1 in an undisturbed site at Perumakkai shola, Kolli hills, Eastern
Ghats, India. In the Vindhyan region of the northern deciduous forest of India,
Chaturvedi et al. (2011a) reported a range of 3.1 m2 ha−1 at the driest site to 18.0 m2
ha−1 at the moist site for woody species having ≥10 cm circumference, with mean
tree basal area being 14.3 m2 ha−1. Chaturvedi et al. (2011a) enumerated 48 species
within a 4.5 ha area for woody species ≥10 cm circumference. The values reported
from other large-scale permanent plot inventories (for trees ≥10  cm DBH, i.e.
≥31.4 cm circumference) were as follows: 37 species in Kampong Thom Province,
Cambodia (Top et al. 2009); 49 species in Vindhyan dry tropical forest, India (Sagar
et  al. 2003); 996 species in 52  ha plot of Lambir Hills National Park, Malaysia
(Condit et al. 2000a); 660 species in a 50 ha plot of Pasoh Forest Reserve, Malaysia
(Kochummen et al. 1990); 229 species in a 50 ha plot in Barro Colorado Island,
Panama (Condit et al. 1996); 146 species in a 30 ha plot at Varagaliar, Anamalais,
Western Ghat, India (Ayyappan and Parthasarathy 1999); and 164 species in a 25 ha
plot of Sinharaja Biosphere Reserve, Sri Lanka (Condit et al. 2000a).
The most species-rich family found by Chaturvedi et al. (2011a) for the Vindhyan
dry tropical forests is Leguminosae, which is also reported by Gentry (1995) as a
common family in TDFs. Singh et al. (1991a) and Sagar et al. (2003) found Acacia
catechu, Anogeissus latifolia, Diospyros melanoxylon, Lagerstroemia parviflora
and Hardwickia binata as the most frequent tree species in the northern Indian
TDF. Due to the dominance of only a few species (Terminalia tomentosa, Shorea
robusta, Diospyros melanoxylon, Buchanania lanzan and Lagerstroemia parvi-
flora), the Shannon-Wiener index was low (Chaturvedi et  al. 2011a; Sagar et  al.
2008a; Raghubanshi and Tripathi 2009) in these forests. According to Martijena and
Bullock (1994), extreme environmental conditions such as shallow soils and low
water availability allow only a few species to dominate the forest as also indicated
by Chaturvedi et al. (2011a).
The number of species and stem density recorded by Chaturvedi and Raghubanshi
(2014) for a northern tropical dry forest in India are given in Table 3.6. The presence
of Diospyros melanoxylon, Emblica officinalis, Lagerstroemia parviflora and
Lannea coromandelica in the adult trees; Flacourtia indica, Lagerstroemia parvi-
flora, Lannea coromandelica and Ziziphus glaberrima in the sapling population;
and Ziziphus glaberrima and Ziziphus oenoplia in the seedling population are
40 3  Vegetation Attributes

Table 3.6  Number of species and stem density in the five study sites located in the forest of
Vindhyan highlands
Site Parameters Adult Sapling Seedling Total
Hathinala Total no. of sp. 31 33 38 41
No. of stems (ha−1) 490 (±60) 1383 (±139) 14,167 (±736) 18,970 (±866)
Gaighat Total no. of sp. 18 22 30 32
No. of stems (ha−1) 330 (±21) 810 (±38) 9167 (±410) 12,070 (±276)
Harnakachar Total no. of sp. 21 26 33 34
No. of stems (ha−1) 337 (±18) 707 (±67) 10,333 (±722) 11,887 (±697)
Ranitali Total no. of sp. 12 20 26 29
No. of stems (ha−1) 203 (±35) 540 (±108) 7600 (±1193) 9083 (±1590)
Kotwa Total no. of sp. 10 11 6 14
No. of stems (ha−1) 126 (±36) 554 (±14) 5040 (±340) 5083 (±659)
Source: Chaturvedi and Raghubanshi (2014)

p­ ossible characteristics of the northern Indian dry deciduous forest composition.


The number of species common to all sites was remarkably low indicating that the
dry deciduous forest vegetation is actually quite diverse in space.
According to Engelbrecht et al. (2007), water availability and the drought sensitiv-
ity influence species distribution in the TDF. The study by Chaturvedi and Raghubanshi
(2014) revealed marked variations in the distribution of species and individuals at
various scales (Fig. 3.3); this characteristic of TDF drew little attention in the past (cf.
Jha and Singh 1990). Sagar and Singh (2006) examined species diversity on several
sites of a TDF in relation to tree basal area and reported that species diversity increased
linearly with an increase in total tree basal area (Fig. 3.4). In the species-poor dry
forest, the recurring disturbance does not allow concentration of biomass or stems in
only a few competitors. The relationships in Fig. 3.4 indicate that additional species
are encountered as the number of stems increases and distribution of individuals
among species is more equitable contributing to high species diversity. Jha et  al.
(2005) assessed the change in vegetation cover over a decade (1988–1998), using the
normalized difference vegetation index (NDVI) for a TDF area, and reported a
decrease in the number of species in the case of negative-change (reduced NDVI)
areas compared to positive-change (increased NDVI) areas. This observation has also
been supported by the study of various diversity indices (Jha et al. 2005).
Sagar et al. (2008a) compared community composition and species diversity of
the understorey and overstorey vegetation among five TDF sites (same as in Sagar
and Singh 2005) in northern India. Individuals ≥30 cm height and up to 9.6 cm DBH
were counted by species as understorey elements. This category included estab-
lished seedlings (≥30 cm height and <3.2 cm diameter) and saplings (≥3.2 cm diam-
eter to <9.6  cm DBH). All individuals >9.6  cm DBH were tallied as overstorey
elements. The uneven distribution of species and individuals across sampling plots
reflects the prevalence of coppicing habit. Upadhyay and Srivastava (1980) and
Harikant Ghildiyal (1982) have already commented on the prevalence of the coppic-
ing habit in the dry tropical species. Pronounced β-diversity also indicates different
levels of habitat heterogeneity (Sagar and Singh 2005). Further, most of the under-
storey species had only one or one to ten individuals at each forest site indicating the
3.1  The Tropical Dry Forest 41

Fig. 3.3  Dispersion patterns in (a) adult, (b) sapling and (c) seedlings at the five study sites.
Different letters for a dispersion pattern indicate statistically significant difference on the basis of
post-hoc analysis. HN Hathinala, GG Gaighat, HK Harnakachar, RT Ranitali, KT Kotwa (Source:
Chaturvedi and Raghubanshi 2014)

mixed nature of the forest. Diospyros melanoxylon, Holarrhena antidysenterica and


Lagerstroemia parviflora exhibited increasing presence along the disturbance gradi-
ent indicating resistance of these species to the human pressure. The four species
consistently present in the overstorey on all the studied sites were Acacia catechu,
Anogeissus latifolia, D. melanoxylon and Lannea coromandelica. Thus, it is evident
42 3  Vegetation Attributes

Fig. 3.4  Relationship between basal area (BA) and different diversity indices. (a) α-diversity
according to α = 0.67 + 0.38BA (r2 = 0.65, p <0.001), (b) species richness (S) according to S = 6.41
+ 1.33BA (r2  =  0.68, p  <0.001), (c) evenness (E) according to E  =  1.97 + 0.24BA (r2  =  0.47,
p  <0.001) and (d) stem density (D) according to D  =  83.57 + 24.08BA (r2  =  0.51, p  <0.001)
(Source: Sagar and Singh 2006)

that the understorey and overstorey species differed in sensitivity to disturbance. The
recurrent human interventions in terms of collection of fuel wood and minor forest
products and grazing and trampling change the habitat fitness for many species
(Pandey and Shukla 1999). On the basis of relative importance of overstorey tree
species, Sagar et al. (2008a) identified Acacia-Shorea, Tectona-Anogeissus, Shorea-
Terminalia, Hardwickia-Shorea and Hardwickia-Butea communities at the five
study sites. Sagar et al. (2008a) suggested the probability of a change with time in
the overstorey species composition under the existing disturbance regime. Therefore,
3.1  The Tropical Dry Forest 43

the Acacia-Shorea community is likely to be replaced by the Briedelia-Terminalia


community, Tectona-Anogeissus community by Diospyros-Holarrhena community,
Shorea-Terminalia community by Diospyros-Terminalia community, Hardwickia-
Shorea community by Diospyros-Hardwickia community and Hardwickia-Butea
community by Holarrhena-Diospyros community in the future.

3.1.2  S
 mall-Scale Variations in Environmental Factors
and the Distribution of Woody Species

Distribution of plant communities and their species and structural diversity is highly
affected by soil water and soil nutrient status (Palmiotto et al. 2004). In the environ-
ment with plenty of soil nutrients, species grow rapidly, allocation to above-ground
parts is more and uptake of nutrients per gram root biomass is higher compared to
the species growing in nutrient-poor environment (Tilman 1987a). Soil organic car-
bon (C) is a major constituent of soil organic matter which has a major effect on
forest productivity and sustainability by influencing soil chemical and biological
properties (Goma-Tchimbakala 2009). Soil nitrogen has been reported to exhibit
strong association with structure of forest and distribution of tree species (Tateno
and Takeda (2003). Various reports also highlight phosphorus as the principal nutri-
ent influencing tree growth and forest productivity in tropics (Palmiotto et al. 2004).
Chaturvedi et al. (2011a) showed marked differences in woody species distribution
in response to small variation in environmental variables (Fig. 3.5); thus in TDF, the
presence or absence of several species is determined by their edaphic preferences.

3.1.3  C
 omposition and Dynamics of Tropical Dry Forest
in Relation to Soil Texture

Jha and Singh (1990) examined the distribution of communities in relation to soil
texture and subjected the data to multivariate analysis. DCA (detrended corre-
spondence analysis) axis 1 (Fig.  3.6) was inversely related to the proportion of
sand (fine sand, r = −0.399; p < 0.01, and coarse sand, r = −0.516; p < 0.03). DCA
axis 2 was negatively related to silt content (r = −0.526; p < 0.03). Thus, it is evi-
dent that the soil texture is an important factor in the constitution and distribution
of TDF communities. The correspondence of soil texture with identified commu-
nities was further supported by discriminant analysis (DA) using the soil textural
parameters and organic carbon (see Jha and Singh 1990). Among the three dis-
criminant functions identified by DA, only functions 1 and 2, which accounted for
63.4% and 30.6% of the total variance, respectively, were important. In the study
of Jha and Singh (1990), 76% of the sites were correctly classified (Fig. 3.7). The
44 3  Vegetation Attributes

Axis 2
T. tomentosa D. melanoxylon

A. latifolia
B. lanzan
S. robusta

L
SMC L. parviflora
G. latifolia Clay
S. anacardium
B. retusa
A. cordifolia
S. swietenioides E. glaucum W. fruticosa
P. marsupium P B. serrata
A. auriculiformis
M. parvifolia O. oogenesis S. febrifuga
M. tomentosa A. indica
C. fistula B. racemosa H. integrifolia
C. siamea
A. odoratissima
S. oleosa G. serrulata N G. turgida M. longifolia
H. excelsum
F. indica T. chebula L. camara H. binata
Z. nummularia
Axis 1
A. catechu C. spinarum Silt

Sand
Rockiness
E. officinalis Z. oenoplea

H. antidysenterica

S. urens
D. strictus

Z. glaberrima

N. arbortristis
L. coromandelica
A. precatorius
F. racemosa

Fig. 3.5  Ordination of the woody species in the forests of Vindhyan highlands by canonical cor-
respondence analysis. SMC soil moisture content, N total nitrogen, P total phosphorus, L light
attenuation (Source: Chaturvedi et al. 2011a)

soil group-plant community relationships are clearly shown in Fig. 3.7, indicating


the position of the nodum centroids in the plane of the first two discriminant
functions.
Jha and Singh (1990) suggested that the patches represented by the stands of the
various communities are dynamic and that new species combinations may arise and
result in new communities. According to Aubréville (1938), combinations of domi-
nant species at a given place and time may be succeeded, not by the same combina-
tions but by different ones. In this way a mosaic of spatially and temporally dynamic
communities may arise in TDF.
3.1  The Tropical Dry Forest 45

Fig. 3.6  Position of 2·0


nodum centroids in the
plane of discriminant 1·6 I
functions 1 and 2 of a
discriminant analysis 1·2
IV
(Source: Jha and Singh
0·8
1990)
0·4 III

DF II
–0·4 II
–0·8 V

–1·2

–1·6

–2·0
–1·6 –1·2 –0·8 –0·4 0·4 0·8 1·2 1·6 2·0
DF I

200
III
II
160 I 8 13 38 39
DCA axis II

17 29 31
35 V
28 22 23 20
120
30
18 14 5 40
7 3 24 6
80 IV 1 41
16 9 12
11 10 37
27
15 32 36
40
33
25
42
34 21

40 80 120 160 200 240 280 320 360 400


DCA axis I

Fig. 3.7  DCA ordination of tropical dry forest stands, axes 1 and 2. Nodum indication as in
Fig. 3.6 (Source: Jha and Singh 1990)

3.1.4  E
 ffect of Woody Plant Canopies on Species Composition
and Diversity of Ground Vegetation

The presence of woody plant canopies alters the microenvironment and the soil prop-
erties (Weltzin and Coughenour 1990). The decreased soil and air temperatures,
wind speed and irradiation due to the presence of trees reduce evaporation of soil
water and increase relative humidity (Jose et al. 2008; Rao et al. 1998). The deep-
rooted woody components transport water from deeper soil layers to drier surface
soils through hydraulic lift and can thus benefit the understorey vegetation during dry
46 3  Vegetation Attributes

periods (Burgess et al. 1998; Ong et al. 1999). In addition, trees also extract nutrients
from deeper soil layers and through litterfall deposit these at the surface enhancing
soil carbon and nutrients in the surface soil layers which benefit the understorey
plants. In the tropics, solar radiation, which is incident perpendicular to the ground,
is intercepted by the forest canopy and exhibits a decreasing gradient from the centre
of the canopy gap to under-canopy locations (Chazdon 1986; Denslow 1987).
According to Vetaas (1992), trees can either reduce or increase grass production by
modifying the resource availability. Studies on the effect of tree canopies on herba-
ceous diversity in tropical forests (Denslow 1987; Gillet et al. 1999; Joshi et al. 2001;
Rodriguez-Echeverria and Perez-Fernandez 2003; Zobel et al. 1994) have yielded
equivocal results. According to Breshears (2006), the ground beneath woody plant
canopy and that located in the inter-canopy areas experience different patterns of
energy, water and soil chemistry. The presence of architecturally contrasting tree
canopies can thus influence the diversity of ground vegetation (Sagar et al. 2008b).
In the study of Sagar et al. (2008b), reduced light infiltration to ground in some cano-
pies was reflected in lesser number of unique species and also lower species richness,
evenness and alpha diversity compared to other canopies (Fig.  3.8). Sagar et  al.

a b
21 2.8 a
a
Mean number of species

18
2.4
Alpha diversity

15
b
b b
12 2.0
b
9
1.6
6

3 1.2

c
10
a
8
Evenness

6
b b
4

0
Canopy-1 Canopy-2 Canopy-3

Fig. 3.8  Herbaceous species diversity under three woody plant canopies (a) mean number of spe-
cies, (b) alpha diversity and (c) species evenness. The bars within a diagram affixed with different
letters are significantly different from each other at P = <0.05 (Source: Sagar et al. 2008b)
3.2  Seasonality, Phenology and Deciduousness 47

(2008b) reported distinct herbaceous communities, each with a characteristic species


composition under the different woody canopy types. Several studies have recog-
nized the light availability as an important plant resource (Blankenship 2002;
Maximov 1929), which in combination with other resources may affect plant perfor-
mance (Cole 2003). Gillet et al. (1999) reported that the number of herbaceous spe-
cies decreased directly under the tree canopy and up to a mean distance of 40 m from
the nearest trees, whereas the number of herbaceous species was optimum at a dis-
tance of about ca.30 m. Light availability on the forest floor is a key factor that influ-
ences intrinsic traits of inhabiting species (Jones et  al. 1994; Walters and Reich
1996). Canopy shade creates a light regime at the ground level which may, depend-
ing upon its intensity, be photosynthetically quite inactive (Fetcher et al. 1983; Turton
and Duff 1992), and since different species of canopy trees intercept different levels
of light (Canham et al. 1994), tree species growing over a sapling population may
influence the growth of these saplings. According to Breshears (2006), “the physical
and biological effects of woody plant canopies on the environment can create sub-
stantial heterogeneity in several ecological properties between canopy and inter can-
opy patches”. This heterogeneity, which will be a function of height and architecture
of the trees, may increase the diversity of herbaceous species as different species may
respond differently to the mosaic of light condition.

3.2  Seasonality, Phenology and Deciduousness

Most phenological changes in TDFs are caused by the following factors: (1) sea-
sonal variation in rainfall (Daubenmire 1972; Bullock and Solis-Magallanes 1990;
Borchert 1994a; Eamus and Prior 2001); (2) seasonal variation in stem water status,
which in turn is determined by environmental factors such as soil water availability
and temperature; and (3) endogenous factors such as leaf age and leaf area (LA),
root size and distribution and stem wood density (Borchert 1994a, b, 1998).
Phenological differences among species promote species coexistence and therefore,
influence the biodiversity (González and Loreau 2009; Godoy and Levine 2013;
Wolkovich et  al. 2014a). The temporal niche differences and the associated trait
variation especially in diverse communities such as tropical forests, however, have
remained understudied (Kelly and Bowler 2005; Zimmerman et al. 2007; Pau et al.
2013). Although tropical ecosystems are strongly affected by the seasonality of
rainfall (Frankie et al. 1974; van Schaik et al. 1993; Pau et al. 2010; Hulshof et al.
2011; Guan et al. 2014), the role of seasonality in mediating community phenology
is not well understood.
Organisms in seasonal environments face the problem of weathering harsh peri-
ods and exploit the advantages of the benign periods. The harsh conditions could
limit the recruitment of offspring or could reduce the allocation to reproduction in
adults (Wheelwright 1985; van Schaik et al. 1993; Khurana and Singh 2001a; Hulshof
et al. 2011), favouring phenological synchrony among species (Borchert et al. 2004;
Vasseur et al. 2014) (Fig. 3.9). Extreme climate seasonality may also lead to strong
phenological variation (Janzen 1967; Frankie et al. 1974; Reich and Borchert 1984;
Murali and Sukumar 1994; Lechowicz 1995). Interactions in biological organisms
48 3  Vegetation Attributes

Fig. 3.9  Patterns of species reproductive phenology (green lines) with respect to phenology of
other species ((a), (b), solid, dashed and dotted lines correspond to different species), rainfall (blue
line), and traits we expect will be associated with phenological variation (c), (d). In (c) there is a
species that responds to seasonal rainfall (solid green line) and is constrained to reproducing in the
rainy season and a species that also responds to intra-seasonal rainfall variation (dashed green
line). Note that multiple mechanisms may act simultaneously and multiple patterns may emerge at
distinct temporal scales (Source: Lasky et al. 2016)

can favour both positive temporal covariation (i.e. synchronous) and negative tempo-
ral covariation or anti-synchrony (i.e. compensatory phenology), which depends on
whether interactions exhibit positive or negative density dependence (Janzen 1967;
Gentry 1974; Stiles 1977; Rathcke and Lacey 1985; Curran and Leighton 2000;
Elzinga et al. 2007; Botes et al. 2008; González and Loreau 2009; Jones and Comita
2010; Albrecht et al. 2015) (Fig. 3.9). Evidence for specific abiotic and biotic drivers
3.2  Seasonality, Phenology and Deciduousness 49

of population phenology exists (Rathcke and Lacey 1985, Elzinga et al. 2007), but
studies on whole community reproductive phenology and associated drivers are lack-
ing (Herrera 1998, Olesen et al. 2008, Yang et al. 2013).
Seasonal water limitation in TDFs can be severe, promoting synchronous repro-
duction during rainy periods (Borchert et  al. 2004; Singh and Kushwaha 2006).
Occurrence of reproductive peaks around the onset of rains (McLaren and McDonald
2005) may lead to a reduction in water stress in adults as well as seedlings (van
Schaik et al. 1993). However, simple explanations may not explain the phenological
trends in very dry regions (Reich and Borchert 1984; Murphy and Lugo 1986;
Murali and Sukumar 1994; Guan et al. 2014), as several TDFs exhibit peak repro-
duction during dry periods (Janzen 1967; Frankie et  al. 1974; Selwyn and
Parthasarathy 2006).
Variation in phenology among species may be associated with their resource
allocation strategies and relative tolerance to drought/water stress (Reich and
Borchert 1984; Borchert et  al. 2004; Wolkovich et  al. 2014b) (Fig.  3.9). Species
reproducing under favourable conditions or that showing high sensitivity to intra-
seasonal fluctuations in abiotic environment (these two distinct criteria may suit
different species; Vicente-Serrano et al. 2013) may possess acquisitive strategy, i.e.
having traits that allow for rapid resource exploitation. In dry forests, acquisitive
species rapidly exploit soil moisture, whereas conservative species are more tolerant
of dry periods (Markesteijn et al. 2011; Sterck et al. 2011), which allow these spe-
cies to reproduce during the dry season. Operation of acquisitive species may be
closer to the margins of safe resource levels, and they have traits which allow for
rapidly exploiting the resource pulses (Markesteijn et  al. 2011; Ouédraogo et  al.
2013; Muscarella and Uriarte 2016). Soils which are nutrient poor generally favour
species with a conservative strategy, i.e. with an evergreen leaf habit and thus a long
leaf life span, whereas the soils which are relatively rich in nutrients should favour
fast-growing acquisitive species with deciduous leaf habit and a short leaf life span
for avoiding the unfavourable growth period (Aerts 1995; Givnish 2002). Thus, in
climates characterized by seasonal drought, there occurs a trade-off between the
deciduous leaf habit to cope with drought and an evergreen leaf habit enabling the
plants to overcome nutrient shortages.
Acquisitive strategies in dry forest trees are associated with (i) low wood density,
(ii) rapid transport and storage of water and (iii) fast growth, notwithstanding the risk
of cavitation and mortality. On the other hand, conservative strategies are associated
with low specific leaf area (SLA), evergreen leaves, low leaf area, low leaf N and P
and a tap root which facilitates water access and storage (Poorter and Markesteijn
2008; Markesteijn et al. 2011; McCulloh et al. 2011; Sterck et al. 2011; Méndez-
Alonzo et al. 2012). Drought stress on seedlings may also influence phenology, such
that species that fruit during rainy seasons may show high germination, while those
that fruit during the dry seasons remain dormant until the wet periods (van Schaik
et al. 1993; Lechowicz 1995; Soriano et al. 2011). However, according to Powers
and Tiffin (2010), among others, “tree functional traits often do not neatly collapse
into a single axis of resource acquisitive versus conservative strategies.” The impacts
of drought on dry forests are little understood (but see Enquist and Leffler 2001;
50 3  Vegetation Attributes

Borchert et al. 2002; Soriano et al. 2011; Maza-Villalobos et al. 2013) due to a lack
of high frequency and long-term studies, with the exception of McLaren and
McDonald (2005) who recorded monthly observations for 2 years in Jamaica.
The phenological patterns exhibited by TDF plants vary widely at spatial and
temporal scales (Kushwaha et  al. 2015). Clumped phenologies, according to van
Schaik et al. (1993), result from two kinds of biotic processes: (i) plants may time
their activities in relation to the seasonally varying availability of biotic agents, and
(ii) they may attract pollinators and dispersers or satiate predators by doing so. At
sites marked by intermediate intensity of drought or across moisture gradients, spe-
cies may display a range of behaviours, from asynchronous to highly synchronized
(Borchert 1980; Reich and Borchert 1984, 1988; Wright and Cornejo 1990); some
of these behaviours are facultative, and individuals of the same species growing on
different sites, or even the same site, may display marked differences among each
other (Borchert 1980; Reich and Borchert 1984). Where dry seasons are moderate,
individuals of a tree species may exhibit asynchrony (Borchert 1980; Reich and
Borchert 1988) and have canopies of different leaf ages with markedly different
physiology.
Among important phenological events, the deciduous leaf habit of trees in TDFs
is quite common. Many tree species are distributed according to water availability
(Balvanera et al. 2011), and lower rainfall sites have a greater proportion of decidu-
ous trees (Singh and Kushwaha 2005). Water availability limits the start and end of
the growing period of deciduous trees (Borchert 1994c; Reich and Borchert 1984).
In contrast to trees with evergreen leaf habit, trees with deciduous leaf habit are not
able to take advantage of intermittent rainfall during dry season, and when such
rains cause early bud break, the deciduous species are damaged (Becknell et  al.
2012). Deciduous trees in wetter areas can retain their leaves longer and are likely
to have lower mortality. After years of higher productivity and lower mortality, for-
ests on the wet end of the TDF spectrum attain a greater biomass than those in the
dry end of the spectrum (Becknell et al. 2012).
There is a strong tendency for species of a given leaf life span to possess a defined
set of structural and functional leaf traits (Reich et al. 1991, 1992; Reich 1993), as
well as several common whole plant attributes (Reich et  al. 1992; Gower et  al.
1993). Short-lived leaves generally have higher photosynthetic rates (especially on
a mass basis), leaf diffusive conductance, leaf nutrient concentrations and specific
leaf areas (Reich et al. 1991, 1992) but are more fragile (Reich et al. 1991; Coley
1988) and more heavily preyed upon by insect herbivores (Coley 1988). In a TDF,
Sobrado (1991) found that deciduous species with shorter-lived leaves (average of
9 months) also had to invest lesser for construction and maintenance than evergreen
species having longer-lived leaves (average of 12  months). Species with shorter-
lived leaves also exhibit greater proportional allocation to leaf area as young plants
(Reich et  al. 1992), which enables them to achieve high whole plant and height
growth rates (Coley 1988; Reich et al. 1992) during early stages of tree growth or
stand development.
Deciduous trees do not have to sustain the costs of leaf respiration or face the
pressure of herbivory in the dry season (Eamus 1999; Givnish 2002); however, they
3.2  Seasonality, Phenology and Deciduousness 51

have to construct new leaves, which is expensive both in terms of nitrogen and car-
bon. Deciduous leaves are richer in N content, have higher specific leaf area and
photosynthetic rates compared to evergreen leaves and are therefore able to fix sub-
stantial amounts of carbon in the short growing period (Eamus 1999; Givnish 2002;
Franco et  al. 2005). Deciduous trees can also conserve N by resorbing nutrients
from senescing leaves (Chapin and Kedrowski 1983; Pallardy 2008; Ratnam et al.
2008; Santa Regina et al. 1997). In the TDF, the upper 10 cm layer of the soil where
the majority of fine roots occur are richest in N content (February and Higgins 2010;
February et al. 2013); that means plants in these systems allocate their roots prefer-
entially in areas where nitrogen concentrations are highest, but this nitrogen is avail-
able only for a brief period to plants, and it is supplied as a pulse at the start of the
growing season (Williams et al. 2009; Jarvis et al. 2007; Higgins et al. 2015).
Worbes et al. (2013) have identified four functional groups of species on the basis
of leaf-fall behaviour: deciduous, brevi-deciduous, stem succulent and evergreen.
Figure 3.10 shows the annual pattern of canopy fullness for the four guilds (ever-
green, deciduous, semi-deciduous and brevi-deciduous). Evergreen species retain a
full canopy all the year round and reduce the canopy fullness in the dry season by
less than 10%. Deciduous species, in contrast, lose all leaves for at least one but
usually 2–4 months of each year. Semi-deciduous species lose at least 50% of their
leaves every year. Brevi-deciduous species on the other hand never lose more than
50% of their leaves. Most trees in TDFs shed their leaves during the dry season, and

Fig. 3.10  Annual pattern of canopy fullness for four phenological guilds in north Australian
savannas. The four guilds are present in equal numbers in Australian savannas but not elsewhere.
The dry season extends from April to October. Key: circles, evergreen; triangles, brevi-deciduous;
squares, semi-deciduous; diamonds, deciduous (Source: Eamus 1999)
52 3  Vegetation Attributes

many species flower or flush soon after leaf shedding (Borchert 1994c). Many trees
rehydrate after leaf shedding, and the rehydration always precedes flowering or
flushing (Borchert 1980, 1991; Reich and Borchert 1982, 1984).
In the majority of TDF species, leaf emergence from buds and subsequent expan-
sion precede the start of the wet season in Africa, Central America, India, Australia
and Costa Rica (Seghieri et al. 1995; Chidumayo 1990; Monasterio and Sarmiento
1976; Rundel and Becker 1987). But as Borchert (1994b) reports, this pattern is not
universal and leaf flushing can occur at the start of the wet season or before and after
the start of the wet season (Borchert 1994b). The advantage of leafing before the
onset of rains is that the plants can expand the leaves rapidly once the rains arrive,
by photosynthesizing vigorously. Murali and Sukumar (1993) observed that the
trees which flushed their leaves later in the rainy season suffered significantly higher
damage by insects and in most dry tropical species, leaf growth stops by the middle
of the wet season.
Compared to rain forests, the phenology of dry forests may be more sensitive to
elevated temperature and elevated CO2. Higher temperatures are expected to lead to
more rapid water depletion, earlier leaf drop, longer leafless periods and more
strongly synchronized phenology. If elevated CO2 leads to an acceleration of organ
development and thus reduced life span, leaves might be shed during otherwise rela-
tively distressful conditions, leading to renewed leaf flushing in some cases prior to
the onset of seasonal drought.
Studies indicate that forest deciduousness increases by decrease in annual rain-
fall or increase in the length and severity of the dry season (Reich 1995; Borchert
1998; Condit et al. 2000b; Bohlman 2010; Gond et al. 2013). Further, geology of the
habitat can significantly alter this relationship (Condit et al. 2000b; Bohlman 2010),
probably because the soil resource availability differs between geological sub-
strates. For instance, on seven volcanic and ten sedimentary geological substrates in
central Panama, Bohlman (2010) correlated annual rainfall with the average decidu-
ousness and compared the forest deciduousness on a limestone substrate (the Tau
geological formation) with other substrates exposed to the same rainfall level and
observed a drastic difference in deciduousness.
Ouédraogo et al. (2016) analysed the environmental determinants of forest decidu-
ousness in mixed moist semi-deciduous forests in the Congo Basin and found evi-
dence that deciduousness provides competitive advantage to deciduous species in
climates with high rainfall seasonality, while the evergreen species persist on resource-
poor soils. Moreover, the relative contribution of geology to the variation in forest
deciduousness was three times stronger than that of the climate, highlighting the
importance of geology and soil for predicting vegetation functional characteristics.
In eastern, central and southern Africa, Faidherbia albida is characterized by a
reverse phenology (Wickens 1969), i.e. trees bear leaves, grow and fruit during the
dry season, whereas they shed leaves after the first rains and resume growth at the
end of the wet season. According to Roupsard et al. (1999), F. albida displays effi-
cient growth during the dry season, when they tap groundwater from deeper soil
layers. The root system in species such as F. albida is complex, extending to several
soil compartments, and the plants utilize soil water from different depths at different
times in the year.
3.2  Seasonality, Phenology and Deciduousness 53

Further, a correlation has been reported between vulnerability to embolism and


water availability (Franks et al. 1995; Alder et al. 1996; Mencuccini and Comstock
1997; Sparks and Black 1999), and it has been reported that the trees from drier
environments being less vulnerable to embolism than those from wetter environ-
ments. Resistance to embolism may be linked with unusual leaf phenology of
Cordia alliodora on Barro Colorado Island; the high resistance to embolism would
allow for the maintenance of a canopy during the dry season.

3.2.1  P
 henology of Seasonally Tropical Dry Forest: Two Case
Studies

Singh and Singh (1992a) selected three sites located in Vindhyan region, Sonbhadra
district, Uttar Pradesh, India. Site 1 comprised a north-facing slope, comparatively
moist due to the presence of a stream, and the forest was dense (stems ≥30 cm cir-
cumference at breast height = 2210 per ha). Sites 2 and 3 occurred on a plateau,
were drier and had low tree densities (950 and 760 stems per ha), respectively.
Permanent plots of 1 ha each were established at each of the three sites. Five indi-
viduals of each of the 26 tree species (Table 3.7) were marked and tagged. On each
individual, four branches were marked with metal tags. Observations at 30-day
intervals were made for each of the tagged individual for the initiation and comple-
tion of the following phenophases: leaf flushing, leaf-fall, flowering, fruiting and
fruit-fall.
This study showed that there was a considerable diversity among species and
sites, in the timing and duration of phenophases. Diversity in the initiation and
completion of different phenophases and their monthly trends among the plant
species are shown in Table 3.7 and Fig. 3.11. Singh and Singh (1992a) observed
that the dry period was characterized by intense phenological activity. Initiation of
leaf-fall coincided with the onset of the post-monsoon, low temperature dry period
and could be a mechanism of maintaining shoot turgidity (Fig. 3.11a, c). However,
leaf flushing began with the rise of temperature and peaked in the hottest and driest
month (May) of the year (Fig. 3.11b); thus, rise in temperature seems to trigger
leaf initiation. Other studies in tropical forests also indicated a lack of correlation
between leaf flushing and the onset of the rainy season, the leaf flushing usually
preceding the onset of rainy season (Walter 1971). The emerging leaves initially
draw nutrients from the pool of nutrients withdrawn earlier from senescing leaves
(see later); however, nothing is definitely known about the water demand and sup-
ply relationships of the new foliage during the dry period, although it can be
argued that most of the needed water may come from deeper soil layer, while
stomatal conductance of the emerging leaves may be reduced. The dry season leaf
flushing, nevertheless, permits the renovation of the canopy before the monsoon
rainfall begins, such that the plants are able to take full advantage of the short rainy
season for primary production and growth. Further, at the community level, leaf
flushing continues for several months which ensures continued availability of pho-
tosynthetically active tissue in the community. This also results into a sustained
54

Table 3.7  Diversity among plant species in the periods of initiation and completion of phenophases resulting into extended phenophases at the community
level
Leaf Completion of Leaf-fall Leaf-fall Flowering Completion Fruit Fruit-fall Completion
Species initiation leaf initiation initiation completion initiation of flowering initiation initiation of fruit-fall
Acacia catechu August
Aegle marmelos August May
Anogeissus latifolia September October
Bauhinia racemosa June
Boswellia serrata June August January February January April
Bridelia retusa August December February
Buchanania lanzan February March April
Butea monosperma June March February March
Diospyros melanoxylon June
Hardwickia binata August February–March April
Lannea coromandelica August November December–January January February January March
Madhuca longifolia August February March August September
Miliusa tomentosa June October November
Pterocarpus marsupium August September November January
Shorea robusta February– June April May
March
Soymida febrifuga June April May
Terminalia alata August
Wrightia tomentosa August
Community peak May July January February–March May June– June January– April–May
period August February
Source: Singh and Singh (1992a)
3  Vegetation Attributes
3.2  Seasonality, Phenology and Deciduousness 55

Fig. 3.11 (a) Ombrothermic diagram for the study area: solid curve, rainfall; broken curve, mean
temperature; hatched area shows xeric conditions; (b) initiation and completion of leafing; (c)
initiation and completion of leaf-fall; (d) initiation and completion of flowering; (e) initiation and
completion of fruiting; and (f) initiation and completion of fruit-fall. Data for site I are represented
by circles, for site II by triangles and for site III by squares. Solid symbols are for initiation and
open symbols for completion of phenological events. The smooth curves (solid for initiation and
broken for completion of various phenological activities) are fitted by using Harvard Graphics
(Source: Singh and Singh 1992a)

supply of tender food resource to phyllophagous insects which constitute an


important component of community food web and benefit the plants by speeding
up the nutrient cycling. The coincidence of peaks in leafing and flowering is inter-
esting. Leafless and low-foliage periods favour wind pollination and permit floral
display to attract pollinators. Massive pre-monsoon flowering (Fig.  3.11d) also
means a continued availability of energy in the subsequent rainy season to devel-
oping fruits when the foliage is actively photosynthesizing. Koelmeyer (1960)
reported flowering in TDFs of Sri Lanka when soil moisture was depleted, and
56 3  Vegetation Attributes

Malaisse (1974) also reported tree flowering in dry deciduous woodland in


Tanzania during the pre-rains, warm period. For a Costa Rican dry forest, two peak
periods of flowering, one extensive period during the long dry season and the other
subdued peak period at the onset of rainy season, were reported by Frankie et al.
(1974).
Experimental pollination showed a high proportion of tree species in the dry for-
est of Costa Rica to be self-incompatible, requiring animals for effective transfer of
pollen for outcrossing (Bawa 1974). All the species studied by Singh and Singh
(1992a), whose pollination mechanism is known, are also animal pollinated, and
flowering in the community was staggered covering a period of 10 months, with a
main May peak and two secondary peaks (Fig. 3.11d). Staggered flowering may be
an evolutionary mechanism to avoid (i) interspecific competition for pollinators, (ii)
interspecific pollen transfer which otherwise could result in reduced pollen donation
and (iii) consequently reduced seed set due to stigma clogging (Rathcke and Lacey
1985). Frankie et al. (1974) found, in a Costa Rican dry forest that the flowering
period of a large number of species was spread over 3.5 months in order to minimize
the flowering overlap. A long period of community-level flowering permits sus-
tained flow of floral food stuffs to different groups of pollinators with populations
peaking in the different periods in an annual cycle. Asynchrony in flowering within
species as exhibited by Acacia catechu and Lagerstroemia parviflora in the study of
Singh and Singh (1992a) promotes outcrossing, reduces intraspecific competition
for pollinators and reduces geitonogamy (Rathcke and Lacey 1985).
The fruiting phenology in the dry forest investigated by Singh and Singh (1992a)
closely followed the flowering phenology (Fig. 3.11d,e); consequently, the fruiting
activity was also quite staggered and prolonged at the community level (Singh and
Singh 1992a). In the dry forest of southern Puerto Rico, no more than 50% of the 33
species were noted by Murphy and Lugo (1986) to fruit at the same time. Sequential
fruiting in the trees of a forest allows full utilization of the dispersal agents resulting
from minimized competition for their service (Snow 1971). The close correspon-
dence between flowering and fruiting phenology minimizes the exposure to flower
and seed predators. In the study of Singh and Singh (1992a), at the community level,
fruit-fall which began in January–February was completed in most species before or
just at the beginning of the monsoon season (Fig. 3.11f). This ensures availability of
sufficient moisture to seeds for germination and seedling establishment.
In another case study, Pandey et al. (2014a) studied leaf area (LA), leaf water
content (LWC), leaf fresh weight (LFW), leaf dry weight (LDW), specific leaf area
(SLA) and chlorophyll content of eight woody species (viz. Shorea robusta,
Buchanania lanzan, Diospyros melanoxylon, Lagerstroemia parviflora, Lannea
coromandelica, Terminalia tomentosa, Holarrhena antidysenterica and Lantana
camara), dominant at four sites in the tropical dry deciduous forest of the Vindhyan
region. These authors recognized four classes of species:
1. Semi-evergreen (SE) (species which are hardly ever without green leaves and if
so only for a relatively short period; Shorea robusta was the species in this
category).
3.2  Seasonality, Phenology and Deciduousness 57

2. Short-deciduous (SD) (species that maintain comparatively low canopy (15–


50%) throughout the year)—two species (Buchanania lanzan and Diospyros
melanoxylon) fell in this group.
3. Mid-deciduous (MD) species, which remain leafless for a period of 2–3 months,
were represented by Lagerstroemia parviflora and Terminalia tomentosa.
4. Long-deciduous (LD) species, which remain leafless for a period of 4–5 months,
were represented by Holarrhena antidysenterica, Lannea coromandelica and
Lantana camara.
Paradoxically, species with varying length of leafless period occur under the
same climatic conditions. The LD species get the shortest growing period to synthe-
size and accumulate biomass. The LD species exhibit high SLA (Fig. 3.12). Studies
have indicated a positive relation between SLA and net photosynthesis (Hoffmann
et al. 2005; Chaturvedi et al. 2011b), and high SLA may compensate for the extended
leaflessness. Furthermore, deciduous species are known to have a low leaf construc-
tion cost (Baruch and Goldstein 1999), assimilate with high photosynthetic rate
(Eamus and Prichard 1998) and can produce the required amount of energy through
photosynthesis in relatively shorter time (Feng et al. 2008). These species also have
high leaf nutrient content (Eamus and Prichard 1998; Feng et al. 2008), particularly
nitrogen, supporting high photosynthetic activity during the growing season.
Figure 3.12 shows seasonal pattern of leaf traits of woody species in the four classes
of deciduousness, and Fig. 3.13a shows differences in leaf traits among the impor-
tant woody species. Pandey et al. (2014a) reported significant variations in leaf phe-
nological events of woody species in moist and dry sites (Fig.  3.13b). The peak
period of leaf-fall initiation in most of the species at dry sites was 2 weeks before
that in the relatively moist sites (Pandey et al. 2014a). Plant species growing at the
dry sites evidently tolerate water stress to a greater extent as compared to those
growing at moist sites. Early leaf flushing helps them to take the advantage of short
wet growing season, as they experience early leaf-fall at the start of the dry season.
Evidently, the deciduous species sprout new leaves during dry summer, when day
length is increasing and temperatures are rising.

3.2.2  Reproductive Phenology

Several authors have hypothesized the reproductive phenology to be an axis of niche


partitioning (Janzen 1967; Gentry 1974; Stiles 1977; Wheelwright 1985; Murali
and Sukumar 1994; Sakai 2001; Botes et al. 2008; Lasky et al. 2016). Lasky et al.
(2016) used multi-scale wavelet analysis to demonstrate that TDF tree species
exhibit (i) synchronous, (ii) asynchronous and (iii) compensatory dynamics in
reproduction, depending on temporal scale and stage of reproduction. Trait differ-
ences may correspond to phenological diversity and may support either acquisitive
or conservative trade-offs in drought response (Markesteijn et al. 2011; Sterck et al.
2011). However, multiple opposing processes may obscure each other’s effects as
they may be driving phenology at distinct scales (Elzinga et al. 2007; Keitt 2008).
58 3  Vegetation Attributes

Fig. 3.12  Seasonal pattern of leaf traits of deciduous species in the tropical deciduous forest
(Source: Pandey et al. 2014a)

3.2.3  Seasonal Variation in Tree Water Status

Temporal correlations between stem water potential (Ψstem) and the time course of
tree development during the dry season confirm the proposed role of water status as
the principal determinant of tree phenology in TDFs (Reich and Borchert 1984;
Borchert 1991, 1992). With few exceptions, buds remained arrested as long as Ψstem
3.2  Seasonality, Phenology and Deciduousness 59

Fig. 3.13  Mean values of leaf traits of different species averaged across months, sites and years
(a), leaf traits at different sites averaged across species, months and years (b). LA leaf area, LDW
leaf dry weight, LFW leaf fresh weight, SLA specific leaf area, CHL chlorophyll concentration,
LWC leaf water content. Sh Shorea robusta, Bu Buchanania lanzan, Di Diospyros melanoxylon, La
Lagerstroemia parviflora, Lan Lannea coromandelica, Ter Terminalia tomentosa, Ho Holarrhena
antidysenterica, Lc Lantana camara. R Ranitalli, N Neruiadamar, B Bokarakhari, H Hathinala.
Bars affixed with different letters are significantly different from each other (P < 0.05) (Source:
Pandey et al. 2014a)

was low, and flowering or bud break is preceded by an increase of Ψstem to saturation
values (> −0.2 MPa). Rate and degree of desiccation have been observed more in
dense hill forests having high root competition compared to the widely spaced
savanna trees (Borchert 1994a). At moist lowland sites, drought mortality in trees is
less because roots have access to the water table (Borchert 1994a) and the trees
exchange leaves during the dry season. These site-dependent differences in soil water
availability, not seasonal rainfall, are the principal environmental cause of variation
in tree water status, phenology and distribution of tree species (Borchert 1994a).
60 3  Vegetation Attributes

3.3  Seasonality, Leaf Attributes and Growth Rates

Chaturvedi et al. (2011c) studied the relationships of eight leaf attributes, viz. spe-
cific leaf area (SLA), leaf carbon concentration (LCC), leaf nitrogen concentration
(LNC), leaf phosphorus concentration (LPC), chlorophyll concentration (Chl),
mass-based stomatal conductance (Gsmass), mass-based photosynthetic rate (Amass)
and intrinsic water use efficiency (WUEi) and relative growth rate (RGR) of six
dominant tree species with change in soil moisture content (SMC) under field con-
ditions at four sites in Vindhyan forest of India. We summarize below the important
findings of the study.
Peak SLA occurred during the pre-monsoon season, and LCC and WUEi were
maximum in the post-monsoon season, whereas the rest of the attributes exhibited
maximum values during the monsoon season (Fig. 3.14). Nevertheless, these obser-
vations suggest that the effect of seasonality on leaf attributes can be confounded
with leaf phenology at least to some extent. For example, the pre-monsoon peak in
SLA coincided with leafing in that season, and new leaves are known to have higher
SLA than older leaves. The build-up of cell wall material and chloroplasts in the
older leaves may reduce their SLA (Poorter et al. 2009). Chaturvedi et al. (2011c)
also reported that about 50–80% of mean annual growth occurred during the mon-
soon season (Fig.  3.15); the average monsoon season growth of Shorea robusta,
Anogeissus latifolia, Diospyros melanoxylon, Lagerstroemia parviflora, Buchanania
lanzan and Terminalia tomentosa accounted for, respectively, 52%, 55%, 57%,
64%, 75% and 76% of annual RGR. Prior et al. (2004) also found relatively greater
stem growth of the savanna trees of northern Australia in the wet season. In the cer-
rado of central Brazil also, maximum radial stem growth in forest and savanna trees
occurred in the wet season (Rossatto et al. 2009). In another study on the Venezuelan
semi-deciduous forest, Schöngart et al. (2002) reported that diameter growth of tree
species was prolonged, occurring for 9–10  months of the year, which could be
because this forest experiences a less severe dry season.
In a similar study, Chaturvedi et al. (2013) studied five leaf attributes, viz., spe-
cific leaf area (SLA), leaf dry matter content (LDMC), concentrations of leaf nitro-
gen (leaf N), phosphorus (leaf P) and chlorophyll (Chl); and two physiological
processes, viz., stomatal conductance (gsnet) and photosynthetic rate (Anet);, and
RGR, of four dominant tree seedling species (viz., Buchanania lanzan, Diospyros
melanoxylon, Shorea robusta, and Terminalia tomentosa) on four sites in the
Vindhyan highlands over a 2-year period. This study reported that the leaf attri-
butes and physiological processes were strongly related to soil water availability on
the one hand and seedling growth on the other. Chaturvedi et al. (2013) suggested
that gsnet is the most important variable which accounted for the greatest amount
of variability (62%) in RGR, emphasizing the role of stomatal conductance in
shaping growth patterns across spatial and temporal gradients of soil water avail-
ability. Further, gsnet and SMC together explained 64% variability in RGR of tropi-
cal tree seedlings. Later, Chaturvedi et al. (2014) analysed eight leaf attributes, viz.
specific leaf area (SLA), leaf dry matter content (LDMC), leaf nitrogen concentra-
tion (leaf N), leaf phosphorus concentration (leaf P), chlorophyll concentration
Fig. 3.14  Mean values of leaf traits across sites, seasons and species. SLA specific leaf area, LCC
leaf carbon concentration, LNC leaf nitrogen concentration, LPC leaf phosphorus concentration,
Chl chlorophyll concentration, HN Hathinala, GG Gaighat, HK Harnakachar, RT Ranitali, PM pre-
monsoon, M monsoon, EPM early post-monsoon, LPM late post-­monsoon, Anla Anogeissus lati-
folia, Bula Buchanania lanzan, Dime Diospyros melanoxylon, Lapa Lagerstroemia parviflora,
Shro Shorea robusta, Teto Terminalia tomentosa. Different letters above bars indicate significant
differences after Tukey’s post-hoc test (α = 0.05) (Source: Chaturvedi et al. 2011c)
62 3  Vegetation Attributes

Fig. 3.15  Seasonal variations in leaf attributes and relative growth rates (RGR) in six dry tropical
tree species across four sites. Gsmass mass-based stomatal conductance, Amass mass-­based photosyn-
thetic rate, WUEi leaf water use efficiency, PM pre-monsoon, M monsoon, EPM early post-mon-
soon, LPM late post-monsoon (Source: Chaturvedi et al. 2011c)

(Chl), mass-based photosynthetic rate (Amass), mass-based stomatal conductance


(Gsmass) and intrinsic water use efficiency (WUEi) and three growth attributes (rela-
tive diameter increment (RDI), relative height increment (RHI) and relative growth
rate (RGR)) of the ten dominant tree saplings (viz. Acacia catechu, Anogeissus
latifolia, Boswellia serrata, Buchanania lanzan, Diospyros melanoxylon,
Hardwickia binata, Lagerstroemia parviflora, Lannea coromandelica, Shorea
robusta and Terminalia tomentosa) in the Vindhyan forest and reported the effects
of site, season and species for a period of 2 years. These authors found that the spe-
cies response varied across sites and seasons. They reported variable plasticity in
attributes, differing among species across the SMC gradient. Among the ten sap-
lings, Boswellia serrata, a highly deciduous tree, exhibited the maximum plasticity
in seven functional attributes. Stepwise multiple regressions showed that the 65%
variability in RDI and 67% variability in RGR were due to Gsmass, and for RHI,
3.4  Plant Strategies Against Drought 63

61% variability was due to Amass. SMC and the other attributes, viz. SLA, Chl,
WUEi and LDMC, in combination could account for only ~2–6% of the variability
in RDI, RHI and RGR, which indicates that other traits/factors, not accounted in
this study, are also important in modulating the growth of tree saplings in TDFs.
Chaturvedi et al. (2014) concluded that growth of the tree saplings in the tropical
dry environment is determined by soil moisture, whereas the response of saplings
of different tree species is modulated by alterations in key functional attributes
such as SLA, Chl, WUEi and LDMC.

3.4  Plant Strategies Against Drought

Drought has been defined as the reduction in rainfall or the change of timing or
distribution of rainless or rainfall periods, which can potentially affect the plant and
microbial processes at the community or ecosystem level (Allen et  al. 2017).
Drought is the major stress experienced by the TDF species and tropical forests
have experienced more frequent and more severe droughts specially during the last
decades (Malhi and Wright 2004), and this trend is expected to continue in future
(Dai 2013). High diversity of tree species and the extreme differences in climate
conditions have given rise to a variety of tolerances and strategies in tropical plants
to drought stress (Worbes et al. 2013). Mendivelso et al. (2014) suggested that most
TDF species are tolerant of short-term droughts, but they vary in their sensitivity to
multi-annual droughts. Deciduousness in TDF components is a phenological attri-
bute expressing adaptation to seasonality and drought, resulting in reduced activi-
ties during the unfavourable dry season and resumption of growth during the short
favourable rainy season (Singh and Singh 1992a). The deciduousness of TDF spe-
cies is affected by rainfall, temperature and solar radiation and, in turn, affects intra-
and interannual pattern of water, carbon and energy balance of the TDFs (Bohlman
2010). Numerous publications have reported on the physiological and structural
adaptations and tolerances of trees to drought stress (e.g. Sobrado 1993; Choat et al.
2003, 2005; Baas et  al. 2004; Brodribb and Holbrook 2005; Motzer et  al. 2005;
Engelbrecht et al. 2007; Fallas-Cedeno et al. 2010; Worbes et al. 2013). Studies on
tropical forests in India show that along progressively drier areas, there are a rising
degree and duration of deciduousness during the dry season and an accompanying
decrease in tree height, leaf size, community complexity and the number of families
and genera (Table 3.8). The degree of drought exposure varies widely, depending on
(i) temperature, (ii) soil water and (iii) rooting depth of the trees (van Schaik et al.
1993). Thus, the extent and intensity of seasonal drought would vary with the geo-
graphical location; for example, Costa Rican dry tropics (having low latitude of
~10°N, low annual temperature variability (to <2  °C) and 5  months dry period)
would experience lower seasonal drought than the Indian Vindhyan dry tropical for-
est (having higher latitude of ~24°N, higher annual temperature variation (to
>20 °C) and 8 months dry period (Singh and Kushwaha 2005).
It is clear that the plant species in TDFs are subjected to water stress during the
dry season (Eamus 1999), and the length of dry season is the controlling factor of
64 3  Vegetation Attributes

Table 3.8  Climo-vegetational characteristics of tropical forest in India


Wet Dry
Climate/characteristics evergreen Moist deciduous deciduous Thorn forest
Mean annual temperature 23–27 20–29 20–29 24–29
(°C)
Mean January temperature 15–21 12–26 16–25 13–26
(°C)
Annual rainfall (cm) 240–320 120–300 75–140 25–90
Annual dry months 3–5 4–8 5–8 7–10
Vegetation deciduousness Entirely or Predominantly Entirely Entirely
nearly absent deciduous; deciduous or deciduous
sub-­canopy nearly so
evergreen
Species richness Extremely Rich Poor Extremely
rich poor
Canopy height (m) 40–50 25–40 8–20 <10
Basal area (m2 ha−1) 40–55 35–50 15–20 <5
Number of ligneous layer 4–6 3 2 1
Source: Singh and Singh (1988)

vegetation structure and patterns (Gritti et al. 2010). Due to locational differences in
the extent and intensity of seasonal drought, TDFs are composed of species with a
mosaic of different plant functional traits (PFTs) showing varying adaptations to
seasonal drought (Borchert 2000). In the dry season, drought-deciduous trees which
drop all of their canopy leaves are phenologically drought avoiding, while the ever-
green trees which maintain their canopy leaves during the dry season are drought
tolerant. Comparative studies between coexisting evergreen and deciduous trees in
nitrogen use, water use and photosynthesis in Neotropical (Sobrado 1986, 1991,
1993, 1997; Martin et al. 1994), Australian (Eamus et al. 1999) and Mediterranean
(Mediavilla and Escudero 2003) dry forests or savannas have shown that evergreen
trees use water more conservatively (i.e. have lower hydraulic conductivity and
lower stomatal conductance) than drought-deciduous trees, even in the wet season.
In the TDFs, evergreen and deciduous species coexist. They have different rooting
patterns and experience distinctive water and nutritional regimes (Sobrado and
Cuenca 1979). The ecological success of evergreen and deciduous species in spe-
cific regions of the habitat is partly related to differences in the rooting patterns
(Sobrado and Cuenca 1979). Deep-rooted evergreen species are restricted to areas
with deeper soil profiles, while the shallow-rooted deciduous species are widely
distributed. Thus, during dry periods, while evergreen species are able to use water
stored in the soil profile, the deciduous species run out of water as soon as the water
in the soil surface layers is depleted. Evergreen species have roots penetrating
deeper in the wetter part of the soil enabling the plants to use water during the dry
season (Sobrado and Cuenca 1979). Shallow-rooting deciduous species may have a
growth pattern closely coupled to the wet season when water and nutrients are avail-
able in the soil (Sobrado and Cuenca 1979; Sobrado 1986).
3.5  Relations Between Water Balance, Wood Traits and Phenological Behaviour 65

Paradoxically, evergreen plants in tropical ecosystems produce leaves with


higher phosphorus and nitrogen contents, on leaf area basis, than do deciduous spe-
cies (Medina 1984). Since nitrogen is limiting to plant growth and since leaves
contain a large proportion of the plant’s nitrogen stock, it would be more expensive
to produce an evergreen than a deciduous leaf of the same area. Sobrado (1991)
estimated the relative costs and benefits of evergreen and deciduous leaves by
assessing their construction and maintenance costs as well as their carbon assimila-
tion capability on an area basis. Results of the study showed that the deciduous and
evergreen species, coexisting in TDF, differ greatly in their energy investment for
leaf construction and maintenance. Further, the long residence time of nitrogen is
favourable for evergreen species as their roots occur in nutrient-poor soil microhabi-
tat, while roots of the deciduous species occur in relatively nutrient-rich conditions,
and therefore, they have high nitrogen-use efficiency.
Ishida et al. (2006) examined the physiological and anatomical bases underlying
the contrasting drought responses of drought-avoiding deciduous and drought-toler-
ant evergreen trees in the TDFs of Thailand. They observed that leaves of the decid-
uous species, growing in nutrient-rich soils, gained more carbon during the short,
wet season, while the evergreens occurring in nutrient-poor soils allocated low N
for leaf carbon assimilation, tolerated photoinhibition and exhibited low carbon
gain and conservative water use.

3.5  R
 elations Between Water Balance, Wood Traits
and Phenological Behaviour

Worbes et al. (2013) measured vessel features, wood density, stomatal conductance,
water potential and δ13C in the leaves and wood in the TDF species of the Hacienda
La Pacifica, situated near Cañas in the Guanacaste region, Costa Rica (10°49′N,
85°15′W), and subjected the data to PCA. They selected 12 tree species (Table 3.9)
in order to include several representatives of each of the 4 functional ecotypes (viz.
stem succulent, deciduous, brevi-deciduous and evergreen). The first two axes of the
principal components together explained almost 77% of the total variance
(Fig. 3.16). The first axis was negatively connected with vessel size, vessel area and
carbon isotope content, and water potential of the leaves, i.e., variables which rep-
resented the “hydraulic conductivity”. According to Tyree et al. (1994), vessel size
and vessel density are inversely related demonstrating a trade-off between efficiency
and safety of water transport (large vessels for effective water transport and small
vessels for safety against embolism). It is known that the trees exposed to intensive
drought stress tend to have small vessels and those occurring in humid regions have
large and effective vessels (Baas et al. 2004). The vessel size and density are impor-
tant features of the water transport in the xylem and are related to the negative water
potential in the leaves, and thereby the vessel size is negatively correlated with the
wood density (Bucci et  al. 2004). High wood stability (i.e. wood density) is
66 3  Vegetation Attributes

Table 3.9  The studied species and their phenological behaviour


Species Family Phenology Growth zone types
Bombacopsis quinata Malvaceae SUC 1
Bursera simaruba Burseraceae SUC 1
Cedrela odorata Meliaceae D 2, 4
Cochlospermum vitifolium Bixaceae SUC 1, 2, 3
Enterolobium cyclocarpum Fabaceae BD 1, 2
Ficus sp. Moraceae EV 3
Guazuma ulmifolia Malvaceae BD 1
Luehea candida Tiliaceae D 1, 2
Simarouba glauca Simaroubaceae EV 1, 2
Swietenia macrophylla Meliaceae EV 1, 2
Tabebuia ochracea Bignoniaceae D 2, 3
Tectona grandis Lamiaceae D 2, 4
Source: Worbes et al. (2013)
The classification of growth zone types follows Worbes (1995) with 1 = density variation, 2 =
terminal parenchyma band, 3 = pattern of parenchyma and fibre bands, 4 = ring porous vessel
distribution
SUC stem succulent, D deciduous, BD brevi-deciduous, EV evergreen, following Borchert 1994a
and the growth zone types of the wood

necessary to establish a high tension in the water conduction system (Borchert and
Pockman 2005). On the second axis of the PCA, stomatal conductance and δ13C of
the wood were oppositely connected with each other and represented “control of
transpiration and water loss”.
Stem-succulent species shed the leaves at the end of the rainy season, when all
other species still maintain their foliage. The leaf shedding represents a sensitive
stomatal control mechanism. Stem succulents close the stomata even under moder-
ate external stress situation at midday in the rainy season, when other trees exhibit
maximum transpiration (see also Brodribb and Holbrook 2005; Chapotin et  al.
2006). Although stem succulents are widespread in several tropical regions as colo-
nizers or components of secondary vegetation, their leaf-fall behaviour is not well
studied (but see Olivares and Medina 1992; Borchert 1998; Schöngart et al. 2002;
Fallas-Cedeno et al. 2010). The climatic vagaries have prompted the TDF species to
develop several adaptational features, such as deciduousness, interesting phenologi-
cal patterns, pronounced seasonal variations, e.g. in tree water status, etc.
In conclusion, the TDF vegetation exhibits high structural and physiological
diversity in life forms and high levels of endemism. In India, the dry deciduous for-
est is the most extensive forest type and comprises a variety of plant communities.
Although the forest is relatively poor in species richness, the species combinations
exhibit variations on different sites, showing a unique level of diversity. Studies in
Indian TDF have found positive association between diversity and productivity.
Small variations in the environmental variables have been found to determine the
3.5  Relations Between Water Balance, Wood Traits and Phenological Behaviour 67

Fig. 3.16  Principal component analysis including functional ecotype and the measured variables,
stomatal conductance, water potential of the leaves, δ13C of the wood, δ13C of the leaves, vessel
size, vessel area, vessel density and wood density. For stomatal conductance and leaf δ13C, the data
from the rainy season were used since the stem-succulent species were leafless in the dry period
(Source: Worbes et al. 2013)

species distribution resulting into patchiness in the composition of the TDF.  The
differences in the dominance of species in the overstorey and understorey and also
variations in diversity of these two life stages indicated the dynamic nature of the
forests. Studies have reported that the community phenology of TDF is character-
ized by synchronous, asynchronous and compensatory dynamics, partly due to sea-
sonal fluctuations in the abiotic constraints. Leafing and flowering in many plants
occur around or before the beginning of the rainy season and precede fruiting. The
community takes full advantage of the rainy season, by producing and recruiting
plants through germination. Both flowering and fruiting activities were quite stag-
gered and prolonged at the community level, providing food to pollinators, frugivo-
res and seed predators, which in return serve as pollen transfer and seed dispersal
agents. Studies also indicated that the growth of species in the TDF is markedly
influenced by soil moisture availability and species tend to adapt themselves by
modulating the leaf traits.
Chapter 4
Plant Traits and Regeneration

Plant functional traits represent the morphological, physiological or phenological


features which exhibit strategies that determine the response of plants to environ-
mental conditions and affect ecosystem services. These traits could be expressed at
different levels, e.g. total plant traits, leaf traits, stem traits, root traits and the regen-
erative traits. Occurring in harsh and relatively unpredictable environment and on
nutrient-poor soils, the tropical dry-deciduous species are endowed with a variety of
functional traits, at all levels, which enable them to survive and grow under uncer-
tain environments and adapt to the changing climate. In this chapter we provide a
brief description of plant functional traits, plant functional types, seed and seedling
ecology, response to changes in ambient CO2, soil water and nutrients and provide
a discussion on regeneration of tropical dry-deciduous forest.

4.1  Plant Functional Traits

Gitay and Noble (1997) defined plant functional traits (PFTs) as “plant characteris-
tics that respond to the dominant ecosystem processes”. Later, Geber and Griffen
(2003) described a functional trait as the phenotypic character influencing the fit-
ness of the organism. According to Pérez-Harguindeguy et al. (2013), PFTs are the
morphological, physiological and phenological plant characteristics which exhibit
ecological strategies and determine the response of plants to variations in environ-
mental factors, influence other trophic levels and affect the ecosystem properties.
The PFTs are functionally related to the plant population and ecosystem processes
(Hillebrand and Matthiessen 2009). According to Fajardo and Siefert (2016), PFTs
are a reflection of the plant adaptation to variations in biotic interactions and their
physical environment. A shift in the species composition can lead to shift in the
functional trait composition, which can influence the functioning of the ecosystem,
and this change is affected not only by the traits of the declining or disappearing
species, but also by the traits of species replacing them (Díaz et al. 2003; Suding

© Springer Nature Singapore Pte Ltd. 2017 69


J. S. Singh, R. K. Chaturvedi, Tropical Dry Deciduous Forest: Research Trends
and Emerging Features, https://doi.org/10.1007/978-981-10-7260-4_4
70 4  Plant Traits and Regeneration

et al. 2006; Lavorel et al. 2007). Thus, the correlations between the ecosystem func-
tion and the structural and ecophysiological traits could possibly predict the future
changes in ecosystem services due to changes in functional trait composition.
Pérez-Harguindeguy et al. (2013) have suggested for the quantification of the inter-
specific trait variations, particularly for the traits which are able to easily predict the
processes at plant and ecosystem level and for the traits which are relatively easy to
measure. Recently, the PFTs important for the ecosystem processes in tropical dry
forest (TDF) components have been reviewed by Chaturvedi et al. (2011d).
Several investigations have documented that despite having low species richness,
TDFs contain high diversity of phenological and functional traits (Eamus 1999;
Sandquist and Cordell 2007). For example, different growth forms in TDFs exhibit
adaptive strategies in response to drought and grazing by different kinds of herbi-
vores. While the perennating tissues in different life forms are adaptations for the
survival of plant species subjected to unpredictable disturbances, plant height is
considered as an important attribute of fecundity and competitive vigour (Cornelissen
et al. 2003). Among the other important adaptive strategies, clonality, spininess and
flammability are important vegetative plant traits against browsing and fire (Saha
and Howe 2003; Raherison and Grouzis 2005). The plant communities in TDF have
drought-avoiding strategies early in succession, and abundance of evergreen strate-
gies, with larger seeds, late in succession (Lohbeck et al. 2015a).
The ecological investigations in TDFs nowadays are mostly concerned for the
understanding of relationships between plant physiological processes and plant (or
leaf) morphological and phenological traits (Chaturvedi et al. 2011d), for example,
relationships of leaf life span with photosynthesis or leaf conductance, associations
of specific leaf area with photosynthesis (e.g. Reich et al. 1998a) and the tree height
and photosynthesis relations (Zhang et al. 2009; Brienen et al. 2010). Quantitative
information about the important ecophysiological relationships on tropical decidu-
ous plant species is currently available with respect to physiological processes/attri-
butes or functional relationships such as (i) nitrogen and water use efficiency
(Sobrado 1991); (ii) photosynthetic attributes and associated traits (Kitajima et al.
1997; Niinemets and Tenhunen 1997; Niinemets 1999; Niinemets et  al. 2009;
Posada et al. 2009); (iii) water and trait relations (Olivares and Medina 1992; Medina
and Francisco 1994; Franco et al. 2005; Wright et al. 2007; Gotsch et al. 2010); (iv)
light-dependent leaf trait variations (Rozendaal et al. 2006; Markesteijn et al. 2007);
(v) nitrogen fixation and capacity of nitrate/ammonium assimilation (Schulze et al.
1991; Högberg 1992; Högberg and Alexander 1995; Freitas et al. 2010); (vi) effects
of drought conditions on leaf conductance, leaf water status and photosynthesis
(Eamus 1999; Brodribb et al. 2003; Brodribb and Holbrook 2003a, b); (vii) herbi-
vore damage (Janzen 1970; Coley and Barone 1996; Coley 1998; Arnold and
Asquith 2002; Campo and Dirzo 2003; Brenes-Arguedas et al. 2009); and (viii) leaf
flushing and flowering phenology (Opler et al. 1980; Seghieri et al. 2009; Hayden
et al. 2010; Fallas-Cedeño et al. 2010). However, still there are many shortcomings
in the existing data sets (for details see Díaz et al. 2004; Wright et al. 2007).
4.1  Plant Functional Traits 71

4.1.1  Leaf Traits

Leaf habit has been considered as the most common FT for categorizing species
into ecologically relevant PFTs (e.g. deciduous vs. evergreen). However, Powers
and Tiffin (2010) suggested that the leaf habit alone has a little importance for cat-
egorizing PFTs in the TDFs of Costa Rica. Westoby (1998) and Weiher et al. (1999)
suggested that a database of FTs should include all traits which are important for the
acquisition and use of resources; the three such traits include SLA (specific leaf
area, i.e. the ratio of leaf area to dry mass), LDMC (leaf dry matter content, i.e. the
ratio of leaf dry mass to saturated fresh mass) and LNC (leaf nitrogen concentra-
tion) (Cunningham et al. 1999; Reich et al. 1999; Wilson et al. 1999; Markesteijn
et al. 2007). Among these three traits, SLA and LNC have been considered to have
a significant influence on primary productivity and cycling of nutrients at the eco-
system level (Reich et al. 1992; Cornelissen et al. 1999a; Aerts and Chapin 2000),
and therefore have already been collected and compiled into a database (Wright
et al. 2004a). Reich et al. (1997) argued that a SLA and LNC together are capable
to accurately determine and predict the maximum photosynthetic rate in wide range
of plant species. A similar suggestion has been given for LDMC (Ryser and Urbas
2000). Studies have also reported that SLA, LNC and LDMC are the traits com-
monly associated with a trade-off between biomass production capacity and nutri-
ent conservation efficiency at the whole plant level (Grime 1997; Poorter and
Garnier 1999; Lavorel et al. 2007).
Data available on the leaf traits of tree species in TDFs are scarce. Table  4.1
shows the data on leaf traits of tree species in tropical deciduous forests summarized
by Chaturvedi et al. (2011d). The leaf traits in TDFs show high interspecific and
interregional variations (Table 4.1). However, the interspecific variations in most of
the cases are greater than the interregional variation (Table 4.1). Generally, the spe-
cies growing in desert, tropical evergreen forests, temperate evergreen forests and
tundra vegetation exhibit lower SLA than observed in the TDF. While, in TDF spe-
cies, few leaf trait values are close to those observed in tropical evergreen species
(e.g. CC, Villar and Merino 2001), other traits exhibit conspicuous differences
between the deciduous and evergreen leaf habits. In the TDF, the deciduous trees
have higher SLA, LNC and photosynthetic rates as observed in evergreen trees (see
references in Powers and Tiffin 2010). The average values of leaf life span (LL) for
desert species (14 months), tropical evergreen species (20 months), temperate ever-
green species (44  months), temperate deciduous species (13  months) and tundra
species (7 months) (Reich et al. 1998a) are higher than the value (6 months) reported
by Prior et al. (2003) for the deciduous species of Australia. Therefore, in TDF spe-
cies, the mean LL is shorter, and LMA is lower, as compared to evergreen woody
species (Wright et al. 2004b). In Argentina, the desert species showed higher aver-
age leaf thickness and leaf water content (LWC) compared to those observed for the
deciduous species (Vendramini et al. 2002). The stomatal conductance (gc) of TDF
species also exhibited substantial variations among the regions and species
(Table 4.1). However, on the contrary, area-based leaf maximum photosynthetic rate
72 4  Plant Traits and Regeneration

Table 4.1  List of leaf traits of tree species in tropical deciduous forests
Species/forest Mean value Range ±1SE
SLA (mm2 mg−1)
Dry forest of Panama 8.5 (n = 7) 5.0–12.5 0.9
Deciduous forest of Australia 9.9 (n = 2) 9.0–10.7 0.8
Dry forest of Costa Rica 10.2 (n = 87) 5.5–23.0 3.6
Deciduous forest of India 10.9 (n = 54) 6.4–19.1 0.4
Deciduous forest of central-western Argentina 11.2 (n = 16) 4.2–25.4 1.5
Deciduous forest of India 11.7 (n = 6) 7.6–14.4 1.3
Dry forest of Panama 12.6 (n = 6) 7.3–21.9 1.1
Deciduous forest of India 11.4 (n = 7) 8.8–16.9 1.0
Deciduous forest of central-western Argentina 15.0 (n = 23) nk 1.2
Dry forest of Venezuela 18.6 (n = 6) 11.6–26.8 2.9
Dry deciduous forest of Bolivia 20.7 (n = 12 11.5–26.9 6.0
LA (cm2)
Dry monsoon forest of Australia 40.0 (n = 4) 31.0–64.0 8.0
Dry deciduous forest of Bolivia 45.1 (n = 12) 2.5–201 13.0
Deciduous forest of India 75.9 (n = 7) 26.4–143 17.6
Chl (mg g−1)
Deciduous forest of India 0.93 (n = 7) 0.72–1.26 0.08
Deciduous forest of India 1.27 (n = 6) 1.04–1.83 0.12
LDMC (% saturated wt.)
Deciduous forest of central-western Argentina 35.9 (n = 13) 13.4–47.4 3.6
Dry deciduous forest of Bolivia 32.2 (n = 12) 23.0–48.0 9.3
Nmass (%)
Savanna of Africa 1.3 (n = 4) 1.2–1.6 0.2
Deciduous forest of Australia 1.4 (n = 2) 1.2–1.6 0.2
Dry forest of Brazil 1.8 (n = 3) 1.4–2.1 0.3
Deciduous forest of Australia 1.8 (n = 27) 0.8–4.0 0.3
Deciduous forest of India 2.0 (n = 54) 0.9–3.2 0.1
Deciduous forest of Central Ethiopia 2.1 (n = 7) 1.7–3.2 0.3
Deciduous forest of India 2.2 (n = 6) 1.9–2.5 0.1
Dry forest of Panama 2.3 (n = 8) nk 0.2
Dry forest of Costa Rica 2.3 (n = 87) 1.2–3.6 0.7
Deciduous forest of Australia 2.4 (n = 6) 1.1–3.2 0.4
Dry forest of Panama 2.4 (n = 6) 1.5–3.8 0.3
LL (mo)
Deciduous forest of Australia 6.3 (n = 6) 4.8–8.2 0.5
Dry forest of Venezuela 8.4 (n = 6) 6.0–10.0 1.3
Aarea (μmol m−2 s−1)
Lowland forest of Panama 12.9 (n = 16) 9.7–18.3 0.7
Dry forest of Panama 13.4 (n = 6) 7.8–19.9 1.8
Deciduous forest of Australia 13.9 (n = 6) 9.6–18.7 1.3
Deciduous forest of Australia 14.8 (n = 2) 14.0–15.6 0.8
(continued)
4.1  Plant Functional Traits 73

Table 4.1 (continued)
Species/forest Mean value Range ±1SE
Amass (nmol g−1 s−1)
Dry forest of Panama 122.3 (n = 7) 58.1–200.4 17.5
Deciduous forest of India 130.6 (n = 6) 70.6–179.1 17.5
Deciduous forest of Australia 143.8 (n = 2) 126.4–161.3 17.5
Deciduous forest of Australia 175.8 (n = 6) 129.0–250.0 16.7
Rdarea (μmol m−2 s−1)
Deciduous forest of Australia 0.61 (n = 2) 0.62–0.60 0.01
Dry forest of Panama 1.90 (n = 6) 2.90–0.80 0.20
gc (mmol m−2 s−1)
Dry forest of Venezuela 204.8 (n = 5) 140.7–274.8 23.0
Deciduous forest of India 337.2 (n = 6) 252.1–406.2 26.2
Dry forest of Panama 456.6 (n = 7) 199.3–670.2 91.4
Dry forest of Panama 573.4 (n = 6) 249.1–1306.6 127.5
Deciduous forest of Australia 680.0 (n = 2) 620.0–750.0 65.0
WUEi (μmol mol−1)
Deciduous forest of India 35.3 (n = 6) 25.6–48.7 3.5
Savanna of central Venezuela 37.5 (n = 3) 30.4–46.0 4.6
Tropical dry forest of Venezuela 44.8 (n = 6) 36.0–53.0 2.7
Pmass (%)
Savanna of Africa 0.07 (n = 4) 0.05–0.09 0.01
Dry forest of Costa Rica 0.11 (n = 87) 0.06–0.20 0.04
Deciduous forest of Australia 0.15 (n = 6) 0.08–0.19 0.02
Deciduous forest of Central Ethiopia 0.18 (n = 7) 0.15–0.30 0.03
Deciduous forest of India 0.19 (n = 6) 0.16–0.26 0.01
Deciduous forest of India 0.21 (n = 54) 0.08–0.52 0.01
LSCmax (mmol m−1 s−1 MPa−1)
Lowland forest of Panama 53.8 (n = 16) 15.8–120.7 6.8
Dry deciduous forest of Bolivia 42.7 (n = 12) 8.0–81.0 12.3
Tropical dry forest of Costa Rica 19.3 (n = 3) 15.0–25.0 11.2
E (mmol m−2 s−1)
Dry forest of Venezuela 6.5 (n = 4) 5.2–7.7 0.5
Deciduous forest of Australia 17.2 (n = 2) 16.9–17.5 0.3
CC (g glu.g−1)
Deciduous forest of Australia 1.2 (n = 2) 1.1–1.3 0.05
Dry forest of Charallave, Venezuela 1.5 (n = 7) 1.4–1.6 0.02
LWC (%)
Dry forest of Costa Rica 58.8 (n = 87) 40.2–73.9 8.0
Deciduous forest of central-western Argentina 60.4 (n = 16) 38.0–83.0 3.0
Cmass (%)
Savanna of Africa 37.3 (n = 4) 35.8–39.7 1.3
Deciduous forest of India 42.8 (n = 54) 32.0–47.7 0.4
Dry forest of Brazil 43.0 (n = 2) 42.7–43.3 0.6
(continued)
74 4  Plant Traits and Regeneration

Table 4.1 (continued)
Species/forest Mean value Range ±1SE
Deciduous forest of India 43.3 (n = 6) 41.6–44.3 0.4
Dry forest of Costa Rica 45.9 (n = 87) 37.6–52.3 3.3
Deciduous forest of India 46.3 (n = 8) 44.8–47.4 0.7
Dry forest of Panama 49.0 (n = 8) nk 0.8
Deciduous forest of Australia 49.2 (n = 27) 0.41–57.6 1.4
Namass (%)
Deciduous forest of India 0.07 (n = 25) nk 0.01
Kmass (%)
Deciduous forest of Central Ethiopia 1.4 (n = 7) 1.0–1.9 0.2
Camass (%)
Deciduous forest of India 1.2 (n = 8) 0.6–2.3 0.3
Thickness (mm)
Deciduous forest of Australia 0.3 (n = 6) 0.2–0.5 0.04
Deciduous forest of central-western Argentina 0.4 (n = 16) 0.2–0.6 0.03
n number of species, nk not known
SLA specific leaf area, LA leaf area, Chl chlorophyll concentration, LDMC leaf dry matter content,
LSCmax maximum leaf-specific hydraulic conductivity, Aarea area-based leaf maximum photosyn-
thetic rate, Amass mass-based leaf maximum photosynthetic rate, Rdarea area-based dark respiration
rate, E leaf transpiration rate, gc leaf stomatal conductance, WUEi intrinsic water use efficiency,
CC leaf construction cost, LWC leaf water content, Nmass mass-based nitrogen concentration, LL
leaf life span, Pmass mass-based phosphorus concentration, Cmass mass-based carbon concentration,
Namass mass-based sodium concentration, Kmass mass-based potassium concentration, Camass mass-­
based calcium concentration. For references see Chaturvedi et  al. (2011d) (Source: Chaturvedi
et al. 2011d)

(Aarea), mass-based nitrogen concentration (Nmass), mass-based phosphorus concen-


tration (Pmass) and mass-based carbon concentration (Cmass) showed little differences
with respect to regions and species (Table  4.1). According to Eamus (1999), the
Nmass in deciduous trees is generally greater compared to evergreen species and con-
sequently shows greater light-saturated assimilation rate. Area-based dark respira-
tion rate (Rdarea) observed in TDF species is lesser compared to that of desert species
(Reich et al. 1998a). Santiago (2003) reported greater average value of maximum
leaf specific hydraulic conductivity (LSCmax) in the evergreen species compared to
the deciduous species of Costa Rica.
Bertiller et al. (2006) observed that the leaf traits associated with carbon fixation
and the decomposition pathway exhibited significant variation with humidity.
While, in deciduous species, plants are leafless during the dry period and the Amass
(the photosynthetic assimilation rate measured under high light, ample soil moisture
and ambient CO2) approaches to zero, the decline in Amass in mature evergreen tree
species (15–50%) and semi-deciduous species (25–75%) is very low under low
water conditions (Eamus 1999). Wright et al. (2004b) reported that the mean N/P
ratio in leaves of deciduous woody species is lower compared to that for evergreens.
Majority of the leaf traits exhibit plasticity which is comparatively lower in the dry
4.1  Plant Functional Traits 75

forest and higher in the moist forest tree species (Markesteijn et al. 2007) and there-
fore follow a gradient of soil moisture.

4.1.2  Stem and Root Traits

A minimum seasonal water potential is required for maintaining the physiological


activity of TDF species (Bhaskar and Ackerly 2006). Also, FTs which determine
acquisition of limiting nutrients, for example, extensive root foraging and/or the
association with mycorrhizal fungi, can be important for carbon dynamics (de Deyn
et al. 2008). Important functional traits of stem comprise stem-specific density (SSD)
or wood specific gravity which exhibit direct association with the availability of soil
water (Preston et  al. 2006) and above-ground biomass productivity together with
species maximum height (Baker et  al. 2008). According to Borchert (1994a), the
increasing wood density is related to a decline in water content of wood, and for the
stem water storage potential, SSD is considered to be important for the rehydration
processes in species exhibiting variations in the duration of deciduousness (Eamus
and Prior 2001); it has also been reported as a good predictor of resistance to drought-
driven embolism (Sperry 2003). Higher SSD is commonly associated with smaller
leaf and twig sizes (Westoby and Wright 2006). Twig dry matter content (TDMC)
and bark thickness are important traits for plants to survive the fire events.
For the drought tolerance and resistance, the plant species in TDF possess vari-
ous other important structural traits such as plant stature (herb, shrub, tree), leaf area
index (LAI) or crown depth, bark thickness, rooting depth, tree/root architecture,
tree life span, plant architecture (DBH to height curve) and size/height. Other stem
and root traits such as nitrogen content (Nmass), phosphorus content (Pmass), carbon
content (Cmass), sodium content (Namass), potassium content (Kmass) and calcium con-
tent (Camass) are also related to the physiology of the TDF tree species (Tables 4.2
and 4.3). Compared to other traits listed in Tables 4.2 and 4.3, SSD exhibits high
variations among regions and among species (e.g. dry forest of Costa Rica, 0.19–
1.20 g cm−3). Borchert (1994a) also reported a higher mean value of SSD in decidu-
ous trees as that in evergreen trees. Traits associated with regeneration, particularly
resprouting capacity, are very important for the TDF experiencing a high population
load (pers. observation).

4.1.3  Reproductive Traits

Reproductive traits, for example, number and biomass of fruits, seed weight, seed
viability, time required for seed germination, time of seed production, number of
seeds, dispersal distance, specialized pollination/dispersal mechanism, maturity
age/size and reproductive phenology, are very important for the success of a TDF
species (Table 4.4). The dispersal ability of TDF species is strongly affected by the
76 4  Plant Traits and Regeneration

Table 4.2  List of stem traits of tree species in tropical deciduous forests
Species/forest Mean value Range ±1SE
SSD (g cm−3)
Dry deciduous forest of Bolivia 0.37 (n = 12) 0.19–0.52 0.11
Dry forest of Costa Rica 0.40 (n = 18) 0.16–0.78 0.04
Deciduous forest of Panama 0.51 (n = 16) 0.35–0.70 0.02
Deciduous forest of India 0.58 (n = 7) 0.37–0.75 0.04
Dry forest of Costa Rica 0.72 (n = 26) 0.19–1.20 0.05
Deciduous forest of Mexico 0.73 (n = 21) 0.27–1.39 0.09
Nmass (%)
Dry forest of Brazil 0.25 (n = 2) 0.20–0.30 0.10
Savanna of Africa 0.25 (n = 4) 0.16–0.45 0.10
Deciduous forest of Central Ethiopia 0.28 (n = 7) 0.20–0.34 0.03
Pmass (%)
Deciduous forest of Central Ethiopia 0.02 (n = 7) 0.01–0.03 0.04
Savanna of Africa 0.03 (n = 4) 0.02–0.04 0.01
Cmass (%)
Savanna of Africa 37.6 (n = 4) 36.5–38.8 0.31
Dry forest of Brazil 43.0 (n = 2) 42.7–43.3 0.61
Deciduous forest of India 47.3 (n = 8) 46.5–49.2 0.41
Namass (%)
Deciduous forest of India 0.08 (n = 25) nk 0.03
Kmass (%)
Deciduous forest of Central Ethiopia 0.22 (n = 7) 0.14–0.46 0.06
Camass (%)
Deciduous forest of India 0.35 (n = 8) 0.07–1.12 0.16
Bark Cmass (%)
Deciduous forest of India 38.8 (n = 8) 34.4–41.2 1.13
Bark Camass (%)
Deciduous forest of India 2.17 (n = 8) 1.00–4.30 0.54
n number of species, nk not known
SSD stem-specific density, Nmass mass-based nitrogen concentration, Pmass mass-based phosphorus
concentration, Cmass mass-based carbon concentration, Namass mass-based sodium concentration,
Kmass mass-based potassium concentration, Camass mass-based calcium concentration. For refer-
ences see Chaturvedi et al. (2011d) (Source: Chaturvedi et al. 2011d)

seed size (Khurana et al. 2006); smaller seeds exhibit greater dispersal ability as
observed in heavier and larger seeds. However, seedlings developed by heavier and
larger seeds possess greater competitive ability enabling them to better establish and
survive under environmental stresses such as competition, moisture conditions,
shading, disturbances, defoliation and herbivory. Khurana et al. (2006) reported that
the structure of the TDF is generally determined by the species with medium- to
large-sized seeds.
4.1  Plant Functional Traits 77

Table 4.3  List of root traits of tree species in tropical deciduous forests
Species/forest Mean value Range ±1SE
Nmass (%)
Deciduous forest of India 0.9 (n = 25) nk 0.1
Dry forest of Brazil 1.2 (n = 3) 0.8–1.5 0.3
Pmass (%)
Deciduous forest of India 0.06 (n = 25) nk 0.02
Cmass (%)
Dry forest of Brazil 42.7 (n = 3) 39.7–47.4 0.61
Namass (%)
Deciduous forest of India 0.09 (n = 25) nk 0.02
Kmass (%)
Deciduous forest of India 0.31 (n = 25) nk 0.06
Camass (%)
Deciduous forest of India 0.56 (n = 25) nk 0.13
n number of species, nk not known
Nmass mass-based nitrogen concentration, Pmass mass-based phosphorus concentration, Cmass mass-­
based carbon concentration, Namass mass-based sodium concentration, Kmass mass-based potassium
concentration, Camass mass-based calcium concentration. For references see Chaturvedi et  al.
(2011d) (Source: Chaturvedi et al. 2011d)

Table 4.4  List of reproductive traits of tree species in tropical deciduous forests. For references
see Chaturvedi et al. (2011d)
Species/forest Mean value Range ±1SE
Fruit no. per tree
Dry forest of Costa Rica 271 (n = 7) 4–1182 219
Fruit wt. (g)
Dry forest of Costa Rica 108 (n = 7) 1.3–385 88.2
Seed wt. (mg)
Deciduous forest of Mexico 184 (n = 22) 0.66–1622 116
Deciduous forest of India 316 (n = 37) 0.10–2224 84.6
Deciduous forest of India 1229 (n = 99) 1.0–20,000 670
Seed viability (days)
Deciduous forest of India 256 (n = 99) 5–720 38.1
Period for seed germination (days)
Deciduous forest of India 16.9 (n = 98) 8–45 1.4
Source: Chaturvedi et al. (2011d)
n number of species

4.1.4  F
 unctional Trait Syndrome for Maintaining Growth
in Seasonally Dry Environments

The TDF species possess important functional traits for conserving water, which
help their survival in the environment experiencing seasonal drought. For ­conserving
water and minimizing transpiration, plants generally reduce LL and modulate leaf
78 4  Plant Traits and Regeneration

FTs in a way that the plant can maintain photosynthesis at a high rate during the
short period of water availability during the rainy season. TDF species generally
possess higher SLA, LNC, Nmass, gc and light-saturated photosynthetic rate and a
lower LMA, LL and LSCmax compared to that observed in the tree species of other
biomes (Chaturvedi et al. 2011d). SLA and LNC exhibit strong association with LL
and, together, accurately predict the maximum photosynthetic rate across the species
(Reich et al. 1997). Santiago et al. (2004) observed that nitrogen content per unit
mass and light- and CO2-saturated photosynthetic rate per unit mass in the leaves of
upper canopy decline with annual precipitation, while leaf thickness becomes higher
and SLA becomes lower in lowland Panamanian forests. Relatively high gc (205–
680  mmol m−2 s−1), combined with relatively short LL (6–8  months) and LSCmax
(19–54 mmol m−1 s−1 MPa−1), indicates a selective advantage to plants experiencing
seasonally dry environments. A high gc may be important for the deciduous species
exhibiting short LL for the maximum utilization of resources in a limited duration
when the available soil moisture is favourable. A review of literature by Markesteijn
et al. (2010) suggests that plant tolerance to water stress is codetermined by a range
of FTs such as high cavitation resistance, strong stomatal control or the maintenance
of tissue turgor pressure at low leaf water potentials. Li et al. (2009) observed the
decline in leaf relative water content, SLA, leaf area ratio and WUE, whereas
increase in the biomass allocation to roots in Sophora davidii seedlings, which
resulted in a higher root/stem mass ratio under drought conditions.

4.2  Plant Functional Types

For understanding the potential response of plants and the ecosystems to the changes
in global environmental conditions, species groups with shared characteristics or
functional traits, known as plant functional types (PFTs), are needed to be identified
and investigated. Hoorens et al. (2010) suggested that groupings of plant species
according to their functional traits can provide information about the relative contri-
bution of different PFTs to the total plant biomass of the ecosystem. Interspecific
variation in functional traits can help in the classification of plant species into PFTs
(Von Willert et al. 1990, 1992; Díaz and Cabido 1997; Lavorel et al. 1997; Westoby
1998; Gitay et  al. 1999; Semenova and van der Maarel 2000; Powers and Tiffin
2010). The grouping of species according to their common biological parameters
minimizes a large variety of species into a small number of functional groups that
can help for making predictions about the functioning of organisms (Duru et  al.
2009). However, the current information about the plant physiological attributes is
restricted which limits the identification and characterization of tree PFTs through
deductive or inductive approaches (Gitay and Noble 1997). This is especially true
for TDFs, where there can be occurrence of a mosaic of different PFTs exhibiting
wide variations in adaptations to seasonal drought (Borchert 2000).
Since the morphological and physiological traits of TDF species vary according
to leaf habit (i.e. evergreen, deciduous or semi-deciduous), this important plant char-
4.3  Seed Size: A Key Trait Determining Species Distribution 79

acteristic can be utilized for identification of ecologically meaningful PFTs (Powers


and Tiffin 2010). Borchert (1994a) has identified different PFTs, on the basis of leaf
phenology and wood density in the TDF of Costa Rica, which vary from deciduous
hardwood and water-conserving woody trees in dry upland forest region to evergreen
softwood trees commonly found in moist lowland sites. Baker et al. (2003) classified
plant species in the semi-deciduous forest of Ghana into two groups: (i) dry forest
pioneers and (ii) wet forest pioneers, particularly on the basis of plant growth rate
variations according to changes in soil moisture conditions. In Mexican cloud forest,
Saldaña-Acosta et al. (2008) classified 33 tree species according to their SLA, height
at maturity, wood density and seed mass into two functional groups.
Sagar and Singh (2003) classified trees growing in Indian TDF according to their
leaf size, leaf texture, deciduousness and bark texture and observed that the per cent
of species and the importance values were greater for medium or low deciduous
classes compared to the highly deciduous class, exhibiting a trade-off between the
water loss and the duration of dry matter synthesis. Further, the tree vegetation was
classified also by the predominance of the species with a combination of small-­
sized leaves (below 201 cm2 leaf area), medium leaf texture, rough bark texture and
medium deciduousness (2–3  months deciduous). Nevertheless, these authors
reported that the distribution of species among the trait categories was not consis-
tent across the five sites examined (Fig. 4.1). The sites represented different levels
of disturbance, and evidently disturbance did not affect the predominant traits,
although the composition of species and their relative importance exhibited varia-
tions along the disturbance gradient.
According to Sagar and Singh (2003), the predominant phenotypic traits com-
monly exhibited by TDF, notwithstanding site-wise differences such as disturbance
intensity, were small leaf size, rough bark texture, medium leaf texture and medium
deciduousness. In this study, instead of highly deciduous trait, both the per cent of
species (Fig. 4.1) and importance value (Fig. 4.2) were larger for medium to less
deciduous trait categories. Although all species possessed one or more of the ideal
traits (viz. small leaf size, rough bark texture, medium leaf texture and medium
deciduousness), only five species possessed all four of the most desirable traits, and
none of these species dominated a site. As many as 13 species possessed a combina-
tion of three most desirable characters, and 15 species possessed a combination of
two of these characteristics. Thus, a considerable amount of redundancy (i.e. high
species richness) occurs within a phenotypic trait category.

4.3  Seed Size: A Key Trait Determining Species Distribution

Species having higher dispersal capacity should be more widely distributed com-
pared to those with low dispersal ability (Hanski et al. 1993). Dispersal capacity in
turn is significantly affected by the seed size of the plant species (Rees 1995).
Dispersal agents readily transport smaller seeds, but as discussed earlier, larger and
heavier seeds produce stronger seedlings which can survive under stressful
80 4  Plant Traits and Regeneration

Fig. 4.1  Distribution of species into various phenotypic traits; (a) leaf size, (b) leaf texture, (c)
deciduousness, (d) bark texture. Thin bars represent ±1SE. HT, KH, MJ, BK, KT and All, respec-
tively, stand for Hathinala, Khatabaran, Majhauli, Bhawani Katariya, Kota and all sites (Source:
Sagar and Singh 2003)

conditions (Venable and Brown 1988; Greene and Johnson 1993; McConnaughay
and Bazzaz 1987; Leishman and Westoby 1994a, b; Hammond and Brown 1995;
Armstrong and Westoby 1993). Further, TDFs usually have a greater proportion of
wind-dispersed species compared to the moist forest (Bullock 1995; Gentry 1995).
For example, in Neotropical dry forest, wind-dispersed genera, for example,
4.3  Seed Size: A Key Trait Determining Species Distribution 81

Fig. 4.2  Mean relative importance values for various phenotypic traits; (a) leaf size, (b) leaf tex-
ture, (c) deciduousness, (d) bark texture. Thin bars represent ±1SE. HT, KH, MJ, BK, KT and All,
respectively, stand for Hathinala, Khatabaran, Majhauli, Bhawani Katariya, Kota and all sites
(Source: Sagar and Singh 2003)

Serjania and Lonchocarpus, are highly abundant and account for about one-third to
one-quarter of tree species (Gentry 1995). In the dry forest of Charallave, Venezuela,
about 30.5% of phanerophytes, such as Ceiba pentandra, Tabebuia ochracea and
Terminalia spp., are anemochorous (Wikander 1984). However, in a Guanacaste dry
forest, Janzen (1988a) reported that about 65% of the tree species were commonly
dispersed by animals, whereas only 25% by wind. Mammals and birds, which are
82 4  Plant Traits and Regeneration

Fig. 4.3  Mean seed mass of species as a function of (a) their respective dispersal mode and (b)
their shade tolerance across five sites in tropical dry forest. Bars represent ±1S.E. “Intolerant” =
relatively shade-intolerant, “moderate” = moderately shade-tolerant, “tolerant” = highly shade-­
tolerant (Source: Khurana et al. 2006)

characteristic of dry forest, rarely visit pastures having little or no remnant woody
vegetation (Guevara 1986), while the secondary succession on pastures which are
abandoned may be enhanced after the introduction of tree seedlings of the species
which are dispersed by animals (Gerhardt 1993). Janzen (1988a) reported that in a
12-year-old abandoned pasture in Guanacaste National Park, at a distance of 200 m
from a species-rich mature forest, the natural regeneration was commonly domi-
nated by wind-dispersed tree species. Compared to the fruits of wind-dispersed spe-
cies, the fruits of animal-dispersed species contain four to six times more proteinase
inhibitors in order to prevent dispersers from collecting premature fruits, but as the
fruits reach ripened stage, the quantity of inhibitors becomes low or negligible
which encourage animals for the removal of fruits for dispersal (Hebber et al. 1993).
Khurana et al. (2006) have analysed the relationships among the size of seed,
distribution of plants and their abundance in a TDF of northern India. The average
seed mass for the tree species of the dry forest was 0.32 ± 0.08 g which is compa-
rable to that (4.32  ±  0.89  g) documented by Hammond and Brown (1995) for a
Neotropical forest located in Mabura Hill, Guyana. Most of the species investigated
by Hammond and Brown (1995) were either dispersed by mammals or gravity dis-
persed. According to Khurana et al. (2006), the average seed mass was significantly
related to both dispersal mode and the shade tolerance (Fig. 4.3), with large number
of small-seeded species exhibiting wind-dispersal mechanism, while all the
­large-­seeded species were dispersed by mammals. According to Guo et al. (2000),
the small-seeded species having greater wind-dispersal capacity exhibit a wider
range of distribution and abundance, which confirms the general rule for the assem-
bly of different plant communities. Chazdon et al. (2003) also observed that plants
exhibiting non-animal dispersal mechanism (i.e. wind, explosive and gravity) were
highly associated with smaller seed mass, while those showing animal dispersal
4.3  Seed Size: A Key Trait Determining Species Distribution 83

Fig. 4.4 Triangular
relationship between seed
mass and number of
individuals (all individuals
of the species in the total
15 ha area) (Source:
Khurana et al. 2006)

mechanism were predominantly showing association with larger seed mass, as


observed in a Mexican rain forest.
Small-seeded species studied by Khurana et al. (2006), generally, were relatively
shade-intolerant, while the large-seeded species mostly had a moderate or a high
tolerance to shade. Foster and Janson (1985) also reported that seeds of the tropical
forest trees that could establish under the closed canopies or in small gaps were
larger than those found in species which require large gaps. Seedlings of large-­
seeded species growing under closed canopies could reach relatively earlier into a
better light environment than those of small-seeded species and could survive for
longer period in the thick litter layers as they have enough food reserve to support
respiration longer under conditions of net carbon deficit (Westoby et al. 1992). Most
of the species investigated by Khurana et al. (2006) were from the relatively shade-­
intolerant category.
The triangular relationship observed between seed mass and the per cent fre-
quency and number of individuals by Khurana et al. (2006) (Fig. 4.4) suggested that
tree individuals of small-seeded species exhibit more wide distribution compared to
those of large-seeded species. It was also observed that the smaller seeds enter into
the soil often more easily compared to large seeds and therefore have a higher prob-
ability of being available in persistent soil seed banks (Khurana and Singh 2001a).
Further, large-sized seeds comprise commonly the first crop of germinants observed
in the early rainy season where the insect outbreaks could lead to their death
(Garwood 1983). Small seeds, therefore, play a role of a back-up crop of seedlings
that may then dominate the site.
Khurana et al. (2006) compared composition of adult and seedling individuals of
the species under the study at several sites. They reported that at the less-disturbed
sites, the recruitment of large-seeded species was disproportionately high and the
recruitment of small-seeded species was low. In contrast, at the heavily disturbed
sites, seedlings germinating from the large-seeded species were less in number,
while the abundance of seedlings developing from small-seeded species was higher.
84 4  Plant Traits and Regeneration

During the study on ten wet tropical forests of Costa Rica, Chazdon et al. (2003)
observed that especially in second-growth forests (frequently disturbed), the rela-
tive abundance of species exhibiting explosive and wind-dispersed seeds, or that
showing insect pollination mechanism, was greater, and the relative abundance of
species showing animal dispersal mechanism, or those exhibiting mammal pollina-
tion mechanism, was lower. According to Khurana et al. (2006), at the sites experi-
encing least disturbance, the maximum number of mature trees can act as dispersal
barriers for the coming of wind-dispersed small seeds in the forest, resulting in a
decline in the proportion of small-seeded, pioneer species within the forest com-
munity. Khurana et al. (2006) also reported that the seedlings of large-seeded, late-­
successional species, for example, Terminalia tomentosa and Diospyros
melanoxylon, were numerous at least disturbed sites and were detected totally
absent, or their density was very low at the sites which were highly disturbed. These
findings suggested that the different seed sizes allow the occupation of different
types of niches available in the TDF along the gradients of light, soil moisture and
temperature, thereby contributing to the maintenance and development of overall
species diversity (Khurana et al. 2006).

4.4  Seed and Seedling Ecology

The seed production pattern, germination capacity, survival and development of


seedlings are influenced by the pronounced seasonality in the environmental condi-
tions in TDFs. The most favourable period for the germination of seedlings and their
establishment in the dry tropics is very short and limited to the rainy season.
However, majority of the canopy of dry forest becomes renovated during the pre-­
monsoon period (Singh and Singh 1992a), and therefore the new seedlings emerg-
ing in the rainy season experience lower intensity and modified quality of sunlight
under relatively closed tree canopies, while those developing in the open experience
high intensity of light, but they also face competition from the fast-growing popula-
tions of herbs. The germination, establishment and development of seedlings, there-
fore, face marked heterogeneous environmental conditions in the dry forest as
observed in the rain forest.

4.4.1  Seed Viability and Dormancy

Trees in the TDF have been observed to produce both orthodox and recalcitrant
seeds. Orthodox seeds (e.g. those of Acacia nilotica and Gmelina arborea) can eas-
ily be stored at low temperatures for a longer time period without losing viability.
Nevertheless, many dry forest tree seeds are found to remain viable only for a short
time period under ordinary conditions (Ray and Brown 1995). For example, the
seeds of Shorea robusta, which is a common tree species in both moist and dry
4.4  Seed and Seedling Ecology 85

tropical forests of India, remain viable for a very short period (only for 7–10 days)
(Saha et al. 1992). The probable reason for this short viability has not been precisely
known; however, the short viability is expected to be due to the loss of moisture at
a very high rate which is associated with the loss of hair present on the seed coat.
According to Thapliyal and Connor (1997), elevated leachate conductivity and low-
ering of fatty acid content mostly due to ageing in certain seeds (such as seeds of
Dalbergia sissoo) are other reasons behind the loss of viability. In the Ethiopian dry
Afromontane forest, seeds of some species such as Bersama abyssinica, Myrsine
africana and Pittosporum viridiflorum are recalcitrant, whereas those of Ekebergia
capensis exhibit a decline in the percentage of germination with increase in the time
of storage (Teketay and Granstrom 1997).

4.4.2  S
 eed Size, Germination, Water Stress and Seedling
Growth

Seed size has been reported to affect several aspects of plant life (Milberg and
Lamont 1997), for example, the dispersal of seeds and water relations, and the ger-
mination, establishment, survival and development of seedlings. Small seeds gener-
ally facilitate the build-up of persistent soil seed bank, which is crucial for the
regeneration of species after any kind of disturbance. On the other hand, a high
quantity of seed reserve of the large-seeded species can enhance their capacities to
survive by providing for the requirements of metabolism during the quiescence
period, until the availability of suitable light or moisture conditions. The large-­
seeded species of primary forest, as reported by Wunderle (1997) are relatively
immobile due to their low dispersal efficiency. As discussed earlier also, large and
bulky seeds contain higher quantities of reserves to stimulate the process of germi-
nation, seedling growth and survival (Milberg and Lamont 1997). Within a species
also, seedlings from larger seeds (as in Terminalia ivorensis) exhibit better growth
compared to those from smaller seeds (Oni and Bada 1992). Moreover, according to
Milberg and Lamont (1997), the effect of large seeds is particularly pronounced in
the nutrient-impoverished soil.
Table 4.5 shows the close relationship between the size and/or weight of seed and
the percentage of germination within species in several tropical tree seedlings.
Although the large seed size commonly accounts for better seed germination and
survival of seedlings, the medium-sized seeds, for example, in Acacia mellifera,
also germinated faster (Srimathi et al. 1991), and similarly did the small seeds of
Santalum album and Lagerstroemia parviflora (Prasad et  al. 1988). Khurana and
Singh (2004a) observed reduced percentage of germinations as well as speed of
germination under the water-stressed conditions (Table 4.6). The percentage of ger-
minations observed in the large-seeded, late-successional species, Terminalia che-
bula and T. arjuna, were reduced, and a higher proportion of seeds were converted
to the dormant category under conditions of high water stress, as also found in the
86 4  Plant Traits and Regeneration

Table 4.5  Effect of seed size and/or seed weight on germination of certain dry tropical tree
species. For references see Khurana and Singh (2001a)
% Germination
Species Heavy/large seeds Light/small seeds
Cassia glauca 74 64
Cassia siamea 36 25
Ceiba pentandra 15 11
Dillenia indica 32 21
Holoptelea integrifolia 39 18
Kydia calycina 47 13
Leucaena leucocephala 76 55
Peltophorum ferrugineum 14 5
Terminalia bellirica 100 23
Terminalia chebula 47 33
Terminalia tomentosa 57 13
Source: Khurana and Singh (2001a)

Table 4.6  Per cent germination of viable seeds and GVI (germination velocity index) at 0 MPa
and per cent reduction in germination and GVI under four water stress levels as compared with
0 MPa, for seeds of five tree species
Species Seed size 0 MPa −0.1 MPa −0.5 MPa −0.9 MPa −1.5 MPa
% germination % reduction in germination
P. emblica Large seed 71.4 ± 7.0 5.2 24.1 61.7 80.8
Small seed 34.4 ± 2.9 9.7 19.4 29.0 39.8
A. procera Large seed 88.0 ± 0.5 26.8 35.8 45.3 85.8
Small seed 82.7 ± 3.9 21.6 36.7 52.5 53.2
A. nilotica Large seed 70.6 ± 2.4 15.7 30.6 53.0 94.8
Small seed 44.8 ± 2.9 7.4 23.1 46.3 53.7
T. chebula Large seed 77.8 ± 0.0 42.9 57.1 81.0 100.0
Small seed 57.9 ± 2.3 33.6 47.2 53.6 100.0
T. arjuna Large seed 88.0 ± 0.5 49.5 67.9 77.4 90.5
Small seed 58.3 ± 4.2 35.7 50.6 64.3 64.3
GVI % reduction in GVI
P. emblica Large seed 1.15 ± 0.04 21.8 40.3 70.2 95.7
Small seed 0.83 ± 0.05 37.3 42.8 70.8 62.9
A. procera Large seed 2.22 ± 0.2 38.6 69.6 77.7 96.8
Small seed 1.55 ± 0.4 45.3 54.8 48.6 49.1
A. nilotica Large seed 0.54 ± 0.01 22.8 31.7 69.8 91.4
Small seed 0.32 ± 0.05 37.3 42.8 70.8 62.9
T. chebula Large seed 0.36 ± 0.03 69.4 75.2 88.8 100.0
Small seed 0.09 ± 0.01 21.9 27.7 40.4 100.0
T. arjuna Large seed 0.31 ± 0.01 47.3 77.1 85.3 94.5
Small seed 0.19 ± 0.02 53.3 73.4 78.4 78.1
Source: Khurana and Singh (2004a)
4.4  Seed and Seedling Ecology 87

small-seeded species, for example, Phyllanthus emblica, Albizia procera and Acacia
nilotica. Therefore, the late-successional species were observed to have the capacity
to sense and to respond to the changes in environmental conditions more success-
fully. Even within the species, Khurana and Singh (2004a) found a tendency for
large seeds to minimize their germination capacity and germination velocity index
(GVI) under high water stress conditions more efficiently than the small seeds. This
may be probably due to a reduction in the seed surface to volume ratio with increas-
ing seed size; and the large seeds might not be able to accumulate adequate water
for their germination in drier soils that could however be suitable for the germina-
tion of smaller seeds (Foster 1986; Harper and Benton 1966).
As reported by Singh and Singh (1992a), seeds of most of the dry tropical spe-
cies become mature in summer and are generally dispersed at the start of the rainy
season when the forest soil has sufficient available moisture for the germination and
growth of seedlings. If, however, seeds do not germinate during the first rainy sea-
son, their germination is delayed and again occurs in the second wet season, for
example, in Diospyros melanoxylon (Ghosh et al. 1976). Because the tree species in
TDF have been historically subjected to unpredictable droughts, at least few of them
exhibit the capacity to germinate under the conditions of substantial seasonal water
stress, for example, Acacia farnesiana (−0.10 MPa; Scifres 1974), Acacia senegal
(−1.38 MPa; Palma et al. 1995), Acacia tortilis (−1.0 MPa; Coughenour and Detling
1986), Cassia obtusifolia (−0.5  MPa; Daiya et  al. 1980) and C. occidentalis
(−0.5 MPa; Daiya et al. 1980). According to Ray and Brown (1995), the germina-
tion of seeds and survival of young seedlings growing in Caribbean dry forest spe-
cies were observed to be closely associated with amount and timing of rainfall;
therefore, in regions experiencing a long dry season, seedling desiccation is expected
to be a major obstacle to tree recruitment, such as, according to Skoglund (1992),
the recruitment of Acacia tortilis seedlings was limited to years with relatively high
rainfall. Moreover, Lieberman and Li (1992) detected highly seasonal seed germi-
nation and seedling mortality in the 53 species from the TDF of Ghana, where they
also reported highest mortality during the dry season.
Several species do exhibit plasticity and capacity to acclimatize in the moisture-­
deficient conditions. An elevation in root to shoot biomass ratio and a higher alloca-
tion of biomass to lateral roots, in species, for example, Drypetes parvifolia and
Teclea verdoornia (Lieberman and Li 1992), increased the root surface area respon-
sible for water uptake (Wright et al. 1992). Tree seedlings growing in open forests
generally face competition from the annual herbaceous species which also germi-
nate during the rainy season; however, in the subsequent dry season, the herbaceous
plants die off and act as mulch, thereby helping to conserve the soil moisture and to
provide protection to the tree seedlings from intense sunlight.
Khurana and Singh (2000) reported an elevation in root/shoot (R/S) ratio and a
decline in stem weight ratio (SWR) in seedlings suggesting a shift in the resource
allocation from shoot to roots at low soil moisture. According to McConnaughay
and Coleman (1999), plants exhibit response to low water availability by moving
the partitioning of carbohydrates to the physiological processes associated with soil
water uptake and conservation in lieu of carbon acquisition. Higher R/S ratios as
88 4  Plant Traits and Regeneration

commonly found in water-stressed plants show relatively higher investment of car-


bohydrates to below-ground foraging structures in agreement with the optimal for-
aging behaviour (McConnaughay and Coleman 1999). Khurana and Singh (2004a)
also recorded a consistent change in allocation of resources from shoot to root at
low moisture conditions as indicated by an elevation in R/S ratio (Fig. 4.5). Findings
of Khurana and Singh (2004a) suggested that the decrease in plant biomass after
imposing water stress was higher in fast-growing tree species as compared to the
slow-growing species (Table 4.7). This was clearly observed in the study of Khurana
and Singh (2004a), where Terminalia chebula, which is a slow-growing late-­
successional tree, and Acacia nilotica, a slow-growing early-successional tree, reg-
istered minimum reduction in biomass after imposing water stress. Incidentally,
these two species also exhibited higher drought tolerance index as compared to the
other forest species. Therefore, the response after experiencing water stress in terms
of seedling growth was not associated with the successional status of the species;
nevertheless, the potential growth rate or relative growth rate (RGR) was more use-
ful for analysing such processes. Leishman and Westoby (1994a) have also sug-
gested that the slow RGR is associated with plant characteristics which confer
drought tolerance.

4.4.3  M
 ajor Environmental Factors Influencing Seedling
Recruitment
4.4.3.1  Temperature and Fire

The dry forest species are much tolerant to moderately high temperatures generally
at all life cycle stages as they are found in areas characterized by a markedly warm
to much hot season each year and are exposed to frequent natural or anthropologi-
cally induced fires (Goldammer 1993). Teketay (1994, 1996) studied in many
Ethiopian leguminous tree species which are found from 0 to 2400 m elevation and
reported that seeds of those species are commonly released from dormancy after
getting heat shock when they are exposed to fire. This phenomenon is also observed
for Acacia sieberiana and A. gerrardii of the dry savanna in Uganda (Mucunguzi
and Oryem-Origa 1996). The seed germination stimulated by the heat shock is
apparently a common feature in Fabaceae, Rhamnaceae, Convolvulaceae and
Sterculiaceae (Bell et al. 1993). Certain species exhibit higher percentage of germi-
nation even when they are exposed to fire-generated smoke. According to Brown
and van Staden (1997), the technique of smoke priming can potentially be applied
to elevate the germination of seeds for ex situ conservation of rare and threatened
African fynbos species which are subjected to repeated fires.
Most of the tree species of TDF show important strategies for avoiding the pos-
sible deleterious consequences of forest fire events on their regeneration. The study
of historical information about the Indian forests indicated that during the previous
century and early period of this century, almost every deciduous forests of the coun-
4.4  Seed and Seedling Ecology 89

Fig. 4.5  Root to shoot ratio and net assimilation rate (NAR) of 5-month-old seedlings of five tree
species from tropical dry forest under four water levels. Narrow bars represent ±1S.E (Source:
Khurana and Singh 2004a)
90 4  Plant Traits and Regeneration

Table 4.7  Height, leaf area and biomass of 5-month-old seedlings of five species grown at four
water stress levels
Species Seed size −0.01 MPa −0.1 MPa −0.61 MPa −1.2 MPa
Height (cm)
P. emblica Large seed 43.2 ± 0.3 29.3 ± 0.9 23.6 ± 0.3 –
Small seed 29.4 ± 0.5 17.8 ± 0.6 14.7 ± 0.3 –
A. procera Large seed 52.5 ± 4.1 42.8 ± 1.2 28.4 ± 0.9 22.8 ± 2.9
Small seed 35.7 ± 5.0 24.7 ± 2.03 21.4 ± 2.6 –
A. nilotica Large seed 35.1 ± 1.6 28.9 ± 2.6 22.4 ± 1.0 18.7 ± 0.4
Small seed 27.6 ± 1.5 20.3 ± 0.3 13.4 ± 0.3 –
T. chebula Large seed 25.9 ± 1.2 16.9 ± 3.3 9.1 ± 0.7 6.1 ± 0.3
Small seed 11.5 ± 1.2 6.8 ± 0.7 4.7 ± 0.3 4.4 ± 0.1
T. arjuna Large seed 73.8 ± 0.9 54.0 ± 0.3 44.0 ± 1.3 32.9 ± 0.6
Small seed 34.8 ± 0.6 28.3 ± 0.9 19.0 ± 1.0 11.3 ± 0.1
Total leaf area (cm2)
P. emblica Large seed 337.1 ± 1.4 239.7 ± 0.8 150.7 ± 1.0 –
Small seed 224.3 ± 0.5 125.0 ± 0.4 72.9 ± 0.4 –
A. procera Large seed 1032.0 ± 6.6 665.8 ± 1.0 167.7 ± 0.3 28.8 ± 0.7
Small seed 534.5 ± 1.3 151.1 ± 0.4 37.7 ± 1.4 –
A. nilotica Large seed 153.8 ± 1.2 80.2 ± 0.6 70.0 ± 0.6 35.6 ± 0.5
Small seed 87.9 ± 0.1 49.1 ± 0.1 24.7 ± 0.5 –
T. chebula Large seed 173.5 ± 2.2 104.1 ± 3.4 43.1 ± 2.2 28.3 ± 1.1
Small seed 68.0 ± 1.3 18.3 ± 0.2 7.1 ± 0.9 4.8 ± 0.2
T. arjuna Large seed 4599.7 ± 15.8 2021.7 ± 18.1 940.3 ± 6.2 329.2 ± 1.6
Small seed 2160.1 ± 2.1 1300.6 ± 2.0 187.7 ± 1.9 46.9 ± 2.1
Plant biomass (g)
P. emblica Large seed 1.4 ± 0.03 1.2 ± 0.1 0.9 ± 0.02 –
Small seed 1.01 ± 0.03 0.8 ± 0.03 0.8 ± 0.02 –
A. procera Large seed 3.8 ± 0.3 2.8 ± 0.3 1.9 ± 0.1 1.4 ± 0.2
Small seed 2.2 ± 0.1 1.5 ± 0.1 1.2 ± 0.1 –
A. nilotica Large seed 0.9 ± 0.1 0.8 ± 0.01 0.7 ± 0.02 0.6 ± 0.02
Small seed 0.7 ± 0.04 0.6 ± 0.02 0.5 ± 0.03 –
T. chebula Large seed 1.4 ± 0.1 1.3 ± 0.02 1.2 ± 0.1 1.02 ± 0.04
Small seed 0.6 ± 0.02 0.6 ± 0.02 0.6 ±  ± 0.01 0.5 ± 0.02
T. arjuna Large seed 19.1 ± 0.1 15.3 ± 0.3 11.1 ± 0.1 9.6 ± 0.5
Small seed 11.0 ± 0.1 10.9 ± 0.1 7.3 ± 0.3 6.1 ± 0.3
Source: Khurana and Singh (2004a)

try were burned annually (Goldammer 1992). It has been documented that the
regeneration of Indian teak (Tectona grandis) forest is much high in burnt areas as
compared to the fire-protected regions, as fire is claimed as an important agent to
stimulate germination of seed and to facilitate seedling establishment. In several
TDF species, it has been reported that after exposure to fire, seeds generally take up
water and start to germinate, but when the time of germination synchronizes with
4.4  Seed and Seedling Ecology 91

the habitat conditions which are deleterious to seedlings, seeds sprout and die, lead-
ing to depletion of the natural seed bank.
As documented by Swaine (1992) and Eriksson et al. (2003), mortality due to
fire, particularly in small-sized trees, is higher compared to the large-sized trees.
There are several interacting top-down and bottom-up factors which spatially deter-
mine the occurrence as well as spread of fire in a landscape (Falk et al. 2007). These
factors may be determined anthropogenically (e.g. ignition, prevention or suppres-
sion of fire) or by other biophysical factors (e.g. topography, vegetation and weather)
(Moreira et al. 2011; Curt et al. 2015). Adult trees are generally resistant to fire due
to the presence of insulating thick bark; however, at the younger sapling stage, they
are sensitive to damage due to fire and require protection (Backéus et  al. 1994;
Eriksson et al. 2003). Disturbance related to frequent wildfires has been supposed to
result in a long-term change in the structure and composition of TDF vegetation
(Griscom and Ashton 2011; Devisscher et al. 2016).

4.4.3.2  Light

Responses of seed germination, survival and growth of seedling to light intensity


exhibit marked variations among species. Ray and Brown (1995) classified the
Caribbean TDF species according to their response to light conditions into three
groups: (i) species which are unlikely to survive without being subjected to medium
shade (25% of full sunlight), for example, Guettarda parviflora and Coccoloba
microstachya; (ii) species exhibiting weak survival rate without being subjected to
medium shade and showing increased survival rate in medium shade, for example,
Guaiacum officinale and Sabinea florida; and (iii) species which exhibit higher sur-
vival rate without experiencing medium shade but show enhanced survival rate in
medium shade, for example, Plumeria alba and Bursera simaruba. These authors
further suggested that the shade demand for survival also integrates tolerance to
drought conditions and light requirement, which allows seedlings to avoid greater
light intensity and also very low moisture. Rincon and Huante (1993) investigated
seedlings of several TDF species found in Mexico and observed that all of them
exhibited higher growth rates as well as net assimilation rates when subjected to the
high light treatment (400 μmol m−2 s−1, i.e. light available in a medium-sized gap)
compared to the low light intensity (80 μmol m−2 s−1, i.e. the light underneath the
vegetation canopy during the rainy season). Their study also indicated that the pio-
neer species in the study region, such as Heliocarpus pallidus and Apoplanesia
paniculata, which produced smaller seeds, showed a higher growth rate under the
high light intensity compared to the shade-tolerant species, for example,
Amphipterygium adstringens, Caesalpinia eriostachys and Caesalpinia platyloba,
which produced larger seeds. It was also observed that the seedling growth in
Dalbergia sissoo and Acacia catechu was maximum in low shade treatment,
whereas that of Casuarina equisetifolia was greatest in the unshaded treatment
(Saxena et al. 1995).
92 4  Plant Traits and Regeneration

4.4.3.3  Soil Nutrients

Nitrogen amendment enhanced the growth rate of tree seedlings of majority of spe-
cies investigated in India, including Populus ciliata (Deol and Khosla 1983),
Eucalyptus tereticornis, Eucalyptus camaldulensis, Eucalyptus grandis and
Eucalyptus citriodora (Prasad and Rawat 1994), Acacia mollissima (Rao 1985),
Acacia catechu (Prasad and Rawat 1992), Bauhinia variegata (Koul et al. 1995) and
Leucaena leucocephala (Sivasupiramanium et al. 1988). Rachmawali et al. (1996)
observed that urea application increases the growth of several seedlings of the
Indonesian species, for example, Casuarina junghuhniana, Albizia lebbeck, A. pro-
cera, Dalbergia latifolia and Acacia sinensis. However, there appears to have a limit
to the quantity of N application that can increase the growth of seedlings. For exam-
ple, N application at relatively low doses increased the biomass of Eucalyptus tereti-
cornis seedlings; however, when the dose was increased more than a threshold level,
the biomass of seedlings decreased (Prasad and Rawat 1994). The reason behind
this trend might be that the high doses of N could lead to the lower uptake of P and
K (Koul et al. 1995). In other studies, P amendment improved growth characteris-
tics in Sesbania sesban (Dutt and Pathania 1986), Bauhinia variegata (Koul et al.
1995), Dalbergia nigra (Chaves et al. 1995) and Leucaena leucocephala (Prasad
and Rawat 1992), and increased P content in Terminalia ivorensis (Aluko and
Advayi 1983), but led to reduction in N and K contents in Eucalyptus grandis
(Prasad et al. 1984), Acacia mollissima and Celtis australis (Rao 1985).
The deposition of atmospheric nitrogen (N) in various ecosystems commonly
from the anthropogenic sources has been observed to be rapidly increasing in the
recent years. Zheng et al. (2002) observed that the rate of deposition of oxidized N,
particularly in the terrestrial ecosystems of Asia, showed increment of about seven
times, and the rate of deposition of reduced N increased three times between 1961
and 2000. Increasing trend of N deposition has been expected to induce a pro-
nounced impact on the structure and functioning of the terrestrial biomes, and this
factor ranks third in the list of major drivers responsible for the biodiversity change
(Sala et al. 2000). Seedlings of species from different successional status (pioneer,
non-pioneer) may show differential responses to increment in the N input, as they
exhibit marked variations in physiological and life-history traits (Tilman 1987b).
For example, the pioneer species have characteristic features associated with high
competitive ability, high relative growth rate and related traits, which promote rapid
growth (Reich et al. 1997, 1998b), whereas the non-pioneer species exhibit slower
growth rates and may show narrower responses to variations in nutrient availability.
Similarly, the evergreen species are supposed to exhibit a lower response to the
nutrient availability compared to the deciduous species (Aerts and Chapin 2000).
The growth of seedlings produced from the seeds of variable sizes may be influ-
enced differently by the increment in N deposition (Bazzaz and Miao 1993; Khurana
and Singh 2000). Higher amounts of carbohydrate present in the endosperm or coty-
ledons in the seedlings growing from large seeds compared to those from small
seeds may lead to an early emergence of an enlarged resource accumulating system
(root or photosynthetic tissue) and faster growth (Hewitt 1998; Khurana and Singh,
4.4  Seed and Seedling Ecology 93

Table 4.8  Seed characteristics (mean ± 1SE) for five tropical tree species under study
Seed Length Width Seed mass Habitat
Study species size (mm) (mm) (mg·seed−1) Trait preference
Albizia procera Large 8.4 ± 0.5 6.0 ± 0.1 55.0 ± 2.0 PL Disturbed
(Mimosaceae) seed
Small 4.9 ± 0.2 3.8 ± 0.1 26.0 ± 0.2
seed
Acacia nilotica Large 8.7 ± 0.5 7.1 ± 0.4 170.0 ± 7.0 PL Disturbed
(Mimosaceae) seed
Small 6.2 ± 0.8 5.4 ± 0.7 110.0 ± 3.0
seed
Phyllanthus emblica Large 6.4 ± 0.6 3.1 ± 0.9 30.0 ± 1.0 PNL Disturbed
(Euphorbiaceae) seed
Small 4.9 ± 0.4 2.4 ± 0.4 20.0 ± 3.0
seed
Terminalia arjuna Large 39.3 ± 1.5 22.3 ± 2.0 2260.0 ± 40.0 NPNL Undisturbed
(Combretaceae) seed
Small 25.5 ± 4.0 14.5 ± 3.0 1310.0 ± 87.0
seed
Terminalia chebula Large 20.7 ± 0.8 10.8 ± 3.0 1030.0 ± 33.0 NPNL Undisturbed
(Combretaceae) seed
Small 15.6 ± 2.0 8.4 ± 1.0 500.0 ± 80.0
seed
Source: Khurana and Singh (2004b)
Mean seed size and mass were determined from 100 seeds. Trait and habitat preference informa-
tion after Troup (1921), Champion and Seth (1968)
PL pioneer, leguminous; PNL pioneer, nonleguminous; NPNL non-pioneer, nonleguminous

2001a, b). Plant species of the N-poor, tropical dry deciduous ecosystems exhibit
marked variations in the life-history traits and thus can be expected to exhibit differ-
ences in response to N inputs. Khurana and Singh (2004b) selected five important
dry tropical tree species, viz. Albizia procera, Acacia nilotica, Phyllanthus emblica,
Terminalia arjuna and Terminalia chebula, which differed in seed size, successional
status, leaf habit and ability of fixing nitrogen, and studied relationships of the
response of seedlings to elevated N inputs by analysing functional types, such as
pioneer versus non-pioneer, legumes versus non-legumes, large- versus small-­
seeded, evergreen versus deciduous leaf habit, etc. They observed the effect of
increasing N input upon the growth traits of the tree seedlings within a species
producing from seeds of variable size. Among the selected species, Albizia procera,
Acacia nilotica and Phyllanthus emblica are pioneer species (Table  4.8). Albizia
procera and Acacia nilotica are N-fixing legumes, while Terminalia arjuna and
Terminalia chebula are non-pioneer, non-leguminous species. Albizia procera,
Phyllanthus emblica and Terminalia arjuna are fast growing, whereas Acacia nilot-
ica and Terminalia chebula are slow-growing species. Increased N input led to the
decline in the R/S ratio (Fig. 4.6), and the seedlings exhibited significant plasticity
in growth as well as allocation of biomass to both above-ground and below-ground
plant components (Fig. 4.7). Enhanced N input increased SLA and RGR (Fig. 4.8),
94 4  Plant Traits and Regeneration

Fig. 4.6  Height, leaf area, root mass, shoot mass and root-to-shoot ratio (R/S) of LS (large-seed)
and SS (small-seed) seedlings of five dry tropical tree species under four levels of nitrogen applica-
tion. Bars represent ±1SE. Ap Albizia procera, An Acacia nilotica, Pe Phyllanthus emblica, Ta
Terminalia arjuna, Tc Terminalia chebula (Source: Khurana and Singh 2004b)
4.4  Seed and Seedling Ecology 95

Fig. 4.7  Per cent increase in biomass of 5-month-old seedlings of five dry tropical species at four
levels of nitrogen application (Source: Khurana and Singh 2004b)

and these two traits exhibited positive relationship with each other (Fig. 4.9a). The
positive relationship was also observed between RGR and net assimilation rate
(NAR) (Fig.  4.9b), suggesting that metabolism per unit biomass invested in
the  leaves was a determining factor controlling RGR.  In these selected species,
the net rate of CO2 assimilation per unit leaf area, which showed positive associa-
tion with foliar N (Fig. 4.9c), was not related with RGR. The investment of N to
96 4  Plant Traits and Regeneration

Fig. 4.8  Relative growth rate (RGR), specific leaf area (SLA), net assimilation rate (NAR) and per
cent foliar N of LS (large-seed) and SS (small-seed) seedlings of five dry tropical tree species
under four levels of nitrogen application. Bars represent ±1SE. Ap Albizia procera, An Acacia
nilotica, Pe Phyllanthus emblica, Ta Terminalia arjuna, Tc Terminalia chebula (Source: Khurana
and Singh 2004b)
Fig. 4.9  Relationships between seedling growth and life-history traits for five dry tropical tree
species grown at four levels of nitrogen application. (a) Relationship between relative growth rate
(RGR) and specific leaf area (SLA): RGR = 0.849SLA + 325.8; r2  =  0.81, P  <  0.0001. (b)
Relationship between RGR and net assimilation rate (NAR): RGR = 0.22NAR + 184.53; r2 = 0.67,
P < 0.0001. (c) Relationship between net CO2 assimilation rate and foliar N (%): net CO2 assimila-
tion rate = 0.55 foliar N (%) + 1.098; r2 = 0.27, P < 0.0001. Each symbol represents the mean value
for a species at a particular treatment (Source: Khurana and Singh 2004b)
98 4  Plant Traits and Regeneration

photosynthetic system and the resultant increment in the net CO2 assimilation rate
were not exhibited proportionally in the form of growth in the two legumes.
Therefore, there might be some other factors and not only the net CO2 assimilation
rate that express the net growth rate of a tree seedling. Further, the enhanced
N-caused growth increment exhibited variations among species; the highest
response to N amendments was observed for the slowest-growing species Terminalia
chebula which is also a non-pioneer, whereas the other non-pioneer, T. arjuna,
exhibited the least response. When observed for the three pioneer species,
Phyllanthus emblica, which is a non-­leguminous species, responded more strongly
as compared to the two leguminous species. The study suggested that the response
to N fertilization was not associated with the successional status of the tree
species.

4.4.3.4  Ambient CO2

The concentration of CO2 in the atmosphere has been increasing, and according to
the data of 2014, the annual atmospheric CO2 concentration exceeded 400  μmol
mol−1 at Mauna Loa (Keeling et  al. 2014). The prediction of global atmospheric
model simulations indicated that this incremental trend will continue (IPCC 2013).
Thus, the evaluation of the responses of the elevated CO2 by tree species is an
important subject for prediction of ecosystem response to elevated CO2. It has been
reported that the elevated CO2 causes increment of the biomass and leaf area in
seedlings of the dry tropical trees; the short-term exposures to elevated CO2 have led
to 5–1000% increase in the leaf area and 10–460% increase in the biomass of seed-
lings of various TDF species (Table 4.9).
Initially, elevated CO2 increases photosynthetic CO2 assimilation rates, but sub-
sequently photosynthetic rates decline due to biochemical as well as morphological
acclimation and other environmental constraints, for example, water and nutrient
supply (Leakey et al. 2009; Watanabe et al. 2011; Warren et al. 2015). The decrease
in photosynthesis (i.e. photosynthetic downregulation) is generally observed when
there is decrease in the maximum carboxylation rate of Rubisco (Vcmax) and the rate
of electron transport, which results into RuBP regeneration (Jmax) (Leakey et  al.
2009). Such types of changes are related with decline in the nitrogen concentrations
and the increment of leaf mass per unit area (LMA) as a result of the assimilation of
photosynthetic products (Leakey et al. 2009).
Several authors, such as Dyckmans and Flessa (2005), Onoda et al. (2005) and
Kitao et al. (2007), have observed a significant correlation between the response of
seedlings to a high CO2 environment and the traits expressing nitrogen use in the
tree species. Enhanced CO2 concentration influences the nitrogen content of leaf
and other biochemical traits and physiological processes such as nitrogen allocation
to the photosynthetic proteins (Nakano et  al. 1997; Hikosaka and Hirose 1998;
Onoda et al. 2005). Kitaoka et al. (2016) evaluated the traits associated with photo-
synthetic and biochemical processes in the leaves of selected four deciduous broad-­
leaved tree species with variable successional traits growing under ambient
(370 μmol mol−1) and elevated (720 μmol mol−1) CO2 concentrations and observed
4.4  Seed and Seedling Ecology 99

Table 4.9  Increase in biomass and leaf area of certain dry tropical seedlings under elevated CO2.
For references see Khurana and Singh (2001a)
Species Duration of exposure (days) Biomass (%) Leaf area (%)
Acacia auriculiformis 90 44 85
Annona squamosa 90 35.5 320
Artocarpus integrifolia 90 155 340
Dalbergia latifolia 90 36 71
D. sissoo 90 193 270
Derris indica 90 135 235
Eucalyptus camaldulensis 84 75 21
E. citriodora 90 51 35
E. cypellocarpa 84 85 27
E. grandis 42 275 –
E. pauciflora 100 212 117
E. pulverulenta 84 158 71
Fagus grandifolia 60 83 92
F. sylvatica 105 54 40
Feronia elephantum 90 17 45
Ficus obtusifolia 90–100 10 35
Hevea brasiliensis RRII 105 17 32
Liriodendron tulipifera 168 22 −15
Piper auritum 111 15 5
Spathodea campanulata 90 88 216
Swietenia macrophylla 90 101 53
Tabebuia rosea – 164 83
Tamarindus indica 90 12.3 27
Tectona grandis 90 460 1012
Ziziphus jujuba 90 125 256
Source: Khurana and Singh (2001a)

that tree species exhibiting different successional traits showed different stomatal as
well as non-stomatal responses to elevated CO2.
Several reviews have suggested that the dry mass of tree seedlings exhibits incre-
ment under the elevated concentration of CO2, as the CO2 affects various growth-­
related traits (Eamus and Jarvis 1989; Ceulemans and Mouseau 1994; Drake et al.
1997). Seed size, successional status as well as the nitrogen-fixing ability may
exhibit differential response in seedlings exposed to elevated CO2 concentration
(Bazzaz and Miao 1993), and analysis of such type of responses could help in the
prediction of the structure and composition of a forest community experiencing
changing climate scenario. The inherent potential of growth existing in the plant
species has major implications regarding their activity and performance in the real
ecosystems, and the geographical distribution areas of many ecosystems could
likely shift due to differences in the response of individual species to changes in
CO2 concentration. Khurana and Singh (2004c) analysed the impact of seed size as
well as successional status on the growth of seedlings under elevated CO2 for the
100 4  Plant Traits and Regeneration

selected five dry tropical tree species, viz. Albizia procera, Acacia nilotica,
Phyllanthus emblica, Terminalia arjuna and Terminalia chebula. In their study,
seedlings germinating from large seeds (LS) and small seeds (SS) were separately
grown at the two CO2 levels (i.e. ambient and elevated, 700–750 ppm). After 30 days
exposure to the elevated CO2, they measured CO2 assimilation rate, stomatal con-
ductance, water use efficiency and foliar N content. Harvesting of the seedlings was
done after the exposure periods of 30 and 60 days. Seedlings were also measured for
height, diameter, leaf area, biomass and other growth traits (viz. RGR, NAR, SLA,
R/S). These authors observed that for all tree seedlings, the elevated CO2 concentra-
tion significantly increased the overall growth, although they differed in magnitude
(Figs. 4.10, 4.11 and 4.12). It was also reported that after 60 days exposure to CO2,
per cent increment in biomass and other growth-related traits was downregulated.
The probable reason behind this downregulation might be the physical distortion of
chloroplast, sucrose phosphate synthetase inhibition, competition among the phos-
phorylated sugars for acquiring binding sites on Rubisco enzyme, increment in the
activity of carbonic anhydrase enzyme and the inability of the Calvin cycle for
regeneration of ribulose bisphosphate or orthophosphate (Eamus and Jarvis 1989;
Drake et al. 1997).
According to the findings of Khurana and Singh (2004c), the specific leaf area
(SLA) was reduced in the seedlings treated with elevated CO2 (Table 4.10). This
decrease in SLA under increased CO2 was also observed by Cornelissen et  al.
(1999b) for the eight woody species and by Zak et al. (2000) for Populus tremuloi-
des. This could be probably due to an increment in the total composition of non-
structural carbohydrates and the leaf thickness (Zak et al. 2000). High CO2 levels
increased root biomass and the R/S ratio helping the plants to acquire more N from
soil for supporting the enhanced growth. According to the report of Eamus and
Jarvis (1989), elevated CO2 enhances the rate of carbon supply, whereas it decreases
the proportion of leaf biomass production.
Khurana and Singh (2004c) also observed that the increased CO2 affected the
quality of foliage by decline in foliar N (Table  4.11). Similar findings were also
reported by Conroy et al. (1992), Cornelissen et al. (1999b), Coleman et al. (1993),
Curtis et  al. (2000) and Zak et  al. (2000). The reduction in N concentration was
mostly as a result of the dilution effect of leaf starch and not due to the decrease in
the absolute N quantity in leaf. In contrast, the rate of CO2 assimilation as well as
the rate of net assimilation was enhanced despite a decrease in foliar N concentra-
tion. The reduction in the Rubisco content is commonly assumed to be the conse-
quence of lower leaf N content and is considered an important physiological
mechanism which is associated with the acclimation of photosynthesis at high CO2
concentration (Stitt 1991). Luo et al. (1994) reported that plants can compensate for
low concentration of leaf N (mass basis) with their additional mesophyll cells,
which supports in maintaining or even enhancing photosynthetic capacity per unit
of leaf area. Further, CO2 also has the capacity to regulate the kinetic property as
well as activity of Rubisco. Under the ambient atmospheric CO2 concentration, the
function of Rubisco is below its Michaelis–Menten constant (Km). Therefore, at the
elevated CO2, the carboxylation rate enhances sharply. In the study of Khurana and
4.4  Seed and Seedling Ecology 101

Ambient CO2 Elevated CO2


LS Seedlings SS Seedlings
A. procera
11
10
9
8
7
6
5
4
3
A. nilotica
70
60
50
40
30
20
10

P. emblica
Plant Height (cm)

35
30
25
20
15
10
5
T. arjuna
50

40

30

20

10

T. chebula
35
30
25
20
15
10
5
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Days

Fig. 4.10  Height of seedlings of five dry tropical tree species at ambient and elevated CO2 levels
over a 60-day period. Vertical bars represent ±1S.E. Note that scale for Y axis is different for dif-
ferent species; this was necessary for maintaining clarity (Source: Khurana and Singh 2004c)
102 4  Plant Traits and Regeneration

LS Seedlings SS Seedlings LS Seedlings SS Seedlings

A. procera A. nilotica
30d
600
60d
400
30d 60d
60d 30d
200 30d 60d

200

400

600
Percent increase in biomass

P. emblica T. arjuna

600
30d
400 30d
30d 60d
200 60d 30d 60d 60d

200

400

600
T. chebula
Leaves
60d
60d
600 Stem
30d

400 30d Root

200

200

400

600

Fig. 4.11  Per cent increase in biomass of 2- and 3-month-old seedlings of five dry tropical species
at elevated CO2 levels (Source: Khurana and Singh 2004c)

Singh (2004c), stomatal conductance as well as transpiration rate decreased, while


water use efficiency increased under the elevated CO2. This reduction in the ­stomatal
conductance as a result of partial closure of stomata may result into a depression in
the leaf transpiration (see reviews of Eamus and Jarvis 1989, Drake et  al. 1997,
Ceulemans and Mouseau 1994). Further, maximum increment in WUE observed in
4.4  Seed and Seedling Ecology 103

Ambient CO2 Elevated CO2


LS Seedlings SS Seedlings
150 A. procera
125
100
75
50
25
0
A. nilotica
150
125
100
75
50
25
Relative growth rate ( mg g-1 d-1)

0
P. emblica
150
125
100
75
50
25
0

T. arjuna
150
125
100
75
50
25
0

T. chebula
150
125
100
75
50
25
0
30 60 30 60
Days

Fig. 4.12  Relative growth rate (mg g−1 day−1) of large-seed (LS) and small-seed (SS) seedlings of
five dry tropical tree species under ambient and elevated CO2 levels. Bars represent ±1SE (Source:
Khurana and Singh 2004c)
104

Table 4.10  Specific leaf area (cm2 g−1) and root/shoot ratio for seedlings of five dry tropical tree species grown at two CO2 levels
Large-seed seedlings Small-seed seedlings
Ambient CO2 Elevated CO2 Ambient CO2 Elevated CO2
Species 30 days 60 days 30 days 60 days 30 days 60 days 30 days 60 days
Specific leaf area (cm2 g−1)
Albizia procera 879.86 ± 18.87 910.0 ± 642.29 593.97 ± 73.48 626.80 ± 39.06 1387.91 ± 173.02 991.92 ± 20.10 1230.37 ± 77.92 928.75 ± 109.52
Acacia nilotica 296.78 ± 31.68 180.03 ± 13.97 101.32 ± 8.55 114.60 ± 18.29 598.36 ± 127.8 329.77 ± 18.75 383.09 ± 62.88 246.50 ± 19.43
Phyllanthus 357.50 ± 16.68 325.10 ± 29.43 237.98 ± 19.83 268.37 ± 34.23 583.53 ± 25.84 298.75 ± 45.35 410.36 ± 35.41 222.80 ± 22.50
emblica
Terminalia 470.54 ± 11.36 487.94 ± 13.87 287.59 ± 20.64 351.15 ± 10.01 234.61 ± 44.17 211.48 ± 23.54 158.97 ± 13.05 202.13 ± 12.81
arjuna
Terminalia 344.81 ± 19.81 189.80 ± 9.01 167.53 ± 3.47 66.83 ± 9.03 501.14 ± 14.99 373.18 ± 14.82 276.40 ± 32.84 186.01 ± 24.05
chebula
R/S ratio
Albizia procera 0.48 ± 0.02 0.59 ± 0.04 0.62 ± 0.14 0.61 ± 0.05 0.64 ± 0.04 0.61 ± 0.02 0.67 ± 0.04 0.63 ± 0.04
Acacia nilotica 0.31 ± 0.05 0.27 ± 0.005 0.35 ± 0.01 0.28 ± 0.01 0.50 ± 0.03 0.41 ± 0.05 0.61 ± 0.13 0.48 ± 0.02
Phyllanthus 0.70 ± 0.03 0.65 ± 0.03 0.72 ± 0.02 0.67 ± 0.03 0.94 ± 0.04 0.69 ± 0.01 0.99 ± 0.05 0.80 ± 0.14
emblica
Terminalia 0.51 ± 0.03 0.47 ± 0.001 0.58 ± 0.02 0.55 ± 0.01 0.85 ± 0.05 0.84 ± 0.03 0.92 ± 0.02 0.95 ± 0.08
arjuna
Terminalia 0.35 ± 0.04 0.33 ± 0.04 0.45 ± 0.02 0.45 ± 0.02 0.56 ± 0.05 0.91 ± 0.09 0.72 ± 0.09 1.03 ± 0.04
chebula
Source: Khurana and Singh (2004c)
4  Plant Traits and Regeneration
4.4  Seed and Seedling Ecology 105

Table 4.11  Foliar N and total leaf N for 1-month-old seedlings of five dry tropical species grown
at two CO2 levels
Foliar N (mg g−1 leaf) Total foliar N (mg plant−1)
Seed Ambient
Species size CO2 Elevated CO2 Ambient CO2 Elevated CO2
Albizia Large 16.93 ± 0.09 14.59 ± 0.53 4963.41 ± 309.77 9167.96 ± 1678.56
procera seed
Small 16.94 ± 1.79 12.83 ± 2.41 1795.14 ± 147.01 2401.15 ± 512.90
seed
Acacia Large 28.72 ± 1.58 20.62 ± 0.30 2279.45 ± 360.89 7804.33 ± 326.83
nilotica seed
Small 23.10 ± 0.67 16.00 ± 0.31 532.78 ± 111.53 767.40 ± 143.56
seed
Phyllanthus Large 13.73 ± 0.44 11.38 ± 0.43 1596.55 ± 106.67 3222.73 ± 207.20
emblica seed
Small 13.0 ± 0.28 11.53 ± 0.52 826.81 ± 22.69 1331.72 ± 88.63
seed
Terminalia Large 19.24 ± 0.43 16.97 ± 0.078 43005.17 ± 533.55 104526.20 ± 2892.4
arjuna seed
Small 17.35 ± 0.51 16.15 ± 0.36 12822.29 ± 1751.37 32982.82 ± 2817.36
seed
Terminalia Large 16.90 ± 0.58 16.64 ± 0.17 3006.70 ± 286.43 7912.69 ± 155.69
chebula seed
Small 14.80 ± 0.41 14.8 ± 0.93 1147.96 ± 6.33 3357.51 ± 259.91
seed
Source: Khurana and Singh (2004c)

the two fast-growing species (Table  4.12) suggested their enhanced resistance to
water stress which is commonly known to be a major factor for the seedling mortal-
ity in the TDF of India (Khurana and Singh 2000).
Khurana and Singh (2004c) reported that the increment in biomass and other
growth traits due to increased CO2 concentration could be associated with the dif-
ferences in seed size. According to their study, species producing bigger seeds
(Terminalia arjuna and T. chebula) showed a greater response to increased CO2
levels compared to the species with smaller seeds (Albizia procera and Phyllanthus
emblica). Even within species, seedlings produced from larger seeds exhibited a
higher response to the increased CO2 levels than that observed in the seedlings from
smaller seeds. The species producing bigger seeds contain more energy as well as
nutrient reserves compared to smaller seeds and therefore might be capable to uti-
lize increased CO2 more efficiently.
Bazzaz and Miao (1993) found that in the environments with nutrient-poor soils,
the late-successional seedlings exhibited better performance under elevated CO2
conditions compared to the early successionals. Based on the results from the simu-
lation modelling analysis of the effect of elevated CO2 on the plant growth, Lloyd
and Farquhar (1996) reported that the species exhibiting inherently slow growth
will show more strong response to elevated CO2 compared to those showing a high
106 4  Plant Traits and Regeneration

Table 4.12  Transpiration rate, stomatal conductance, CO2 assimilation rate and WUE of 2-month-­
old five dry tropical tree seedlings grown at two CO2 levels
Large seed seedlings Small seed seedlings
Species Ambient CO2 Elevated CO2 Ambient CO2 Elevated CO2
Transpiration rate (m mol m s )
−2 −1

Albizia procera 12.37 ± 0.2 11.18 ± 0.4 12.07 ± 0.27 9.49 ± 0.5


Acacia nilotica 13.52 ± 0.01 13.28 ± 0.01 13.34 ± 0.06 11.90 ± 0.63
Phyllanthus emblica 13.32 ± 0.90 12.50 ± 0.35 13.80 ± 0.6 10.74 ± 0.30
Terminalia arjuna 11.16 ± 0.2 10.34 ± 0.29 11.82 ± 0.25 11.52 ± 0.47
Terminalia chebula 11.60 ± 0.30 10.00 ± 0.50 10.70 ± 0.49 10.20 ± 0.58
Stomatal conductance (μmol m−2 s−1)
Albizia procera 1.28 ± 0.04 0.99 ± 0.03 1.21 ± 0.06 0.91 ± 0.04
Acacia nilotica 1.15 ± 0.03 0.99 ± 0.26 1.09 ± 0.03 0.77 ± 0.02
Phyllanthus emblica 1.24 ± 0.05 1.04 ± 0.01 1.21 ± 0.07 1.03 ± 0.03
Terminalia arjuna 3.11 ± 0.41 2.64 ± 0.57 8.39 ± 0.50 7.32 ± 0.48
Terminalia chebula 0.69 ± 0.02 0.66 ± 0.02 0.70 ± 0.02 0.59 ± 0.11
CO2 assimilation rate (μmol m−2 s−1)
Albizia procera 4.90 ± 0.04 9.50 ± 0.03 3.88 ± 0.06 9.64 ± 0.035
Acacia nilotica 4.66 ± 0.03 9.83 ± 0.03 4.08 ± 0.03 10.27 ± 0.02
Phyllanthus emblica 5.85 ± 0.05 8.19 ± 0.01 5.02 ± 0.06 10.16 ± 0.03
Terminalia arjuna 3.70 ± 0.41 11.09 ± 0.57 2.54 ± 0.52 7.34 ± 0.48
Terminalia chebula 2.76 ± 0.03 5.39 ± 0.02 2.58 ± 0.02 7.73 ± 0.11
Water use efficiency (μmol mol−1)
Albizia procera 396.53 ± 34.64 846.59 ± 58.40 342.78 ± 48.86 1017.83 ± 41.32
Acacia nilotica 344.81 ± 26.82 739.66 ± 38.11 305.82 ± 26.51 874.46 ± 140.4
Phyllanthus emblica 449.13 ± 72.86 656.19 ± 24.31 367.11 ± 41.75 943.52 ± 61.48
Terminalia arjuna 333.19 ± 34.90 978.57 ± 146.10 216.19 ± 36.25 627.72 ± 175.43
Terminalia chebula 236.88 ± 11.67 448.77 ± 59.41 241.31 ± 7.17 654.97 ± 203.45

potential growth rate. Several other studies have also reported large growth stimula-
tion in the N-fixing plants exposed to elevated CO2 (e.g. Zanetti et al. 1996; Vogel
et al. 1997). Khurana and Singh (2004c), however, reported that legumes, as a func-
tional group, did not show better response compared to the non-legumes. Among
the two legumes studied by them, Acacia nilotica (comparatively slow growing)
exhibited a more strong positive response to increased CO2 compared to the faster-­
growing Albizia procera. Therefore, according to the findings of Khurana and Singh
(2004c), the response of elevated CO2 was highly marked in the legumes with
slower growth rate.

4.4.3.5  Predators and Parasites

Based upon the type, intensity, predation site and also the kind of seed, the impact
of predators and pathogens upon seeds and seedlings may be either harmful or
advantageous. Janzen (1981) observed that among the 975 species found in
4.5 Regeneration 107

Guanacaste Province, almost 100 species were regularly eaten with particular beetle
larvae which develop and kill the seeds. According to Mohamed-Yasseen et  al.
(1994), thick seed coats prevent penetration of insect ovipositors, while phenolic
compounds in the seed coat inhibited the growth of microorganisms. In
Mesoamerican TDFs, vertebrates can function as pre-dispersal/post-dispersal pred-
ators (Dirzo and Dominguez 1995). For example, Gryj (1990) performed an experi-
ment by placing the seeds of Erythroxylum havanense on the soil with protection
and without protective exclosures for the vertebrates; he found that the protected
seeds exhibited 3.7 times greater chance of successful establishment. According to
Howe and Smallwood (1982), at the places exposed to frequent frugivores, the natu-
ral selection could favour plant species exhibiting animal-dispersible fruits; while
the habitats where fruit-eating mammals are rare, competitive seedlings may
become abundant.

4.5  Regeneration

It is apparent that the duration and the seasonality of the dry period influence the
structure and ecophysiological processes of the TDF vegetation (Lieberman 1982;
Reich and Borchert 1984; Lieberman and Lieberman 1984). The seasonality in dry
and wet periods restricts the time period of the important ecosystem processes such
as seed germination, seedling survival, establishment and development of the plants.
In Guanacaste, Costa Rica, and in Yucatán, Mexico, an abrupt decline in the survival
of seedlings during the dry season has been reported in few studies (e.g. Gerhardt
1993; Ceccon et  al. 2004). The disturbance history, primarily the extent of the
removal of forest trees, the management system after the disturbance, the introduc-
tion of any exotic species and the proximity of intact forest are the factors which
impact the regeneration and biomass of secondary forests (Griscom and Ashton
2011). The disturbances created by tree falls also influence the composition and
structure of tropical forests (Bongers et al. 2009; Denslow 1985; Uriarte et al. 2009)
as several tropical species depend on canopy gaps for regeneration (Kellner et al.
2009; Schnitzer and Carson 2001; White 1979). For TDFs, however, various studies
have reported that gaps play a minor role in the regeneration. There is evidence that
the rate of gap formation in TDF is lower compared to the average tropical humid
forest (Durán 2004). Also, the gap size observed in these forests tends to be smaller
compared to the tropical humid forest (Whigham et al. 1999). According to Swaine
et al. (1990) and Mooney et al. (1995), the phenomenon of gap-phase dynamics is
not applicable for the TDF ecosystems.
Formation of canopy gaps in TDFs is infrequent because (i) size of trees are
small, and therefore wind is not able to throw (Durán 2004) or snap them (Putz et al.
1983); (ii) the root/shoot ratios of trees are relatively large (Richards 1996); (iii)
trees generally tend to die while standing, which leads to the formation of small or
ill-defined gaps (Durán 2004); and (iv) TDFs experience a short duration of wet
season and thus are less frequently exposed to storms which are capable for causing
108 4  Plant Traits and Regeneration

structural damage (Dickinson et al. 2001, but see Dunphy et al. 2000 for counterex-
ample). However, few studies have reported some role of canopy gaps in TDFs, for
example, Dechnik-Vázquez et al. (2016) investigated 75 canopy gaps in a Mexican
TDF and observed their potential consequences for the regeneration of the forest
species. They indicated a modest but more active influence of tree-fall gaps in the
regeneration of TDF species and confirmed the role of wind throw as the main cause
of tree-fall gaps in the TDF of Nizanda, in conformity with other studies (Schliemann
and Bockheim 2011). The finding highlights the impact of hurricanes, strong wind
gusts or gales in determining the forest dynamics in regions prone to these phenom-
ena (Baker et al. 2005; Bullock 1986; Dittus 1985; Imbert et al. 1996; Lewis and
Bannar-Martin 2012; Lin et al. 2011; Lugo 2008; Quigley and Platt 2003; Whigham
et al. 2003).
Several authors have reported that dry fruits having small seeds with low water
content are commonly associated with wind-dispersal mechanism which is an
important characteristic of TDF species (Bullock 1995; Griz and Machado 2001;
Figueiredo 2002). It has also been documented that compared to species producing
animal-dispersed seeds, the species producing small, wind-dispersed seeds are not
highly affected by fragmentation or hunting (Gillespie 1999; Janzen 2002). Wind-­
dispersed seeds are produced in markedly higher numbers and, due to small size, are
dispersed to a few hundred meters away from the source of propagules as compared
to the vertebrate-dispersed seeds (Willson and Crome 1989; Zimmerman et  al.
2000; Cubiña and Aide 2001). For example, according to an observation by Holl
(1999), anemochoric seeds were about 47 times more frequent compared to animal-­
dispersed seeds (3 vs. 141 seeds m−2 year−1) in open pastures at a distance of up to
250 m from the adjacent forest. Therefore, wind-dispersal mechanism allows spe-
cies to overcome an important limitation for forest regeneration, i.e. lack of the
arrival of seeds in open areas. In few tropical pastures (Posada et al. 2000), there is
an overdominance of wind-dispersed species which is also found in some tropical
as well as subtropical secondary forests (Finegan and Delgado 2000; Marcano-Vega
et al. 2002). In contrast, trees or shrubs growing in abandoned agricultural land usu-
ally provide perches for birds and bats, which leads to the increase in the number of
seeds which are dispersed by vertebrates (Willson and Crome 1989; da Silva et al.
1996; Nepstad et  al. 1996; Slocum and Horvitz 2000). Moreover, some species
colonize pastures because they are dispersed by cattle or horses (Janzen 1981; Aide
et al. 2000); while due to extinction or reduction in the population of seed dispers-
ers, many other species are not able to colonize even after several decades of sec-
ondary succession (Aide et  al. 2000; Marcano-Vega et  al. 2002). Although the
seasonally dry forests exhibit the characteristic feature of the presence of a high
quantity of wind-dispersed, dry fruits, as compared to the moist forests, they do
have many species producing fleshy fruits and large seeds (Gentry 1995; Justiniano
and Fredericksen 2000; Griz and Machado 2001; Figueiredo 2002).
Regeneration in TDFs exhibits substantial differences from the process of regen-
eration in tropical humid forests; for example, sprouting in TDF is considered to be
relatively more important as they experience germination constraints and high mor-
tality of seedlings (Vieira and Scariot 2006). Most TDF species resprout after dis-
4.5 Regeneration 109

turbances, start with a vigorous shoots and do not carry the burden of vulnerable life
stages (Miller and Kauffman 1998; Kammesheidt 1999; Bond and Midgley 2001;
Kennard et al. 2002). Hardwood species are specially considered as better sprouters
as their roots and stems exhibit slower decay compared to softwood species (Vieira
et al. 2013). Also, the species containing high root starch concentrations and those
exhibiting high root to shoot ratios are generally strong resprouters (Bond and
Midgley 2001). However, due to frequent harvesting, fire and intensive use of trac-
tors, several species have been observed to lose resprouting ability (Uhl et al. 1988;
de Rouw 1993; Sampaio et al. 1993; Nepstad et al. 1996), while few TDF species
are either weak sprouters or non-sprouter (Bond and Midgley 2001; Vesk and
Westoby 2004).
In conclusion, plant functional traits (PFTs) can integrate the evolutionary and
ecological history of a species and can potentially predict its response to as well as
the effect of environmental conditions on the functioning of the TDF ecosystem.
Nevertheless, ecological studies are needed for the development of a comprehensive
list of ecologically important functional traits. Studies in Indian TDFs have reported
that the tree species share the important phenotypic attributes of the dry deciduous
forest. Small leaf area, medium leaf texture, rough bark texture as well as medium
deciduousness are the characteristic traits of the tropical dry deciduous forest veg-
etation. Study in TDF has reported that a number of species as well as importance
values were greater for the medium- or less-deciduous trait classes as compared to
the high-deciduous trait category, exhibiting a trade-off between transpiration and
the period of biomass synthesis. Study has indicated existence of the colonization—
competition trade-offs among species, leading to the interplay of biotic interactions,
intensity of disturbances and dispersal events, which together control local com-
munity structure and species diversity. The variation in seed sizes allows them to
occupy different types of niches available in the dry forest according to gradients of
light, moisture and temperature, which contributes to the maintenance of the total
species diversity. Healthy seeds and seedlings are the most important assets for any
restoration programme. Further studies, however, are needed on seed longevity,
viability and dormancy, suitable temperatures for seed germination and conditions
for seed storage in order to reduce the loss of viability. It has been reported that the
short-term exposures of seeds to elevated CO2 together with exogenous inputs of N
and P can produce bigger seedlings for introduction into forests. Small seeds could
easily enter into the soil compared to large seeds and therefore facilitate the forma-
tion up of a persistent seed bank. Moreover, large seeds, within species, become the
first crop which germinate early in the beginning of rainy season as they have high
germinability. However, the early rainy season insect outbreaks lead to the die-off
of majority of the early germinating seeds due to predation. Under such situations,
small seeds act as the back-up crop of seedlings. Studies in Indian TDF have
revealed a wide variation in response to the exogenous N input. The responses of
species did not follow the functional types (namely, legumes, non-legumes; pioneer,
non-pioneer; evergreen, deciduous; large-seeded, small-seeded; etc.); rather, the
response of species differed more or less individualistically. As none of the species
exhibited an adverse effect of N input, we can expect a higher forest biomass in the
110 4  Plant Traits and Regeneration

future due to N deposition. Another study has revealed that due to elevated CO2, the
rate of biomass accumulation in forest stands is likely to increase in the near future.
This could promote the speed of succession process and ecosystem development in
the TDF regions subjected to frequent disturbance. However, the study also found
that the positive influence of elevated CO2 was only for a limited time period and the
impact was downregulated if the exposure time increased more than 30  days.
Relatively more increase in growth rates of slow-growing species compared to fast-­
growing ones may disturb the competition hierarchy and could lead to changes in
species composition. Species having larger seeds exhibited a more robust response
to elevated CO2, and therefore it may be expected that in the future, such species
would dominate the forest. Even within a species, individuals emerging from larger
seeds would dominate over those with smaller seeds. These could result in a decline
in the ecosystem diversity in TDFs.
Chapter 5
Productivity and Nutrient Cycling

Primary productivity and nutrient cycling are the two fundamental functions of
­ecosystems which form the basis for a majority of ecosystem services important to
humans. Occurring in relatively harsh environment and frequently exposed to
drought and nutrient limitation, the tropical dry deciduous forest still realizes mark-
edly high net primary productivity, which is stored both above and below ground.
At global level, a high level of carbon sequestration is exhibited by the widely
­distributed tropical dry deciduous forest. The net primary productivity of these for-
ests is expected to depend on the amount and seasonal distribution of rainfall, and
summer rainfall variations could alter the production and respiration rates and thus
affect the annual carbon balance. Studies in many cases have indicated the aggrad-
ing nature of the forest and marked resorption of nutrients during senescence. The
interaction between soil fertility as well as foliar nutrient resorption has a p­ ronounced
effect on litter quality and, therefore, on the rate of decomposition and nutrient
cycling. Obviously the nutrient cycling is conservative. In this chapter we briefly
describe the productivity levels, litterfall and decomposition, nutrient conserving
strategies and the pattern of nutrient cycling in the tropical dry deciduous forests.

5.1  Carbon Stock and Biomass Accumulation Pattern

Forests account for approximately 70–90% of terrestrial above- and belowground


biomass (Houghton et al. 2009), and they therefore exhibit a major global carbon
sink (Brienen et al. 2015). The total carbon stock in the forest ecosystems has been
estimated to be about 1150 Gt, of which 49% is contributed by the boreal forests,
14% by temperate forests and 37% by tropical forests (Malhi et al. 1999). According
to the published estimates of above-ground biomass for TDF, the quantity varies by
an order of magnitude ranging from 28 to 390  Mg ha−1 (Jaramillo et  al. 2011;
Martínez-Yrízar 1995). Pan et al. (2011) observed that in tropical forest, the sum of
above and below ground live biomass accounts to 262 Pg C. On the basis of the

© Springer Nature Singapore Pte Ltd. 2017 111


J. S. Singh, R. K. Chaturvedi, Tropical Dry Deciduous Forest: Research Trends
and Emerging Features, https://doi.org/10.1007/978-981-10-7260-4_5
112 5  Productivity and Nutrient Cycling

Table 5.1  Global area and carbon stocks in seasonally dry tropical forest (SDTF) grouped by
region. SDTF area was estimated using the WWF ecoregion map. Current SDTF was calculated by
summing C stock estimates from Saatchi et al. (2011) within SDTF area delineated by the WWF
biome map. Potential SDTF C stocks were estimated using the precipitation-biomass relationship
developed from the literature review
SDTF area Mean SDTF MAP Current SDTF C (Pg) Potential SDTF
Region (million km2) by area (mm) (Saatchi et al. 2011) C (Pg)
Americas 1 1088 3.28 8.30
Asia 1.5 1119 5.05 11.86
Africa 0.2 1403 0.34 1.89
World 2.7 1187 8.67 22.06
Source: Becknell et al. (2012)

carbon stock data estimated by Saatchi et al. (2011), 8.7 Pg of carbon stock are pres-
ently being stored in the above-ground biomass of the vegetation in seasonally dry
forests worldwide (Becknell et al. 2012). This total quantity of carbon worldwide
can be simplified down into 3.28, 5.05 and 0.34 Pg C contributed by the Americas,
Asia and Africa, respectively (Table 5.1). The regional contribution of potential bio-
mass reported for Americas, Asia and Africa, respectively, stands at 8, 12 and 2 Pg
C (Table 5.1). The annual net carbon sequestration accounted by tropical forests and
savanna combined together could register for 60% of the global terrestrial photo-
synthesis (Field et  al. 1998). In tropical forests, the carbon stock is distributed
almost in equal amounts between vegetation and soil, while in the forests located at
the high latitudes, about 84% of the carbon is in the form of soil organic matter, and
only 16% carbon is contributed by the active living biomass (Malhi et al. 1999).
According to Saatchi et al. (2011), tropical forests register for about 247 Gt vegeta-
tion carbon stocks, in which around 193 Gt is stored in the above ground and only
54 Gt belowground. The tropical dry forests (TDFs) cover about 42% landmass in
the tropics (Miles et al. 2006) and account for around 110 Gt vegetation carbon as
compared to 134 Gt vegetation carbon stock of tropical rainforest (Foley 1995).
In the Amazon basin, biomass is precisely correlated with the annual rainfall
recorded across a gradient ranging from 1000 to 2400 mm (Saatchi et al. 2007), and
generally also the biomass measured in mature tropical forest tends to increase
along the gradient of precipitation (Brown and Lugo 1982; Murphy and Lugo 1986;
Malhi et al. 2006). Since deciduous trees are a dominant component of TDF, and the
duration of the leafless period is well correlated with the amount of annual precipi-
tation, more precipitation results into more annual biomass accumulation and, ulti-
mately, higher overall biomass stocks, therefore precipitation generally determines
an upper limit for the storage of carbon in the TDF (Brown and Lugo 1982).
Although in the wetter forests, nutrients and the availability of light could be more
important factor for limiting plant growth and carbon storage (Graham et al. 2003;
Cleveland et al. 2011), the strong and long period of droughts in the TDFs mean that
the availability of water generally constrains the growth (Eamus 1999).
The biomass accumulation observed in secondary dry forests is principally a
­function of the age of stand, but the pattern of accumulation is also influenced by
changes in the mean annual precipitation (MAP), and, as reported by Becknell et al.
5.1  Carbon Stock and Biomass Accumulation Pattern 113

(2012), the maximum biomass accumulated according to the age of sites is greater in
the wettest sites (1500–2000 mm MAP) compared to the drier or intermediate sites.
Few studies have also indicated that fertility of soil or disturbances are not as impor-
tant as are the climatic variables for determining the growth rates of tree and forest
(Toledo et  al. 2011). Three processes have been suggested to account for the net
changes in the above-ground biomass of forest: (1) biomass increment from recruit-
ing trees, (2) biomass increment in the existing trees, and (3) biomass decline due to
tree mortality (Brienen et al. 2015; Rozendaal and Chazdon 2015). The availability
of soil nutrients (Quesada et  al. 2012), quantity or biomass of vegetation (i.e. the
initial stand biomass) and quality of vegetation (i.e. species richness and functional
attributes of the species; Finegan et  al. 2015; Lohbeck et  al. 2015b) have been
reported to act simultaneously for shaping the biomass change and productivity. The
presence of conservative trait values which enhance drought tolerance, for example,
dense wood (Pineda-García et  al. 2013) and lower specific leaf area (Poorter and
Bongers 2006), could enhance productivity in the TDF; however they may reduce the
biomass productivity in wet forests (Malhi et al. 2004; Finegan et al. 2015). Since the
effects of niche complementarity are significant in highly stressful environments
(Paquette and Messier 2011), it should be more commonly detected in the dry forests
compared to that in wet forests. The biomass dynamics of vegetation in TDFs are
generally shaped by the quantity of vegetation followed by the quality of vegetation
(Prado-Junior et al. 2016). The abundance of conservative species (i.e. species per-
forming better in drought-stressed environment) in TDF is correlated with high bio-
mass increment and storage.
Ecosystem respiration, comprised of heterotrophic (i.e. decomposers) as well as
autotrophic (i.e. plants, roots) components, is considered to be the largest compo-
nent exhibiting net ecosystem production (NEP) in the forested ecosystems and may
therefore determine the carbon sequestration potential of an ecosystem (Schlesinger
and Andrews 2000). In the TDF, amount of carbon exchange is highly associated
with the annual as well as seasonal variations of rainfall due to a tight association of
rainfall with ecosystem processes, for example, net primary production (NPP) (Sala
et al. 1988; Méndez-Barroso et al. 2009), rate of litter decomposition and respira-
tory fluxes (Fierer and Schimel 2003; Anaya et al. 2007), and transpiration (Scott
et al. 2009; Reyes-García et al. 2012; Peng et al. 2013). Variations in summer rain-
fall may alter the production as well as respiration rates and therefore influence the
annual carbon balance (Hutyra et al. 2007). Ecosystem respiration has been reported
to substantially influence the net carbon balance in the TDF (Goulden et al. 1998;
Valentini et al. 2000). According to Yépez et al. (2007), the CO2 released from res-
piration in seasonally dry ecosystems, particularly during the first phase of summer
rains can be greater than 45% of the carbon accumulated by plants during the whole
growing season. The variations in rainfall distribution (i.e. timing, duration and
magnitude) may alter respiration and, therefore, can i­nfluence the overall carbon
balance in the TDF (Rohr et al. 2013; Cable et al. 2013; Reichstein et al. 2013).
The time lag due to the impact of previous wetness conditions influences the
ecosystem carbon fluxes, and therefore, the responses due to the precipitation events
observed during the growing season are non-linear (Reynolds et  al. 2004; Cable
114 5  Productivity and Nutrient Cycling

et al. 2013; Lupascu et al. 2014). Reichmann et al. (2013) showed that greater pri-
mary production than the expected value results in periods with previous wet condi-
tions, while lower production than the expected value occurs in periods with
previous dry conditions. A longer preceding period of dry condition enhances the
rates of decomposition, mineralization, as well as microbial activity and therefore
favours carbon accumulation (Jarvis et al. 2007; Yépez et al. 2007; Cable et al. 2008;
Yan et al. 2014).
Verduzco et al. (2015) measured the carbon and water fluxes for a continuous
period of about 4.5 years in a TDF and observed that variations in seasonal precipi-
tation can markedly affect carbon flux dynamics. Positive correlations of wet winter
conditions with increment in the normalized difference vegetation index (NDVI)
values indicated an extended period of vegetation processes during end of winter
and spring season (Verduzco et al. 2015). The increased respiration flux before the
vegetation green-up also suggested the presence of winter soil moisture, as well as
increase in labile carbon source for heterotrophic respiration.
As discussed earlier also, enhanced rates of microbial activity are observed after
rewetting of dry soils (Jarvis et al. 2007). The increased rates are governed by the
soil aggregate breakdown (Birch 1958; Appel 1998; Xiang et al. 2008), resulting
into an increment in the microbial populations (Griffiths and Birch 1961), and
osmoregulant turnover (Fierer and Schimel 2003), leading to the triggering of the
microbially driven processes of decomposition, mineralization and CO2 emission
pulses (Anaya et al. 2012; Austin et al. 2004; Carbone et al. 2011). Seasonally dry
forests located in Brazil also exhibited a similar relationship among the period of
previous wetness, availability of soil organic matter and the subsequent rate of
decomposition (Sanches et al. 2008). Robles-Morua et al. (2015) have suggested
that drier winter conditions as well as an intense summer monsoon due to the impact
of climate change scenario might well be responsible for promoting a stronger car-
bon sink in the TDFs. Evapotranspirational water exchanges in TDF play a signifi-
cant role in the transfer of water vapour to the atmosphere which can potentially
generate precipitation for recycling of water (Dominguez et  al. 2008; Méndez-­
Barroso and Vivoni 2010). Increasing availability of water may or may not result in
increasing productivity but restricting the water availability certainly lowers the
productivity (Cavelier et al. 1999; Yavitt and Wright 2001). Water scarcity leads to
increase in the tree mortality (Segura et  al. 2003) which is negatively associated
with annual precipitation with a 2- or 3-year delay (Suresh et al. 2010). Water scar-
city also stimulates the occurrence of leaf abscission as well as dormancy in several
TDF species. Longer dry period corresponds to increased mortality and reduction in
productivity, accumulation of biomass and the maximum attainable biomass.
Several workers have studied the impact of rainfall regime on ecosystem carbon
cycling in tropical forests (e.g. Schuur 2003; Powers et al. 2009b; Posada and Schuur
2011; Malhi et al. 2015), but the total understanding of the impact of precipitation
remains incomplete (Powers et  al. 2011; Marín-Spiotta and Sharma 2013). The
availability of water controls majority of the dominant ecosystem processes in TDFs
(Campo et  al. 1998, 2001); the longer drought period results in the reduction of
decomposition rates that leads to greater carbon storage on the forest floor as well as
5.1  Carbon Stock and Biomass Accumulation Pattern 115

higher accumulation of particulate organic matter in the mineral soils, i.e. the free
light fraction (Cuevas et  al. 2013; Roa-Fuentes et  al. 2013). This free light and
organic matter (e.g. Schmidt et al. 2011; Lehmann and Kleber 2015) has important
implications for carbon sequestration. Under such conditions, organic matter dynam-
ics of the forest might be mainly determined by the chemical recalcitrance as well as
the environmental conditions influencing the microbial activity. These observations
indicate that the soil organic matter (SOM) pool as well as the carbon storage capac-
ity may indeed be much susceptible to climatic changes.

5.1.1  D
 ry Matter Dynamics, Storage and Flux of Nutrients
in TDF of India: A Case Study

Singh and Singh (1991a) studied net primary productivity of vegetation on three
sites of a TDF located in northern India. Estimates of net production recorded for
different components of trees, shrubs and herbs are summarized in Table 5.2. Total
net production (average for the two annual cycles) was greater for Site 1, located on
the topmost position of the hill (19.2 t ha−1 year−1), followed by the Site 2, located
on the slope of the hill (14.8), and Site 3, located at the base of the hill (11.3), and
averaged 15.1 t ha−1 year−1. Total above-ground net production of trees on different
sites ranged from 2.8 to 7.7 t ha−1 year−1. The contribution of foliage production was
38–57% to the total above-ground net production of trees. Among the perennial
aerial parts, the contribution of branches and boles registered 25–38% and 18–26%,

Table 5.2  Net production in trees, shrubs and herbs and total vegetation on the three study sites
Net production (t ha−1 year−1 ± 1SE)
Components Site 1 Site 2 Site 3 Average
Trees
 Foliage 2.9 ± 1.7 1.6 ± 0.1 1.5 ± 0.1 2.0 ± 0.2
 Branch 2.9 ± 0.7 0.7 ± 0.1 1.1 ± 0.1 1.6 ± 0.3
 Bole 1.9 ± 0.4 0.5 ± 0.1 0.9 ± 0.1 1.2 ± 0.2
 Coarse root 0.7 ± 0.2 0.2 ± 0.03 0.3 ± 0.03 0.4 ± 0.1
 Total 8.4 ± 1.7 3.0 ± 0.3 3.8 ± 0.3 5.2 ± 0.6
Shrubs
 Foliage 2.6 ± 0.02 2.9 ± 0.1 2.2 ± 0.01 2.6 ± 0.01
 Stem 0.7 ± 0.l 0.8 ± 0.1 0.5 ± 0.1 0.6 ± 0.1
 Coarse root 0.2 ± 0.09 0.2 ± 0.03 0.1 ± 0.01 0.2 ± 0.01
 Total 3.5 ± 0.1 3.9 ± 0.1 2.8 ± 0.01 3.4 ± 0.1
Above-ground herbs 1.1 ± 0.3 0.9 ± 0.2 0.8 ± 0.2 0.9 ± 0.2
Wood and miscellaneous litter 1.8 ± 0.1 2.4 ± 0.4 1.4 ± 0.1 1.8 ± 0.2
Stand fine root 4.4 ± 0.3 4.6 ± 0.9 2.5 ± 0.6 3.8 ± 0.8
Total vegetation 19.2 14.8 11.3 15.1
Source: Singh and Singh (1991a)
116 5  Productivity and Nutrient Cycling

respectively. The contribution of roots (coarse + fine root) was substantial and
ranged from 29 to 53 t ha−1 year−1. The contribution of fine roots in the study of
Singh and Singh (1991a) was 40% of the total input of dry matter to the forest floor,
while in various other forests, the contribution of fine root has been reported within
the range 20–77% (Vogt et al. 1986). The belowground transfers through the roots
(i.e. root turnover) were 0.7 times that of aerial inputs (i.e. litterfall). Perry et al.
(1989) observed that plants in TDFs allocate a large proportion of photosynthesised
biomass to roots. In other studies, roots and mycorrhizal fungi have been reported
to contribute about 70–80% of total net primary production (Vogt et al. 1982; Fogel
and Hunt 1983). Therefore, regarding the pathway for carbon and nutrient flux, it is
evident that the fine roots are equivalent or may even exceed in importance com-
pared to the above-ground litterfall.
In the study of Singh and Singh (1991a), the net above-ground shrub production
varied from 2.7 to 3.7 t ha−1 year−1 and exhibited 23% of the total net biomass pro-
duction (Table 5.2). Above-ground herbaceous biomass production varied between
0.8 and 1.1 t ha−1 year−1, which accounted for 6% of the total net production of the
forest, which varied between 11.3 and 19.21 t ha−1 year−1 with the mean value of
15.1 t ha−1 year−1.
The dry matter values reported by Singh and Singh (1991a) for the standing
crops, net production, amount of litterfall, etc. which were converted to carbon and
further averaged across the sites have been presented in Fig. 5.1. The proportion of
above-ground carbon storage  in the forest was 83% of the total carbon stored in
vegetation and 51% of the carbon stored in the stand (i.e. vegetation plus soil). The
forest stand received a carbon input mainly through the net primary production
(NPP), at the rate of 6.7 t ha−1 year−1, and of this value, 72% was contributed by the
above-ground and 28% by the root NPP (Singh and Singh 1991a). Wood (i.e. bole
plus branch) and roots contributed for 64% of the total vegetation carbon input and
the remaining 36% of the production being channelled for the development and
elaboration of the different photosynthetic parts (i.e. tree and shrub foliage together
with herbaceous shoots). The input of carbon from the foliage compartment into the
litter compartment was observed to be 2.1 t ha−1 year−1, which is equivalent to 100%
of the amount of carbon produced by herbaceous shoot together with 86% of foliage
carbon production. About 14% (i.e. 0.3  t ha−1 year−1) of the carbon produced by
foliage did not find its path into the leaf litterfall, and it was anticipated to be lost
during the senescence. On the basis of the data for 90 woody species, Lal (1990),
observed an average decrease of 23% in the dry matter and 26% in the quantity of
carbon during the senescence of leaves. According to Singh and Singh (1991a), the
addition of carbon to the soil through the means of turnover of roots as well as
above-ground litter accounted to 1.7 and 2.3 t ha−1 year−1, respectively (turnover of
fine root was assumed to be ≤1% year−1 and turnover of litter to be 78% year−1). The
total carbon deposition from different components of vegetation to the forest floor
accounted 4.6 t ha−1 year−1 (0.3 t ha−1 year−1 carbon was lost during the senescence),
while the total net production of carbon in the forest accounted 6.7 t ha−1 year−1,
indicating that the 27% of carbon was retained in the vegetation for the short term.
The total carbon input into soil through the turnover of litter and roots was 4.0 t ha−1
5.1  Carbon Stock and Biomass Accumulation Pattern 117

Fig. 5.1  Compartment model showing annual carbon budget for the three sites of the tropical dry
forest. Values in “tanks” represent average carbon content (t ha−1 ± 1 SE). The foliage compartment
also includes above-ground standing crop of herbs, and the root compartment includes coarse roots
as well as fine roots. Net annual fluxes between the compartments are given on arrows (t ha−1
year−1 ± 1 SE) (Source: Singh and Singh 1991a)

year−1, indicating that around 40% of the carbon retained in the system over the total
annual cycle. Thus, Singh and Singh (1991a) suggested that the forest studied by
them was an aggrading system.
Singh and Singh (1991b) studied the magnitude of the storages of N, P, Ca, K
and Na on three sites in the TDF of India and investigated the partitioning of nutri-
ents in the vegetation and soil, uptake and internal recycling of nutrients, efficiency
of litterfall in nutrient cycling and contribution of fine roots to the uptake and
­transfer of nutrients. The study sites of Singh and Singh (1991b) were the same as
that of Singh and Singh (1991a). N, P, Ca, K and Na concentrations in different
components of trees, shrubs, herbs, fine roots and litter layer are summarized in
Table 5.3. The pattern of nutrient concentration in different growth forms was in the
order herb > shrub > tree. The concentration of all nutrients was highest for leaves
118 5  Productivity and Nutrient Cycling

Table 5.3  Nutrient concentrations in different components of tree, shrub and herb
Nutrient concentration (% ±1SE)
Components N P Ca K Na
Trees
 Bole 0.600 ± 0.121 0.041 ± 0.010 0.260 ± 0.160 0.196 ± 0.040 0.080 ± 0.030
 Branch 0.710 ± 0.140 0.048 ± 0.010 0.455 ± 0.040 0.399 ± 0.020 0.056 ± 0.001
 Coarse root 0.820 ± 0.100 0.043 ± 0.010 0.574 ± 0.079 0.286 ± 0.050 0.060 ± 0.003
 Foliage 2.040 ± 0.250 0.134 ± 0.020 0.401 ± 0.070 0.749 ± 0.180 0.074 ± 0.005
Shrubs
 Stem 0.770 ± 0.080 0.053 ± 0.00 3 0.460 ± 0.064 0.464 ± 0.070 0.067 ± 0.004
 Coarse root 0.790 ± 0.085 0.049 ± 0.006 0.406 ± 0.110 0.450 ± 0.050 0.055 ± 0.002
 Foliage 2.340 ± 0.300 0.159 ± 0.003 0.521 ± 0.043 l.001 ± 0.120 0.079 ± 0.006
Herbs 1.540 ± 0.110 0.131 ± 0.020 0.460 ± 0.070 0.969 ± 0.140 0.094 ± 0.010
Litter 1.160 ± 0.110 0.065 ± 0.011 0.383 ± 0.128 0.467 ± 0.159 0.071 ± 0.003
Fine root l.045 ± 0.150 0.075 ± 0.025 0.548 ± 0.174 0.327 ± 0.073 0.111 ± 0.030
Source: Singh and Singh (1991b)

(except for Ca) followed by branches and bole. The foliar N:P ratio averaged 15.
Medina (1980) reported an N:P ratio of between 12 and 16 for tropical dry decidu-
ous and evergreen vegetation. In the belowground components, fine roots (<5 mm)
had a higher concentration than the coarse roots. Shrub foliage was more nutrient-
rich than tree foliage. The quantities of nutrients (kg ha−1) in the vegetation ranged,
among the sites, between 454 and 764 N, 30 and 50 P, 222 and 384 Ca, 199 and 342
K, and 39 and 65 Na. The total vegetation biomass was highest on the site where the
standing state of nutrients was maximum (Table  5.4). Singh and Singh (1991b)
found the greatest amount of nutrient in branch and bole components, although the
bole and branch had the lowest and the foliage the highest nutrient concentrations.
The relative contribution of different tree components to the total standing state of
nutrients was in the order: branch > bole > roots > foliage. In the shrub layer, it fol-
lowed the order: stem >foliage > roots. The estimates of nutrient storage for N, P
and K in tropical forests vary widely (Table 5.5) due to differences in nutrient con-
centrations and the amount of biomass.
The estimates by Singh and Singh (1991b) for the gross uptake (i.e. amount
associated with the net primary production) of nutrients by the vegetation and by
different components of vegetation are given in Table 5.6, and those after correction
for internal cycling (on the basis of percent retranslocation of nutrients during
senescence), i.e. net uptake, are given in Table 5.7. Nutrient resorption or retranslo-
cation during senescence is discussed in more detail later in the chapter. The relative
share of shrubs in the total vegetation uptake was more substantial. Fine roots also
contributed substantially to the total uptake.
Compartment models of nutrient cycling in the forest studied by Singh and Singh
(1991b) are depicted in Figs. 5.2 and 5.3. The mean standing states of nutrients are
given in “tanks” and the net annual flux on arrows. In these figures the direction of
nutrient flux from soil to foliage indicates a one-way movement, although the ­nutrients
5.1  Carbon Stock and Biomass Accumulation Pattern 119

Table 5.4  Standing state of nutrients in tree, shrub, herb and litter layers and in fine roots on the
three study sites in TDF
Nutrient standing state (kg ha−1 ±1SE)
Components N P Ca K Na
Site 1
Trees
 Bole 163 ± 30 11 ± 1.9 71 ± 12 53 ± 9 22 ± 4
 Branch 234.8 ± 50 16 ± 3 150 ± 30 131.9 ± 27 19 ± 4
 Foliage 55 ± 9 3.6 ± 0.6 11 ± 1.8 20.1 ± 4 1.9 ± 0.3
 Coarse root 82.9 ± 14 4 ± 0.7 58.0 ± 9.6 28.95 ± 5 6.06 ± l.0
 Total 535.5 34.6 290.00 233.95 48.96
Shrubs
 Stem 69 ± 20 5±2 41 ± 14 42 ± 14 6±2
 Foliage 64.35 ± 2 4 ± 0.2 14.3 ± 0.5 27 ± 1 2 ± 0.08
 Coarse root 17 ± 5 1.1 ± 0.3 8.7 ± 2 9.7 ± 3 1.2 ± 0.3
 Total 150.35 10.1 64.0 78.7 9.2
Above-ground herb 6.16 ± 1.2 0.5 ± 0.l 1.84 ± 0.4 4 ± 0.8 0.4 ± 0.07
Stand fine root 34.38 ± 3 2.47 ± 0.2 18.0 ± 1.6 10.77 ± 1.0 3.66 ± 0.3
Litter mass 37.35 ± 2.0 2.09 ± 0.l 12.3 ± 0.6 15.0 ± 0.7 2.3 ± 0.l
Total vegetation 763.94 50.08 384.4 342.42 64.52
Site 2
Trees
 Bole 54.8 ± 10 3.7 ± 0.9 23.7 ± 6.0 17.9 ± 4 7±2
 Branch 73.2 ± 24 5±2 47 ± 16 41 ± 14 6 ± 1.9
 Foliage 32 ± 2 2 ± 0.1 6 ± 0.4 12 ± 0.7 1 ± 0.07
 Coarse root 27 ± 7 1.4 ± 0.4 19 ± 5.0 9.6 ± 2 2 ± 0.5
 Total 187 12.1 95.7 80.5 16
Shrubs
 Stem 110.64 ± 33 7.6 ± 2 66.1 ± 19 66.7 ± 19 9.6 ± 3
 Foliage 78 ± 8 5 ± 0.5 17 ± 2 33 ± 3 3 ± 0.3
 Coarse root 26 ± 7 2 ± 0.4 13 ± 4 15 ± 4 2 ± 0.5
 Total 214.64 14.60 96.1 114.7 14.6
Above-ground herb 5 ± 1.4 0.45 ± 0.l 2 ± 0.4 3.29 ± 0.87 0.3 ± 0.08
Stand fine root 41.17 ± 0.2 2.96 ± 0.02 21.59 ± 0.1 12.89 ± 0.06 4.4 ± 0.02
Litter mass 35.26 ± 0.1 1.98 ± 0.06 11.64 ± 0.3 14.18 ± 0.4 2.16 ± 0.01
Total vegetation 483.04 32.09 227.0 3 225.5 37.5
Site 3
Trees
 Bole 101.8 ± 20 7 ± l.0 44 ± 7.0 33 ± 5.5 13.6 ± 2
 Branch 130 ± 30 9 ± 2.0 83.3 ± 18 73 ± 16 10.3 ± 2
 Foliage 32 ± 2 2 ± 0.1 6 ± 0.3 11.6 ± 0.6 1.1 ± 0.06
 Coarse root 48 ± 7 3 ± 0.4 34 ± 5 17 ± 3 4 ± 0.5
 Total 311.8 21.0 167.3 134.6 29.0
(continued)
120 5  Productivity and Nutrient Cycling

Table 5.4 (continued)
Nutrient standing state (kg ha−1 ±1SE)
Components N P Ca K Na
Shrubs
 Stem 31 ± 7 2.1 ± 0.5 18 ± 4 18 ± 4 3 ± 0.6
 Foliage 50 ± 10.9 3 ± 0.7 11 ± 2.0 22 ± 0.5 1.7 ± 0.4
 Coarse root 9±2 0.5 ± 0.l 4 ± 0.9 4 ± 0.96 0.5 ± 0.1
 Total 90.0 5.6 33.0 44.0 5.2
Above-ground herb 4 ± 1.2 0.4 ± 0.l 1.3 ± 0.4 3 ± 0.77 0.3 ± 0.08
Stand fine root 22.68 ± 2 1.63 ± 0.l 11.9 ± 0.8 7.1 ± 0.5 2.4 ± 0.2
Litter mass 25.98 ± 1 1.5 ± 0.06 8.58 ± 0.4 10.4 ± 0.5 1.6 ± 0.07
Total vegetation 454.46 30.13 222.08 199.14 38.5
Source: Singh and Singh (1991b)

Table 5.5  Nutrient (N, P and K) storages in vegetation and soil compartments of certain tropical
forests. For references see Singh and Singh (1991b)

Forest type and Nutrient content (kg ha−1)


depth of soil Vegetation Soil
sampling Location N P K N P K
Wet subtropical El Verde, 1198 – 167 – – 31
(25 cm depth) Puerto Rico
Wet premontane Turrialba, 370 118 355 3310 14 85
(8 cm depth) Costa Rica
Subtropical lower Puerto Rico – 80.6 – – 705 –
montane rain
(100 cm depth)
Montane rain New Guinea 643 34 632 19,200 16 403
(30 cm depth)
Tropical wet terra San Carlos, 1084 – – 3507 – –
firme (16 cm Venezuela
depth)
Tropical wet Venezuela 336 – – 785 – –
caatinga (16 cm
depth)
Tropical moist Panama – 144 2981 – 11–33 197–508
(30 cm depth)
Subtropical dry Guanica, 370 13 205 9100 1820 7460
(85 cm depth) Puerto Rico
Dry deciduous India 717 56 462 2906 126 377
(30 cm depth)
Dry tropical Belize 670 – – 7500 – –
Dry deciduous India 455–764 30–50 199–342 1217–2365 398–482 279–449
Source: Singh and Singh (1991b)
5.1  Carbon Stock and Biomass Accumulation Pattern 121

Table 5.6  Quantities of nutrients associated with net production in different layers of vegetation
and their components (i.e. gross uptake). The wood and miscellaneous litter were regarded as a
separate component because of their varied origin
Nutrient gross uptake (kg ha−1 year−1 ± 1SE)
Components N P Ca K Na
Site 1
Trees
 Foliage 59 ± 11 4.0 ± 0.1 11.6 ± 2.0 21.7 ± 3.9 2.1 ± 0.4
 Branch 21 ± 5.0 1.4 ± 0.3 13.5 ± 3.3 12.0 ± 3.0 1.7 ± 0.4
 Bole 11.7 ± 2.5 0.8 ± 0.2 5.1 ± 1.0 4.0 ± 0.8 2.0 ± 0.3
 Coarse root 6.0 ± 1.3 0.3 ± 0.1 4.0 ± 0.9 2.1 ± 0.4 0.4 ± 0.1
 Total 97.7 6.5 34.2 39.8 6.2
Shrubs
 Foliage 62.0 ± 5.0 4.0 ± 0.03 14.0 ± 0.1 26.4 ± 0.2 2.0 ± 0.2
 Stem 6.0 ± 0.7 0.4 ± 0.1 3.4 ± 0.4 3.0 ± 0.4 0.5 ± 0.1
 Coarse root 1.5 ± 0.7 0.09 ± 0.04 0.8 ± 0.4 0.9 ± 0.4 0.1 ± 0.05
 Total 69.5 4.5 18.2 30.3 2.6
Above-ground herb 16.6 ± 4.0 1.4 ± 0.3 4.9 ± 1.2 10.5 ± 2.4 1.0 ± 0.2
Wood and miscellaneous 15.6 ± 0.9 l.0 ± 0.1 7.8 ± 0.4 4.5 ± 0.3 1.2 ± 0.1
litter
Stand fine root 46.0 ± 7.7 3.3 ± 0.3 24.0 ± 4.0 14.0 ± 2.4 4.9 ± 0.8
Total vegetation 245.4 16.7 89.1 99.1 15.9
Site 2
Trees
 Foliage 32.2 ± 2.0 2.1 ± 0.1 6.0 ± 0.4 11.8 ± 0.7 1.2 ± 0.1
 Branch 4.7 ± 0.9 0.3 ± 0.1 3.0 ± 0.6 3.0 ± 0.6 0.4 ± 0.1
 Bole 3.2 ± 0.6 0.2 ± 0.04 1.4 ± 0.2 1.0 ± 0.2 0.4 ± 0.1
 Coarse root 1.6 ± 0.2 0.08 ± 0.0 l 1.1 ± 0.2 0.5 ± 0.1 0.1 ± 0.02
 Total 41.7 2.7 11.5 16.3 2.1
Shrubs
 Foliage 67.0 ± 4.2 4.6 ± 0.3 15.0 ± 0.9 29.0 ± 1.8 2.3 ± 0.1
 Stem 5.8 ± 0.7 0.4 ± 0.1 3.5 ± 0.5 3.0 ± 0.5 0.5 ± 0.1
 Coarse root 1.6 ± 0.2 0.09 ± 0.01 0.8 ± 0.1 0.9 ± 0.1 0.1 ± 0.02
 Total 74.4 5.1 19.3 32.9 2.9
Above-ground herb 14.3 ± 3.7 1.2 ± 0.3 4.2 ± 1.1 9.0 ± 2.3 0.9 ± 0.2
Wood and miscellaneous 20.8 ± 3.1 1.3 ± 0.2 11 ± 1.6 6.1 ± 0.9 1.6 ± 0.2
litter
Stand fine root 48.0 ± 10.3 3.5 ± 0.7 25.5 ± 5.0 14.9 ± 3.0 5.0 ± 1.1
Total vegetation 199.2 13.8 71.5 79.2 12.5
Site 3
Trees
 Foliage 31.4 ± 1.9 2.1 ± 0.1 6.2 ± 0.4 11.5 ± 0.7 1.1 ± 0.02
 Branch 7.9 ± 0.9 0.5 ± 0.1 5.0 ± 0.6 4.0 ± 0.9 0.6 ± 0.1
 Bole 6.0 ± 0.6 0.4 ± 0.04 2.5 ± 0.3 1.9 ± 0.2 0.8 ± 0.1
 Coarse root 3.0 ± 0.2 0.1 ± 0.02 1.9 ± 0.2 0.9 ± 0.1 0.2 ± 0.02
 Total 48.3 3.1 15.6 18.3 2.7
(continued)
122 5  Productivity and Nutrient Cycling

Table 5.6 (continued)
Nutrient gross uptake (kg ha−1 year−1 ± 1SE)
Components N P Ca K Na
Shrubs
 Foliage 50.3 ± 0.1 3.4 ± 0.01 11.0 ± 0.02 21.5 ± 0.04 2.0 ± 0.01
 Stem 3.5 ± 0.5 0.2 ± 0.03 2.0 ± 0.3 2.0 ± 0.3 0.3 ± 0.04
 Coarse root 0.9 ± 0.1 0.05 ± 0.01 0.4 ± 0.1 0.5 ± 0.1 0.06 ± 0.01
 Total 54.7 3.7 13.4 24.0 2.4
Above-ground herb 13.0 ± 3.0 1.0 ± 0.3 4.0 ± 0.9 7.9 ± 2.0 0.8 ± 0.2
Wood and miscellaneous 12.2 ± 0.9 0.8 ± 0.1 6.1 ± 0.5 3.6 ± 0.3 0.9 ± 0.1
litter
Stand fine root 26.5 ± 6.0 1.9 ± 0.5 13.9 ± 3.5 8.3 ± 2.0 2.8 ± 0.7
Total vegetation 154.7 10.5 53.1 62.1 9.6
Source: Singh and Singh (1991b)

Table 5.7  Nutrient uptake after adjustment for internal recycling by different layers of vegetation
and their components. Wood and miscellaneous litter is regarded as a separate component because
of its varied origin
Nutrient net uptake (kg ha−1 year−1 ± 1SE)
Components N P Ca K Na
Site 1
Trees
 Foliage 19.5 ± 3.6 1.2 ± 0.02 11.6 ± 2.0 4.8 ± 0.8 2.1 ± 0.4
 Branch 21.0 ± 5.0 1.4 ± 0.3 13.5 ± 3.3 12.0 ± 3.0 1.7 ± 0.4
 Bole 11.7 ± 2.5 0.8 ± 0.2 5.1 ± 1.0 4.0 ± 0.8 2.0 ± 0.3
 Coarse root 6.0 ± 1.3 0.3 ± 0.1 4.0 ± 0.9 2.1 ± 0.4 0.4 ± 0.1
 Total 58.2 3.7 34.2 22.9 6.2
Shrubs
 Foliage 20.5 ± 1.7 12 ± 0.02 14.0 ± 0.1 5.8 ± 0.04 2.0 ± 0.2
 Stem 6.0 ± 0.7 0.4 ± 0.1 3.4 ± 0.4 3.0 ± 0.4 0.5 ± 0.1
 Coarse root 1.5 ± 0.7 0.10 ± 0.04 0.8 ± 0.4 0.9 ± 0.4 0.1 ± 0.05
 Total 28.0 17 18.2 9.7 2.6
Above-ground herb 16.6 ± 4.0 1.4 ± 0.3 5.0 ± 1.2 10.5 ± 2.4 1.0 ± 0.2
Wood and miscellaneous 15.6 ± 0.9 l.0 ± 0.1 7.8 ± 0.4 4.5 ± 0.3 1.2 ± 0.1
litter
Stand fine root 46.0 ± 7.7 3.3 ± 0.5 24.0 ± 4.0 14.0 ± 2.4 4.9 ± 0.8
Total vegetation 164.4 11.1 89.2 61.6 15.2
Site 2
Trees
 Foliage 10.6 ± 0.7 0.6 ± 0.03 6.0 ± 0.4 2.6 ± 0.2 1.2 ± 0.1
 Branch 4.7 ± 0.9 0.3 ± 0.1 3.0 ± 0.6 3.0 ± 0.6 0.4 ± 0.1
 Bole 3.2 ± 0.6 0.2 ± 0.04 1.4 ± 0.2 1.0 ± 0.2 0.4 ± 0.1
 Coarse root 1.6 ± 0.2 0.1 ± 0.0 l 1.1 ± 0.2 0.5 ± 0.1 0.1 ± 0.02
 Total 20.1 1.2 11.5 7.1 2.1
(continued)
5.1  Carbon Stock and Biomass Accumulation Pattern 123

Table 5.7 (continued)
Nutrient net uptake (kg ha−1 year−1 ± 1SE)
Components N P Ca K Na
Shrubs
 Foliage 22.1 ± 1.4 1.4 ± 0.1 15.0 ± 1.0 6.4 ± 0.4 2.3 ± 0.1
 Stem 5.8 ± 0.7 0.4 ± 0.1 3.5 ± 0.5 3.0 ± 0.5 0.5 ± 0.1
 Coarse root 1.6 ± 0.2 0.1 ± 0.02 0.8 ± 0.1 0.9 ± 0.1 0.1 ± 0.02
 Total 29.5 1.9 19.3 10.3 2.9
Above-ground herb 14.3 ± 3.7 1.2 ± 0.3 4.2 ± 1.1 9.0 ± 2.3 0.9 ± 0.2
Wood and miscellaneous 20.8 ± 3.1 1.4 ± 0.2 11 ± 1.6 6.1 ± 0.9 1.6 ± 0.2
litter
Stand fine root 48.1 ± 10.3 3.5 ± 0.7 25.5 ± 5.0 14.9 ± 3.0 5.0 ± 1.1
Total vegetation 132.8 9.2 71.5 47.4 12.5
Site 3
Trees
 Foliage 10.4 ± 0.7 0.6 ± 0.03 6.2 ± 0.1 2.5 ± 0.2 1.1 ± 0.01
 Branch 7.9 ± 0.9 0.5 ± 0.1 5.0 ± 0.1 4.0 ± 0.6 0.6 ± 0.1
 Bole 6.0 ± 0.6 0.4 ± 0.04 2.5 ± 0.3 1.9 ± 0.2 0.8 ± 0.1
 Coarse root 3.0 ± 0.2 0.2 ± 0.02 1.9 ± 0.2 0.9 ± 0.1 0.2 ± 0.02
 Total 27.3 1.7 15.6 9.3 2.7
Shrubs
 Foliage 16.6 ± 1.1 1.0 ± 0.03 11.0 ± 0.02 4.7 ± 0.03 2.0 ± 0.01
 Stem 3.5 ± 0.5 0.2 ± 0.03 2.0 ± 0.3 2.0 ± 0.3 0.3 ± 0.04
 Coarse root 0.9 ± 0.1 0.1 ± 0.01 0.4 ± 0.1 0.5 ± 0.1 0.1 ± 0.01
 Total 21.0 1.3 13.4 7.2 2.4
Above-ground herb 13.0 ± 3.0 1.0 ± 0.3 4.0 ± 0.9 7.9 ± 2.0 0.8 ± 0.2
Wood and miscellaneous 12.2 ± 0.9 0.8 ± 0.1 6.1 ± 0.5 3.6 ± 0.3 0.9 ± 0.1
litter
Stand fine root 26.5 ± 6.0 1.9 ± 0.5 13.9 ± 3.5 8.3 ± 2.0 2.8 ± 0.7
Total vegetation 100.0 6.7 53.0 36.3 9.6
Source: Singh and Singh (1991b)

utilized by the foliage in organic matter synthesis are redistributed among different
components during the assimilate transfers. As reported by Singh and Singh (1991b),
above-ground nutrient storage was 82% (N), 83% (P), 76% (Ca), 85% (K) and 79%
(Na) of that stored in the vegetation and 21% (N), 6% (P), 4% (Ca), 36% (K) and 25%
(Na) of that stored in the stand (including soil). The total net vegetation uptake (kg
ha−1 year−1) was 143 N, 10 P, 78 Ca, 52 K and 12 Na. 52% N, 50% P and K, 51% Ca
and 42% Na of the net uptake by the above-ground vegetation was returned to the
forest floor in litterfall (Table 5.8). Annually, 56 kg N, 4 kg P, 33 kg Ca, 19 kg K and
4 kg Na ha−1 year−1 were recycled to the soil through litter turnover. Of the total nutri-
ent released to the soil, the contribution of fine roots was 42% N, 43% P, 39% Ca,
41% K and 51% Na. This emphasizes the importance of fine roots in nutrient cycling
of TDFs. Singh and Singh (1991b) observed that of the total plant uptake, 20% N and
P, 22% Ca and 25% K and Na were retained by the standing vegetation over the com-
plete annual cycle. This budgeting indicates that the forest site studied by Singh and
Singh (1991b) was an aggrading system, at least in the short term.
124 5  Productivity and Nutrient Cycling

Fig. 5.2  Compartment model showing the distribution and cycling of N and P in the three sites of
the tropical dry forest. Internal recycling is represented by the broken line. Values in “tanks” rep-
resent average nutrient content (kg ha−1 ± 1 SE). The foliage compartment also includes above-­
ground standing crop of herbs. The root compartment includes coarse roots as well as fine roots.
Net annual fluxes between the compartments are given on arrows (kg ha−1 year−1 ± 1 SE) (Source:
Singh and Singh 1991b)

Singh and Singh (1993a, b) analysed the biomass, net production and nutrient
cycling data to elucidate the significance of short-lived forest components for the
productivity as well as nutrient cycling processes of the TDF ecosystems. Their
findings reported that the storage of dry matter in the short-lived components of the
forest, viz. foliage, herbs and fine roots, accounted only 7%, 0.6% and 4%, respec-
tively, as compared to the very high figures observed for bole, branch as well as
coarse roots, i.e. 40%, 31% and 13%, respectively (Table 5.9). The short-lived com-
ponents registered for 62% of the total net forest production (i.e. 15.2 t ha−1 year−1,
30% by foliage, 26% by fine roots and 6% by herbaceous shoots), as compared to
38% contribution by the long-lived components. Table 5.10 shows that the nutrient
storage in the short-lived forest components was also very low as compared to their
contribution in the total uptake of nutrients. The nutrient storage in the foliage com-
ponent accounted 18% N, 19% P, 8% Ca, 16% K and 6% Na, while the percentage
5.1  Carbon Stock and Biomass Accumulation Pattern 125

Fig. 5.3  Compartment model showing the distribution and cycling of Ca, K and Na in the three
sites of the tropical dry forest. Internal recycling is represented by the broken line. Values in
“tanks” represent average nutrient content (kg ha−1 ± 1 SE). The foliage compartment also includes
above-ground standing crop of herbs. The root compartment includes coarse roots as well as fine
roots. Net annual fluxes between the compartments are given on arrows (kg ha−1 year−1 ± 1 SE)
(Source: Singh and Singh 1991b)

nutrient uptake by foliage registered 25%, 22%, 30%, 18% and 28% for N, P, Ca, K
and Na, respectively. The storage of different nutrients observed in above-ground
herbaceous material accounted between 0.6% and 1.3%, and their uptake was
6–19%, i.e. many times greater compared to the storage. The fine root contribution
to total amount of nutrient uptake was also registered higher (26–34%) as observed
for the storage (4.0–6.0%). With respect to the total nutrient return, the short-lived
components contributed in the range of 31–46% through the foliage, 7–24% by
herbaceous species and 33–45% through the fine roots. According to Singh and
Singh (1993a, b), the fine root contribution was 37% for the dry matter deposition
and was 33–45% for the nutrient return as compared to the leaf litter whose contri-
bution accounted 38% for dry matter deposition and 31–46% for the nutrient return.
The study by Singh and Singh (1993a, b) revealed that quite opposite to their struc-
tural value (i.e. relative storage of biomass), the short-lived components present in
126 5  Productivity and Nutrient Cycling

Table 5.8  Mean concentrations and contents of nutrients in litterfall and nutrient return through
mortality of fine roots
Nutrients
N P Ca K Na
Concentrations (%, mean ± SE)
 Leaf litterfall 1.12 ± 0.16 0.07 ± 0.01 0.71 ± 0.08 0.30 ± 0.07 0.08 ± 0.01
 Wood litterfall 0.89 ± 0.17 0.06 ± 0.01 0.45 ± 0.09 0.26 ± 0.02 0.07 ± 0.02
Contents (kg ha−1 year−1, mean ± SE)
Site 1
 Leaf litterfall 44.8 ± 4.03 2.67 ± 0.24 28.27 ± 2.5 11.9 ± l.07 3.13 ± 0.28
 Wood litterfall 15.57 ± 0.87 0.99 ± 0.06 7.82 ± 0.44 4.53 ± 0.25 1.21 ± 0.07
 Total litterfall 60.37 3.66 36.09 16.4 4.34
 Herbaceous litter 16.63 ± 4.0 1.42 ± 0.3 5.0 ± 1.2 10.5 ± 2.4 1.01 ± 0.24
 Fine root mortality 46 ± 7.7 3.3 ± 0.5 24.0 ± 4.0 14.0 ± 2.4 4.9 ± 0.8
 Total return 123.0 8.38 65.09 40.90 10.25
Site 2
 Leaf litterfall 48.83 ± 4.26 2.92 ± 0.25 30.8 ± 2.68 12.98 ± 1.13 3.4 ± 0.29
 Wood litterfall 20.82 ± 3.1 1.34 ± 0.19 10.50 ± 1.56 6.08 ± 0.91 1.62 ± 0.24
 Total litterfall 69.65 4.26 41.32 19.06 5.03
 Herbaceous litter 14.30 ± 3.70 1.20 ± 0.30 4.20 ± 1.10 9.01 ± 2.30 0.87 ± 0.23
 Fine root mortality 48.0 ± 10.3 3.50 ± 0.70 25.0 ± 5.0 14.9 ± 3.0 5.0 ± l.10
 Total return 131.95 8.96 70.52 42.97 10.9
Site 3
 Leaf litterfall 39.42 ± 2.13 2.35 ± 0.13 24.88 ± 1.34 10.48 ± 0.57 2.75 ± 0.15
 Wood litterfall 12.19 ± 0.98 0.78 ± 0.06 6.12 ± 0.4 9 3.55 ± 0.28 0.94 ± 0.08
 Total litterfall 51.61 3.13 31.0 14.03 3.69
 Herbaceous litter 13.0 ± 3.0 1.0 ± 0.27 4.0 ± 0.97 7.9 ± 2.0 0.77 ± 0.19
 Fine root mortality 26.54 ± 6.0 1.9 ± 0.47 13.9 ± 3.45 8.3 ± 0.2 2.83 ± 0.7
 Total return 91.15 6.03 48.9 30.23 7.29
Source: Singh and Singh (1991b)

the TDF ecosystem, viz. tree foliage, fine roots and herbs, play an important and
dominant role in the ecosystem functioning (i.e. relative contribution for the nutri-
ent cycling) of the tropical dry deciduous forest.
The study of Singh and Singh (1991a, b, 1993a, b) further revealed that the TDF is
characterized by three levels of superimposed mineral cycling, viz. internal, short-
term through short-lived components and long-term through long-lived components.

5.1.2  C
 arbon Stock and Accumulation in Woody Species
of TDF in India

Chaturvedi et  al. (2011b) observed a large variation in the above-ground carbon
density as well as accumulation in the stem biomass, among the five sites studied by
them (Fig. 5.4). The density of carbon reported by Chaturvedi et al. (2011b) at the
5.1  Carbon Stock and Biomass Accumulation Pattern 127

Table 5.9  Structure and dry matter of different components of a tropical dry forest in Uttar
Pradesh, India. Values are means for the three sites and two subsequent years. Percentage values
refer to subtotals
Mean ± SE Percentage (%)
Tree density (stems ha−1) 489 ± 80
Shrub density (stems ha−1) 4764 ± 51
Tree basal cover (m2 ha−1) 6.7 ± 1.9
Shrub basal cover (m2 ha−1) 5.1 ± 1.4
Total basal cover (m2 ha−1) 11.8 ± 1.7
Biomass (t ha−1)
 Foliage 4.7 ± 0.5 7.0
 Branches 20.6 ± 3.6 30.7
 Bole 26.9 ± 4.5 40.1
 Coarse roots 8.6 ± 1.2 12.8
 Fine roots 3.1 ± 0.2 4.6
 Above-ground herbs 0.4 ± 0.1 0.6
 Litter 2.8 ± 0.02 4.2
 Total 67.1
Net production (t ha−1 year−1)
 Foliage 4.6 ± 0.5 30.3
 Branches 3.4 ± 0.5 22.4
 Bole 1.8 ± 0.3 11.8
 Coarse roots 0.6 ± 0.1 3.9
 Fine roots 3.9 ± 0.8 25.7
 Above-ground herbs 0.9 ± 0.2 5.9
 Total 15.2
Return (t ha−1 year−1)
 Leaf litter 4.0 ± 0.2 37.7
 Woody litter 1.8 ± 0.2 17.0
 Above-ground herbs 0.9 ± 0.2 8.5
 Fine root mortality 3.9 ± 0.8 36.8
 Total 10.6
Retention in vegetation compartment (t ha−1 4.6
year−1)
Source: Singh and Singh (1993a, b)

wettest and comparatively most protected site, Hathinala was higher than nine times
the value of minimum carbon density estimated for the driest and comparatively
most disturbed site, Kotwa. Their study therefore indicated a significant patchiness
in the spatial distribution of carbon stocks in the TDF. Evidently, the soil fertility
conditions of the study sites were reflected in the status of above-ground carbon
density. The live carbon accumulation varied significantly among the species as well
as sites as was observed for the above-ground carbon stock. The driest site exhibited
lowest accumulation of live carbon in stems (0.05 t-C ha−1 year−1), and among the
sites, the wettest, Hathinala site, registered the highest (5.3  t-C ha−1 year−1), the
Table 5.10  Contribution of short-lived and long-lived components in storage, net uptake and return of nutrients in the tropical dry forest. Values are averages for three
sites and for two consecutive years
(% of total ± 1 S.E.)
Vegetation Storage Annual uptake Annual return
components N P Ca K Na N P Ca K Na N P Ca K Na
Foliage 18.3 ± 2.1 18.6 ± 2.0 7.9 ± 1.1 16.4 ± 1.7 5.6 ± 1.3 25.2 ± 0.8 22.2 ± 0.7 30.1 ± 1.2 18.2 ± 0.8 27.8 ± 1.9 38.5 ± 2.1 33.9 ± 2.2 45.5 ± 2.4 30.9 ± 1.7 32.4 ± 2.3
Branch 25.8 ± 4.9 26.3 ± 5.2 33.6 ± 5.9 32.1 ± 6.5 24.8 ± 4.1 20.8 ± 0.9 20.0 ± 0.9 21.7 ± 1.3 22.5 ± 2.3 17.5 ± 0.8 14.0 ± 1.0 13.3 ± 0.9 13.2 ± 0.9 12.3 ± 0.9 13.3 ± 0.9
Bole 31.2 ± 1.5 32.2 ± 1.5 31.5 ± 3.7 30.2 ± 3.7 43.8 ± 3.9 8.9 ± 1.2 8.9 ± 1.5 8.2 ± 0.8 10.7 ± 0.9 10.3 ± 2.5
Coarse root 12.3 ± 0.6 10.4 ± 0.4 16.4 ± 1.0 10.9 ± 0.2 13.2 ± 0.3 3.6 ± 0.6 3.3 ± 0.7 4.8 ± 0.8 4.1 ± 0.6 3.2 ± 0.5
Above- 0.9 ± 0.1 1.2 ± 0.1 0.6 ± 0.1 1.3 ± 1.1 0.6 ± 0.1 10.9 ± 0.9 13.3 ± 0.7 6.0 ± 0.6 18.6 ± 1.4 7.1 ± 0.6 12.6 ± 1.0 15.7 ± 1.1 7.0 ± 0.7 23.8 ± 1.7 9.3 ± 0.8
ground
herbs
Litter 5.8 ± 0.7 4.9 ± 0.6 3.9 ± 0.6 5.2 ± 0.5 4.3 ± 0.7
Fine root 5.8 ± 1.2 6.4 ± 1.4 6.2 ± 1.5 4.2 ± 0.8 7.6 ± 1.9 30.5 ± 3.0 32.2 ± 3.0 29.7 ± 3.0 25.8 ± 2.9 34.1 ± 3.4 34.9 ± 2.6 37.0 ± 2.6 34.3 ± 2.6 32.9 ± 2.3 45.0 ± 2.7
Total (kg 566.7 ± 98.6 37.6 ± 6.4 278.7 ± 53.3 255.7 ± 44.0 46.3 ± 8.8 131.9 ± 18.6 9.0 ± 1.3 71.1 ± 10.4 48.8 ± 7.3 12.6 ± 1.8 115.0 ± 12.4 7.8 ± 0.9 61.5 ± 6.5 38.2 ± 3.9 9.5 ± 1.1
ha−1)
Source: Singh and Singh (1993a, b)
5.1  Carbon Stock and Biomass Accumulation Pattern 129

Fig. 5.4  Carbon density in stem biomass at the five sites. HN Hathinala, GG Gaighat, HK
Harnakachar, RT Ranitali, KT Kotwa. Narrow bars represent ±1S.E. Bars tagged with different
letters are significantly different (P < 0.05) from each other (Source: Chaturvedi et al. 2011b)

highest value therefore was 106 times greater than the lowest value. Carbon accu-
mulation rates in the TDF sites of Chaturvedi et al. (2011b) were significantly asso-
ciated with soil moisture (R = 0.73, P < 0.001, n = 45), total soil organic carbon
(R = 0.39, P < 0.01, n = 45) and total soil nitrogen (R = 0.50, P < 0.001, n = 45). The
authors demonstrated a significant influence of soil moisture availability on the
growth of TDF trees. Carbon accumulation when analysed per unit stem basal area
(BA) exhibited that the species which are not common (very low number of indi-
viduals) at a given site can potentially contribute more than their fair contribution to
the total accumulation of carbon at the forest site. Further, Chaturvedi et al. (2011b)
reported that the contribution of leguminous species regarding the above-ground
carbon accumulation was greater at the comparatively drier sites as compared to
other sites. Similar to their findings, Macedo et al. (2008) also reported that nitro-
gen-fixing leguminous trees contribute more effectively in sequestering CO2 com-
pared to the nonleguminous species.
Chaturvedi et al. (2011b) also reported greater value of soil organic carbon, total
nitrogen and total phosphorus at the comparatively wetter site which also accounted
greater clay content (Fig. 5.5). Their studies have also observed positive relation-
ship of soil organic carbon with clay content in a variety of soils and vegetation
types (see Jobbágy and Jackson 2000 for references). Similar with the distribution
of carbon in vegetation, the quantity of carbon stored in the soil of TDF also exhib-
ited marked patchiness; the sites which accounted high density of carbon in the
vegetation also registered high density of carbon in the soil (Chaturvedi et  al.
2011b).
130 5  Productivity and Nutrient Cycling

Fig. 5.5  Organic C, total N and total P in t ha−1 stored in soil to a depth of 30 cm at the five sites.
Narrow bars represent ±1S.E. HN Hathinala, GG Gaighat, HK Harnakachar, RT Ranitali, KT
Kotwa (Source: Chaturvedi et al. 2011b)

5.2  Nutrient Cycling and Nutrient Conservation Strategies

5.2.1  Effect of Precipitation Regime on N and P Cycles

Such features as low range of mean annual precipitation (MAP), annual ratio of
precipitation to potential evapotranspiration of less than 1.0 and four to seven dry
months in a year (Dirzo et al. 2011) contribute in shaping the control of biogeo-
chemical cycles (Campo et al. 2001; Gei and Powers 2014; Verduzco et al. 2015) in
the TDF. However, there is a lack of studies on the effect of rainfall variability on
biogeochemical processes in the TDFs (Campo 2016), limiting our understanding
of the potential effects of predicted increases in drought frequency and duration and
the long-term reduction in mean annual precipitation (Meir and Pennington 2011).
Although the positive effect of lower rainfall in wet-to-mesic climates may be appli-
cable along a wide range of tropical forests (Schuur and Matson 2001; Schuur 2003;
Luyssaert et  al. 2007; Posada and Schuur 2011), it is not clear whether rainfall
decrease will have the same effect in TDFs, where water availability even in the
highest rainfall areas is relatively low (Vicente-Serrano et al. 2013).
TDFs account for a large proportion of global terrestrial cycling and storage of C
(Pan et al. 2011). However, as per the predictions of the Earth system models, the C
fluxes in tropical forests can become highly vulnerable with reduced MAP during
the twenty-first century (Neelin et al. 2006; IPCC 2013). Experiments on nutrient
5.2  Nutrient Cycling and Nutrient Conservation Strategies 131

manipulation have indicated that C cycling in tropical forests is limited by N and/or


P (Elser et  al. 2007; LeBauer and Treseder 2008; Bejarano-Castillo et  al. 2015;
Powers et al. 2015) and it is expected that such limitations will increase in the future
(Vitousek et al. 2010; Wieder et al. 2015). Nevertheless, the nature and extent of
controls of ecosystem functioning in the tropical forest biome remain poorly under-
stood (Cleveland et al. 2011).
Campo and Merino (2016) found that, while litterfall decreased only slightly
(about 10%) between stands experiencing 1240 mm rainfall year−1 and those expe-
riencing 642  mm rainfall year−1, the decomposition decreased by 56%. Campo
(2016) selected a gradient of 12 mature TDFs located in subhumid, intermediate
and semiarid climates in the Yucatan Peninsula, Mexico, where the forest function
is primarily driven by seasonal drought (Campo and Vázquez Yanes 2004) and ana-
lysed the storage and turnover of N and P in the forest floor and mineral soil and
explored the relationship between these processes and pools and the level of rain-
fall. These authors indicated that with decreasing rainfall, the litterfall decreases
slightly (10%), while nutrient use efficiency increases by 20% for N and by 40% for
P.  In the TDFs dry conditions inhibit uptake of nutrients from soil and influence
mineralization and nutrient release by reducing the decomposition rate (Campo
et  al. 1998; Saynes et  al. 2005; Bejarano et  al. 2014); thus nutrient limitation is
related to the water limitation. TDF function may be limited by P because P is
immobilized by calcium in carbonate-rich soils such as those of the Yucatan
Peninsula (Lugo and Murphy 1986; Read and Lawrence 2003; Krasilnikov et al.
2013). The tendency of decreasing litterfall, only slightly with decreasing MAP
(Table  5.11), as noted by Campo (2016) has been observed in other TDFs also
(Martínez-Yrízar and Sarukhán 1990; Read and Lawrence 2003). The low sensitiv-
ity of litterfall to changes in rainfall does not represent a lack of response of the
above-ground productivity to changes in water availability because plants are able
to adjust to the increasing water shortage (Doughty et al. 2014).
Campo (2016) also found that nutrient use efficiency increased with reduced
MAP; N and P are cycled more efficiently with greater water deficit (N use effi-

Table 5.11  Mean ± SE of amounts of nutrients in litterfall and litter (Source: Campo 2016)
Semiarid Intermediate Subhumid
Litterfall
Dry mass (kg m−2 year−1) 0.61b ± 0.01 0.65a ± 0.03 0.66a ± 0.01
N (g N m−2 year−1) 8.6b ± 0.9 11.5a ± 1.9 11.8a ± 1.0
P (g P m−2 year−1) 0.506b ± 0.111 0.627b ± 0.108 0.917a ± 0.067
N:P ratio 17a ± 1.8 18a ± 2.1 13b ± 0.9
Litter
Dry mass (kg m−2 year−1) 1.09a ± 0.07 0.89b ± 0.08 0.53c ± 0.02
N (g N m−2 year−1) 19.2a ± 1.5 16.6a ± 1.1 9.1b ± 0.5
P (g P m−2 year−1) 0.933a ± 0.078 1.148a ± 0.101 0.744b ± 0.049
N:P ratio 21a ± 1.9 14b ± 2.3 11b ± 0.9
Different letters within a line indicate significant differences (P < 0.05) across sites
132 5  Productivity and Nutrient Cycling

ciency increases from 56 at the subhumid site to 71 at the semiarid site, and P use
efficiency increases from 720 at the subhumid site to 1205 at the semiarid site as
estimated from Table 5.11). Large increases in P use efficiency (an increase of 40%)
than in N use efficiency (an increase of 21%) with decrease in MAP show that the
extent of the dry season and/or intensity of drought may have created P limitation in
these ecosystems (Campo and Vázquez-Yanes 2004). In the TDF of the Yucatan
Peninsula, P may be immobilized by calcium (Gamboa et al. 2010; Cuevas et al.
2013), the condition being more pronounced in drier sites with shallow soils. The
marked decrease in rainfall from subhumid to semiarid climates limits ecosystem
function on the Yucatan Peninsula (Bejarano et al. 2014; Campo and Merino 2016)
and promotes biomass investments belowground (Roa-Fuentes et  al. 2013). The
results of the study of Campo (2016) indicated the prevalence of nutrient immobili-
zation and revealed that “the sub-humid site had neither N nor P limitation, while
ecosystem function at the intermediate site appeared to be regulated by P limitation,
and at the semi-arid site it appeared to be regulated by N limitation”.

5.2.2  Foliar Nutrient Concentration and Temporal Variation

As discussed earlier, in several TDF trees, the water loss due to transpiration is
eliminated by shedding of leaves, enabling the rehydration of physiologically active
stem tissues for supporting the next flushing of leaves during the dry season
(Borchert 1994a). The capacity of the soil for supplying nutrients to plants (Singh
et al. 1989; Raghubanshi 1992; Roy and Singh 1995), and the capacity of roots to
capture soil nutrients are, however, significantly lowest in the dry season each year
because of a substantial low soil water availability (Pandey and Singh 1992a). Lal
et al. (2001a) collected and analysed time-series data on the expansion of leaves and
concentrations of nutrients for the eight TDF tree species exhibiting wide difference
in leaf traits (Table 5.12). The concentrations of N-, P- and K mass per leaf enhanced
with leaf development and thereafter stabilized for a certain period of time which is
consistent with the temporal pattern of the quantity of leaf mass (Figs. 5.6, 5.7, and
5.8). Lal et al. (2001a) reported a subsequent decline in the concentration of nutri-
ents with the onset of leaf senescence leading to the withdrawal of nutrients from
leaves and accumulation in perennial tissues. In majority of species, the Ca-mass
exhibited an increasing trend starting from leaf initiation to the leaf senescence with
some intermediary fluctuations (Fig. 5.7). An increment in Ca-mass and a decline in
the N content were apparently running parallel with the pattern of leaf senescence.
The process of nutrient resorption provides sufficient nutrient and makes the species
less dependent on the current uptake of nutrients and is therefore a most important
mechanism for nutrient conservation, which has important implications both at the
level of population as well as at the ecosystem level. The mechanism of foliar
resorption can account for a significant portion of N and P utilized annually by the
forest (Ryan and Bormann 1982). The study of Lal et  al. (2001a) indicated that
resorbed nutrients, mainly N, P and K, could become a significant source of
Table 5.12  Nutrient concentrations of leaves of the selected tree species at mature and senescent stages of leaves during the growing season (mgg−1 ± 1 S.E.)
Nutrient/ Ficus Syzygium Shorea Adina Buchanania Butea Diospyros Terminalia Mean (±1
stage racemosa cumini robusta cordifolia lanzan monosperma melanoxylon alata SE)
x x x x x x
N: mature x21.0a ± 1.0 20.6a ± 1.0 19.7a ± 1.2 x21.0a ± 1.3 19.9a ± 1.2 25.4b ± 1.3 14.7c ± 0.9 14.7c ± 0.7 19.6 ± 1.3
y y y y y y y
N: 8.9a ± 0.5 10.0b ± 0.7 10.2b ± 0.6 y13.6c ± 0.8 14.4c ± 0.9 16.1c ± 1.1 11.3b ± 0.6 11.0b ± 0.8 11.9 ± 0.9
senescent
x x x x x x x x
P: mature 1.9a ± 0.1 1.1b ± 0.1 1.1b ± 0.1 1.5a ± 0.1 1.3c ± 0.1 1.3c ± 0.1 1.1b ± 0.1 1.3c ± 0.1 1.3 ± 0.1
y y y y y y x y
P: 0.9a ± 0.0 0.4b ± 0.0 0.7c ± 0.0 1.1a ± 0.1 0.7c ± 0.0 0.8c ± 0.1 0.9a ± 0.1 0.9a ± 0.0 0.8 ± 0.1
senescent
x x x x x x x
K: mature x20.0a ± 1.0 12.4b ± 0.6 8.8c ± 0.5 10.6b ± 0.6 11.2b ± 0.6 8.5c ± 0.5 12.1b ± 0.7 8.5c ± 0.5 11.5 ± 1.4
y y y y y y y y
K: 14.5a ± 0.9 8.9b ± 0.5 6.9c ± 0.3 5.2d ± 0.3 7.4b ± 0.5 4.5d ± 0.3 7.6b ± 0.5 5.9d ± 0.4 7.6 ± 1.2
senescent
x x x x x x x
Na: mature x0.4a ± 0.0 0.7b ± 0.0 0.3a ± 0.0 0.2a ± 0.0 0.6b ± 0.0 0.2a ± 0.0 0.2a ± 0.0 0.3a ± 0.0 0.4 ± 0.1
5.2  Nutrient Cycling and Nutrient Conservation Strategies

x x x x x x x x
Na: 0.5a ± 0.0 0.7b ± 0.0 0.3a ± 0.0 0.1a ± 0.0 0.6b ± 0.0 0.2a ± 0.0 0.2a ± 0.0 0.3a ± 0.0 0.4 ± 0.1
senescent
x x x x x x x
Ca: mature x22.9a ± 1.1 14.5b ± 1.0 11.1c ± 0.8 10.9c ± 0.6 17.3b ± 1.0 18.6b ± 1.3 10.0c ± 0.5 23.8a ± 1.2 16.1 ± 2.0
x y x y x x y x
Ca: 27.5a ± 1.9 18.5b ± 1.1 12.0c ± 0.8 13.4c ± 0.9 20.2b ± 1.0 20.6b ± 1.0 12.1c ± 0.8 24.8a ± 1.2 18.6 ± 2.2
senescent
Source: Lal et al. (2001a)
Values a–c within a column with different letters and values x, y within a row with different letters are different from each other at P < 0.05
133
134 5  Productivity and Nutrient Cycling

Fig. 5.6  Temporal variation in N- and P-mass in leaves of eight tree species from a tropical dry
forest site. Vertical bars indicate ±1 SE (Source: Lal et al. 2001a)

nutrients supplied to plants and are supposed to meet a remarkable proportion of the
nutrient utilized by the developing leaves in most of the tree species studied by
them. For example, the retranslocated nutrients support 50–100% of leaf area devel-
opment and 46–80% of leaf mass development in tropical deciduous species
(Table 5.13). This internal cycling of nutrients, thereby supporting the pre-monsoon
leaf initiation as well as expansion, enables the forest trees to take sufficient advan-
tage of the rainy season during which the water and soil nutrients are abundant and
are utilized to support a high forest productivity. According to Killingbeck (1996),
5.2  Nutrient Cycling and Nutrient Conservation Strategies 135

Fig. 5.7  Temporal variation in K-, Na- and Ca-mass in leaves of eight tree species from a tropical
dry forest site. Vertical bars indicate ±1 SE (Source: Lal et al. 2001a)

“Resorption proficiency is a parameter describing the minimum level to which a


nutrient is reduced during senescence, and has also been used to quantify nutrient
resorption”. Greater proficiencies are associated with lower final concentration of
nutrients in the senesced leaves. Resorption proficiency and resorption efficiency
continue to receive widespread attention because of their strong relationship with
the fertility of soil, plant nutritional condition, plant functional groups and the
changes in climatic conditions (Lal et al. 2001a; Wright and Westoby 2003; Kobe
et al. 2005; Rentería et al. 2005; Drenovsky and Richards 2006; Ratnam et al. 2008;
136 5  Productivity and Nutrient Cycling

Fig. 5.8  Temporal variation in leaf area and mass in eight tree species from a tropical dry forest
site. Vertical bars indicate ±1 SE (Source: Lal et al. 2001a)

Yuan and Chen 2009a, b). The complications in methodological and conceptual
issues for the assessment of nutrient resorption efficiency (see Killingbeck 1996;
van Heerwaarden et al. 2003) led to the development of another alternative method
for the estimation of nutrient resorption with a statistical modelling approach (Kobe
et al. 2005) that relates nutrients in green and senesced leaves by the application of
allometric model. After comparison of 297 perennial species, it has been observed
5.2  Nutrient Cycling and Nutrient Conservation Strategies 137

Table 5.13  Percent area and mass of mature leaf supported by N-mass retranslocated in the
previous season from senescing leaves in selected dry tropical species
Retranslocation Percent support
Species Mass (mg leaf−1) Percent Area basis Mass basis
F. racemosa 8.3 66 97 65
S. cumini 10.8 62 100 76
S. robusta 21.4 53 100 80
A. cordifolia 29.2 49 100 48
B. lanzan 9.0 39 74 63
B. monosperma 17.9 52 100 59
D. melanoxylon 5.8 42 60 46
T. alata 10.1 45 50 50
Based on Lal et al. (2001a)

that N and P resorption efficiency in plants decreased with increasing N and P nutri-
ent status in green leaves.
Rentería et al. (2005), studying with the six dry deciduous tree species, reported
that the levels of concentrations of green and senesced-leaf P, but not N, varied in
response to the topography-associated changes in nutrient and availability of water
mainly due to changes in annual precipitation. Study by these authors indicates that
the availability of water determines nutrient resorption greater than the availability
of soil nutrients. Rentería and Jaramillo (2011) observed leaf mass per area (LMA),
concentrations of the leaf N and P and the nutrient resorption efficiency in 21 woody
species growing in a TDF ecosystem in the western Mexico and reported that the (1)
resorption efficiency and proficiency of P increased in low rainfall years, (2) the
costs of resorption of nutrients relative to the acquisition from soil vary between N
and P and (3) conservation of P increases as a result of decreases in rainfall
conditions.
There is lack of studies on the resorption of nutrients in tropical forests. Few
existing studies have reported a negative (Vitousek 1998; Cordell et  al. 2001) or
lack of correlation (Lal et al. 2001b) between the efficiency of resorption and avail-
ability of soil nutrients. Lal et  al. (2001b) found no significant difference in the
average foliar P content between their two study sites exhibiting contrasting soil
fertility in a TDF located in India. According to a study of Rentería et al. (2005) on
a TDF in Mexico, the availability of soil water more than the availability of soil
nutrient controls efficiency of resorption in the Chamela TDF, whereas resorption
proficiency could be interactively determined by both nutrient and water
availability.
The interaction found between the soil fertility and nutrient resorption by leaves
has a significant effect on the quality of litter and, therefore, on the rate of decom-
position and cycling of nutrients (Hobbie 1992). Lal et al. (2001b) investigated leaf
traits, such as leaf habit, specific leaf mass (SLM, dry mass of leaf per unit leaf
area), leaf chemistry (concentration of nutrients and C/N ratios) and resorption
capacity of nutrient for 90 TDF tree species growing on the habitats with contrasting
138 5  Productivity and Nutrient Cycling

soil types (viz. comparatively infertile Ultisol and relatively fertile Inceptisol). The
selected tree species differed in the concentrations of mineral nutrients resulting
into differential efficiency of nutrient uses. The resorption efficiencies ranged from
26% to 83% (mean  =  58.32 ± 1.20%) for nitrogen and from 16% to 80%
(mean = 49.57 ± 1.48%) for phosphorus. Mean proficiency in the study of Lal et al.
(2001b), across all tree species, was 1.05 ± 0.04% for nitrogen and 0.14 ± 0.01% for
phosphorus. These values were similar to the mean value of 0.87 ± 0.04% for nitro-
gen and 0.06 ± 0.01% for phosphorus documented by Killingbeck (1996), suggest-
ing a somewhat lower proficiency value for the dry tropical tree species. Killingbeck
(1996) and Eckstein et al. (1999) reported that the evergreens are more proficient for
reducing phosphorus in their senescing leaves compared to the deciduous species;
however, the study of Lal et al. (2001b) reported no differences in the resorption
proficiencies of nitrogen and phosphorus between the leaf habits (Table  5.14),
although significantly higher pools of carbon, nitrogen and phosphorus were
resorbed during the leaf senescence in deciduous species compared to the evergreen
species, particularly on the Inceptisol site (Table 5.15). Greater efficiency of nutri-
ent resorptions observed in deciduous species may correspond with a shorter foliar
retention time.
The common hypothesis that species growing in the low-nutrient environments are
comparatively more efficient in their nutrient use than the species from high-­nutrient

Table 5.14  Comparison of specific leaf mass (mg cm−2) and leaf chemistry between Ultisol and
Inceptisol sites for deciduous species (site differences) and between evergreen and deciduous
species growing on the Inceptisol (leaf habit differences) site
Ultisol, deciduous Inceptisol, deciduous Inceptisol, evergreen
Leaf traits (n = 19) (n = 35) (n = 33)
SLMm 10.76 ± 0.53a 9.30 ± 0.42bx 9.15 ± 0.53x
SLMs 8.19 ± 0.54a 6.54 ± 0.31bx 7.90 ± 0.52y
Cm concn. (%) 45.43 ± 0.39a 40.54 ± 0.74bx 43.92 ± 0.65y
Cs concn. (%) 43.64 ± 0.44a 39.14 ± 0.74bx 42.58 ± 0.72y
Nm concn. (%) 1.88 ± 0.09a 2.04 ± 0.09ax 1.97 ± 0.12x
Ns concn. (%) 1.23 ± 0.07a 0.98 ± 0.06bx 1.03 ± 0.06x
Pm concn. (%) 0.16 ± 0.03a 0.22 ± 0.02ax 0.23 ± 0.02x
Ps concn. (%) 0.12 ± 0.03a 0.14 ± 0.01ax 0.16 ± 0.02x
C/Nm (%) 24.83 ± 1.27a 21.19 ± 1.01bx 24.69 ± 1.38y
C/Ns (%) 38.57 ± 2.73a 43.99 ± 2.43bx 45.95 ± 2.72x
Source: Lal et al. (2001b)
Values are given as mean ± 1 SE. Values followed by different letters in a row are significantly
different at P < 0.05 (letters a and b indicate differences between columns 1 and 2, and letters x and
y indicate differences in columns 2 and 3)
5.2  Nutrient Cycling and Nutrient Conservation Strategies 139

environments (Hobbie 1992; Demars and Boerner 1997) has not been supported by
the findings of Lal et  al. (2001b). According to their study, the deciduous species
growing on the Inceptisol site showed a significantly higher resorption efficiency for
nitrogen and phosphorus compared to the deciduous species growing on the nutrient-
poor Ultisol site (Table 5.15). Regarding the resorption proficiency, the species grow-
ing on the Inceptisol site exhibited a higher capacity to reduce nitrogen but a similar
capacity also for reducing phosphorus in senescing leaves compared to the species
growing on the Ultisol site (Table 5.14). In contrast, many studies have suggested that
on nutrient-rich sites, a higher mass of leaf nutrients is withdrawn; however, at the
same time, a higher quantity of nutrient is also left behind in senesced leaves as com-
pared to the leaves of plants developing on infertile sites (see Lambers et al. 1998).
According to Lal et al. (2001b), nitrogen and phosphorus resorption efficiencies were
not absolutely related to the status of nutrient in the mature leaves (Table 5.16), and this
observation was also supported by the findings of Aerts (1996) from an investigation of
a temperate data set. A negative association between phosphorus concentration in green
leaves and phosphorus resorption efficiency was observed only for evergreen species
growing on the Inceptisol site (Table 5.15), thus suggesting that leaf habit exerts a sig-
nificant influence on the relationship between phosphorus concentration and phospho-
rus resorption. The efficiencies of nitrogen and phosphorus resorption were positively
associated in the deciduous species (Table 5.16). This relationship suggests that nitrogen
and phosphorus resorption may be determined by similar biochemical metabolic pro-
cesses. Lal et al. (2001b) observed that a substantial quantity of carbon was also resorbed
during leaf senescence, and the carbon resorption efficiency exhibited positive relation-
ship with nitrogen resorption efficiency in all species groups (Table 5.16). Fahey et al.
(1998) reported that greater carbohydrate resorption was linked with higher nitrogen and
phosphorus resorption, and the findings of Lal et al. (2001b) support the significance of
carbon as a vehicle for the mobilization of nutrients, particularly nitrogen.

Table 5.15  Comparison of resorbed nutrient pools and resorption efficiencies between Ultisol and
Inceptisol sites for deciduous species (site differences) and between evergreen and deciduous
species (leaf habit differences) growing on the Inceptisol
Leaf Ultisol, deciduous Inceptisol, deciduous Inceptisol, evergreen
trait (n = 19) (n = 35) (n = 33)
Resorbed nutrient pools (μg·cm−2)
C 1315.71 ± 82.34a 1201.54 ± 85.48ax 640.83 ± 65.54y
N 101.82 ± 8.71a 119.96 ± 5.75ax 91.80 ± 5.40y
P 7.58 ± 0.81a 11.16 ± 1.02bx 7.78 ± 0.58y
Resorption efficiency (%)
C 27.86 ± 1.94a 31.81 ± 1.68ax 17.26 ± 1.72y
N 51.81 ± 3.04a 65.81 ± 1.47bx 54.73 ± 1.68y
P 48.07 ± 2.89a 56.34 ± 2.04bx 42.44 ± 2.30y
Source: Lal et al. (2001b)
Values are given as mean ± 1 SE. Values followed by different letters in a row are significantly
different at P < 0.05 (letters a and b indicate differences between columns 1 and 2, and letters x and
y indicate differences in columns 2 and 3)
140 5  Productivity and Nutrient Cycling

Table 5.16  Correlations between selected leaf chemistry parameters and resorption efficiencies
for tropical dry forest species
Inceptisol, deciduous Inceptisol, evergreen Ultisol, deciduous
Relation (n = 35) (n = 33) (n = 19)
Nm (%) vs. 0.18 ns 0.29 ns 0.24 ns
Nre
Pm (%) vs. −0.25 ns −0.44** −0.33 ns
Pre
Cre vs. Nre 0.34* 0.46*** 0.56**
Cre vs. Pre 0.35* 0.16 ns 0.53**
Nre vs. Pre 0.35* 0.14 ns 0.60***
C/Nm vs. Nre −0.21 ns −0.38* −0.14 ns
C/Nm vs. C/ 0.68*** 0.84*** 0.67***
Ns
Source: Lal et al. (2001b)
ns not significant
*P < 0.05, **P < 0.02, ***P < 0.01

A current general ecological conception states that the species growing in


nutrient-­poor ecosystems exhibit high C/N ratios, i.e. low quality of residue and
high nitrogen use efficiency (Wali et al. 1999), and Eckstein et al. (1999) observed
that the concentration of nutrients in litter was greater on more fertile sites. Lal et al.
(2001b) reported higher carbon concentration in the species developing on the
Ultisol site compared to the species growing on the nutrient-rich Inceptisol site
(Table 5.14). The nitrogen concentration in mature tree leaves did not vary between
the two study sites, while that in the senesced leaves was higher in the Ultisol spe-
cies (Table 5.15); therefore, it may be concluded that the nutrient-poor ecosystems
cannot commonly produce low-quality residue.

5.2.3  Litter Decomposition

Litter decomposition has been considered as an important process which is associ-


ated with many above-ground and belowground processes of the forest ecosystem
and is one of the most important pathways through which the carbon fixed in the
process of photosynthesis is returned back to the atmosphere (Coûteaux et al. 1995;
Schimel 1995). Litter decomposition also releases nutrients for reuse by the plants.
By the decomposition of litter and the degradation of soil organic matter, more than
ten times high carbon dioxide (CO2) is emitted to the atmosphere annually com-
pared to the fossil fuel burning and industrial sources (Prentice et  al. 2001).
Therefore, it has been suggested that the directional changes in the decomposition
of organic matter could have a remarkable impact on the feedback of CO2 flux
between the terrestrial ecosystems and atmosphere (Schlesinger and Andrews 2000).
5.2  Nutrient Cycling and Nutrient Conservation Strategies 141

Primarily, litter decomposition rates are controlled by climate, quality of litter


and the decomposer communities (Lavelle et al. 1993; Coûteaux et al. 1995; Aerts
1997). At global and regional scales, apart from biotic factors, temperature and the
availability of water appear to be the most important determining factors for the lit-
ter decay rate (Liski et al. 2003). Effect of precipitation on litter decomposition has
been discussed in detail in the next section.
Several studies have reported that the combined effect of climate and the quality
of litter is the primary factor controlling litter decomposition (Coûteaux et al. 1995;
Aerts 1997; Moorhead et al. 1999; Gholz et al. 2000; Silver and Miya 2001), while
the C/N ratio is the strongest predictor of the decay rate (Perez-Harguindeguy et al.
2000). Complexes of bacteria, fungi (combining saprobic micro fungi) and the soil
fauna play a significant role in the decomposition of leaf litter (Persiani et al. 2011).
The decomposition of living logged trunk and fallen trees, including the litter are
mostly affected by fungi, particularly basidiomycetes (Saitta et al. 2011). Therefore,
there are three key determining factors influencing litter decomposition, which
operate in the sequence: (1)  environment (e.g. light, temperature, precipitation,
etc.), (2) litter chemistry (e.g. nitrogen content, C/N ratio), and (3) soil fauna and
microbes (Swift et al. 1979; Meentemeyer 1984; Lavelle et al. 1993). Major propor-
tion of the rainfall in TDF occurs for a short period, during the rainy season which
is flanked by a dry hot summer season and a dry cool winter season. Therefore, the
environmental conditions which the leaf litter experience vary from month to month.
Pandey et al. (2014b) monitored the rate of litter decomposition at the interval of
1 month over an annual cycle at five sites (viz. Hathinala, Bokrakhari, Neruiyadamar,
Ranitali, Kotwa) in a tropical dry deciduous forest located in the Vindhyan region,
India. The study reported that the mean litter weight loss exhibits a strong positive
relationship with the soil surface temperature (SST) as well as soil moisture content
(SMC) (Fig.  5.9). Their study also reported lower weight loss during the winter
season and higher during the rainy season (Fig. 5.10), which may represent cold and
dry conditions during winter season inhibiting weight loss and the warm and humid
conditions during the rainy season leading to enhancement in the decomposition
rate. Even under favourable temperature and precipitation conditions throughout the
year, the decomposition rates differ strongly across the tropical forests (Stephan
et al. 2011). During summer season, the high temperature and low moisture may
strongly limit growth as well as the activity of microbial community (Criquet et al.
2004; Fioretto et al. 2005), whereas, during the spring and autumn seasons when the
temperature is milder and conditions are wetter, the metabolic rates of microbes
become higher, temporarily increasing the litter decay rates (Coûteaux et al. 1995).
The availability of moisture in the form of rainfall is considered as the natural key
climatic factor which directly or indirectly determines the structure and function of
the tropical forest ecosystems (Richards and Caldwell 1987; Clark et  al. 2003;
Feeley et al. 2007). The litter weight loss is enhanced by high nitrogen and low C/N
ratio. Therefore, it is very certain that the concentration of nutrients combined with
SMC and SST controls the rate of litter decomposition and, therefore, organic mat-
ter turnover as well as nutrient cycling in the TDF.  In the study of Pandey et  al.
(2014b), the mean carbon content as well as the C/N ratio of the residual litter
142 5  Productivity and Nutrient Cycling

Fig. 5.9  Monthly pattern of weight loss (%) across species and sites, SST (°C) and SMC (%).
Narrow bars represent ±1 SE (Source: Pandey et al. 2014b)

declined, whereas nitrogen and phosphorus contents increased, exhibiting nutrient


immobilization (Fig. 5.11).
Comparisons among sites exhibiting differences in total annual rainfall have
generated information regarding the effect of rainfall on the rate of litter decomposi-
tion in tropical forest ecosystems (Austin and Vitousek 2000; Schuur 2001; Powers
et  al. 2009b), as did the results from experiments on dry-season irrigation and
through fall (Wieder and Wright 1995; Vasconcelos et al. 2007; Wieder et al. 2009).
Studies referred above have shown that higher amount of total rainfall may influ-
ence the rate of litter decomposition both in positive and in negative directions.
Increased leaching of soluble compounds present in the litter combined with the
microbial activity triggered by suitable soil moisture conditions are important fac-
tors responsible for the positive impact of high annual rainfall on the rate of litter
decomposition. However, soil saturation together with low oxygen diffusion under
the conditions of high rainfall may influence the activity of decomposers. Maass
and Burgos (2011) reported that the marked interannual difference in rainfall
resulted in almost fivefold difference between the decomposition rates recorded in
wettest and the driest year across a 50-year period. Moreover, the rainfall in TDF
typically represent as discrete, infrequent and highly unpredictable rainfall events,
together with high variability among years (Murphy and Lugo 1986; Maass and
Burgos 2011).
5.2  Nutrient Cycling and Nutrient Conservation Strategies 143

Fig. 5.10  Decomposition pattern of leaf litter of eight species at five sites: (a) S. robusta; (b) B.
lanzan; (c) D. melanoxylon; (d) L. parviflora; (e) L. coromandelica; (f) T. tomentosa; (g) H. anti-
dysenterica; (h) L. camara. Vertical bars represent ±1 SE (Source: Pandey et al. 2014b)

The impact of rainfall variability on the ecosystem processes has been generally
assessed following two major approaches: (1) through the long-term investigation of
ecosystem processes and regional history of precipitation and (2) through the short-
term experimental manipulations in the soil moisture conditions (Wieder and Wright
1995; Knapp et al. 2002; Weltzin et al. 2003; Sanches et al. 2008; Austin et al. 2009).
Anaya et al. (2012) evaluated the patterns of litter decomposition in the TDF located
in Mexico, in response to differences in rainfall, by investigating a long-term (8-year)
data set collected for litterfall, surface litter and rainfall. Seasonal variations in sur-
face litter carbon mass as observed by Anaya et  al. (2012) (Figs.  5.12 and 5.13)
exhibited the characteristic trend observed in TDFs across the globe (Murphy and
Lugo 1986; Jaramillo et al. 2011). The lower litter mass in the rainy season compared
to the dry season is associated with both a decrease in litterfall and an increase in the
rate of litter decomposition during the rainy season, as earlier reported by Martínez-
144 5  Productivity and Nutrient Cycling

Fig. 5.11  Leaf litter carbon content (%), leaf litter nitrogen content (%), leaf litter phosphorus
content (%) and leaf litter C/N ratio. The values are mean across species and sites. Vertical bars
represent ±1 SE (Source: Pandey et al. 2014b)

Fig. 5.12  Monthly variation of litterfall C (g m−2 month−1) in four long-term plots at the tropical
dry forest in Chamela, Mexico. Data are means of 8 years ±1 SE (Source: Anaya et al. 2012)
5.2  Nutrient Cycling and Nutrient Conservation Strategies 145

Fig. 5.13  Seasonal variation of surface litter C (g C m−2) in four long-term plots at the tropical dry
forest in Chamela, Mexico. Data are means of 8 years for each collection date ±1 SE. Bars repre-
sent the dry (dark grey; May), the rainy (white; October) and the early-dry (light grey; January)
sample collection dates (Source: Anaya et al. 2012)

Yrízar and Sarukhán (1993) and Anaya et al. (2007). It has been reported that during
the wet season, soluble carbon and nitrogen are leached, and favourable soil moisture
conditions promote the microbial activity leading to increase in the decomposition
rate (Roy and Singh 1994; Wieder and Wright 1995; Anaya et al. 2007), while the
litter input to the surface of soil is lowest during the wet season (Martínez-Yrízar and
Sarukhán 1990). The analysis of litter decomposition data by Anaya et al. (2012)
revealed conspicuous variation in average annual surface litter carbon among years
(Fig. 5.14). The frequency of high rainfall events (>10 mm) had a remarkable impact
on the decomposition rates, which supports the hypothesis that a larger frequency of
high rainfall events could have a positive effect on the rates of decomposition in TDF,
at least at a landscape scale; however the relevance of low rainfall events (<5 mm) for
the processes in soil mediated by the activity of microbes has also been observed for
the arid and semiarid regions (Austin et al. 2004). The structure of ecosystem may
also influence the functional response to the quantity of rainfall (Huxman et  al.
2004); the interception of rainfall in canopy and litter layer in forests could diminish
the impact of low rainfall events. Thus, Anaya et al. (2012) concluded that the influ-
ence of water input for litter decomposition in TDF is determined by the rainfall
pulse size: low- (<5 mm) as well as medium (5.1–9.9 mm)-sized rainfall events were
not able to trigger litter decomposition as much as large rainfall pulses (>10 mm),
and there occurred a significant linear relationship between the frequency of large
rainfall events (>10 mm) and the rates of rainy season decomposition (Fig. 5.15).
These studies suggest that changes in the precipitation regime which alter the fre-
quency of the rainfall pulses would affect litter C fluxes in this TDF ecosystem. The
TDF flora, thus, has developed pronounced primary productivity and nutrient conser-
146 5  Productivity and Nutrient Cycling

Fig. 5.14  Annual variation of surface litter C (g C m−2) during an 8-year period in four long-term
plots at the tropical dry forest in Chamela, Mexico. Data are means ±1 SE (n = 24) (Source: Anaya
et al. 2012)

2.0
Decomposition rate (g C m-2 d-1)

Y = 0.0605x - 0.0523
R2 = 0.56
1.6

1.2

0.8
Upper
Middle
0.4 Low
IV

0
10 15 20 25 30
Frequency of rainfall events ≥ 10mm

Fig. 5.15  Regression model between the frequency of rainfall events >10 mm and rainy season
decomposition rates in four long-term plots at the tropical dry forest of Chamela, Mexico. The
regression model was significant at P < 0.01; df = 1, 30 (Source: Anaya et al. 2012)

vation strategies in order to cope up with the nutrient poverty of the substratum, and
limited and fluctuating growing season.
In conclusion, the global variation in biomass among TDFs is mainly explained
by precipitation. In such ecosystems, the long-term decline in precipitation is likely
to result in an elevation in organic layer and mineral soil carbon storage, mostly due
to decreasing decomposition and increasing chemical recalcitrance of organic mat-
5.2  Nutrient Cycling and Nutrient Conservation Strategies 147

ter, as a result of the change in litter composition and alteration in wildfire patterns.
The carbon storage in the soil could increase as the predicted drought period
becomes longer, provided primary productivity is maintained. Study in Indian TDF
has reported a disproportionately larger contribution of short-lived components (i.e.
foliage, herbaceous species and fine roots) to the carbon turnover, in comparison to
their proportion in the total standing crop of the forest. Nutrient return through fine
roots was equally important as nutrient return from leaf litter. Studies indicated that
nutrients resorbed from senescing leaves could be an important source of nutrients
supplied to plants and cover a significant proportion of the nutrient required by the
developing leaves; the nutrients retranslocated support 50–100% of leaf area devel-
opment and 46–80% of leaf mass development. The TDF is characterized by three
levels of superimposed mineral cycling, viz. internal, short-term through short-lived
components and long-term through long-lived components. Carbon and nutrient
budgeting in the Indian TDF indicated that the forest was an aggrading system.
Litter weight loss in TDF is usually high and is influenced by environmental param-
eters such as surface soil temperature and soil moisture content. The influence of
large rainfall events (>10 mm) for the litter decomposition indicates that variations
in the precipitation conditions would influence the vulnerability of the litter carbon
and nutrient pools to extreme events. There appears to be a considerable scope to
increase the carbon stock and accumulation in TDFs through management implica-
tions such as protection and proper management of the high carbon stock sites and
the old-growth trees from deforestation and fire and increasing the forest stock by
planting species having potential for high carbon accumulation.
Chapter 6
Influence of Biotic Pressure and Land-Use
Changes

The tropical dry forests (TDFs) occur in regions which, historically, have had cli-
matic and soil characteristics suitable for human occupancy and development of
various kinds of infrastructure including agriculture. As a consequence, these for-
ests have suffered from high level of biotic disturbance including gradual and abrupt
land-use changes causing fragmentation and replacement of biotic communities.
The forests have been converted to other ecosystem types, such as savanna, treeless
grassland and cultivated land. These changes have resulted in marked alteration in
species composition, diversity, soil characteristics and ecosystem functioning and
also in the invasion by exotic species. In this chapter we describe the general effects
of biotic disturbance, decrease in the forested area and susceptibility to climate
change, changes in species composition and diversity, fragmentation and edge
effect, recruitment pattern of saplings and juveniles, impact of village ecosystems,
savannization and changes in soil properties with emphasis on soil microbial bio-
mass and microbial processes.

6.1  Biotic Disturbance in Tropical Dry Forests

Since the climatic and edaphic characteristics of TDFs have been attractive for
human settlement, they have historically supported large density of human popula-
tions (Tosi Jr and Voertman 1964; Sánchez-Azofeifa et al. 2005). Dry forests are
considered to be more threatened and comparatively less protected than the moist
and wet forests (Gerhardt 1993; Powers and Tiffin 2010) and have declined in area
substantially during the past decades. Particularly, in central and northern India,
TDFs are threatened by lopping, forest fires, overgrazing and clearing mostly for
cultivation (Jha and Singh 1990), and this has led to conversion of the forest cover
into different vegetation types, such as dry deciduous scrub, savanna and grasslands
which are progressively species-poor (Champion and Seth 1968; Sagar and Singh
2003). Therefore, as a result of such destructions, the TDF is now frequently

© Springer Nature Singapore Pte Ltd. 2017 149


J. S. Singh, R. K. Chaturvedi, Tropical Dry Deciduous Forest: Research Trends
and Emerging Features, https://doi.org/10.1007/978-981-10-7260-4_6
150 6  Influence of Biotic Pressure and Land-Use Changes

observed as a mosaic of variably disturbed secondary vegetation, together with the


patches of comparatively undisturbed primary vegetation (Trejo and Dirzo 2000;
Steininger et al. 2001; Gordon et al. 2004; Gove et al. 2005). These transformed
vegetation types experience marked heterogeneity regarding conditions of solar
radiation, temperature, moisture as well as the rate of nutrient release (Khurana and
Singh 2001a). As repeatedly said, the TDFs rank among the highly endangered ter-
restrial ecosystems (Murphy and Lugo 1986; Lerdau et al. 1991; Gerhardt 1993;
Laurance 1999; Trejo and Dirzo 2000; Li et al. 2006) and, as reported by Janzen
(1988b), the most threatened among all major tropical forest types. It has been
reported that the high rates of the loss and degradation of forests continue in several
tropical countries possessing dry forests (FAO 2001). Miles et al. (2006) have pre-
sented a new global distribution map of TDF depicting the per cent cover of trees at
a resolution of 500 m together with exposure of TDFs to various kinds of threats,
such as climate change, habitat fragmentation, fire, population density of humans
and conversion to cropland. The expansion of protected TDF was estimated from
the analysis of the distribution of protected areas. According to Miles et al. (2006),
the percentage cover of forested area in 2001, relative to the percentage cover delin-
eated by Olson et al. (2000), varies from c.16% in TDFs of South and Southeast
Asia to higher than 40% in Latin America. According to the estimated forest loss,
the TDF of Latin America experienced the highest decline in the percentage cover
of forested area between 1980 and 2000 (12% or 0.22 × 106 km2). For Asia, the cor-
responding value was relatively low, at c. 2%. In Africa, the forest deforestation
rates were comparatively low, with the exception of the deforestation in Madagascar,
where a loss of 18% (0.06 × 106 km2) was documented for this period (Fig. 6.1).
Miles et al. (2006) also suggested that a very high proportion of dry forest areas
in the Americas were at risk of severe impact of climate change (39.8% and 37.0%,
respectively, recorded for North and Central America, as well as for South America).
In other regions, the percentage of TDF at risk of severe change in forest area was
<20% (Fig. 6.2). It may be mentioned that a large decline in precipitation is antici-
pated in many parts of the tropical America. The TDFs, as shown by the investiga-
tion of Miles et al. (2006), occur either as comparatively intact blocks of habitat or
as largely fragmented patches. These authors also observed that the fire occurrence
in TDF varied relatively low between the geographic regions (Fig. 6.3), with per-
centage values ranging from 17.4% (Eurasia) to 26.9% (Africa) of forest regions
affected by fire. Given that the data collected by Miles et al. (2006) was within a
short period of only 3 years, the occurrence of fire appears to be a widely distributed
and significant factor. The observed forest regions varied significantly regarding the
areas at high risk of transformation to agricultural land (Fig. 6.4); forest areas exhib-
iting high suitability for cultivation varied from 5.5% (Southeast Asia and
Australasia) to 20.2% (Eurasia). To summarize, about 97% of the remaining TDF
area is currently experiencing a high level of one or more kinds of the threats con-
sidered by Miles et al. (2006) (excluding deforestation, which was evaluated sepa-
rately). Lamb et  al. (2005) suggested that the damaging consequences after
deforestation and degradation of TDFs include the loss of diversity of biological
6.1  Biotic Disturbance in Tropical Dry Forests 151

Fig. 6.1  Estimated percentage area forested in the year 2001 relative to total area of forest habitat
(top) and estimated decreases in percentage forest area from 1980 to 2000 relative to total area
(bottom) within the administrative boundaries of the protected areas (dark grey), within the 50 km
buffer surrounding the protected areas (striped) and total area (dotted). Abbreviations, with total
habitat area as delineated by Olson et al. (2000) in 106 km2 given in parentheses, are LA DRY Latin
American Dry Forests (1.8), LA MOIST Latin American Moist Forests (9.2), ASIA DRY South and
Southeast Asia Dry Forests (3.7), ASIA MOIST South and Southeast Asia Moist Forests (6.0),
AFRICA MOIST African Moist Forests (3.3), MAD DRY Madagascar Dry Forests (0.3), MAD
MOIST Madagascar Moist Forests (0.3) (After DeFries et al. 2004). Error bars represent the range
of estimates for decrease in forest cover based on correction factors for calibrating AVHRR- and
Landsat-derived estimates (See DeFries et al. 2002) (Source: Miles et al. 2006)
152 6  Influence of Biotic Pressure and Land-Use Changes

Fig. 6.2  Estimated distribution of tropical dry forest at risk of serious climate change (2040–
2069) (Source: Miles et al. 2006)

Fig. 6.3  Estimated percentage of 10 km cells containing tropical dry forest to have experienced
fire, 1998–2000 (Source: Miles et al. 2006)
6.1  Biotic Disturbance in Tropical Dry Forests 153

Fig. 6.4  Estimated percentage of 10  km cells containing tropical dry forest with greater than
marginal suitability for rain-fed crops in the year 2000 (Source: Miles et al. 2006)

Table 6.1  The stand structure of the tropical dry forest along a disturbance gradienta
Bhawani
Variables Hathinala Khatabaran Majhauli Katariya Kota
Mean basal area 8.50 (1.50) 13.78 (2.10) 8.72 (0.66) 6.17 (0.72) 1.30
(m2 ha−1) (0.56)
Mean no. of stems ha−1 419.00 279.33 395.67 215.33 35
(64.01) (50.61) (17.68) (23.13) (10.69)
Mean no. of species ha−1 23.00 (2.89) 22.00 (0.00) 18.33 (0.67) 16.67 (1.73) 3.67
(0.33)
Mean no. of genera ha−1 22.33 (2.33) 22.00 (0.00) 17.67 (0.88) 16.00 (0.58) 3.67
(0.33)
Total no. of unique 6 13 1 1 1
species
Total no. of species per 0.025 0.036 0.019 0.034 0.067
individual
Source: Sagar et al. (2003)
The values in parenthesis are ±1SE
a
154 6  Influence of Biotic Pressure and Land-Use Changes

organisms, decline in watershed protection and reduction of non-timber forest prod-


ucts, leading to the loss of means of survival for the forest-dwelling people.

6.2  T
 ree Species Composition, Dispersion and Diversity:
A Case Study from a Tropical Dry Forest Region of India

Sagar et al. (2003) investigated the impact of different levels of disturbances on the
species composition as well as on dispersal pattern of some selected tree species
occurring at five sites located in a dry tropical environment along a gradient of dis-
turbance. The five forest sites registered a total of 4033 stems (Table 6.1) of 49 spe-
cies having trees ≥30 cm CBH. The number of tree species and total number of
stems per quadrat (10 × 10 m) varied from 1 to 9 species and 1 to 12 individuals,
respectively, with most of the quadrats having 1–3 tree species and 1–4 stems
(Table 6.2), exhibiting a patchy distribution of tree species and their individuals in
the forest at the five study sites.

Table 6.2  Number of quadrats with varying number of adult tree species and individuals at each
of the five tropical dry forest sites arranged according to increasing level of disturbance
No. of species No. of individuals
Sites 1–3 4–6 7–9 1–4 5–8 9–12
Hathinala 194 91 5 167 110 13
Khatabaran 228 44 0 221 51 0
Majhauli 230 65 1 185 106 5
Bhawani Katariya 246 14 0 234 25 1
Kota 75 0 0 75 0 0
Source: Sagar et al. (2003)

Table 6.3  Pattern of tree species diversity in tropical dry forest sites arranged according to
increasing level of disturbance
Variables Hathinala Khatabaran Majhauli Bhawani Katariya Kota
Species richness (Margalef index) 4.204 4.308 3.108 3.245 1.289
Evenness (Whittaker index) 5.656 5.876 3.994 4.126 1.808
Simpson’s diversity (D = 1 – λ) 0.904 0.902 0.857 0.838 0.698
expH′ 13.86 13.695 9.826 9.347 4.047
β-Diversity 1.348 1.364 1.255 1.320 1.909
Source: Sagar et al. (2003)
6.3  Impact of Forest Fragmentation and Edge Effects on Species Diversity 155

Table 6.3 exhibits the pattern of tree species diversity along the gradient of dis-
turbance suggesting that the species richness, evenness and the parameters of
α-diversity (D and eH′) declined as the level of disturbance enhanced. Margalef’s
species richness varied from 1.289 to 4.308, among the study sites and species even-
ness from 1.808 to 5.876. Alpha-diversity indices reported by Sagar et al. (2003)
were also greatest at the least disturbed site and lowest at the highly disturbed site,
values being 0.698–0.904 for D and 4.047–13.860 for eH′. β-Diversity was greatest
for the extensively disturbed Kota site. β-Diversity varied from 1.255 to 1.909. In
the study of Sagar et al. (2003), β-diversity was not found as a very sensitive indica-
tor for the determination of disturbance, as almost similar values were observed for
the four of the five sites, although the number of species recorded per individual site
had a direct positive impact on the β-diversity (Table 6.3).

6.3  I mpact of Forest Fragmentation and Edge Effects


on Species Diversity

Forests are reported to be increasingly fragmented particularly in the tropical


regions across the globe (Laurance 1999; Broadbent et al. 2008), and loss of habitat
and the attendant fragmentation are suggested to be the main contributing factors
for the loss of biodiversity across the globe (Brooks et  al. 2002; Maxwell et  al.
2016). Many studies have reported the relationship between the area of forest patch
and species diversity in the tropical region (Pimm and Raven 2000; Hill and Curran
2001; Wagner and Edwards 2001), between the patch area and structure of com-
munity (Lovejoy et  al. 1983, 1986; Bunge and Fitzpatrick 1993; Colwell and
Coddington 1994; Turner and Corlett 1996; Kemper et al. 1999); these observations
emphasized the concern about the fragmentation of forest and its impact on plant
diversity. Recently, Jha et al. (2005) investigated the impact of declining patch size
of a highly fragmented TDF on the plant diversity by satellite data featuring two
time periods 06-12-1988 (Landsat-TM) and 05-12-1998 (LISS III). These authors
categorized the existing vegetation of the forest area into broad categories, i.e. Sal-­
dominated forest and mixed forest cover types. Among the two forest categories, the
mixed forests were abundant with dry deciduous species like Lagerstroemia parvi-
flora, Butea monosperma, Diospyros melanoxylon, Anogeissus latifolia, Acacia cat-
echu, Lannea coromandelica, Boswellia serrata, Zizyphus sps. etc. Jha et al. (2005)
analysed the forest patches under these two categories of communities separately.
The forest landscape of the study region consisted of 14.70% area of Sal-dominated
forest patches or fragments and 85.30% of mixed forest patches. From Table 6.4, we
observe that except for the Khatabaran site and Hathinala site, Kota, Bhavani Kataria
and Majhauli sites commonly exhibited larger mean patch area (MPA) in the mixed
forest compared to that in the Sal forest. Jha et al. (2005) also observed a related
decline in the mean patch perimeter (MPP) in mixed forests located in Khatabaran
and Sal forests located in Hathinala region. Kota, Bhavani Kataria and Majhauli
156 6  Influence of Biotic Pressure and Land-Use Changes

Table 6.4  The fragmentation indices in Shorea and mixed forest categories analysis
Mean patch size (in
Number of patches ha) Mean patch perimeter (km)
Site 1988 1998 1988 1998 1988 1998
Shorea forests
Kota 0 4 0 1.23 0 0.05
Bhavani Kataria 12 44 2.5 1.34 0.08 0.05
Hathinala 18 34 8.99 0.99 0.17 0.04
Majhauli 15 13 0.94 0.39 0.04 0.02
Khatabaran 4 13 64.29 20.76 0.67 0.23
Mixed forests
Kota 5 13 11.66 1.23 0.23 0.16
Bhavani Kataria 8 25 64.29 10.9 0.58 0.16
Hathinala 12 5 18.56 70.42 0.30 0.45
Majhauli 17 1 0.94 0.39 0.04 2.93
Khatabaran 15 44 8.86 2.65 0.16 0.07
Source: Jha et al. (2005)
NP number of patches, MPA mean patch area, MPP mean patch perimeter

9
Reduction in no. of Species

3
0.5 1.0 1.5 2.0
Loss in patch Size (ha)

Fig. 6.5  Reduction in patch size in relation to loss in species number in various categories of
change (Source: Jha et al. 2005)

sites exhibited a trend of larger perimeter in mixed forest as compared to the Sal
forest. Higher fragmentation observed in Sal-dominated areas could be due to the
selective harvesting of Sal trees, since their wood is of very high timber value. The
mixed forest located in Hathinala site was the only exception where around 2.8
times increment in the mean patch size was detected during the study period. This
increment in mean patch size is probably due to enforcement of relatively more
protection measures by the forest department of the state. In the study of Jha et al.
6.3  Impact of Forest Fragmentation and Edge Effects on Species Diversity 157

Table 6.5  Relationship between reduction in patch size (x) and species loss (y) according to the
linear equation: y = a+bx (where y is dependent and x is independent variable; a = intercept and
b = slope)
Area a b r2 p
Positive reference 0.5 0.533 1 <0.01
Positive change 0.457 5.527 0.99 <0.05
Negative reference 1.467 7.085 0.99 <0.05
Negative change 1.75 7.77 1 <0.01
Source: Jha et al. (2005)

(2005), except for the Majhauli site, all the other study sites inhabiting Sal forest
exhibited an increment in the number of patches. For the sites under the mixed for-
est category, the number of patches recorded in the Hathinala and Majhauli sites
was considerably reduced while an increment in the number of patches was recorded
in the Kota, Bhavani Kataria and Khatabaran sites (Table  6.4). Jha et  al. (2005)
observed that the forest fragments of larger area contain the maximum number of
tree species. The larger fragments also exhibited maximum proportion of rare spe-
cies (Hill and Curran 2003). Figure 6.5 and Table 6.5 exhibit a positive relationship
between the decline in patch area and decline in the number of species. Further, Jha
et al. (2005) reported a general trend of increasing number of patches in the study
region. There was a corresponding decline in MPA and MPP with parallel incre-
ment in the number of patches, which indicated that the phenomenon of fragmenta-
tion is taking place in the forest region.
Forest fragmentation creates forest edges and imposes edge effects (Murcia
1995; Ries et al. 2004). Numerous studies have reported the impact of edges on the
abundance of species and the structural and functional aspects of plant and animal
communities (e.g. Bruna 2003; Asquith and Mejia-Chang 2005; Crowley et  al.
2012). Edge effects have been highly documented for moist or wet tropical forests,
while they have been little studied for dry tropical and subtropical forest regions
(Lima-Ribeiro 2008; Benítez-Malvido et al. 2014). In TDF, the contrasting differ-
ences between the edge and interior could be less marked, and the adaptations
exhibited by the biota to harsh conditions of environment may result into relatively
weak edge effects (Werner and Gradstein 2009; Sampaio and Scariot 2011; Benítez-­
Malvido et al. 2014; see also McWethy et al. 2009). The study related to the aspects
of forest fragmentation is “complicated by edge effects because they are confounded
with fragment area (patch size) effects” (Laurance and Yensen 1991; Malcolm 1994;
Laurance 2008) and “becomes even more complicated when locations in fragments
are simultaneously influenced by more than one edge”, representing “edge additiv-
ity” (Malcolm 1994; Fletcher 2005; Ewers et al. 2007; Fig. 6.6). The edge additivity
can occur, for example, when the fragments are narrow and area located within
them are affected by edges on the opposite sides of the fragment or the situations
when the locations are affected by edges from the multiple sides. The edge effects
in these situations are expected to be even more extreme; the effect of edges in the
interior locations of small fragments is more intense compared to the edge effects at
equal distances in the interior locations of the larger fragments (Malcolm 1994,
158 6  Influence of Biotic Pressure and Land-Use Changes

Fig. 6.6  Schematic representation of additive edge effects. In a the edge effects on opposite sides
of a circular fragment do not overlap, but when a fragment is smaller, edge effects overlap and
become additive (b). As a result, net edge effects in the small fragment are stronger than at equal
distances into the larger fragment (c) (Source: Malcolm et al. 2017)

2001; Fig. 6.6). The situations where the perimeter-to-area ratios are large, intense
edge effects could be especially common (Bender et  al. 1998; Malcolm 2001).
Edges at the study site of Malcolm et al. (2017), which is located in the TDFs of
northwestern Madagascar, exhibited fewer woody stems compared to that in the
interior of the continuous forest, and these effects of forest edge extended surpris-
ingly far interior into the forest (>275 m). Reduced density of stems and species
richness were found in the areas near to the edges in the forest fragments (Benítez-­
Malvido and Martínez-Ramos 2003a). The fragments which were smallest in size
exhibited the lowest stem densities as well as species richness in conformity with
that of “edge additivity” (Benítez-Malvido 1998; Scariot 1999; Benítez-Malvido
and Martínez-Ramos 2003b). In western Borneo, Curran et al. (1999) observed that
the recruitment of seedlings of dipterocarp along a 90,000 ha protected area col-
lapsed mostly due to both natural as well as anthropogenic disturbances, such as
drought and logging. A study investigating the selective logging in dry forests
located in western Madagascar also observed low stem densities as well as species
richness (Ganzhorn et al. 1990). The reduction in stem densities in the drastically
logged forests in Uganda was associated with the edge-related increases in the pre-
dation of seedlings (Chapman and Chapman 1997; Struhsaker 1997). These reduc-
tions have significant implications for future regeneration of forest and the
maintenance of biodiversity in forest (Struhsaker 1997; Scariot 1999; Benítez-­
Malvido and Martínez-Ramos 2003b). A decline in the availability of water may be
an important factor for the decrease in stem densities in the habitats close to edges.
For example, “it is common to encounter cattle and pits dug by local people seeking
plant tubers within continuous and fragmented forests, especially near forest edges”
6.4  Diversity, Recruitment and Future Composition of Juveniles and Saplings 159

(Steffens and Lehman 2016), “and cattle have been shown to reduce seedling sur-
vival through browsing and trampling” (for references, see Griscom et  al. 2009;
Benítez-Malvido et al. 2014). In conclusion, the edge model presented by Malcolm
et al. (2017) predicts that the nearest-edge relationships will be comparatively less
steep as there is decline in fragment size and that, when the distance from the near-
est edge is a controlling factor, edge effects will differ with fragment area.

6.4  D
 iversity, Recruitment and Future Composition
of Juveniles and Saplings

The regeneration and growth of trees depend on the intensity and frequency of graz-
ing, stocking rate of the livestock, the response of various species to damage due to
trampling and browsing, and plant phenophase (Braithwaite and Mayhead 1996;
Hester et al. 1996; O’Connor 1996; Drexhage and Colin 2003). Species composi-
tion of a forest may change following disturbances due the differences in modes of
suppression of the regeneration process, and the disturbances may culminate into
degradation of the forest cover (Leiva and Fernandez-Ales 2003; Quézel and Médail
2003; Plieninger et al. 2004; Dufour-Dror 2007). Repeated harvest of only a few
selected species could influence the dynamics of the left over species leading ulti-
mately to a decline in the forest biodiversity (Zida et al. 2007).
Interactions between livestock browsing, disturbance intensity, understorey
growing conditions and historical land-use patterns impact significantly on the for-
est regeneration (Weisberg and Bugmann 2003; Weisberg et al. 2005; Juan-Baeza
et al. 2015). Livestock activity compacts the soil, reduces the litter cover, decreases
water percolation and increases surface runoff (Lull 1959; Holl 1999), reducing
ultimately the subsurface availability of soil water (Martinez and Zinck 2004;
Zimmerman et al. 2006; Germer et al. 2010). According to Han et al. (2013), intense
browsing by livestock for longer period, together with effects of climate change in
the arid ecosystems located in Xinjiang, China, has resulted in soil and vegetation
degradation, leading ultimately to loss of soil nutrients, and a decrease in net pri-
mary production and species diversity.
Tree juveniles and  saplings are considered comparatively more sensitive to
changes in environmental conditions and are developmentally and physiologically
different from large trees (Jarvis 1995), and also under favourable environments the
mortality rates of juveniles and  saplings are greater compared to mature trees
(Grogan et al. 2011). Therefore, assessing the response of juveniles and saplings to
disturbances and comparing the same with that of their adult conspecifics can help
to understand the vulnerability of a tree species to disturbance (Ouédraogo et al.
2015). Chaturvedi et al. (2012) for juveniles (>30 cm to 1.1 m height (H) and 0.5–
3.2 cm stem diameter (D) at 10 cm above the ground) and Chaturvedi et al. (2017a)
for saplings (>1 m height and ≥3.2 to <9.6 cm stem diameter) analysed the effects
of biotic (browsing and harvesting) and abiotic disturbances (drought and fire), on
160 6  Influence of Biotic Pressure and Land-Use Changes

Fig. 6.7  Stump count of new juvenile trees at regular intervals from September 2005 to June 2007
at the five sites (Source: Chaturvedi et al. 2012)

the diversity, recruitment, mortality and future composition on five sites, viz.
Hathinala, Gaighat, Harnakachar, Ranitali and Kotwa, in TDF of India. The sam-
pling period of Chaturvedi et al. (2012, 2017a) was from 2005 to 2007. In these
TDF sites, grazing/browsing commonly by livestock and harvesting by the local
human population are the most important factors for the mortality of juvenile and
sapling trees, along with the long period of droughts within the annual cycle. The
most heavily disturbed site which accounted for the highest number of mortality for
juveniles/saplings also exhibited the lowest species richness and stem density.
Chaturvedi et  al. (2012, 2017a) observed pronounced influence of season on the
disturbance intensity as the quantity of dead juveniles (Fig.  6.7) and saplings
(Fig.  6.8) differed significantly in different seasons of the sampling periods. The
number of dead juveniles and saplings continued to increase from September to
June, while in the sampling of next September, there was again a decline in the
quantity of dead saplings. Local people generally harvest maximum quantity of
small trees in June, i.e. the period before the onset of monsoon season, for using in
the rainy season as firewood. There is increase in the demand for harvested trees in
June also because they are used for fencing crop fields for protection against domes-
tic and wild animals. Therefore, the maximum quantity of juvenile and sapling mor-
tality due to harvest was observed in the month of June, generally at all study sites.
6.4  Diversity, Recruitment and Future Composition of Juveniles and Saplings 161

Fig. 6.8  Harvest index (a) and browse index (b) of tree saplings analysed at regular intervals from
September 2005 to June 2007 at the five sites. S-1, Kotwa; S-2, Ranitali; S-3, Harnakachar; S-4,
Gaighat; S-5, Hathinala. Data are mean across plots. Narrow bars represent 1 SE (Source:
Chaturvedi et al. 2017a, b)

Forest edges as well as village common lands in the tropical dry region, during
September, are thickly covered with dense layer of herbaceous vegetation. Therefore,
the browsing domestic animals do not need to go far inside the interior of the for-
ested land; this leads to a lower number of juvenile and sapling mortality due to
browsing. From December to March, i.e. in the dry winter season, there is a decline
in the herbaceous vegetation (Pandey and Singh 1992a), and therefore, the browsing
162 6  Influence of Biotic Pressure and Land-Use Changes

domestic animals have to move long distances inside the forest leading to increase
in the number of dead juveniles and saplings due to browsing from December to
March. In June, the forest floor vegetation dries out, and due to lack of herbs and
grasses, people start stall-feeding their livestock, resulting in decline in the number
of dead juveniles and saplings due to browsing by livestock. From July to September,
i.e. during rainy season, the forest floor is generally wet; the fire occurrence becomes
nil. During the dry period of May to mid-June, juveniles and saplings are more vul-
nerable to drought and fire, as during this period the forest floor is very dry and has
considerable amount of fuel load in terms of thick litter layer. In the study of
Chaturvedi et al. (2012, 2017a), juvenile and sapling mortality due to the impact of
drought and fire was only detected in June. In the study of Chaturvedi et al. (2017a),
the newly killed saplings observed in the area of recent fire event were supposed to
have been killed due to fire and not because of drought. The results of Chaturvedi
et al. (2012, 2017a) support the hypothesis that the impact of disturbance will be
greater during the dry months of the year.
Among the sites studied by Chaturvedi et al. (2012, 2017a), those sites (such as
Hathinala and Gaighat), which were comparatively richer in clay content, also
accounted higher SMC (Chaturvedi et al. 2011b), and such sites registered lower
mortality and higher recruitment of juveniles (Fig. 6.9) and saplings (Fig. 6.10) than
sites exhibiting lower clay content and SMC, which supports the hypothesis that the
species regeneration after a disturbance would be greater at moist sites than the rela-

Fig. 6.9  Mortality and recruitment indices of juvenile trees at the five sites. HN Hathinala, GG
Gaighat, HK Harnakachar, RT Ranitali, KT Kotwa. Narrow bars represent ±1SE. Bars tagged with
different letters are significantly different (P < 0.05) from each other (Source: Chaturvedi et al.
2012)
6.4  Diversity, Recruitment and Future Composition of Juveniles and Saplings 163

Fig. 6.10  Annual recruitment index (ARI) and mortality index (AMI) of tree saplings at the five
sites. S-1 Kotwa; S-2 Ranitali, S-3 Harnakachar, S-4 Gaighat, S-5 Hathinala. Narrow bars represent
±1 SE. Bars tagged with different letters are significantly different (P  <  0.05) from each other
(Source: Chaturvedi et al. 2017a)

tively dry sites. In the tropical dry environment, presence of shade by canopy trees
which provides favourable environment with rich soil moisture supports early
establishment (Vieira and Scariot 2006). Several field survey studies have
reported juveniles and saplings to be more abundant on moist sites (e.g. Muick and
Bartolome 1987; Standiford et al. 1991; Swiecki et al. 1993).
Chaturvedi et al. (2012, 2017a) found that the species composition of the juve-
niles and saplings was different than the floristic composition of mature trees at the
same sites. Chaturvedi et  al. (2011b) had earlier reported that on account of the
basal area, the mature tree communities at the highly disturbed Kotwa site was
dominated by the drought tolerant tree, Lannea coromandelica, and at the least
disturbed Hathinala site by the moisture loving tree, Terminalia tomentosa, while,
in the study of Chaturvedi et al. (2012, 2017a), small tree population at Kotwa site
was abundant with the species Nyctanthes arbortristis and at Hathinala site by
Diospyros melanoxylon. These authors argued that “We can consider the present
forest canopy to be established long ago with large canopy trees such as Lannea
coromandelica and Terminalia tomentosa, when the anthropogenic disturbance was
minimum”. Since the intensity of anthropogenic disturbance currently is too high,
there is maximum chance of shift in the composition of species and replacement of
the bigger canopy trees by low-statured trees, for example, Nyctanthes arbortristis
and Diospyros melanoxylon.
164 6  Influence of Biotic Pressure and Land-Use Changes

Table 6.6  Requirements and sources of fodder and firewood for the five villages (×105KJ
year−1 ha−1 cultivated land)
Total Per cent of Firewood extraction
requirement fodder
Village of fodder Sources of fodder requirement Own use Marketed Total
Patwadh 1488.83 Forest 1194.73 80.24 388.79 86.20 475.00
Crop 294.10 19.75
residue
Senduria 1503.16 Forest 1313.56 87.38 193.98 81.40 275.40
Crop 179.10 11.91
residue
Import 10.50 0.69
Chakari 1556.90 Forest 1339.20 86.01 308.20 609.50 917.70
Crop 197.20 12.66
residue
Import 20.50 1.31
Khokha 1193.70 Forest 1077.80 90.2 362.79 – 362.79
Crop 115.90 9.7
residue
Baheradol 2446.10 Forest 2325.10 95.0 425.85 – 425.85
Crop 121.00 5.0
residue
Source: Singh and Singh (1992b)

6.5  F
 odder and Fuel Extraction and Functioning of Village
Ecosystems

According to Singh and Singh (2011), production of fuelwood, small quantity of


timber and the other non-timber forest products (NTFPs) is the major objective of
forest management, and in the TDFs of India, firewood, bamboo, leaves of tendu
(Diospyros melanoxylon), sandal (Santalum album) wood, red sanders, fruits and
seeds, medicinal plants, spices, etc. are the most common forest products. It has
been estimated that the NTFPs from these forests generally contribute 77% of the
net production of forests, with the current net value of revenues generated from
NTFPs amounting US $1016–1348  ha−1, which is significantly greater than the
returns generated from alternate land uses (Mahapatra and Tewari 2005). If the use
of forest products, option, and existence value are together taken into account, the
total current value of non-timber products and services available from a TDF in
India ranges from a minimum value of US $4034 to a maximum value of US
$6662 ha−1 (Chopra 1993).
Singh and Singh (1992b) have assessed energy efficiency of the rain-fed crop-
ping systems in five villages situated within the TDF between 24° 26′–24° 32′ N
latitude and 82° 57′–83° 3′ E longitude and investigated the effect of agricultural
activity on the adjoining natural TDF ecosystems. These authors found that the
agricultural yield was not sufficient to satisfy the food requirement of the local
6.6  Effect of Cultivation on Soil Subsystem 165

human population; as a result 11.5–49.7% of the food grains required by the inhab-
itants were imported from the market. Moreover, the energy requirements estimated
for the five agroecosystems were observed to be largely subsidized by the surround-
ing TDF in terms of fodder and firewood (Table 6.6). Thus, according to the inves-
tigation of Singh and Singh (1992b), the forest ecosystem surrounding the villages
satisfies 80–95% of livestock fodder needs and 81–100% of the energy in the form
of fuelwood. These authors reported that, on an average, 4.1 units of actual energy
are expended to gain 1 unit of agronomic energy. Out of this energy, 3.9 units are
taken from the natural ecosystems. About 38% of the firewood collected from the
forest is sold in markets. This excessive burden on the natural ecosystems results
into illegal felling and lopping of forest trees leading to “ever-increasing concentric
circles of forest destruction around the villages”. This illegal felling and lopping
together with intensive grazing cause savannization.

6.6  Effect of Cultivation on Soil Subsystem

6.6.1  V
 iable Nitrifier Community and Nutrient Availability
in TDF and Derived Cropland Sites

Cultivation has been reported to alter the structural integrity of soil and reduces the
quantity of soil organic matter as well as microbial biomass (Srivastava and Singh
1989; Van Veen and Paul 1981). Jha et al. (1996b) studied dynamics of seasonal
variations in the rate of N-mineralization and the size of viable populations of nitri-
fying bacteria for the TDF site and the nearby cropland site. These authors found
that the cropland soil, which exhibited a lower organic carbon content, accounted a
smaller viable nitrifier community and lower rates of N-mineralization and nitrifica-
tion as compared to the TDF soil. In the cropland soils, the annual average quantity
of ammonia- and nitrite-oxidizing bacteria accounted 48 and 31% of that recorded
for forest soils. Differences observed in the ratios between the physiological groups,
ammonia- and nitrite-oxidizing bacteria in the forest as well as cropland soils might
be because of soil heterogeneity and the influence of surface electric charge on the
structural stability and growth of the microbes (Focht and Verstraete 1977; Prosser
1989). In soil, where most of the particles are negatively charged the attachment of
cells of nitrite-oxidizing bacteria may be less strong compared to that of the
ammonia-­oxidizing bacteria (Prosser 1989). Clay minerals have been reported to
strongly adsorb ammonium ions, leading to preferential colonization of the surfaces
of clay particles by ammonia-oxidizing bacteria which, however, exhibit reduced
growth rate mainly due to localized depletion of ammonium ions at the surface
(Powell and Prosser 1992). The high concentration of ammonia adsorbed to clay
particles can promote its oxidation; however, localized increases in the concentra-
tion of H+ ion at the same anionic site counterbalance it (Prosser 1989). Soil mois-
ture content in the study of Jha et al. (1996b) varied from 8% to 22% in the forest
166 6  Influence of Biotic Pressure and Land-Use Changes

25 a 16 b

Mineral – N (µg g–1)


Soil moisture (%)

20 12
15
8
10
5 4

0 0

c 5 d
6

Mineral – P (µg g–1)


4
NH4 / NO3 ratio

4 3

2
2
+

1
0 0
N-mineralization (µgg–1mo–1)

40 e 3 f
Amm / Nitr ratio

30
2
20
1
10

0 0
S O N D J F A M J J A S S O N D J F A M J J A S
Month Month

Forest Cropland

Fig. 6.11  Variation in (a) soil moisture, (b) mineral-N (NH4+-N + NO3−), (c) NH4+ -N/NO3−-N
ratio, (d) available P (NaHCO3-Pi), (e) N-mineralization rate, and (f) ammonification/nitrification
ratio in the tropical dry forest and the derived cropland sites, during the annual cycle. Line bars
represent ±1SE (Source: Jha et al. 1996b)

soil and ranged from 7% to 20% in the cropland soil (Fig.  6.11a). Soil moisture
contents recorded by Jha et al. (1996b) were highest in the rainy season and lowest
during the summer season. In the forest soil, the concentration of mineral-N (NH4+-
N+NO3−-N) varied from 4.64 to 12.33  μg g−1 and from 6.83 to 15  μg g−l in the
cropland soil (Fig.  6.11b). NH4+/NO3− ratio was always found >1, and for the
8 months of the year, this ratio was higher for the cropland soil compared to the for-
est soil (Fig. 6.11c). The available P concentrations were always higher in forest soil
(Fig.  6.11d). Lower PO4−-P concentrations in the cropland soil probably show
greater P-fixation. The rates of net N-mineralization were consistently higher for the
forest soil compared to the cropland soil, but the differences between study sites
were much lower during the dry months each year (Fig. 6.11e). The ammonifica-
tion/nitrification ratio was consistently >1 for the cropland soil, and for 10 months
of the year, this ratio was greater for cropland soil compared to the forest soil
(Fig. 6.11f). The rates of net N-mineralization and nitrification of both adjoining
sites were maximum during the rainy season and minimum during the summer
6.6  Effect of Cultivation on Soil Subsystem 167

Table 6.7  Seasonal averages Season Forest Cropland


of N-mineralization and
N-mineralization
nitrification rates in soils
from forest and cropland sites Rainy 28.77 ± 2.33 16.64 ± 0.76
(μg g−1 mo−1 ± 1 S.E.) Winter 10.85 ± 0.19 8.75 ± 0.64
Summer 5.37 ± 0.74 5.18 ± 0.57
Nitrification
Rainy 18.20 ± 1.48 4.82 ± 0.48
Winter 4.87 ± 0.32 3.41 ± 0.51
Summer 2.67 ± 0.39 1.74 ± 0.27
Source: Jha et al. (1996b)

Table 6.8 Regression parameters, correlation coefficients and significance levels for the
relationships of N-mineralization (Y, μg g−1 mo−1) with soil moisture and mineral-N contents. The
values for soil moisture are in % and for the others in μg g−1
X-variable a b N r2 p
Forest
Soil moisturea −14.5 2.239 12 0.93 0.0000
NO3− -Nb 3.4 −0.356 12 0.38 0.03
Mineral-Nb 4.1 −0.215 12 0.39 0.03
Cropland
Soil moisturea −5.3 1.247 12 0.93 0.0000
NO4+ -Nc 4.8 −1.298 12 0.34 0.04
Mineral-Na 22.9 −1.247 12 0.33 0.05
Pooled data
Soil moisturea −11.9 1.912 24 0.83 0.0000
Soil moisturec −2.8 2.063 24 0.89 0.0000
NO3− -Nb 3.2 −0.317 24 0.27 0.009
NO3− -Nc 2.9 −0.702 24 0.21 0.03
Mineral-Nc 5.2 −1.321 24 0.35 0.002
Source: Jha et al. (1996b)
a
Y = a + bX
b
Y = exp. (a + bX)
c
Y = aXb (a is in log)

season (Table 6.7); however, the mineralization recorded during the wet season was
highly conspicuous for the forest site compared to the cropland site.
Jha et al. (1996b) reported that the rate of N-mineralization was positively cor-
related to soil moisture, while negatively with the mineral nitrogen on both forest as
well as cropland sites (Table 6.8). The MPN counts observed for the free-living cells
of ammonia- as well as nitrite-oxidizing bacteria exhibited marked variation with
season (Table 6.9) and were recorded highest during the rainy season and lowest
during the summer season, each year. However, the increment in the quantity of
168 6  Influence of Biotic Pressure and Land-Use Changes

Table 6.9  MPN counts MPN counts


(×103 g-l dry soil ±1SE) of Nitrite-oxidizing
ammonia- and nitrite-­ Season Ammonia-oxidizing bacteria bacteria
oxidizing bacteria in different
seasons at different sites. Forest
Values in parentheses indicate Rainy 140.00 ± 17.32 120.00 ± 10.00
annual mean population Winter 0.90 ± 0.10 0.76 ± 0.07
Summer 0.72 ± 0.14 0.73 ± 0.03
(40.50) (47.21)
Cropland
Rainy 66.67 ± 11.10 36.67 ± 1.86
Winter 0.60 ± 0.08 0.58 ± 0.07
Summer 0.53 ± 0.08 0.60 ± 0.08
(22.60) (12.62)
Source: Jha et al. (1996b)

free-living cells observed during the wet season was very high for the forest site
compared to the cropland site, in which the latter exhibited smaller viable commu-
nity sizes as observed for both ammonia- as well as nitrite-oxidizing bacteria; how-
ever, the difference between the two study sites was much higher for the
nitrite-oxidizing bacteria. The study by Jha et  al. (1996b) indicated that in the
nutrient-­poor sites (1) N-mineralization was controlled by the number of nitrifier
cells, and (2) the structure of viable of nitrifying bacterial community, and their
activity were limited by the availability of water. During the wet period, higher
N-mineralization was required to satisfy the greater demand of nitrogen by vigor-
ously growing plant communities. In the cropland sites, decreased organic matter
content and more clay content suppressed the original size of the viable nitrifying
bacterial community, especially the nitrite-oxidizing bacteria, and the rates of
N-mineralization. Therefore, the soil moisture availability, organic matter content
and soil texture potentially determine the size of the viable nitrifying bacterial com-
munity and consequently the supply of nutrients to plants growing in the nutrient-­
poor dry tropical regions.

6.6.2  C
 omparative Analysis of Microbial C, N and P in TDF
and the Alternate Land Uses

It is established that the “soil microbial biomass constitutes a transformation matrix


and acts as a labile reservoir of plant-available nutrients” (Jenkinson and Ladd
1981; Singh et  al. 1989). In general, plants supply carbon to the community of
microbial organisms, and in return microbes provide nutrients necessary for plant
growth by processes of mineralization of the residues of plant and animal, as well
as soil organic matter.
6.6  Effect of Cultivation on Soil Subsystem 169

Table 6.10  Soi1 organic C, N and P in the microbial biomass, biomass C: N: P ratio and N and P
concentrations in the microbial biomass in soils of different land uses

N in a P in a
Microbial C Microbial N Microbial P
Microbial microbial microbial
Organic C (%) Organic N (%) Organic P C:N:P biomass biomass
Land uses (%) ratio (%) (%)
Forest 2.8 2.9 11.2 23:3:1 5.3 2.1
Savanna 3.3 3.6 12.2 22:2:1 4.8 2.3
Cropfield 2.4 3.1 19.2 17:2:1 6.6 3.0
Mine spoil 3.9b 3.5 11.9 28:2:1 4.1 1.8
Source: Srivastava and Singh (1991)
a
Calculated by assuming that dry biomass contains 50% C (Brookes et al. 1984)
b
Calculated by assuming a C:N ratio of 10.6 for mine spoil

According to Misra (1983) and Singh et al. (1985), majority of the savanna and
grasslands found in India have been considered to be derived from TDF. When the
forest vegetation is converted to savanna, it leads to various changes in the vegeta-
tion characteristics as well as properties of soil (Srivastava 1989). Cultivation of a
soil which was previously flourished with natural vegetation results into consider-
able losses of the soil organic matter as well as microbial biomass (Bauer and Black
1981; Gupta and Germida 1988; Srivastava and Singh 1989). Surface mining have
been observed to result in severe disturbances and a soil substrate which is very dif-
ferent from the undisturbed soil, particularly in biological activities (Chaturvedi and
Singh 2017). Some studies are currently available which describe the impact of soil
disturbance created by surface mining on common microbiological activities as
well as associated soil microbial biomass (Stroo and Jenks 1982; Visser et al. 1983;
Srivastava et al. 1989). In a related study, Srivastava and Singh (1991) assessed the
effects of alternate land-use types on the composition of microbial C, N and P of a
TDF soil in four land-use types, viz. forest, savanna, cropland and mine spoil.
According to their observations, organic C, N and P composition in the microbial
biomass for the four land-use types, respectively, ranged from 2.4% to 3.9% C,
2.9% to 3.6% N and 11.2% to 19.2% P (Table 6.10).The mean microbial C:N:P ratio
in the soil was 23:2:l (Table 6.10). Concentration of N in the soil microbial biomass
ranged from 4.1% to 6.6% (Table  6.10). The soil biomass P concentrations esti-
mated by Srivastava and Singh (1991) ranged from 2.1% to 3.0%.
According to Srivastava and Singh (1991), the conversion of forest vegetation into
savanna resulted in a decline in the quantity of plant biomass (Table  6.11).
Savannization was also responsible for the reductions in soil organic C and mineral
nutrients as well as the microbial properties, and therefore decreases in the quantity
of microbial biomass in soil would have no exception. Probably, the decline in soil
nutrients and the microbial C, N and P as a result of savannization reflects the decreas-
ing amount of plant biomass in savanna as compared to that observed in the forest.
Studies have indicated that the cultivation of native forest vegetation and grass-
land leads to remarkable changes in levels of soil organic matter as well as available
170 6  Influence of Biotic Pressure and Land-Use Changes

Table 6.11  Above-ground plant biomass and root biomass (up to 10 cm depth) in forest, savanna,
cropfields and mine spoils. Values are means ±1 SE
Above-ground biomassa (kg ha−1) Belowground biomass (kg ha−1) Total
Forest
Sal forest 3460 ± 420 7980 ± 800 11,440
Mixed forest 3310 ± 720 13,760 ± 1900 17,070
Mean 3390 10,870 14,260
Savanna
Ranitali grazed 660 ± 50 5370 ± 560 6030
Ranitali protected 5350 ± 120 4680 ± 930 10,030
Telburva grazed 1010 ± 100 4710 ± 660 5720
Telburva protected 3650 ± 390 7780 ± 1240 11,430
Mean 2670 5640 8310
Cropfields
 15-year-old 1000 ± 160 630 ± 130 1630
 40-year-old 1250 ± 220 820 ± 200 2070
Mean 1130 730 1860
Mine spoils
 5-year-old 1160 ± 100 2110 ± 330 3270
 12-year-old 1280 ± 90 4020 ± 420 5300
 20-year-old 1580 ± 120 4080 ± 200 5660
Mean 1340 3400 4740
Source: Srivastava and Singh (1991)
a
Above-ground biomasses are litter mass for forests, shoot residue + leaf litter for cropfields and
canopy biomass for savanna and mine spoils

nutrients. It has been therefore observed that soils in the cropped lands have very
low organic matter contents as well as mineral nutrients compared to those of the
virgin habitats (Allison 1973; Bauer and Black 1981; Srivastava and Singh 1989),
and the studies in tropical soils have reported that the continuous cultivation in the
forested land result in a rapid decline in the concentration of organic matter and
nitrogen (e.g. Allen 1985; Aweto 1981; Ayanaba et al. 1976; Brams 1971; Nye and
Greenland 1964) which do not have better chances to recover fully during the pro-
cess of forest succession (Aweto 1981; Raich 1983; Werner 1984). Therefore, it
may be suggested that in natural ecosystems majority of the organic matter pro-
duced by the vegetation through the process of photosynthesis is returned back to
the soil; however, in the cultivated region majority of the organic matter is removed
for the consumption of human and animals. In the impoverished soils the fertility is
therefore quickly degraded by cultivation which also influences the physical struc-
ture of the soil and increases the loss of fine particles by the process of erosion
(Srivastava and Singh 1989).
After savannization, the highest decline in microbial C, N and P reported by
Srivastava and Singh (1991) was 34.8%, 41.5% and 30.8%, respectively (Table 6.12).
However, the decline in microbial C, N and P was observed greater due to cultiva-
tion and mining. Visser et al. (1983) suggested that decline in microbial biomass
6.6  Effect of Cultivation on Soil Subsystem 171

Table 6.12  Per cent decline in values for selected soil properties (μg g−1) and microbial C, N and
P (μg g−1) in derived land uses as compared to those of native forest
Savanna Cropfield Mine spoil
Organic C 44.6 51.2 a

Total N 52.1 52.7 65.8


Total P 39.5 51.4 65.9
Organic N 52.3 52.8 66.0
Organic P 36.5 66.5 56.6
Mineral N 13.2b 35.2 37.1
NaHCO3-Pi 26.8b 56.4 30.9b
Microbial C 34.8 58.9 45.5
Microbial N 41.5 49.2 58.4
Microbial P 30.8 42.3 53.8
Source: Srivastava and Singh (1991)
a
Not calculated due to lack of data for organic C for mine spoil
b
Differences are not significant. All other differences are significant at P < 0.05

700 a b
Mean annual microbial biomass C (m g/g)

600

500

400

300

200

100

0
280 560 840 1120 1400 0 360 720 1080 1440 1800
Root biomass (g/m2) Total (root + shoot) biomass (g/m2)

Fig. 6.12  Relationship between mean annual microbial biomass C and (a) root biomass, and (b)
total (above-ground + root) biomass (Source: Srivastava and Singh 1991)

reported for the mine spoils is mostly due to lack of the top soil layer as well as the
related plant components and also because of decrease in a favourable nutrient level.
Srivastava et al. (1989) found a direct association between the age of spoil and com-
position of microbial C, N and P in soil, and indicated that microbial biomass is the
determining factor for the proper restoration of degraded mine spoils, and also the
172 6  Influence of Biotic Pressure and Land-Use Changes

Fig. 6.13  Territorial map


drawn on the basis of
Discriminant Analysis. A,
forest; B, savanna; C, 8
cropfield; D, mine spoils D
(Source: Srivastava and

Function 2
Singh 1991) A
0 B

−8 C

−12 −8 −4 0 4 8 12
Function 1

amount of microbial biomass can be regarded as functional index for the soil rede-
velopment. On the basis of their observations, Srivastava et al. (1989) suggested that
in the dry tropical environment, there is a direct relationship of microbial biomass
with the plant biomass (Fig. 6.12), and it has also been detected that the microbial
biomass is very sensitive to changes in land-use patterns as it is found to decline
remarkably after conversion of TDF to savanna and cultivated cropland and also due
to the disruption of soil layers after surface mining for coal.
Srivastava and Singh (1991) subjected the soil data to Discriminant Analysis
(SPSS, new version) for understanding at what extent the major land-use changes
can be differentiated among themselves according to the microbial biomass and
various other soil characteristics. Soil characteristics used for the analysis were soil
organic C, organic N, organic P, mineral N, NaHCO3-Pi and microbial C, N and
P.  The results of the analysis identified three distinct Discriminating Functions
which combined together exhibited for 100% variance in the recorded data. The
variations accounted for the first, second and third Discriminating Function were
80.99%, 14.25% and 4.77% variation, respectively. A territorial map was developed
by applying the first two Discriminating Functions as well as the group centroid
averages in Fig. 6.13. The scatter plot suggested that the ecosystems derived from
forests such as savanna, cropland and mine spoil exhibit much differences from the
original forest vegetation in terms of soil characteristics and microbial biomass. The
ecosystems derived from original forests were closer together, although somewhat
distinct from each other. The study by Srivastava and Singh (1991), therefore indi-
cates that the land-use changes modify various soil properties, while at the same
time majority of the soil properties stay interrelated.
According to Srivastava and Singh (1991), the flux of C, N and P through the
activity of the microbial biomass calculated from the upper 10 cm soil layer in the
forest (601, 64 and 26 kg ha−1 year−1, respectively) was almost twice as high as the
C, N and P flux measured in the cultivated soils (262, 34 and 16 kg ha−1 year−1,
6.7  Invasive Alien Plants and Their Effect on TDFs in India 173

Table 6.13  Mean annual soil biomass C, N and P and the annual flux of C, N and P through the
microbial biomass
Nutrients in the biomass (kg ha−1) Annual flux (kg ha−1 year−1)a
C N P C N P
Forest 752 81 32 601 64 26
Savanna 484 47 22 387 37 17
Cropfield 328 43 20 262 34 16
Mine spoil 419 34 16 335 27 13
Source: Srivastava and Singh (1991)
Assuming a biomass turn over time of 1.25 year as suggested by Jenkinson and Ladd (1981) for
a

warmer countries

respectively) (Srivastava and Singh 1991; Table 6.13). As compared to the flux of


nutrients from microbial biomass, the uptake of N and P by plants growing in dry
forest was measured as 59–125 kg N ha−1 year−1 and 4–9 kg P ha−1 year−1 (Singh
and Misra 1979) and that in dry grassland as 29.3 kg N ha−1 year−1 and 4.9 kg P
ha−1 year−1 (Singh et al. 1979). On the basis of these observations, it can be argued
that the micro-organisms are most important source of nutrients for plants growing
in the dry forest, savanna, cropland as well as mine spoil ecosystems.

6.7  Invasive Alien Plants and Their Effect on TDFs in India

“From an ecological perspective, any species introduced to an ecosystem beyond its


home range that establishes, naturalizes and spreads is said to be invasive”
(Williamson 1996). Most invasions have occurred due to the economic activities
(Perrings et al. 2002). Invasive weeds have higher growth rates as well as greater
biomass production than the native species, more competitive ability, high effi-
ciency of reproduction including production of very high amount of seeds, efficient
dispersal mechanism, vegetative reproduction, capacity of rapid establishment and
many other traits which help them in adapting to new habitats (Simberloff et  al.
2005; Sharma et al. 2005a). Several invasive species possess allelopathic potential
and have capacity of high tolerance to different types of abiotic disturbances
(Sharma et al. 2005a; Huang et al. 2009).
It has been estimated that about 18% of the total Indian flora constitutes alien
species, in which around 55% are from American countries, 30% from Asian and
Malaysian region and 15% from European and Central Asian countries (Nayar
1997). Singh et al. (2010) documented several invasive alien plant species in the
flora of the dry tropical region of Uttar Pradesh (UP), which is the fifth largest and
most populous state of India. The invasive plant species in this state are mostly con-
tributed by tropical America, which accounted for 62.5%; next is tropical Africa and
South America accounting 10.5% each and also Europe accounting 3.3% species.
The remaining 13.2% invasive species were found to be collectively contributed by
174 6  Influence of Biotic Pressure and Land-Use Changes

11 different regions. Not surprisingly, majority of the invasive species in the state
have their origin in tropical regions.
Among the documented high impacting invasive plant species, Lantana camara
L. (Lantana) ranks at the top position (Batianoff and Butler 2003) and has been con-
sidered as one of the 100 most harmful invasive alien species across the globe (GISP
2003). This species has observed to spread in almost all terrestrial habitats in the
TDF (Sharma and Raghubanshi 2006). Due to its highly effective allelopathic prop-
erties, Lantana has strong potential to interrupt the regeneration process of several
other species growing close to it either by decreasing their germination capacity,
reducing the early growth rates or selectively enhancing the mortality of the plant
species (Sharma et al. 2005a, 2005b). Sharma and Raghubanshi (2007) identified the
declining tree species population at different levels of lantana invasion, viz. low,
medium and high lantana cover. Figure 6.14 shows the proportion of declining tree
species at low (0–30%), medium (31–60%) and high (61–100%) percentages of lan-
tana cover at several locations in a dry deciduous forest of India as presented by two
consecutive census (2002 and 2003). Collection of fuel wood from the forest leads to
the opening of canopy, which allows lantana to flourish and dominate over the neigh-
bouring species by interfering with their regeneration process. Generally the shading
effect together with allelopathic activity of lantana is the main cause for stunted
growth of tree seedlings. The growth of lantana is faster compared to the native trees;
it captures the resources very efficiently and develops commonly in the intercanopy
as well as at the forest edges. Sharma and Raghubanshi (2006) reported that the lan-
tana acts as an umbrella and prevents light penetration to the floor of the forest which

Fig. 6.14  Proportion of declining species at low, medium and high lantana cover within Vindhyan
dry deciduous forest of India for two consecutive census (2002 and 2003) (Source: Sharma and
Raghubanshi 2007)
6.8  Degradation and Savannization 175

decreases the development of tree seedlings growing under the lantana canopy, ulti-
mately leading to decline in the population of important native trees of the dry forest,
such as Adina cordifolia, Boswellia serrata, Bridelia retusa, Buchanania lanzan,
Cassia fistula, Elaeodendron glaucum, Emblica officinalis, Eriolena quinquelaris,
Hardwickia binata, Miliusa tomentosa, Schrebera swietenioides, which need high to
moderate light for their growth (Troup 1921). Moreover, the lantana bushes are
inflammable which attract fire (Hiremath and Sundaram 2005), and their burning
produces intense heat and smoke causing damage to seeds and mortality of seedlings
in the native vegetation (Moore and Wein 1977).
Sharma and Raghubanshi (2007) reported 12 more tree species exhibiting decline
in population in the dry forest, other than the earlier list of species reported by Sagar
and Singh (2004) for the disturbance-driven decline in population size. The species
listed by Sharma and Raghubanshi (2007) showed a decline in population at differ-
ent stages of lantana invasion. The number of tree species exhibiting decline in
population increased when both factors, i.e. disturbance and lantana invasion, were
considered together; therefore, the disturbance and invasion are supposed to syner-
gistically influence the process of tree regeneration in the TDF.  As reported by
Sharma and Raghubanshi (2007), populations of the species which were reduced to
single individuals are Boswellia serrata, Bridelia retusa, Cassia fistula, Emblica
officinalis and Eriolena quinquelaris at the medium lantana invasion and
Elaeodendron glaucum and Miliusa tomentosa at the high lantana invasion. It can
be, therefore, predicted that the species with single surviving individual is highly
vulnerable as “a local population composed of only a few individuals can undergo
catastrophic decline due to environmental change, genetic problem or simple ran-
dom events when isolated in a limited geographic range” (Cunningham and Saigo
1999).

6.8  Degradation and Savannization

Griscom and Ashton (2011) categorized the degradation processes into (1) acute
single-time events and (2) chronic several repeated events. Also within each of these
degradation categories, the processes of degradation can be further categorized
based on (i) changes in structural properties (e.g. floristics, mechanisms of regen-
eration, successional processes) and (ii) changes in functional attributes of the forest
and the landscape (e.g. fertility of soil, hydrological infiltration, subsurface flow of
water) (Ashton et  al. 2001). According to Griscom and Ashton (2011), the effi-
ciency of land clearance in the TDFs of Central America was generally driven by
two important factors: (i) the soil fertility (the higher the fertility, the higher the
intensity of the clearance of land to maximize productivity) and (ii) the topographi-
cal conditions (the higher the degree of rough topographical relief, the greater inef-
ficient the process of clearing the land). Introduction of exotic grass may cause
remarkable barriers to natural process of succession by preventing seed germination
and out-competing several native species for acquiring light and soil resources
176 6  Influence of Biotic Pressure and Land-Use Changes

(Nepstad et al. 1996; Cabin et al. 2002; Hooper et al. 2002; Griscom et al. 2009).
Several studies have documented that African grasses induce an additional negative,
indirect impact on succession process by perpetuating fire (Cabin et al. 2002; Janzen
2002; Lamb et  al. 2005; Calvo-Alvarado et  al. 2009). Fire-adapted grasses have
been reported to catalyse a positive feedback loop, enhancing the frequency and
intensity of wild fire (D’Antonio and Vitousek 1992), thus maintaining themselves,
and prohibiting change to forest regrowth.
Pandey and Singh (1992b) investigated the amount as well as distribution of
rainfall and grazing impacts on the composition of woody and herbaceous species
dynamics of the savanna vegetation which is considered to be derived from
TDF. Ground layer of the savanna is abundant with C4 grasses, viz. Heteropogon
contortus, Chrysopogon fulvus and Bothriochloa pertusa, in different associations.
The herbaceous plant cover experiences heavy use through uncontrolled range graz-
ing throughout the whole year, and the residual woody elements is subjected to
frequent lopping for getting fodder and fuel wood. In TDFs, the grazing pressure
has been reported uneven and increases remarkably during the rainy season due to
immigration of domestic ungulates (e.g. buffalo, cow, goat) from surrounding vil-
lages (Pandey and Singh 1992b). Grazing had been found to have a profound and
differential impact on species composition as exhibited by the higher per cent dis-
similarity among grazed sites compared to that among the permanently protected
sites (Fig. 6.15). Grazing was also observed to promote plant diversity (Fig. 6.15).
According to Pandey and Singh (1992b), the growth of savanna vegetation started
with the initiation of monsoonal rainfall (June) leading to the build-up of live plant
biomass from July to October during which the moisture was sufficient and tem-
peratures favourable (Fig. 6.16). Pandey and Singh (1992b) observed a secondary
lower peak in the live shoot biomass during winter season which occurred as a result
of renewed seasonal growth in the perennial grasses, such as Heteropogon contor-
tus, Chrysopogon fulvus and Bothriochloa pertusa in response to the winter rain
showers coming from NE monsoon.
Live shoot biomass measured in June was positively correlated with rainfall of
the previous rainy season (Fig. 6.17b), with the previous winter season (Fig. 6.17c)
and with the previous rainy + winter seasons rainfall (Fig. 6.17d). These relation-
ships exhibited that the antecedent rainfall affects the biomass of terminal live shoot.
The positive relationship of live shoot biomass differences according to the rainfall
variability (Fig.  6.17a) indicates that the savanna ecosystem is a pulsed system
responding very fast to the  rainfall events. However, as observed in the grazed
savanna, the intensity of grazing is the most important factor determining the vari-
ability of the biomass of live shoot. The relationship of peak plant canopy biomass
of the permanently protected savanna with the total amount of rainfall received dur-
ing rainy season (Fig.  6.18a) as well as with late rainy season rainfall suggests
(Fig. 6.18b) that the rainfall during the early rainy season initiates the growth, while
the rainfall during the late rainy season sustains the biomass accumulation in
savanna. Evidently the droughts in the mid-growing season are expected to be
highly disastrous. Grazing in the preceding years stimulated peak canopy plant
6.8  Degradation and Savannization 177

12 PROTECTED GRAZED
RANITALI
S –W NUMBER

8 HATHINALA
TELBURVA

0
MARGALEF NUMBER WHITTAKER NUMBER

0
5

1
0
0.10
SIMPSON NUMBER

0.10

0.10
0
1986 1987 1988 1989 1986 1987 1988 1989

Fig. 6.15  Species diversity, equitability, species richness and concentration of dominance at the
time of peak biomass in permanently protected and grazed savanna on the three study sites. The
peak biomass was attained in October in permanently protected savanna and in September in the
grazed savanna. The concentration of dominance was calculated as Ʃ(ni/N)2, where ni is the total
biomass of species i and N the total biomass of all species. Formulae for calculating other param-
eters are given in the text. The age of fencing at Ranitali, Hathinala and Telburva sites, respectively,
was 2, 3 and 7 years in 1986 (the first year of observation) (Source: Pandey and Singh 1992b)
178 6  Influence of Biotic Pressure and Land-Use Changes

750
a 250
b 750
c
PERMANENTLY PROTECTED SAVANNA GRAZED SAVANNA TEMPORARILY
FENCED
LIVE SHOOT BIOMASS SAVANNA
DEAD SHOOT MASS
LITTER MASS
600 200 600

450 150 450


gm–2

gm–2

gm–2
300 100 300

150 50 150

0 0 0
J J A S ONDJ F MA MJ J A S ONDJ F MA N J J J A S ONDJ F MA MJ J A S ONDJ F MA N J J J A S ONDJ F MA MJ
1986 1987 1988 1986 1987 1988 1987 1988
MONTHS MONTHS

Fig. 6.16  Herbage dynamics at the savanna site. a Live shoot biomass, dead shoot mass and litter
mass in permanently protected savanna; b live shoot biomass, dead shoot mass and litter mass in
grazed savanna; c live shoot biomass, dead shoot mass and litter mass in temporarily fenced
savanna. Bars represent ±1 SE (Source: Pandey and Singh 1992b)

b­ iomass in temporarily protected savanna. However, the highest stimulation of bio-


mass yield was observed in lightly and medium-grazed savanna (Table 6.14).
Effect of savannization on carbon balance is discussed later.

6.9  W
 ater-Stable Aggregates and Microbial Biomass in TDF
and Derived Ecosystems

The primary soil particles which cohere among each other more strongly compared
to other surrounding soil particles form a group are called soil aggregates (Kemper
and Rosenau 1986). The capacity of soil aggregates to resist particle breakdown
when wetted or after rainfall determines the permeability of soil with water, air and
plant roots; therefore the water-stable aggregates are responsible for making the soil
layer porous and aerobic. The soil organic matter is related to three types of physical
units: (i) the free primary soil particles (i.e. sand, silt and clay), (ii) macro-­aggregates,
and (iii) micro-aggregates (Tisdall and Oades 1982; Oades 1984).
6.9  Water-Stable Aggregates and Microbial Biomass in TDF and Derived Ecosystems 179

95 24
a b
C.V. OF GREEN BIOMASS (%)

LIVE BIOMASS (g m-2) JUNE


90 20

85 16

80 12

75 8

70 4
125 130 135 140 145 150 400 800 1200 1600
C.V. OF RAINFALL (%) RAINY SEASON RAINFALL (mm)

24 24
c d
LIVE BIOMASS (g m-2) JUNE

LIVE BIOMASS (g m-2) JUNE

20 20

16 16

12 12

8 8

4 4
0 50 100 150 200 400 800 1200 1600
WINTER SEASON RAINFALL (mm) RAINY+WINTER SEASON RAINFALL (mm)

Fig. 6.17  Relationships between (a) intra-annual coefficient of variation (c.v.) of green biomass
(Y, %) and intra-annual coefficient of variation of monthly rainfall (X, %) in permanently protected
savanna, across three sites: Y = − 0.0678 + 0.6208 X, (r2 = 0.7506, P < 0.05). (b) Live shoot bio-
mass in June in permanently protected savanna (Y, g m−2) and rainfall of preceding rainy season
(X, mm) across three sites: Y = − 64.0125 + 11.3593 ln (X) (r2 = − 0.5570, P < 0.01). (c) Live
shoot biomass in June in permanently protected savanna (Y, g m−2) and rainfall of preceding winter
season (X, mm) across three sites: Y = 7.5580 + 1.6783 ln (X), (r2 = 0.3523, P < 0.05). (d) Live
shoot biomass in June in permanently protected savanna (Y, g m−2) and rainfall of preceding rainy
+ winter season (X, mm) across three sites: Y = − 63.5385 + 11.2373 ln (X), (r2 = 0.5576, P < 0.01)
(Source: Pandey and Singh 1992b)

As already discussed, the transformation of forest ecosystem into savanna eco-


system and cropland in the TDF decreases total soil organic matter content
(Srivastava and Singh 1991). The composition of soil macro-aggregates averaged
62% in forest ecosystem, 55% in savanna ecosystem and 34% in the cropland
(S.  Singh 1993, unpubl. Ph.D. thesis, Banaras Hindu Univ.), and therefore, the
structure of soil also deteriorates upon conversion of forests to savanna and
­croplands. Singh and Singh (1995) assessed the dynamics and distribution of micro-
bial biomass related to the water-stable macro- and micro-aggregates in the sur-
180 6  Influence of Biotic Pressure and Land-Use Changes

800 800

PEAK CANOPY BIOMASS (g m-2)


PEAK CANOPY BIOMASS (g m-2)
PERMANENTLY PERMANENTLY
PROTECTED PROTECTED
700
700 SAVANNA SAVANNA
600
600
500
500
400

400
300

300 200
600 800 1000 1200 1400 1600 0 200 400 600 800 1000 1200
RAINY SEASON RAINFALL (mm) LATE RAINY SEASON RAINFALL (mm)

140 180
PEAK CANOPY BIOMASS (g m-2)
GRAZED GRAZED
PEAK CANOPY BIOMASS (g m-2)

SAVANNA 160 SAVANNA


120
140

100 120

100
80 80

60
60
40

40 20
0.65 0.70 0.75 0.80 0.85 0.90 400 600 800 1000 1200 1400 1600
RAINY SEASON GRAZING INTENSITY RAINY SEASON RAINFALL (mm)

Fig. 6.18  Relationships between (a) peak canopy biomass in permanently protected savanna (Y, g
m−2) and rainfall of rainy season (X, mm) across three sites: Y = 2349.5639 + 417.1299 ln (X),
(r2 = 0.74, P < 0.001). (b) Peak canopy biomass in protected savanna (Y, gm−2) and rainfall in later
part (August + September) of rainy season (X, mm) across three sites: Y = − 640.0109 + 186.0740
ln (X), (r2 = 0.7776, P < 0.001). (c) Peak canopy biomass in grazed savanna (Y, g m−2) and grazing
intensity of rainy season (X) across three sites: Y = 327.0909–282.2475 X, (r2 = 0.6331, P < 0.10).
(d) Peak canopy biomass in grazed savanna (Y, g m−2) and rainfall of rainy season (X, mm) across
three sites: Y = 65006.1172 (X 0.9362), (r2 = 0.62, P < 0.01). Peak canopy biomass occurred in
October in all years in permanently protected and in September in grazed savanna (Source: Pandey
and Singh 1992b)

rounding forest, savanna and cropland soils. Their observations indicated that there
was considerable site-wise variability in the quantity of microbial C, N and P linked
with the soil macro- and micro-aggregates (Tables 6.15, 6.16, and 6.17). The com-
position of soil organic carbon immobilized into the microbial biomass (i.e. the ratio
of biomass C-to-soil organic C) in both soil aggregate classes also exhibited varia-
tions across the study sites (Table  6.18). Similarly, the composition of total soil
nitrogen and phosphorus immobilized (i.e. the ratios of biomass N-to-total soil N
and biomass P-to-total soil P) in the microbial biomass also exhibited variations
(Table 6.18). According to Singh and Singh (1995), the sites exhibiting lower values
Table 6.14  Density (D, stems ha−1), foliage biomass (FB, g m2) and total shoot biomass (TSB, g m−2) of woody species in permanently protected and grazed
savanna
Protected savanna Grazed savanna
Ranitali Hathinala Telburva Ranitali Hathinala Telburva
Species D FB TSB D FB TSB D FB TSB D FB TSB D FB TSB D FB TSB
Hardwickia binata 0 0 0 0 0 0 70 51.3 1678 0 0 0 0 0 0 10 29.7 106
Diospyros melanoxylon 60 22.0 478 30 11.1 299 10 1.2 17 10 5.2 21 20 1.8 22 10 2.3 7
Anogeissus latifolia 0 0 0 90 31.1 1042 10 0.4 6 70 23.3 474 70 12.3 284 20 1.4 13
Butea monosperma 0 0 0 0 0 0 40 25.7 549 0 0 0 0 0 0 10 11.7 264
Lagerstroemia parviflora 80 126.0 2540 30 23.8 45 20 5.2 68 20 8.7 130 30 4.7 57 50 23.7 111
Acacia catechu 60 40.0 926 50 31.7 1663 30 6.7 148 50 10.3 223 30 5.4 153 90 8.7 156
Ziziphus glaberrima 40 11.7 212 10 0.2 2 0 0 0 10 2.5 38 10 1.4 17 10 1.1 14
Soymida febrifuga 0 0 0 10 27.4 877 0 0 0 0 0 0 0 0 0 0 0 0
Ziziphus mauritiana 70 12.0 172 0 0 0 0 0 0 30 2.9 50 0 0 0 0 0 0
Emblica officinalis 10 3.3 50 0 0 0 0 0 0 10 0.9 9 0 0 0 0 0 0
Boswellia serrata 10 70.0 920 0 0 0 0 0 0 10 19.0 488 10 18.7 479 0 0 0
Wrightia tomentosa 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 30 2.2 73
Total 330 285.0 5298 220 125.3 3928 180 90.5 2466 210 72.8 1433 170 44.3 1012 230 80.8 744
Source: Pandey and Singh( 1992b)
6.9  Water-Stable Aggregates and Microbial Biomass in TDF and Derived Ecosystems
181
182 6  Influence of Biotic Pressure and Land-Use Changes

Table 6.15  Microbial biomass C in aggregate fractions of soil


Macro-aggregates Micro-aggregates

MB - C MB - C
´100 ´100
Sites MB-C(μg g−1) SOC MB-C(μg g−1) SOC
Hathinala forest
Dry 401 ± 15.9a 2.2 375 ± 18.0ab 2.1
Moist 494 ± 16.5b 1.7 382 ± 16.7a 1.3
Marihan forest
Hill top 356 ± 1.0c 1.6 348 ± 9.3b 1.5
Hill base 237 ± 8.7d 4.8 127 ± 8.9c 2.6
Telburva savanna 274 ± 1.2c 1.7 260 ± 5.7d 1.6
Barkachha cropland 274 ± 2.3c 4.2 172 ± 8.7c 2.6
BHU cropland 208 ± 9.3f 3.8 I54 ± 8.7c 2.8
Source: Singh and Singh (1995)
Values are mean (± 1 SE) of three replicates. MB-C microbial biomass C, SOC soil organic C
a–f
Values with same letter in a column are not significantly different at P ≤ 0.05

Table 6.16  Microbial biomass N in aggregate fractions of soil


Macro-aggregates Micro-aggregates

MB - N MB - N
´100 ´100
Sites MB-N (μg g−1) TSN MB-N (μg g−1) TSN
Hathinala forest
Dry 51.3 ± 0.0a 4.1 66.0 ± 5.2ab 5.3
Moist 66.0 ± 1.4b 3.5 72.4 ± 0.9a 3.8
Marihan forest
Hill top 43.4 ± 3.9c 3.3 58.6 ± 5.1bc 4.5
Hill base 24.3 ± 1.7d 8.3 27.7 ± 0.8d 9.5
Telburva savanna 36.7 ± 3.8e 3.4 50.5 ± 3.3c 4.7
Barkachha cropland 28.2 ± 0.6d 4.9 39.1 ± 1.4e 6.8
BHU cropland 25.3 ± 1.3d 4.5 35.1 ± 0.8e 6.3
Source: Singh and Singh (1995)
Values are mean (± 1 SE) of three replicates. MB-N microbial biomass N; TSN total soil N
a–e
Values with same letter in a column are not significantly different at P ≤ 0.05.

of total soil C and N showed greater C and N immobilization rates. The rate of P
immobilization in microbial biomass (i.e. the ratio of microbial P-to-total soil P) of
soils was comparatively greater than the rate of N immobilization. Singh and Singh
(1995) reported consistently higher values of microbial C-to-total soil organic C
ratios for the soil macro-aggregates compared to micro-aggregates. However, the
reverse condition was true for the ratios of microbial N or P to total soil N or P. The
microbial species composition found in the soil biomass linked with the aggregates
6.9  Water-Stable Aggregates and Microbial Biomass in TDF and Derived Ecosystems 183

Table 6.17  Microbial biomass P in aggregate fractions of soil


Macro-aggregates Micro-aggregates

MB - P MB - P
´100 ´100
Sites MB-P (μg g−1) TSP MB-P (μg g−1) TSP
Hathinala forest
Dry 16.7 ± 1.1a 12.5 20.0 ± 2.9a 15.0
Moist 20.3 ± 1.1b 18.0 21.7 ± 1.7a 19.3
Marihan forest
Hill top 12.7 ± 0.0c 9.5 16.7 ± 1.7a 12.5
Hill base 5.3 ± 1.0d 4.2 10.0 ± 0.0b 8.0
Telburva savanna 8.3 ± 1.4d 7.7 15.0 ± 2.9ab 13.8
Barkachha cropland 7.6 ± 1.0d 4.8 11.7 ± 1.7b 7.4
BHU cropland 5.9 ± 0.0d 5.4 10.8 ± 0.8b 10.0
Source: Singh and Singh (1995)
Values are mean (± 1 SE) of three replicates. MB-P microbial biomass P, TSP total soil P
a
Values with same letter in a column are not significantly different at P ≤ 0.05.

Table 6.18  Nutrient ratios in microbial biomass in macro- and micro-aggregates


N in the P in the
Microbial Microbial Microbial microbial microbial
C-to-N ratio C-to-P ratio N-to-P ratio biomass (%)a biomass (%)a
Forest
Macro-­ 8.8 27.0 3.4 6.2 1.9
aggregates
Micro-­ 5.5 18.0 3.3 8.1 2.8
aggregates
Savanna
Macro-­ 7.5 33.0 4.4 6.7 1.5
aggregates
Micro-­ 5.2 17.3 3.4 9.5 2.8
aggregates
Cropland
Macro-­ 9.0 35.4 3.9 5.6 1.4
aggregates
Micro-­ 4.4 14.4 3.3 11.4 3.5
aggregates
Source: Singh and Singh (1995)
The values are the average of four sites of forest and two sites of cropland
a
Calculated by assuming that microbial biomass contains 50% C of the total dry wt (Brookes et al.
1984; Srivastava and Singh 1988, 1991)

of different particle sizes may indeed be different. According to Tisdall and Oades
(1982), fungi commonly dominate in macro-aggregates, while bacteria are abun-
dant in micro-aggregates. The occurrence of bacteria within soil micro-aggregates
184 6  Influence of Biotic Pressure and Land-Use Changes

has been demonstrated with the help of electron microscopy (Foster 1988; Gupta
and Germida 1988). It has been observed that fungi and bacteria have remarkably
different C-to-nutrient ratios, for example, Jenkinson and Ladd (1981) reported that
the C-to-N ratio of the fungal hyphae is generally in the range of 10–12, while that
of bacteria commonly between 3 and 5. Singh and Singh (1995) found that the
microbial biomass C-to-N as well as C-to-P ratios were lower for the soil
­micro-­aggregates compared to that for macro-aggregates, which may indicate the
dominance of bacterial populations in the soil micro-aggregates. The microbial bio-
mass N and P concentrations linked with soil micro-aggregates were also greater
compared to that linked with macro-aggregates. Earlier, Hendrix et al. (1986) have
postulated the presence of two different types of microbial food webs in soil, viz.
fungal-­dominated and bacterial-dominated, and consequently have related these
food webs to, respectively, no-tillage and conventional tillage systems. Although the
empirical evidence have not been reported in the study of Singh and Singh (1995),
however, it is possible that the soil macro- and micro-aggregates may provide sup-
port to the fungal- as well as bacteria-dominated food webs, respectively, in the
studied ecosystems. The abundance of bacterial populations in micro-aggregates
results into greater turnover and carbon losses (Coleman et al. 1988), whereas, the
fungal-domination in soil macro-aggregates could favour higher activity of fungal-
based organisms in the food web and lead to higher retention of microbial carbon.
In the study of Singh and Singh (1995), the quantity of microbial C, N and P
concentration in both macro- and micro-aggregates decreased across the forest-­
savanna-­cropland gradient. This decrease was considerably due to a decline of total
soil organic matter in soil, as the total soil C and N exhibited positive relationship
with microbial C and N measured in both macro- as well as micro-aggregates
(Fig. 6.19). Evidently, the biomass of microbial organisms in these soils was con-
trolled by the quantity of soil organic matter. In another study, Ruess and Seagle
(1994) have also observed positive correlations between microbial C and the total
soil C, as well as between microbial N and the total soil N in landscapes of Serengeti.
In the study of Singh and Singh (1995), the composition of microbial biomass N
in the macro- and micro-aggregates measured in cropland soil was, respectively,
58% and 66% compared to that in the forest soil (average across all sites). Similarly,
the microbial P contents of macro- and micro-aggregates measured from cropland
soil were, respectively, 49% and 66% of that observed in the forest soil. It is evident
that higher amount of microbial N and P in the soil were lost due to conversion of
ecosystem from macro-aggregate fraction compared to that from micro-aggregates.
This observation is in contrast to the pattern shown for microbial biomass C. The
composition of microbial biomass C in macro- and micro-aggregates measured
from cropland soil was 65 and 53%, respectively, of that observed in the forest soil.
Gupta and Germida (1988) postulated a significant change in the microbial biomass
quality, especially that related to macro-aggregates, due to the effect of cultivation.
Singh and Singh (1996) observed that the composition of soil macro-aggregate size
fractions was greatest in the forest ecosystem followed by the savanna ecosystem
6.9  Water-Stable Aggregates and Microbial Biomass in TDF and Derived Ecosystems 185

MACROAGGREGATES MICROAGGREGATES
500 a d
MICROBIAL C (µg g-1)
400

300

200

100
0.0 1.0 2.0 3.0 0.0 1.0 2.0 3.0

TOTAL SOIL C (%)

80 b e
MICROBIAL N (µg g-1)

60

40

20

0.00 0.05 0.10 0.15 0.20 0.00 0.05 0.10 0.15 0.20
TOTAL SOIL N (%)

25 c f
MICROBIAL P (µg g-1)

20

15

10

0
80 120 160 200 80 120 160 200

TOTAL SOIL P (µg g-1)

Fig. 6.19  Relationships between microbial biomass C, N and P in macro- and micro-aggregates.
a Y = 176.70+97.95X (r2 = 0.84, P < 0.001), where Y is microbial C μg g−1 in macro-aggregates
and X is total soil organic C (%). b Y = 11.76+271.70X (r2 = 0.94, P < 0.001), where Y is microbial
N μg g−1 in macro-aggregates and X is total soil N (%). c Microbial P in macro-aggregates was
independent of total soil P. d Y = 100.95+108.1X (r2 = 0.88, P < 0.001), where Y is microbial C μg
g−1 in micro-aggregates and X is total soil organic C (%). e Y  =  19.89+295.97X (r2  =  0.95,
P < 0.001), where Y is microbial N (μg g−1) in micro-aggregates and X is total soil N (%). (f)
Microbial P in micro-aggregates was independent of total soil P (Source: Singh and Singh 1995)
186 6  Influence of Biotic Pressure and Land-Use Changes

and cropland ecosystem, whereas the proportion of soil micro-aggregates was great-
est in the cropland ecosystem and lowest in the forest ecosystem. This explains why
the forest ecosystem is capable of C conservation better than the alternate land use
types. Soil organic C and total N contents observed in both macro- as well as micro-­
aggregates were also greatest in the forest ecosystem followed by the savanna eco-
system and cropland ecosystem. While the soil organic C and total N contents were
reported higher in the soil macro-aggregate size fractions, total P content in soil was
greatest in the micro-aggregate size fractions.
Singh et al. (1991b) estimated carbon storage in forest and savanna vegetation
for two consecutive years and found that the total carbon storage in the forest veg-
etation averaged 28.68 t C ha−1 and that in savanna 12.89 t C ha−1 (Table 6.19). In
savanna ecosystem, the contribution of below-ground biomass to total carbon stor-
age was more (37%) compared to that in the forest ecosystem (15%). The contribu-
tion of herbaceous vegetation plus fine roots in total C stock of vegetation was only
4% for forest ecosystem compared to 49% for savannah ecosystem. The net carbon
input (TNP) in forest and savanna ranged between 4.75 and 8.06 t C ha−1 year−1
(mean = 6.27 t C ha−1 year−1) and between 5.49 and 6.45 t C ha−1 year−1 (mean = 5.96 t
C ha−1  year−1), respectively (Table  6.20). Thus, although the carbon storage in
savanna (mean B/P = 2.18) was half that in forest (mean B/P = 4.54), carbon flux
was almost equal to that in forest, resulting in mean P/B ratio of 0.47 for savanna
and 0.22 for forest. The contribution of herbaceous vegetation plus fine roots to
TNP averaged 84% for savanna and only 28% for forest. Evidently, drawing upon
the local biodiversity resources, the converted savanna system maintained the same
level of ecosystem functioning (productivity) as the forest through fast-growing
species of shorter life-span which replaced the longer-lived, slow growing species
of the forest at a cost of reduced C conservation in the biomass. Thus, a low-­standing
crop of C (i.e. small structure) in the savanna, in which functional importance shifts
from woody canopy to herbaceous ground stratum where C4 grasses assume pre-
dominance, was associated with high C flux. On the other hand, a high-standing
crop of C (i.e. larger structure) in the dry forest from which the savanna is derived
was associated with lower C flux. The share of belowground parts in TNP in savanna
was greater (36%) compared to forest (26%). However, this increased below-ground
input was not reflected in soil carbon status. Organic C in savanna soil averaged
1.16%, which is only 53% of that of forest soils (2.18%). High temperatures and
rapid, pulsed release of nutrients in impoverished soil to support flushes of growth
in tropical grasslands result in higher rates of decomposition, negating the potential
effect of increased C flux to roots. Other studies have suggested 20–40% loss in soil
carbon after conversion of forest vegetation to permanent agriculture and pasture
landscape (Detwiler and Hall 1988; Srivastava and Singh 1989). Singh et al. (1991b)
suggest that conversion of TDF into savanna increases C flux relative to storage and
that savanna is a poor carbon-conserving system.
To conclude, the TDFs experience multiple pressures, often simultaneously, and
therefore, for the effective conservation of the distinctive biodiversity of TDF, a key
priority should be expansion of the network of protected areas across the globe to
include particularly TDF area in South, Central and North America. Highest priority
6.9  Water-Stable Aggregates and Microbial Biomass in TDF and Derived Ecosystems 187

Table 6.19  Plant carbon mass (t C ha−1) in tropical dry deciduous forest and derived savanna
Component Above-ground Below-ground Total
Forest
Site 1
Woody 34.32 4.80 39.12
Herbaceous (+ tree fine roots) 0.14 0.97 1.12
Total 34.46 5.77 40.24
Site 2
Woody 18.31 2.65 20.97
Herbaceous (+ tree fine roots) 0.13 1.16 1.29
Total 18.44 3.81 22.26
Site 3
Woody 19.90 2.75 22.65
Herbaceous (+ tree fine roots) 0.10 0.65 0.75
Total 20.00 3.40 23.40
Average for forest
Woody 24.18 3.41 27.58
Herbaceous (+ tree fine roots) 0.12 0.93 1.05
Total 24.30 4.33 28.68
Savanna
Site 1
Woody 7.83 0.91 8.74
Herbaceous (+ tree fine roots) 1.92 3.33 5.25
Total 9.75 4.24 13.99
Site 2
Woody 4.04 0.64 4.68
Herbaceous (+ tree fine roots) 2.26 4.02 6.28
Total 6.30 4.66 10.96
Site 3
Woody 5.61 0.88 6.49
Herbaceous (+ tree fine roots) 2.68 4.54 7.22
Total 8.29 5.42 13.71
Average for savanna
Woody 5.83 0.81 6.64
Herbaceous (+ tree fine roots) 2.29 3.96 6.25
Total 8.11 4.77 12.89
Source: Singh et al. (1991b)
Differences in standing crops of carbon between forest and savanna were significant, P < 0.001

should be given to the forest regions most at risk due to human activity, as well as
those of highest biodiversity value. The study in Indian TDF reported decline in
diversity of plant species with increment in the level of disturbance. Investigations
revealed an increase in the number of forest fragments in the Indian study sites
indicating fragmentation in progress. Further research is needed to investigate the
possible impact of forest fragmentation on the overall species diversity of the TDFs.
188 6  Influence of Biotic Pressure and Land-Use Changes

Table 6.20  Carbon input through net primary production (t C ha−1 year−1) in tropical dry deciduous
forest and derived savanna. TNP Total Net Production
Component Above-ground Belowground TNP
Forest
Site 1
Woody 5.70 0.37 6.07
Herbaceous (+ tree fine roots) 0.38 1.61 1.99
Total 6.08 1.98 8.06
Site 2
Woody 3.83 0.16 3.99
Herbaceous (+ tree fine roots) 0.36 1.68 2.04
Total 4.19 1.84 6.04
Site 3
Woody 3.34 0.18 3.52
Herbaceous (+ tree fine roots) 0.30 0.93 1.23
Total 3.64 1.11 4.75
Average for forest
Woody 4.29 0.23 4.52
Herbaceous (+ tree fine roots) 0.34 1.41 1.75
Total 4.63 1.64 6.27
Savanna
Site 1
Woody 0.96 0.08 1.04
Herbaceous (+ tree fine roots) 2.71 1.74 4.45
Total 3.67 1.82 5.49
Site 2
Woody 0.87 0.10 0.97
Herbaceous (+ tree fine roots) 2.94 2.02 4.96
Total 3.81 2.12 5.93
Site 3
Woody 0.77 0.06 0.83
Herbaceous (+ tree fine roots) 3.24 2.38 5.62
Total 4.01 2.44 6.45
Average for savanna
Woody 0.86 0.08 0.94
Herbaceous (+ tree fine roots) 2.96 2.05 5.01
Total 3.82 2.13 5.96
Source: Singh et al. (1991b)
Differences in carbon-input values between forest and savanna were significant, P < 0.001

Studies on regeneration in TDFs of India found that compared to natural mortality


due to impacts of drought and fire, damage to juveniles due to browsing and mortal-
ity of saplings due to harvesting were remarkably higher. The site with the mini-
mum level of soil moisture content throughout the year exhibited the minimum
number of species and stem density and also had the maximum level of disturbance.
For the protection and increasing the plant diversity and recruitment of juveniles
6.9  Water-Stable Aggregates and Microbial Biomass in TDF and Derived Ecosystems 189

and saplings, harvesting and browsing need to be controlled. The energy require-
ment of the agroecosystems in the TDF region is substantially subsidized by forage
as well as wood inputs from surrounding forests. The study in Indian TDF soil
reported a remarkable seasonality in the viable size of community of ammonia- as
well as nitrite-oxidizing bacteria. The results indicated that in the nutrient-poor for-
est sites, N-mineralization was controlled by the viable nitrifying bacterial commu-
nity, which was water limited. In the croplands, decreased organic matter content
and more clay content tended to suppress the size of the viable nitrifying bacterial
community, particularly the community of nitrite-oxidizing bacteria, and the rates
of N-mineralization. Positive relationships of the soil microbial C, N and P with the
root biomass as well as total plant biomass (above-ground biomass + root biomass)
have been reported. There was remarkable reduction in the amounts of soil nutrient
contents and microbial C, N and P as a result of the conversion of forest into other
land uses. Lantana invasion in the Indian TDF severely depleted tree population.
The lantana shrub exhibits dense understorey in the TDF which perturbs the recruit-
ment of native tree seedlings in the forest, leading to differential depletion in the
population of native trees. The study on a forest-derived savanna vegetation reported
that the savanna is a pulsed system which shows quick response to rainfall events.
The study also found that the conversion of TDF into savanna leads to increase in
carbon flux compared to storage.
Chapter 7
Research Perspectives

Tropical dry forests (TDFs) contain unique and threatened biodiversity (Pennington
et al. 2009) and provide a suite of ecosystem services which are the basis for the
livelihoods of millions  of people; however, the understanding of the linkages
between ecological processes and human well-being is poor (Maass et al. 2005). In
most of the dry tropical regions, ecosystem services and their delivery have not been
quantified. For the sustainable development in the TDF areas, there is an urgent
need of holistic understanding of the interrelationships among various productive
activities, human well-being and ecosystem functioning (Maass et al. 2005).
The low and erratic rainfall, prolonged dry periods and impoverished soil have
resulted in adaptations to drought and nutrient conserving mechanisms; these have
been little studied, particularly through experiments, in several regions. Although
TDF shows some degree of resilience in the face of biotic and abiotic disturbances
such as  wood harvest, grazing, ecosystem conversion and drought (Durán et  al.
2002; Segura et al. 2003), changes do occur in the structure and functioning which
need to be quantified and understood.
In spite of some resistance to the invasion of exotic species, large areas of the
TDF are still being invaded by alien species, adversely affecting the regeneration of
indigenous tree species and interfering with forestry operations. This aspect has
been little studied.
The seasonal tropics are experiencing alterations in rainfall regimes due to cli-
mate change, partly derived by anthropogenic activities (Greve et al. 2014; Chadwick
et al. 2015); even more extreme changes in future precipitation have been projected
(Malhi et al. 2008; Feng et al. 2013) which may be characterized by decreased rain-
fall and increased dry periods (e.g. Maloney et al. 2013; Chadwick et al. 2015; Duffy
et al. 2015). However, it is not yet clear how the ecosystem processes, ecological
dynamics of species and communities will be affected by variations in the timing
and magnitude of precipitation (Vico et al. 2014). Allen et al. (2017) have suggested
that the future droughts might have several possibilities such as reduction in the
annual rainfall, alteration in the timing of rainfall, alteration in the length of dry
season, altered length of dry season coupled with reduction in rainfall or ­multi-­year

© Springer Nature Singapore Pte Ltd. 2017 191


J. S. Singh, R. K. Chaturvedi, Tropical Dry Deciduous Forest: Research Trends
and Emerging Features, https://doi.org/10.1007/978-981-10-7260-4_7
192 7  Research Perspectives

drought with sequential low rainfall; concerted research effort is needed to under-
stand the response of TDF to such drought conditions. Elevated temperatures are
generally associated with the drought, and both the drought and increased tempera-
tures may reduce growth and/or increase tree mortality (Allen et al. 2010). Also, in
drought conditions, the effect of other disturbances may increase (Van Bloem et al.
2006; Imbert and Portecop 2008; Chaturvedi et  al. 2017b) and can enhance the
occurrence, duration or severity of fires or insect/pathogen outbreaks; thus drought
and ecosystem functioning will need in-depth studies. Moreover, topography,
edaphic and vegetation characteristics may modify ecosystem responses to drought
(Powers et al. 2015), and appropriate and comprehensive research strategy needs to
be developed to assess the response of TDFs to drought in various regions.
Changes and variability in precipitation regimes may affect growth, physiology
and phenology of TDF species (Allen et al. 2017), and low precipitation could sig-
nificantly impact productivity and carbon balance. Increased inter-annual variability
in precipitation, high intervals between the wet and dry years and especially for
TDF the predicted decline in precipitation could affect the relative performance of
species possessing different leaf habits and functional traits.
Due to patchy spatial distribution of many species (Balvanera et al. 2002) and
presence of many rare and endemic species (Ceballos 1995; Durán et al. 2002; Lott
and Atkinson 2002; Miranda 2002), a rapid species loss can occur in the TDF in
future due to climate change (Maass et al. 2005), but the loss has to be scientifically
assessed. Despite the extensive distribution of TDFs and the growing global support
against the destruction by humans (Janzen 1988b), the scientists and land managers
have mostly neglected their status (Miles et  al. 2006; Chidumayo and Marunda
2010; Blackie et al. 2014; Sunderland et al. 2015). Some analysis has been done for
the identification of forest threats from the population growing in adjoining areas or
rural belt, but empirical investigation at the global level regarding the dynamics of
TDF deforestation is still warranted (Blackie et al. 2014).
Rudel (2017) suggested few important questions about the dynamics of defores-
tation which need to be investigated in the areas experiencing deforestation. First,
although the land managers or the authorities managing the forest must pay more
attention towards the multiple uses of forests, they should also think how they can
do so in a sustainable way (Pretty 2011; Montpellier Panel 2013). Second, the fire
dynamics in forests have been poorly documented; what is the role of fire in climate
change and how can it be managed so that it does not increase the greenhouse gas
emission? Third, what is the effect of climate change in tropical biomes, and how is
the natural effect of climate change coupled with certain human feedback effects?
The forest loss has been accelerated due to drought induced by climate change
directly because of decreasing precipitation but also indirectly through changes in
rural livelihoods. Onishi (2016) has recently reported an instance in Madagascar
that farmers, after experiencing crop failures due to less rain, decided to produce
charcoal by cutting trees and thus accelerated the deforestation rates in the dry for-
ests of Madagascar.
In Central America, the Caribbean and other tropical countries, the re-­
establishment of the forest through successional processes has been observed after
7  Research Perspectives 193

abandonment of the land use pressure on the deforested landscapes (Rudel et al.
2000; Rudel 2005). In Puerto Rico, a series of island-wide forest inventories
observed that forest succession following abandonment of agricultural use produced
forest stands with the dominance of introduced species (Lugo and Helmer 2004).
Hobbs et al. (2006) categorized such types of forests as novel forests, since they are
developed as a consequence of human activity, leading to formation of new species
combinations, which are expanding in land cover all over the world (Hobbs et al.
2013). Novel forests are categorized as “the new wild” of the Anthropocene Epoch
(Pearce 2015) and represent a “new world order” (Hobbs et  al. 2013); in Puerto
Rico, the anthropogenic effect has converted 75% of the forest cover into novel
forests (Martinuzzi, et al. 2013). These novel forests have been little studied, and
their role in the conservation of TDFs remains undocumented.
The limited studies of the dry tropics are not adequate for defining the real status
of dry forests. Majority of deforestation data sets, such as those of the Food and
Agriculture Organization of the United Nations’ (FAO) Forest Resource Assessment
(FRA) report, do not differentiate between dry and wet forests (Blackie et al. 2014).
After the introduction of updated measurement technologies during the 1990s, the
earlier measures for dry forests and woodlands have now changed. Earlier, when
remote sensing was not commonly used, dry forests in a geographical area were
defined by emphasizing seasonality of the climate and tree stature (see Murphy and
Lugo 1986). However, with the advent of remote sensing technologies after 1990,
canopy cover became the more noticeable measure of dry forests (Miles et al. 2006).
Identification of the regional factors influencing the dynamics of deforestation
gives important insight about the diverse combinations of the forces leading to the
destruction of tropical forests in different places (Rudel 2017). Particularly, the dif-
ferences in rainfall might have cascading effects on the processes of deforestation;
therefore, differences in the deforestation dynamics might be observed in dry and
wet forests. Evidently, investigation of the differences in the dynamics of deforesta-
tion is necessary for the implementation of region- specific policies for reducing the
deforestation rate. The investigation of the differences in wet and dry forest dynam-
ics is also important because the global models of climate change are now expand-
ing the project towards the subtropical dry zones and, probably, dry forests in the
near future (Norris et al. 2016). REDD+ (Reducing Emissions from Deforestation
and Degradation) systems could be implemented to reduce deforestation both in dry
and wet forests; however the organization by which REDD+ programmes is sup-
posed to be implemented should be quite different for dry and wet forests. The
community-based organizations have an important role for the implementation of
REDD+ in the dry forests where generally a small number of local people degrade
the forests, whereas deforestation in wet forests is done by large enterprises; there-
fore for administering REDD+ in wet forests, a more centralized governing body,
including internationally certified groups and national governments, taking the help
of remote sensing tools, could probably play a significant role (Rudel 2017).
Studies in TDFs on forest succession, degradation and restoration are rare, and
many of them have been reported from few sites. Before taking up restoration, it is
important to find out the drivers of deforestation. For example, Rudel (2017) argued
194 7  Research Perspectives

that wet tropical deforestation is “industrial”, while dry tropical deforestation is


“artisanal” in which “small groups, work in labor intensive ways with few tools,
cutting down trees in dry forests and woodlands for collection of foods, wood-based
fuels, and construction poles for local and regional markets”; this results in rapid
depletion of forest cover and pronounced fragmentation with its deleterious effect
on biodiversity. Therefore, the conservation policy for TDFs has to be made at local
levels.
There is a dire need to restore forests in large areas on deforested or degraded
land with the aim to increase the ecosystem functionality and resilience, conserve
biodiversity, sequester carbon and mitigate effects of global climate change
(Brancalion and Chazdon 2017). Brancalion and Chazdon (2017) have proposed
four guiding principles for reforestation: “(1) restoration interventions should have
proper objectives for the enhancement and diversification of local livelihoods; (2)
afforestation should not replace native tropical grasslands or savannah ecosystems;
(3) reforestation approaches should promote landscape heterogeneity and biological
diversity; and (4) residual carbon stocks should be quantitatively and qualitatively
distinguished from newly established carbon stocks”. The restoration approach
requires collection of baseline data at landscape level and monitoring changes in the
landscapes during the restoration process. The principles for restoration proposed
by Brancalion and Chazdon (2017) would provide a platform for the implementa-
tion of management initiatives based broadly on socio-environmental benefits
including, carbon mitigation and wood production.
Some other areas needing attention are faunal and microbial diversity, physiolog-
ical mechanisms underlying the seasonality of leaf senescence and canopy renewal,
pollination, reproduction and breeding systems.
Globally, efforts are being made to combat tropical species extinctions; however
the high-quality, objective information on tropical biodiversity is still lacking which
has hampered the quantitative analysis of conservation strategies (Beaudrot et al.
2016). Researchers are mostly involved in the issues of tropical deforestation, and
they generally neglect the research on forest governance. Since the process of defor-
estation in dry forests is mostly artisanal, it is likely to produce more fragmented
forests. This situation, coupled with the anticipated deleterious impact of forest
fragmentation on biodiversity, would particularly make it more important to analyse
the deforestation-biodiversity crisis in the settings of TDFs (Rudel 2017). A social
dimension is an important feature of conservation solutions, which includes oppor-
tunities and incentives for the local human populations and the private landowners
(Blackie et al. 2014).
For the conservation of Puerto Rico’s dry forests, the protocols suggested by
Lugo and Erickson (2017) include “(1) increases in species dominance and density
at the 1-ha scale; (2) restoring forest conditions on degraded sites; (3) facilitating
the regeneration and growth of native species on degraded sites; (4) establishing
litter patches with diverse concentrations of chemical elements; (5) increasing the
presence of nitrogen-fixing species and the availability of nitrogen in forest stands;
(6) increasing the concentration of some elements in leaf litter above those observed
in native forest stands; (7) decreasing the C/N ratio below those observed in native
7  Research Perspectives 195

dry forest stands; and (8) potentially accelerating the flux of mass and nutrients by
accelerating litter decomposition or primary productivity rates”.
In a recent communication, DRYFLOR (2016) provided a scientific framework
for the conservation of highly threatened seasonally dry tropical forests (SDTFs)
distributed across Latin America and the Caribbean. They reported variation in flo-
ristic composition at a continental scale and identified 12 floristic groups of dry
forest across the Neotropics. These figures are likely to highlight high levels of spe-
cies endemism, which clearly illustrates the high floristic turnover (beta diversity)
as revealed by the data of DRYFLOR (2016). Such high endemism and beta diver-
sity in the dry forests indicate that failure in the protection of the forest would result
in high depletion of unique species diversity. The high floristic turnover suggests
that large numbers of conservation areas spread across many countries are required
for the protection of the total diversity of TDFs. For the protection and preservation
of the TDF ecosystem, there are only a few national parks and biological reserves,
and the real biological research stations bridging the gap between the conservation
biology and ecology in TDFs are also handful. Moreover, the policies and funding
from the local governments and international institutions are mainly implemented
for the protection of pristine national parks or mature secondary forest, located far
from the TDF region.
Under the prevailing biodiversity crisis, about 200 countries have committed to
expand the coverage of effective protected habitats and prohibit further extinctions
of threatened species, and implement measures to improve the status of species
expected to decline by 2020 under the Convention on Biological Diversity (CBD)
Aichi Biodiversity Targets. According to Beaudrot et al. (2016), the two key factors
inhibiting the progress for achieving Aichi Targets 11 (Protected Areas) and 12
(Preventing Extinctions) are (1) significant disparity in the existing collection of
tropical biodiversity data as compared to that of higher latitude regions (Collen
et al. 2008). The magnitude of disparity becomes even stronger due to the fact that
tropical regions contain the highest species richness. Thus, the rigorous assessment
of the response of tropical species to threats gets hampered, leading to prevention of
conservation efforts for the targeted species (Brooks et al. 2009; Feeley and Silman
2011). (2) The quality of the data available for tropical biodiversity is not adequate;
therefore conclusions are mostly drawn on the basis of expert opinion and the aggre-
gated secondary data. Particularly, primary in situ data for the populations in pro-
tected tropical areas are lacking (Geldmann et  al. 2013). For example, the most
recent comprehensive assessment of the effective performance of protected areas in
tropical forests was based mostly on the expert opinion which concluded that many
parks were not effective (Laurance et al. 2012). Moreover, the effectiveness of con-
servation decisions requires more objective evaluation with the help of primary data
and robust predictors of the changing biodiversity (Pereira et al. 2013).
For mitigating climate change, reducing tropical forest deforestation and prevent-
ing biodiversity loss, various national and international policies and ­programmes
have been deployed (Börner et  al. 2016). However, our current knowledge about
achieving forest conservation and other related development goals is still poor and
fragmented (Ferraro 2009, 2011; Pattanayak et al. 2010; Baylis et al. 2016). Therefore,
the robust baseline data needs to be recorded at landscape scales. Moreover, follow-
196 7  Research Perspectives

ing REDD+ guidelines, the targeted conservation goals can be achieved by the estab-
lishment of new forests coupled with conserving residual forests and by implementing
measures for the recovery of existing secondary and degraded forests. Despite vari-
ous socio-environmental benefits of forests promoted by the tropical forest and land-
scape restoration (FLR) movement, the market-driven demand of forest products
such as carbon, timber, pulp and biofuels is potentially disturbing the FLR goals
(Newton et al. 2012). Therefore, the countries and organizations implementing FLR
initiatives should extensively monitor carbon stocks by providing technical and
financial support for multiple surveys organized by FLR programme. Establishment
of scientific monitoring guidelines for evaluating landscape-­scale carbon stocks and
other related FLR outcomes could become an important step towards more signifi-
cant and accurate evaluations (Börner et al. 2016).
FLR programmes do not necessarily follow formal standardized protocols and
are regionally driven by engagement of local stakeholders, and according to country-­
specific requirements and capacities, nevertheless, clear guiding procedures are
needed for proper accountability and for dealing with essential issues for attracting
confidence of “investors” and “entrepreneurs” in FLR initiatives in order to get
proper financial support (Gutierrez and Keijzer 2015). Local FLR initiatives could
be critical in improving and advancing the progress of monitoring approaches.
Capacity of these initiatives could be improved by organizing workshops and pro-
viding technical support and other necessary tools required for simulating temporal
variations in landscape parameters and carbon stocks by the data generated after a
series of different kinds of restorative interventions.
In conclusion, TDFs are in dire need of research projects focussed on integrated
and multidisciplinary conservation which could expand the traditional species- and
niche-based field of research, leading to increase in the ecological and biological
knowledge base and including human dimensions, which inevitably underlie how
ecosystems change over time.
References

Abiven S, Recous S, Reyes V et al (2005) Mineralisation of C and N from root, stem and leaf resi-
dues in soil and role of their biochemical quality. Biol Fertil Soils 42:119–128
Aerts R (1995) The advantages of being evergreen. Trends Ecol Evol 10:402–407
Aerts R (1996) Nutrient resorption from senescing leaves of perennials: are there general patterns?
J Ecol 84:597–608
Aerts R (1997) Climate, leaf litter chemistry and leaf litter decomposition in terrestrial ecosystems:
a triangular relationship. Oikos 79:439–449
Aerts R, Chapin FS (2000) The mineral nutrition of wild plants revisited: a re-evaluation of pro-
cesses and patterns. Adv Ecol Res 30:1–67
Aiba SI, Kitayama K (1999) Structure, composition and species diversity in an altitude-substrate
matrix of rain forest tree communities on Mount Kinabalu, Borneo. Plant Ecol 140:139–157
Aide TM, Zimmerman JK, Pascarella JB et al (2000) Forest regeneration in a chronosequence of
tropical abandoned pastures: implications for restoration ecology. Restor Ecol 8:328–338
Albrecht J, Bohle V, Berens DG et al (2015) Variation in neighbourhood context shapes frugivore-
mediated facilitation and competition among codispersed plant species. J Ecol 103:526–536
Alder NN, Sperry JS, Pockman WT (1996) Root and stem xylem embolism, stomatal conductance,
and leaf turgor in Acer grandidentatum populations along a soil moisture gradient. Oecologia
105:293–301
Allen JC (1985) Soil response to forest clearing in the United States and the tropics: geological and
biological factors. Biotropica 17:15–27
Allen CD, Macalady AK, Chenchouni H et al (2010) A global overview of drought and heat induced
tree mortality reveals emerging climate change risks for forests. For Ecol Manag 259:660–684
Allen K, Dupuy JM, Gei MG et al (2017) Will seasonally dry tropical forests be sensitive or resis-
tant to future changes in rainfall regimes? Environ Res Lett 12:023001
Allison FA (1973) Soil organic matter and its role in crop production. Elsevier, Amsterdam
Aluko AP, Advayi EA (1983) Response of forest tree seedlings (Terminalia ivorensis) to varying
levels of nitrogen and phosphorus fertilizers. Plant Nutr 6:219–237
Alvim PDe T (1960) Moisture stress as a requirement for flowering of coffee. Science 132:354
Alvim PDe T (1964) Tree growth and periodicity in tropical climates. In: Zimmerman MH (ed)
The formation of wood in tropical trees. Academic, New York, pp 479–495
Anaya CA, García-Oliva F, Jaramillo VJ (2007) Rainfall and labile carbon availability control litter
nitrogen dynamics in a tropical dry forest. Oecologia 150:602–610
Anaya CA, Jaramillo VJ, Martínez-Yrízar A et al (2012) Large rainfall pulses control litter decom-
position in a tropical dry forest: evidence from an 8-year study. Ecosystems 15:652–663

© Springer Nature Singapore Pte Ltd. 2017 197


J. S. Singh, R. K. Chaturvedi, Tropical Dry Deciduous Forest: Research Trends
and Emerging Features, https://doi.org/10.1007/978-981-10-7260-4
198 References

Appel T (1998) Non-biomass soil organic N—the substrate for N mineralization flushes following
soil drying–rewetting and for organic N rendered CaCl2-extractable upon soil drying. Soil Biol
Biochem 30:1445–1456
Archibald OW (1995) Ecology of world vegetation. Chapman and Hall, London
Armstrong DP, Westoby M (1993) Seedlings from large seeds tolerate defoliation better: a test
using phylogenetically independent contrasts. Ecology 74:1092–1100
Arnason JT, Lambert JDH (1982) Nitrogen cycling in the seasonally dry forest of Belize, Central
America. Plant Soil 67:333–342
Arnold AE, Asquith NM (2002) Herbivory in a fragmented tropical forest: patterns from islands at
Lago Gatún, Panama. Biodivers Conserv 11:1663–1680
Arunachalam K, Singh ND, Arunachalam A (2003) Decomposition of leguminous crop residues in
jhum cultivation system in northeast India. J Plant Nutr Soil Sci 166:731–736
Ashton MS, Gunatilleke CVS, Singhakumara BMP et  al (2001) Restoration pathways for rain
forests in southwest Sri Lanka: a review of concepts and models. For Ecol Manag 154:1–23
Asquith NM, Mejia-Chang M (2005) Mammals, edge effects, and the loss of tropical forest diver-
sity. Ecology 86:379–390
Aubréville A (1938) La forêt Coloniale: les forêts de l’Afrique occidentale français. Ann Acad Sci
Col (Paris) 9:1–245
Augspurger CK (1980) Mass flowering of a tropical shrub (Hybanthus prunifolius): influence on
pollinator attraction and movement. Evolution 34:475–488
Augspurger CK (1983) Phenology, flowering synchrony, and fruit set of six Neotropical shrubs.
Biotropica 15:257–267
Austin AT, Vitousek PM (2000) Precipitation, decomposition and litter decomposability of
Metrosideros polymorpha in active forests on Hawaii. J Ecol 88:129–138
Austin AT, Vivanco VL (2006) Plant litter decomposition in a semi-arid ecosystem controlled by
photo degradation. Nature 422:555–558
Austin AT, Yahdjian L, Stark JM et al (2004) Water pulses and biogeochemical cycles in arid and
semiarid ecosystems. Oecologia 141:221–235
Austin AT, Araujo PI, Leva PE (2009) Interaction of position, litter type, and water pulses on
decomposition of grasses from the semiarid Patagonian steppe. Ecology 90:2642–2647
Aweto AO (1981) Secondary succession and soil fertility restoration in south-western Nigeria.
II. Soil fertility restoration. J Ecol 69:609–614
Ayanaba A, Tuckwell SB, Jenkinson DS (1976) The effects of clearing and cropping on the organic
reserves and biomass of tropical forest soils. Soil Biol Biochem 8:519–525
Ayyappan N, Parthasarathy N (1999) Biodiversity inventory of trees in a large-scale permanent
plot of tropical evergreen forest at Varagalaiar, Anamalais, Western Ghats, India. Biodivers
Conserv 8:1533–1554
Baas P, Ewers FW, Davis SD et al (2004) Evolution of xylem physiology. In: Hemsley AR, Poole
I (eds) The evolution of plant physiology. Elsevier, Amsterdam, pp 273–295
Backéus I, Rulangaranga ZK, Skoglund J (1994) Vegetation changes on formerly overgrazed hill
slopes in semi-arid central Tanzania. J Veg Sci 5:327–336
Backeus I, Pettersson B, Stromquist L et al (2006) Tree communities and structural dynamics in
miombo (Brachystegia-Julbernardia) woodland, Tanzania. For Ecol Manag 230:171–178
Baker HG (1972) Seed weight in relation to environmental conditions in California. Ecology
53:997–1010
Baker TR, Burslem DFRP, Swaine MD (2003) Associations between tree growth, soil fertility
and water availability at local and regional scales in Ghanaian tropical rain forest. J Trop Ecol
19:109–125
Baker PJ, Bunyavejchewin S, Oliver CD et  al (2005) Disturbance history and historical stand
dynamics of a seasonal tropical forest in western Thailand. Ecol Monogr 75:317–343
Baker TR, Phillips OL, Laurance WF et al (2008) Do species traits determine patterns of wood
production in Amazonian forests? Biogeosciences 5:3593–3621
References 199

Balvanera P, Lott E, Segura G et al (2002) Patterns of ß-diversity in a Mexican tropical dry forest.
J Veg Sci 13:145–158
Balvanera P, Quijas S, Pérez-Jiménez A (2011) Distribution patterns of tropical dry forest trees
along a mesoscale water availability gradient. Biotropica 43:414–422
Barley KP (1970) The configuration of the root system in relation to nutrient uptake. Adv Agron
22:159–201
Baruch Z, Goldstein G (1999) Leaf construction cost, nutrient concentration, and net CO2 assimi-
lation of native and invasive species in Hawaii. Oecologia 121:183–192
Batianoff GN, Butler DW (2003) Impact assessment and analysis of sixty-six priority invasive
weeds in southeast Queensland. Plant Protect Q 18:11–17
Bauer A, Black AL (1981) Soil carbon, nitrogen and bulk density comparison in two cropland till-
age systems after 25 years and in virgin grasslands. Soil Sci Soc Am J 45:1166–1170
Bawa KS (1974) Breeding systems of tree species of a lowland tropical community. Evolution
28:85–92
Baylis K, Honey-Rosés J, Börner J et al (2016) Mainstreaming impact evaluation in nature conser-
vation. Conserv Lett 9(1):58–64
Bazzaz FA, Miao SL (1993) Successional status, seed size, and responses of tree seedlings to CO2,
light and nutrients. Ecology 74:104–112
Bazzaz FA, Pickett STA (1980) Physiological ecology of tropical succession: a comparative
review. Annu Rev Ecol Syst 11:287–310
Beaudrot L, Ahumada JA, O’Brien T et al (2016) Standardized assessment of biodiversity trends in
tropical forest protected areas: the end is not in sight. PLoS Biol 14:e1002357
Becknell JM, Kissing Kucek L, Powers JS (2012) Aboveground biomass in mature and second-
ary seasonally dry tropical forests: a literature review and global synthesis. For Ecol Manag
276:88–95
Bejarano M, Crosby MM, Parra V et al (2014) Precipitation regime and nitrogen addition effects
on leaf litter decomposition in tropical dry forests. Biotropica 46:415–424
Bejarano-Castillo M, Campo J, Roa-Fuentes LL (2015) Effects of increased nitrogen availability
on C and N cycles in tropical forests: a meta-analysis. PLoS One 10:e0144253
Bell DT, Plummer JA, Taylor SK (1993) Seed germination ecology in south western Western
Australia. Bot Rev 76:24–73
Belser LW (1979) Population ecology of nitrifying bacteria. Annu Rev Microbiol 33:309–333
Bender DJ, Contreras TA, Fahrig L (1998) Habitat loss and population decline: a meta-analysis of
the patch size effect. Ecology 79:517–533
Benítez-Malvido J (1998) Impact of forest fragmentation on seedling abundance in a tropical rain
forest. Conserv Biol 12:380–389
Benítez-Malvido J, Martínez-Ramos M (2003a) Influence of edge exposure on tree seedling spe-
cies recruitment in tropical rain forest fragments. Biotropica 35:530–541
Benítez-Malvido J, Martínez-Ramos M (2003b) Impact of forest fragmentation on understory
plant species richness in Amazonia. Conserv Biol 17:389–400
Benítez-Malvido J, Gallardo-Vásquez JC, Alvarez-Añorve MY et al (2014) Influence of matrix
type on tree community assemblages along tropical dry forest edges. Am J Bot 101:820–829
Bertiller MB, Mazzarino MJ, Carrera AL et al (2006) Leaf strategies and soil N across a regional
humidity gradient in Patagonia. Oecologia 148:612–624
Bhaskar R, Ackerly DD (2006) Ecological relevance of minimum seasonal water potentials.
Physiol Plant 127:353–359
Birch HF (1958) The effect of soil drying on humus decomposition and nitrogen availability. Plant
Soil 10:9–31
Birch HF (1960) Nitrification in soil after different periods of dryness. Plant Soil 12:81–96
Blackie R, Baldauf C, Gautier D et al (2014) Tropical dry forests: the state of global knowledge
and recommendations for future research, Discussion paper 2. CIFOR, Bogor
Blankenship RE (2002) Molecular mechanisms of photosynthesis. Blackwell Science, Oxford
200 References

Bloom AJ, Chapin FS, Mooney HA (1985) Resource limitation in plants – an economic analogy.
Annu Rev Ecol Syst 16:363–392
Boeklen WJ, Simberloff D (1987) Area based extinction models in conservation. In: Elliot DK (ed)
Dynamics of extinctions. Wiley, New York, pp 247–276
Boerner REJ (1985) Foliar nutrients dynamics, growth and nutrient use efficiency of Hamamellis
virginiana in three forest microsites. Can J Bot 63:1476–1481
Bohlman SA (2010) Landscape patterns and environmental controls of deciduousness in forests of
central Panama. Glob Ecol Biogeogr 19:376–385
Bond WJ, Midgley JJ (2001) Ecology of sprouting in woody plants: the persistence niche. Trends
Ecol Evol 16:45–51
Bond WJ, Parr CL (2010) Beyond the forest edge: ecology, diversity and conservation of the
grassy biomes. Biol Conserv 143:2395–2404
Bongers F, Poorter L, Hawthorne WD et al (2009) The intermediate disturbance hypothesis applies
to tropical forests, but disturbance contributes little to tree diversity. Ecol Lett 12:798–805
Boom A, Sinninge-Damste JS, de Leeuw JW (2005) Cutan, a common aliphatic biopolymer in
cuticles of drought-adapted plants. Org Geochem 36:595–601
Borchert R (1980) Phenology and ecophysiology of tropical trees: Erythrina poeppigiana
O.F. Cook. Ecology 61:1065–1074
Borchert R (1983) Phenology and control of flowering in tropical trees. Biotropica 15:81–89
Borchert R (1991) Growth periodicity and dormancy. In: Raghavendra AS (ed) Physiology of
trees. Wiley, New York, pp 219–242
Borchert R (1992) Computer simulation of tree growth periodicity and climatic hydroperiodicity
in tropical forests. Biotropica 24:385–395
Borchert R (1994a) Soil and stem water storage determine phenology and distribution of tropical
dry forest trees. Ecology 75:1437–1449
Borchert R (1994b) Water status and development of tropical trees during seasonal drought. Trees
8:115–125
Borchert R (1994c) Induction of rehydration and bud break by irrigation or rain in deciduous trees
of a tropical dry forest in Costa Rica. Trees 8:198–204
Borchert R (1998) Response of tropical trees to rainfall seasonality and its long-term changes.
Clim Chang 39:381–393
Borchert R (2000) Organismic and environmental controls of bud growth in tropical trees. In:
Viemont JD, Crabbe J (eds) Dormancy in plants: from whole plant behavior to cellular control.
CAB International, Wallingford, pp 87–107
Borchert R, Pockman WT (2005) Water storage capacitance and xylem tension in isolated branches
of temperate and tropical trees. Tree Physiol 25:457–466
Borchert R, Rivera G, Hagnauer W (2002) Modification of vegetative phenology in a tropical
semi-deciduous forest by abnormal drought and rain. Biotropica 34:27–39
Borchert R, Meyer SA, Felger RS et al (2004) Environmental control of flowering periodicity in
Costa Rican and Mexican tropical dry forests. Glob Ecol Biogeogr 13:409–425
Börner J, Baylis K, Corbera E et al (2016) Emerging evidence on the effectiveness of tropical for-
est conservation. PLoS One 11(11):e0159152
Botes C, Johnson SD, Cowling RM (2008) Coexistence of succulent tree aloes: partitioning of bird
pollinators by floral traits and flowering phenology. Oikos 117:875–882
Braithwaite TW, Mayhead GJ (1996) The effect of simulated browsing on the growth of sessile oak
Quercus petraea (Matt.) Lieblein. J Arboric 20:59–64
Brams EA (1971) Continuous cultivation of West African soils: organic matter diminution and
effects of applied lime and phosphorus. Plant Soil 35:401–414
Brancalion PHS, Chazdon RL (2017) Beyond hectares: four principles to guide reforestation in the
context of tropical forest and landscape restoration. Restor Ecol 25:491–496
Brearley FQ, Prajadinata S, Kidd PS et al (2004) Structure and floristics of an old secondary rain
forest in Central Kalimantan, Indonesia, and a comparison with adjacent primary forest. For
Ecol Manag 195:385–397
References 201

Brenes-Arguedas T, Coley PD, Kursar TA (2009) Pests vs. drought as determinants of plant distri-
bution along a tropical rainfall gradient. Ecology 90:1751–1761
Breshears DD (2006) The grassland-forest continuum: trends in ecosystem properties for woody
plant mosaics. Front Ecol Environ 4:96–104
Brienen RJW, Zuidema PA, Martínez-Ramos M (2010) Attaining the canopy in dry and moist
tropical forests: strong differences in tree growth trajectories reflect variation in growing condi-
tions. Oecologia 163:485–496
Brienen RJW, Phillips OL, Feldpausch TR et al (2015) Long-term decline of the Amazon carbon
sink. Nature 519:344–348
Broadbent EN, Asner GP, Keller M et al (2008) Forest fragmentation and edge effects from defor-
estation and selective logging in the Brazilian Amazon. Biol Conserv 141:1745–1757
Brodribb TJ, Holbrook NM (2003a) Changes in leaf hydraulic conductance during leaf shedding
in seasonally dry tropical forest. New Phytol 158:295–303
Brodribb TJ, Holbrook NM (2003b) Stomatal closure during leaf dehydration, correlation with
other leaf physiological traits. Plant Physiol 132:2166–2173
Brodribb TJ, Holbrook NM (2005) Leaf physiology does not predict leaf habit; examples from
tropical dry forests. Trees 19:290–295
Brodribb TJ, Holbrook NM, Edwards EJ et  al (2003) Relations between stomatal closure, leaf
turgor and xylem vulnerability in eight tropical dry forest trees. Plant Cell Environ 26:443–450
Brookes PC, Powlson DS, Jenkinson DS (1984) Phosphorus in the soil microbial biomass. Soil
Biol Biochem 16:169–175
Brooks TM, Mittermeier RA, Mittermeier CG et  al (2002) Habitat loss and extinction in the
hotspots of biodiversity. Conserv Biol 16:909–923
Brooks TM, Wright SJ, Sheil D (2009) Evaluating the success of conservation actions in safe-
guarding tropical forest biodiversity. Conserv Biol 23(6):1448–1457
Brown S, Lugo AE (1982) The storage and production of organic matter in tropical forests and
their role in the global carbon cycle. Biotropica 14:161–187
Brown NAC, van Staden J (1997) Smoke as a germination cue: a review. Plant Growth Regul
22:115–124
Bruna EM (2003) Are plant populations in fragmented habitats recruitment limited? Tests with an
Amazonian herb. Ecology 84:932–947
Bucci SJ, Goldstein G, Meinzer FC et al (2004) Functional convergence in hydraulic architecture
and water relations of tropical savanna trees: from leaf to whole plant. Tree Physiol 24:891–899
Bullock SH (1986) Climate of Chamela, Jalisco, and trends in the south coastal region of Mexico.
Arch Meteorol Geophys Bioclimatol (Ser B) 36:297–316
Bullock SH (1990) Abundance and allometrics of vines and self-supporting plants in a tropical
deciduous forest in Mexico. Biotropica 22:22–35
Bullock SH (1995) Plant reproduction in neotropical dry forest. In: Bullock SH, Mooney HA,
Medina E (eds) Seasonally dry tropical forests. Cambridge University Press, Cambridge,
pp 277–296
Bullock SH, Solís-Magallanes JA (1990) Phenology of canopy trees of a tropical deciduous forest
in México. Biotropica 22:22–35
Bullock SH, Mooney HA, Medina E (1995) Seasonally dry tropical forests. Cambridge University
Press, Cambridge, p 455
Bunge J, Fitzpatrick M (1993) Estimating the number of species: a review. J  Am Stat Assoc
88:364–373
Burgess SO, Adams MA, Turner NC et al (1998) The redistribution of soil water by tree roots
systems. Oecologia 115:306–311
Burke IC, Reiners A, Schimel DS (1989) Organic matter turnover in a sage-brush steppe land-
scape. Biogeochemistry 7:11–31
Burnham RJ (1997) Stand characteristics and leaf litter composition of a dry forest hectare in Santa
Rosa National Park, Costa Rica. Biotropica 29:387–395
202 References

Burslem DFRP, Whitmore TC (1999) Species diversity susceptibility to disturbance and tree popu-
lation dynamics in tropical rainforest. J Veg Sci 10:767–776
Cabin RJ, Weller SG, Lorence DH et al (2002) Effects of light, alien grass, and native species addi-
tions on Hawaiian dry forest restoration. Ecol Appl 12:1595–1610
Cable JM, Ogle K, Williams DG et  al (2008) Soil texture drives responses of soil respiration
to precipitation pulses in the Sonoran Desert: implications for climate change. Ecosystems
11:961–979
Cable JM, Ogle K, Barron-Gafford G et al (2013) Antecedent conditions influence soil respiration
differences in shrub and grass patches. Ecosystems 16:1230–1247
Cabrera ML (1993) Modeling the flush of nitrogen mineralization caused by drying and rewetting
soils. Soil Sci Soc Am J 57:63–66
Calder EA (1957) Features of nitrate accumulation in Uganda soil. J Soil Sci 8:60–72
Calvo-Alvarado A, McLennan B, Sanchez-Azofeifa A et al (2009) Deforestation and forest resto-
ration in Guanacaste, Costa Rica: putting conservation policies in context. J For Ecol Manag
258:931–940
Cameron RS, Posner AM (1979) Mineralisable organic nitrogen in soil fractionated according to
particle size. Eur J Soil Sci 30:565–577
Campo J (2016) Shift from ecosystem P to N limitation at precipitation gradient in tropical dry
forests at Yucatan, Mexico. Environ Res Lett 11:095006
Campo J, Dirzo R (2003) Leaf quality and herbivory responses to soil nutrient addition in second-
ary tropical dry forests of Yucatán, Mexico. J Trop Ecol 19:525–530
Campo J, Merino A (2016) Variations in soil carbon sequestration and their determinants along a
precipitation gradient in seasonally dry tropical forests. Glob Chang Biol 22:1942–1956
Campo J, Vazquez-Yanes C (2004) Effects of nutrient limitation on aboveground carbon dynamics
during tropical dry forest regeneration in Yucatan, Mexico. Ecosystems 7:311–319
Campo J, Jaramillo VJ, Maass JM (1998) Pulses of soil phosphorus availability in a tropical dry
forest: effects of seasonality and level of wetting. Oecologia 115:167–172
Campo J, Maas JM, Jaramillo VJ et al (2001) Phosphorus cycling in a Mexican tropical dry forest
ecosystem. Biogeochemistry 53:161–179
Canham CD, Finzi AC, Pacala SW et al (1994) Causes and consequences of resource heteroge-
neity in forests: interspecific variation in light transmission by canopy trees. Can J  For Res
24:337–349
Cao S, Yu Q, Sanchez-Azofeifa A et al (2015) Mapping tropical dry forest succession using mul-
tiple criteria spectral mixture analysis. ISPRS J Photogramm Remote Sens 109:17–29
Carbone MS, Still CJ, Ambrose AR et al (2011) Seasonal and episodic moisture controls on plant
and microbial contributions to soil respiration. Oecologia 167:265–278
Castillo-Campos G, Halffter G, Moreno CE (2008) Primary and secondary vegetation patches as
contributors to floristic diversity in a tropical deciduous forest landscape. Biodivers Conserv
17:1701–1714
Cavelier J, Wright J, Santamaría J (1999) Effects of irrigation on litterfall, fine root biomass and
production in a semideciduous lowland forest in Panama. Plant Soil 211:207–213
Ceballos G (1995) Vertebrate diversity, ecology and conservation in neotropical dry forests.
In: Bullock SH, Mooney HA, Medina E (eds) Seasonally dry tropical forests. Cambridge
University Press, Cambridge, pp 195–220
Ceccon E, Omstead I, Vázquez-Yanes C et al (2002) Vegetation and soil properties in two tropical
dry forests of differing regeneration status in Yucatán. Agrociencia 36:621–631
Ceccon E, Huante P, Campo-Alves J (2003) Effects of nitrogen and phosphorus fertilization on the
survival and recruitment of seedlings of dominant tree species in two abandoned tropical dry
forests in Yucatán, Mexico. J For Ecol Manag 182:387–402
Ceccon E, Sanchéz S, Campo-Alves J (2004) Tree seedling dynamics in two abandoned tropical
dry forests of differing successional status in Yucatán, Mexico: a field experiment with N and
P fertilization. Plant Ecol 170:12–26
References 203

Ceulemans R, Mousseau M (1994) Effects of elevated CO2 on woody plants. New Phytol
127:425–446
Chadwick R, Good P, Martin G et  al (2015) Large rainfall changes consistently projected over
substantial areas of tropical land. Nat Clim Chang 6:177–181
Champion HG, Seth SK (1968) A revised survey of the forest types of India. Government of India
Publication, New Delhi
Chapin FS III (1980) The mineral nutrition of wild plants. Annu Rev Ecol Syst 11:233–260
Chapin FS, Kedrowski RA (1983) Seasonal changes in nitrogen and phosphorus fractions and
autumn retranslocation in evergreen and deciduous taiga trees. Ecology 64:376–391
Chapman CA, Chapman LJ (1997) Forest regeneration in logged and unlogged forests of Kibale
National Park, Uganda. Biotropica 29:396–412
Chapotin SM, Razanameharizaka JH, Holbrook NM (2006) Baobab trees (Adansonia) in
Madagascar use stored water to flush new leaves but not to support stomatal opening before the
rainy season. New Phytol 169:549–559
Chaturvedi RK (2010) Plant functional traits in dry deciduous forests of India. Ph.D. thesis,
Department of Botany, Banaras Hindu University, Varanasi, India
Chaturvedi RK, Raghubanshi AS (2014) Species composition, distribution and diversity of woody
species in tropical dry forest of India. J Sustain For 33:729–756
Chaturvedi RK, Raghubanshi AS (2015) Assessment of carbon density and accumulation in mono-
and multi-specific stands in tropical dry forests of India. For Ecol Manag 339:11–21
Chaturvedi OP, Singh JS (1987) The structure and function of pine forest in Central Himalaya
II. Nutrient dynamics. Ann Bot 60:253–267
Chaturvedi RK, Singh JS (2017) Restoration of mine spoil in a dry tropical region: a review. Proc
Indian Natl Sci Acad. https://doi.org/10.16943/ptinsa/2017/49123
Chaturvedi RK, Raghubanshi AS, Singh JS (2011a) Effect of small scale variations in environ-
mental factors on the distribution of woody species in tropical deciduous forests of Vindhyan
Highlands. India J Bot. https://doi.org/10.1155/2011/297097
Chaturvedi RK, Raghubanshi AS, Singh JS (2011b) Carbon density and accumulation in woody
species of tropical dry forest in India. For Ecol Manag 262:1576–1588
Chaturvedi RK, Raghubanshi AS, Singh JS (2011c) Leaf attributes and tree growth in a tropical
dry forest. J Veg Sci 22:917–931
Chaturvedi RK, Raghubanshi AS, Singh JS (2011d) Plant functional traits with particular refer-
ence to dry deciduous forests: a review. J Biosci 36:963–981
Chaturvedi RK, Raghubanshi AS, Singh JS (2012) Effect of grazing and harvesting on diversity,
recruitment and carbon accumulation of juvenile trees in tropical dry forests. For Ecol Manag
284:152–162
Chaturvedi RK, Raghubanshi AS, Singh JS (2013) Growth of tree seedlings in a dry tropical forest
in relation to soil moisture and leaf traits. J Plant Ecol 6:158–170
Chaturvedi RK, Raghubanshi AS, Singh JS (2014) Relative effects of different leaf attributes on
sapling growth in tropical dry forest. J Plant Ecol 7:544–558
Chaturvedi RK, Raghubanshi AS, Singh JS (2017a) Sapling harvest: a predominant factor affect-
ing future composition of tropical dry forests. For Ecol Manag 384:221–235
Chaturvedi RK, Raghubanshi AS, Tomlinson KW et al (2017b) Impacts of human disturbance in
tropical dry forests increase with soil moisture stress. J Veg Sci 28:997–1007
Chaves L, De F, De C et al (1995) Growth of Jacaranda da Bahia (Dalbergia nigra) seedlings in
response to inoculation with vesicular arbuscular mycorrhizal fungi at different soil phospho-
rus level. Rev Árvore 19:32–49
Chazdon RL (1986) Light variation and carbon gain in rain forest understorey palms. J  Ecol
74:995–1012
Chazdon RL, Denslow J  (2002) Floristic composition and species richness. In: Chazdon RL,
Whitmore TC (eds) Foundations of tropical forest biology, classic papers with commentaries.
University of Chicago Press, Chicago, pp 513–522
204 References

Chazdon RL, Carega S, Webb C et al (2003) Community and phylogenetic structures of reproduc-
tive traits of woody species in wet tropical forests. Ecol Monogr 73:331–348
Chidumayo EN (1990) Above-ground woody biomass structure and productivity in a Zambian
woodland. For Ecol Manag 36:33–46
Chidumayo E, Marunda C (2010) Dry forests and woodlands in Sub-Saharan Africa: contexts and
challenges. In: Chidumayo E, Gumbo D (eds) The dry forests and Woodlands of Africa: man-
aging for products and services. Earthscan, London, pp 1–9
Chittibabu CV, Parthasarathy N (2000) Attenuated tree species diversity in human-impacted tropi-
cal evergreen forest sites at Kolli hills, Eastern Ghats, India. Biodivers Conserv 9:1493–1519
Choat B, Ball MC, Luky JG et al (2003) Pit membrane porosity and water stress-induced cavitation
in four co-existing dry rainforest tree species. Plant Physiol 131:41–48
Choat B, Ball MC, Luky JG et al (2005) Hydraulic architecture of deciduous and evergreen dry
rainforest tree species from northeastern Australia. Trees 19:305–311
Chopra K (1993) The value of non-timber forest products: an estimation for tropical deciduous
forests in India. Econ Bot 47:251–257
Clarholm M (1985) Interactions of bacteria, protozoa and plants leading to mineralization of soil
nitrogen. Soil Biol Biochem 17:181–187
Clark DA, Piper SC, Keeling CD et al (2003) Tropical rain forest tree growth and atmospheric car-
bon dynamics linked to interannual temperature variation during 1984–2000. Proc Natl Acad
Sci USA 100:5852–5857
Cleveland CC, Townsend AR, Taylor P et al (2011) Relationships among net primary productiv-
ity, nutrients and climate in tropical rain forest: a pan-tropical analysis. Ecol Lett 14:939–947
Cole PG (2003) Environmental constraints on the distribution of the non-native invasive grass,
Microstegium vimineum. Ph.D. thesis, Department of Ecology and Evolutionary Biology,
University of Tennessee Knoxville, TN
Coleman DC, Cole CV, Anderson RV et al (1977) An analysis of rhizosphere-saprophage interac-
tions in terrestrial ecosystems. In: Loom U, Persson T (eds) Soil organisms as components of
ecosystems, No. 25. Ecological Bulletin, Stockholm, pp 299–309
Coleman DC, Crossley DA, Beare MH et al (1988) Interaction at root/soil and litter/soil interfaces
in terrestrial ecosystems: II. Interaction between invertebrates and organisms associated with
plant rhizospheres. Agric Ecosyst Environ 24:117–134
Coleman JS, McConnaughay KDM, Bazzaz FA (1993) Elevated CO2 and plant nitrogen use: is
reduced tissue nitrogen concentration size-dependent? Oecologia 93:195–200
Coley PD (1988) Effects of plant growth rate and leaf lifetime on the amount and type of anti-
herbivore defense. Oecologia 74:531–536
Coley PD (1998) Possible effects of climate change on plant/herbivore interactions in moist tropi-
cal forests. Clim Chang 39:455–472
Coley PD, Barone JA (1996) Herbivory and plant defenses in tropical forest. Annu Rev Ecol Syst
27:305–335
Collen B, Ram M, Zamin T et al (2008) The tropical biodiversity data gap: addressing disparity in
global monitoring. Trop Conserv Sci 1(2):75–88
Colwell RK, Coddington JA (1994) Estimating terrestrial biodiversity through extrapolation. Phil
Trans R Soc Lond B Biol Sci 345:101–118
Condit R, Hubbell SP, La Frankie JV et al (1996) Species–area and species–individual relation-
ships for tropical trees: a comparison of three 50-ha plots. J Ecol 84:549–562
Condit RS, Ashton PS, Baker P et al (2000a) Spatial patterns in the distribution of tropical tree
species. Science 288:1414–1418
Condit R, Watts K, Bohlman SA et al (2000b) Quantifying the deciduousness of tropical forest
canopies under varying climates. J Veg Sci 11:649–658
Conroy JP, Milham PJ, Barlow EWR (1992) Effect of nitrogen and phosphorus availability on the
growth response of Eucalyptus grandis. Plant Cell Environ 15:843–847
Cooke R, Ranere AJ (1992) Prehistoric human adaptations to the seasonally dry forests of Panama.
World Archaeol 24:114–133
References 205

Cordell S, Goldstein G, Meinzer FC et al (2001) Regulation of leaf life-span and nutrient-use effi-
ciency of Metrosideros polymorpha trees at two extremes of a long chronosequence in Hawaii.
Oecologia 127:198–206
Cornejo FH, Varela A, Wright SJ (1994) Tropical forest litter decomposition under seasonal
drought: nutrient release, fungi, and bacteria. Oikos 70:183–190
Cornelissen JHC, Pérez-Harguindeguy N, Díaz S et al (1999a) Leaf structure and defence control
litter decomposition rate across species and life forms in regional flora on two continents. New
Phytol 143:191–200
Cornelissen JHC, Carnelli AL, Callaghan TV (1999b) Generalities in the growth, allocation and
leaf quality responses to elevated CO2 in eight woody species. New Phytol 14:401–409
Cornelissen JHC, Lavorel S, Garnier E et al (2003) A handbook of protocols for standardized and
easy measurement of plant functional traits worldwide. Aust J Bot 51:335–380
Coughenour MB, Detling JK (1986) Acacia tortilis seed germination responses to water potential
and nutrients. Afr J Ecol 24:203–205
Coûteaux MM, Bottner P, Berg B (1995) Litter decomposition, climate and litter quality. Trends
Ecol Evol 10:63–66
Criquet S, Ferre E, Farnet AM et al (2004) Annual dynamics of phosphatase activities in an ever-
green oak litter: influence of biotic and abiotic factors. Soil Biol Biochem 36:111–1118
Crowley BE, McGoogan KC, Lehman SM (2012) Edge effects on foliar stable isotope values in a
Madagascan tropical dry forest. PLoS One 7:1–9
Cubiña A, Aide TM (2001) The effect of distance from forest edge on seed rain and soil seed bank
in a tropical pasture. Biotropica 33:260–267
Cuevas RM, Hidalgo C, Payan F et  al (2013) Precipitation influences on active fractions of
soil organic matter in karstic soils of Yucatan: regional and seasonal patterns. Eur J For Res
132:667–677
Cunningham WP, Saigo BW (1999) Environmental sciences: a global concern. The McGraw-Hill
Companies, Boston
Cunningham SA, Summerhayes B, Westoby M (1999) Evolutionary divergences in leaf structure
and chemistry, comparing rainfall and soil nutrient gradients. Ecol Monogr 69:569–588
Curran LM, Leighton M (2000) Vertebrate responses to spatiotemporal variation in seed produc-
tion of mast-fruiting dipterocarpaceae. Ecol Monogr 70:101–128
Curran LM, Caniago I, Paoli GD et al (1999) Impact of El Niño and logging on canopy tree recruit-
ment in Borneo. Science 286:2184–2188
Curt T, Borgniet L, Ibanez T et al (2015) Understanding fire patterns and fire drivers for setting a
sustainable management policy of the New-Caledonian biodiversity hotspot. For Ecol Manag
337:48–60
Curtis PS, Vogel CS, Wang X et al (2000) Gas exchange, leaf nitrogen, and growth efficiency of
Populus tremuloides in a CO2-enriched atmosphere. Ecol Appl 10:3–17
D’Antonio CM, Vitousek PM (1992) Biological invasions by exotic grasses, the grass/fire cycle,
and global change. Annu Rev Ecol Syst 23:63–87
da Silva JMC, Uhl C, Murray G (1996) Plant succession, landscape management, and the ecology
of frugivorous birds in abandoned Amazonian pastures. Conserv Biol 10:491–503
Dai AG (2013) Increasing drought under global warming in observations and models. Nat Clim
Chang 3:52–58
Daiya KS, Sharma HK, Chawan DD et al (1980) Effect of salt solutions of different osmotic poten-
tial on seed germination and seedling growth in some Cassia species. Folia Geobot Phytotaxon
5:149–153
Dash MC, Guru CB (1980) Distribution and seasonal variation in numbers of testacea (protozoa)
in some Indian soils. Pedobiologia 20:325–342
Daubenmire R (1972) Phenology and other characteristics of tropical semi-deciduous forest in
northeastern Costa Rica. J Ecol 60:147–170
Davidson EA, Matson PA, Vitousek PM et al (1993) Processes regulating soil emissions of NO and
N2O in a seasonally dry tropical forest. Ecology 74:130–139
206 References

de Deyn GB, Cornelissen JHC, Bardgett RD (2008) Plant functional traits and soil carbon seques-
tration in contrasting biomes. Ecol Lett 11:516–531
De Queiroz LP (2006) The Brazilian Caatinga: phytogeographical patterns inferred from distri-
bution data of the Leguminosae. In: Pennington RT, Ratter JA, Lewis GP (eds) Neotropical
savannas and seasonally dry forests: plant biodiversity, biogeography and conservation. CRC
Press, Boca Raton, pp 121–158
de Rouw A (1993) Regeneration by sprouting in slash and burn rice cultivation, Taϊ Rain Forest,
Côte d`Ivoire. J Trop Ecol 9:387–408
De Souza JP, Aráujo GM, Haridasan M (2007) Influence of soil fertility on the distribution of tree
species in a deciduous forest in the Trîangulo Mineiro region of Brazil. Plant Ecol 191:253–263
Dechnik-Vázquez YA, Meave JA, Pérez-García EA et al (2016) The effect of treefall gaps on the
understorey structure and composition of the tropical dry forest of Nizanda, Oaxaca, Mexico:
implications for forest regeneration. J Trop Ecol 32:89–106
DeFries R, Hansen A, Newton AC, Hansen MC (2004) Increasing isolation of protected areas in
tropical forests over the past twenty years. Ecol Appl 15:19–26
DeFries R, Houghton RA, Hansen M, Field C, Skole DL, Townshend J (2002) Carbon emissions-
from tropical deforestation and regrowth based on satellite observations for the 1980s and 90s.
Proc Natl Acad Sci U S A 99:14256–14261
Del Arco JM, Escudero A, Garrido MV (1991) Effects of site characteristics on nitrogen retranslo-
cation from senescing leaves. Ecology 72:701–708
Demars BG, Boerner REJ (1997) Foliar nutrients dynamics and resorption in naturalized Lonicera
maackii (Caprifoliaceae) populations in Ohio, USA. Am J Bot 84:112–117
Denevan WM (1992) The pristine myth: the landscape of the Americas in 1492. Ann Assoc Am
Geogr 82:369–385
Denslow JS (1985) Disturbance-mediated coexistence of species. In: Pickett STA, White PS (eds)
The ecology of natural disturbance and patch dynamics. Academic, San Diego, pp 307–323
Denslow JS (1987) Tropical rain forest gaps and tree species diversity. Annu Rev Ecol Syst
18:431–451
Deol GS, Khosla PK (1983) Provenance related growth response of P. ciliata Wall. Ex Royle to N
fertilization. Ind For 109:30–40
Detwiler RP, Hall CAS (1988) Tropical forests and the global carbon cycle. Science 239:42–47
Devisscher T, Malhi Y, Landívar VDR et  al (2016) Understanding ecological transitions under
recurrent wildfire: a case study in the seasonally dry tropical forests of the Chiquitania, Bolivia.
For Ecol Manag 360:273–286
Díaz S, Cabido M (1997) Plant functional types and ecosystem function in relation to global
change. J Veg Sci 8:463–474
Díaz S, Symstad AJ, Chapin FS et al (2003) Functional diversity revealed by removal experiments.
Trends Ecol Evol 181:40–46
Díaz S, Hodgson JG, Thompson K et al (2004) The plant traits that drive ecosystems: evidence
from three continents. J Veg Sci 15:295–304
Dickinson MB, Hermann SM, Whigham DF (2001) Low rates of background canopy-gap distur-
bance in a seasonally dry forest in the Yucatan Peninsula with a history of fires and hurricanes.
J Trop Ecol 17:895–902
Dirzo R, Dominguez CA (1995) Plant herbivore interactions in Mesoamerican tropical dry for-
est. In: Bullock SH, Mooney HA, Medina E (eds) Seasonally dry tropical forests. Cambridge
University Press, Cambridge, pp 304–325
Dirzo R, Young HS, Mooney HA et al (2011) Introduction. In: Dirzo R, Young HS, Mooney HA
et al (eds) Seasonally dry tropical forests: ecology and conservation. Island Press, Washington,
DC, pp xi–xiii
Dittus WP (1985) The influence of cyclones on the dry evergreen forest of Sri Lanka. Biotropica
17:1–14
References 207

Dominguez F, Kumar P, Vivoni ER (2008) Precipitation recycling variability and ecoclimatologi-


cal stability – a study using NARR data. Part II: North American Monsoon region. J Climatol
21:5187–5203
Dommergues YR (1966) Biologie du sol. Presses Universitaire de France, Paris
Doughty CE, Malhi Y, Araujo-Murakami A et  al (2014) Allocation trade-offs dominate the
response of tropical forest growth to seasonal and interannual drought. Ecology 95:2192–2201
Drake BG, Gonzalez-Meler MA, Long SP (1997) More efficient plants: a consequence of rising
atmospheric CO2? Annu Rev Plant Physiol Plant Mol Biol 48:609–639
Drenovsky RE, Richards JH (2006) Low leaf N and P resorption contributes to nutrient limitation
in two desert shrubs. Plant Ecol 183:305–314
Drew MC, Saker LR (1975) Nutrient supply and the growth of the seminal root system in barley,
II. Localized compensatory increases in lateral root growth and rates of nitrate uptake when
nitrate supply is restricted to only part of the root system. J Exp Bot 29:435–451
Drexhage M, Colin F (2003) Effects of browsing on shoots and roots of naturally regenerated ses-
sile oak seedlings. Ann For Sci 60:173–178
DRYFLOR (2016) Plant diversity patterns in neotropical dry forests and their conservation impli-
cations. Science 353:1383–1387
Duffy PB, Brando P, Asner GP et al (2015) Projections of future meteorological drought and wet
periods in the Amazon. Proc Natl Acad Sci USA 112:13172–13177
Dufour-Dror JM (2007) Influence of cattle grazing on the density of oak seedlings and saplings in
a Tabor oak forest in Israel. Acta Oecol 31:223–228
Dunphy BK, Murphy PG, Lugo AE (2000) The tendency for trees to be multiple-stemmed in tropi-
cal and subtropical dry forests: studies of Guanica forest, Puerto Rico. Trop Ecol 41:161–167
Durán E (2004) Estructura, diversidad y mortalidad del componente arbóreo en un mosaico ambi-
ental de Chamela, México. Ph.D. dissertation, Universidad Nacional Autónoma de México,
México, DF. p 135
Durán E, Balvanera P, Lott E et al (2002) Composición, estructura y dinámica de la vegetación.
In: Noguera FA, Quesada M, Vega JH et  al (eds) Historia Natural de Chamela. Instituto de
Biología, UNAM, México, pp 443–472
Duru M, Khaled RAH, Ducourtieux C et al (2009) Do plant functional types based on leaf dry
matter content allow characterizing native grass species and grasslands for herbage growth
pattern? Plant Ecol 201:421–433
Dutt AK, Pathania U (1986) Effect of nitrogen and phosphorus on seedling growth and nodulation
in Leucaena leucocephala. Leucaena Res Rep 7:38–41
Dyckmans J, Flessa H (2005) Partitioning of remobilised N in young beech (Fagus sylvatica L.) is
not affected by elevated [CO2]. Ann For Sci 62:285–288
Eamus D (1999) Ecophysiological traits of deciduous and evergreen woody species in the season-
ally dry tropics. Trends Ecol Evol 14:11–16
Eamus D, Jarvis PG (1989) The direct effects of increase in the global atmospheric CO2 concentra-
tion on natural and commercial temperate trees and forests. Adv Ecol Res 19:1–55
Eamus D, Prichard H (1998) A cost-benefit analysis of leaves of four Australian savanna species.
Tree Physiol 18:537–545
Eamus D, Prior L (2001) Ecophysiology of trees of seasonally dry tropics: comparisons among
phenologies. Adv Ecol Res 32:113–197
Eamus D, Myers B, Duff G et al (1999) Seasonal changes in photosynthesis of eight savanna tree
species. Tree Physiol 19:665–671
Eckstein RL, Karlsson PS, Weih M (1999) Leaf lifespan and nutrient resorption as determinants of
plant nutrient conservation in temperate–arctic regions. New Phytol 143:177–189
Edwards PJ (1982) Studies of mineral cycling in a montane rain forest in New Guinea. V. Rates of
cycling in throughfall and litter fall. J Ecol 70:807–827
Ehrlich PR, Wilson EO (1991) Biodiversity studies: science and policy. Science 253:758–762
208 References

Elser JJ, Bracken MES, Gruner DS et  al (2007) Global analysis of nitrogen and phosphorus
limitation of primary producers in freshwater, marine and terrestrial ecosystems. Ecol Lett
10:1135–1142
Elzinga JA, Atlan A, Biere A et al (2007) Time after time: flowering phenology and biotic interac-
tions. Trends Ecol Evol 22:432–439
Engelbrecht BMJ, Comita LS, Condit RS et al (2007) Drought sensitivity shapes species distribu-
tion patterns in tropical forests. Nature 447:80–82
Enquist BJ, Leffler AJ (2001) Long-term tree ring chronologies from sympatric tropical dry-forest
trees: individualistic responses to climatic variation. J Trop Ecol 17:41–60
Eriksson I, Teketay D, Granstrom A (2003) Response of plant communities to fire in an Acacia
woodland and a dry Afromontane forest, southern Ethiopia. For Ecol Manag 177:39–50
Escudero A, Del Arco JM, Sanz IC et  al (1992) Effects of leaf longevity and retranslocation
efficiency on the retention time of nutrients in the leaf biomass of different woody species.
Oecologia 90:80–87
Ewel J (1980) Tropical succession: manifold routes to maturity. Biotropica 12:2–7
Ewers RM, Thorpe S, Didham RK (2007) Synergistic interactions between edge and area effects
in a heavily fragmented landscape. Ecology 88:96–106
Fahey TJ, Battles JJ, Wilson GF (1998) Responses of early successional northern hardwood forests
to changes in nutrient availability. Ecol Monogr 68:183–212
Fajardo A, Siefert A (2016) Phenological variation of leaf functional traits within species.
Oecologia 180:951–959
Falk DA, Miller C, McKenzie D et  al (2007) Cross-scale analysis of fire regimes. Ecosystems
10:809–823
Fallas-Cedeño L, Holbrook NM, Rocha OJ et al (2010) Phenology, lignotubers, and water rela-
tions of Cochlospermum vitifolium, a pioneer tropical dry forest tree in Costa Rica. Biotropica
42:104–111
FAO (2000) Management of natural forests of dry tropical zones. In: FAO conservation guide
no. 32. Food and Agriculture Organization of the United Nations, Rome, p 9. Prepared by
Bellefontaine R, Gaston A, Petrucci Y. Available at: http://www.fao.org/docrep/005/w4442e/
w4442e00.htm
FAO (2001) Global forest resources assessment 2000. In: FAO forestry paper no. 140. Food and
Agriculture Organization of the United Nations, Rome. Available at: http://www.fao.org/for-
estry/fo/fra/index.jsp
FAO (2008) Forests and energy – key issues. In: FAO forestry paper no. 154. Food and Agriculture
Organization of the United Nations, Rome. Available at: www.fao.org/docrep/010/i0139e/
i0139e00.htm
FAO, JRC (2012) Global forest land-use change 1990–2005. In: Lindquist EJ, D’Annunzio R,
Gerrand A et al (eds) FAO forestry paper no. 169. Food and Agriculture Organization of the
United Nations and European Commission Joint Research Centre, Rome. Available at: http://
www.fao.org/docrep/017/i3110e/i3110e.pdf
February EC, Higgins SI (2010) The distribution of tree and grass roots in savannas in relation to
soil nitrogen and water. S Afr J Bot 76:517–523
February EC, Cook GD, Richards AE (2013) Root dynamics influence tree–grass coexistence in
an Australian savanna. Aust Ecol 38:66–75
Feeley KJ, Silman MR (2011) The data void in modeling current and future distributions of tropi-
cal species. Glob Chang Biol 17:626–630
Feeley KJ, Wright SJ, Nur Supardi MN et al (2007) Decelerating growth in tropical forest trees.
Ecol Lett 10:461–469
Feng X, Simpson AJ, Wilson KP et al (2008) Increased cuticular carbon sequestration and lignin
oxidation in response to soil warming. Nat Geosci 1:836–839
Feng X, Porporato A, Rodriguez-Iturbe I (2013) Changes in rainfall seasonality in the tropics. Nat
Clim Chang 3:811–815
References 209

Fernandez C, Acosta FJ, Abella G et al (2002) Complex edge effect fields as additive processes in
patches of ecological systems. Ecol Model 149:273–283
Ferraro PJ (2009) Counterfactual thinking and impact evaluation in environmental policy. N Dir
Eval 122:75–84
Ferraro PJ (2011) The future of payments for environmental services. Conserv Biol 25(6):1134–1138
Fetcher N, Strain BR, Oberbauer SF (1983) Effects of light regime on the growth, leaf morphol-
ogy, and water relations of seedlings of two species of tropical trees. Oecologia 58:314–319
Field CB, Behrnfeld MJ, Randerson JT et al (1998) Primary production of the biosphere: integrat-
ing terrestrial and oceanic components. Science 281:237–240
Fierer N, Schimel JP (2003) A proposed mechanism for the pulse in carbon dioxide production
commonly observed following the rapid rewetting of a dry soil. Soil Sci Soc Am J 67:798–805
Figueiredo IB (2002) Padrões de polinizacxão e dispersão de sementes de espécies arbóreas
de floresta estacional decidual, Brasil Central. Bachelor Monograph. UNESP, Instituto de
Biociências, Rio Claro
Finegan B, Delgado D (2000) Structural and floristic heterogeneity in a 30-year-old Costa Rican
rain forest restored, on pasture through natural secondary succession. Restor Ecol 8:380–393
Finegan B, Pẽna-Claros M, de Oliveira A et al (2015) Does functional trait diversity predict above-
ground biomass and productivity of tropical forests? Testing three alternative hypotheses.
J Ecol 103:191–201
Fioretto A, Nardo C, Di Papa S et al (2005) Lignin and cellulose degradation and nitrogen dynam-
ics during decomposition of three leaf litter species in a Mediterranean ecosystem. Soil Biol
Biochem 37:1083–1091
Fischer AG (1960) Latitudinal variations in organic diversity. Evolution 14:64–81
Fitter AH (1985) Functional significance of root morphology and root system architecture. In:
Fitter AH, Atkinson D, Read DJ et al (eds) Ecological interactions in soil: plants, microbes and
animals, British ecological society, special, publ. 4. Blackwell Scientific Publications, Oxford,
pp 87–106
Fletcher RJ (2005) Multiple edge effects and their implications in fragmented landscapes. J Anim
Ecol 74:342–352
Focht DD, Verstraete W (1977) Biochemical ecology of nitrification and denitrification. Adv
Microb Ecol 1:135–214
Fog K (1988) The effect of added nitrogen on the rate of decomposition of organic matter. Biol
Rev 63:433–462
Fogel R, Hunt G (1983) Contribution of mycorrhizae and soil fungi to nutrient cycling in a Douglas
fir ecosystem. Can J For Res 13:219–232
Foley JA (1995) An equilibrium model of the terrestrial carbon budget. Tellus 47B:310–319
Forman RTT (1995) Land mosaics: the ecology of landscapes and regions. Cambridge University
Press, Cambridge
Foster SA (1986) On the adaptive value of large seeds for tropical moist forest trees: a review and
synthesis. Bot Rev 52:260–299
Foster RC (1988) Microenvironments of soil micro-organisms. Biol Fertil Soils 6:189–203
Foster SA, Janson CH (1985) The relationship between seed size and establishment conditions in
tropical woody plants. Ecology 66:773–780
Franco AC, Bustamante M, Caldas LS et al (2005) Leaf functional traits of Neotropical savanna
trees in relation to seasonal water deficit. Trees Struct Funct 19:326–335
Frankie GW, Baker HG, Opler PA (1974) Comparative phenological studies of trees in tropical wet
and dry forests in the lowlands of Costa Rica. J Ecol 62:881–899
Franks PJ, Gibson A, Bachelard EP (1995) Xylem permeability and embolism susceptibility in
seedlings of Eucalyptus camaldulensis dehnh from 2 different climatic zones. Aust J  Plant
Physiol 22:15–21
Freitas ADS, Sampaio EVSB, Santos CERS et  al (2010) Biological nitrogen fixation in tree
legumes of the Brazilian semi-arid caatinga. J Arid Environ 74:344–349
210 References

Furley PA, Proctor J, Ratter JA (1992) Nature and dynamics of forest-savanna boundaries.
Chapman and Hall, London
Galicia L, Zarco-Arista AE, Mendoza-Robles KI et al (2008) Land use/cover, landforms and frag-
mentation patterns in a tropical dry forest in the southern Pacific region of Mexico. Singap
J Trop Geogr 29:137–154
Gamboa AM, Hidalgo C, de Leon F et al (2010) Nutrient addition differentially affects soil carbon
sequestration in secondary tropical dry forests: early versus late-succession stages. Restor Ecol
18:252–260
Ganzhorn JU, Ganzhorn AW, Abraham J-P et al (1990) The impact of selective logging on forest
structure and tenrec populations in western Madagascar. Oecologia 84:126–133
Garwood NC (1983) Seed germination in a seasonal tropical forest in Panama: a community study.
Ecol Monogr 53:159–181
Gasparri NI, Grau HR (2009) Deforestation and fragmentation of Chaco dry forest in NW
Argentina (1972–2007). For Ecol Manag 258:913–921
Geber MA, Griffen LR (2003) Inheritance and natural selection on functional traits. Int J Plant
Sci 164:21–42
Gei MG, Powers JS (2014) Nutrient cycling in tropical dry forests. In: Sánchez-Azofeifa A, Powers
JS, Fernandes GW et al (eds) Tropical dry forests in the Americas: ecology, conservation, and
management. CRC Press, Boca Raton, pp 141–155
Geldmann J, Barnes M, Coad L et al (2013) Effectiveness of terrestrial protected areas in reducing
habitat loss and population declines. Biol Conserv 161:230–238
Gentry AH (1974) Flowering phenology and diversity in tropical bignoniaceae. Biotropica 6:64–68
Gentry AH (1982) Neotropical floristic diversity: phytographical connections between Central and
South America, Pleistocene climatic fluctuations, or an accident of the Andean orogeny? Ann
Mo Bot Gard 69:557–593
Gentry AH (1988) Changes in plant community diversity and floristic composition on environmen-
tal and geographical gradients. Ann Mo Bot Gard 75:1–34
Gentry AH (1995) Diversity and floristic composition of neotropical dry forests. In: Bullock SH,
Mooney HA, Medina E (eds) Seasonally dry tropical forests. Cambridge University Press,
New York, pp 146–194
Gerhardt K (1993) Tree seedling development in tropical dry abandoned pasture and secondary
forest in Costa Rica. J Veg Sci 4:95–102
Germer S, Neill C, Krusche AV et al (2010) Influence of land-use change on near-surface hydro-
logical processes: undisturbed forest to pasture. J Hydrol 380:473–480
Gholz HL, Wedin DA, Smitherman SM et al (2000) Long-term dynamics of pine and hardwood
litter in contrasting environments: toward a global model of decomposition. Glob Chang Biol
6:751–765
Ghosh RC, Mathur NK, Singh RP (1976) Diospyros melanoxylon: its problems and cultivation.
Ind For 102:326–336
Gibson DJ (1988) The maintenance of plant and soil heterogeneity in dune grassland. J  Ecol
76:497–508
Gillespie TW (1999) Life history characteristics and rarity of woody plants in tropical dry forest
fragments of Central America. J Trop Ecol 15:637–649
Gillespie TW (2006) Diversity, biogeography and conservation of woody plants in tropical dry
forest of south Florida. In: Pennington RT, Ratter JA, Lewis GP (eds) Neotropical savannas and
seasonally dry forests: plant biodiversity, biogeography and conservation. CRC Press, Boca
Raton, pp 383–394
Gillespie TW, Grijalva A, Farris CN (2000) Diversity, composition, and structure of tropical dry
forests in Central America. Plant Ecol 147:37–47
Gillet F, Murisier B, Buttler A et al (1999) Influence of tree cover on the diversity of herbaceous
communities in subalpine wooded pastures. Appl Veg Sci 2:47–54
GISP (2003) The invasive alien species (IAS) problem. The Global Invasive Species Programme
(GISP). Available from URL: http://www.gisp.org/about/IAS.asp. Accessed 12 Feb 2005
References 211

Gitay H, Noble IR (1997) What are functional types and how should we seek them? In: Smith TM,
Shugart HH, Woodward FI (eds) Plant functional types: their relevance to ecosystem properties
and global change. Cambridge University Press, Cambridge, pp 3–19
Gitay H, Noble IR, Connell JH (1999) Deriving functional types for rainforest trees. J Veg Sci
10:641–650
Givnish TJ (2002) Adaptive significance of evergreen vs. deciduous leaves: solving the triple para-
dox. Silva Fenn 36:703–743
Godoy O, Levine JM (2013) Phenology effects on invasion success: insights from coupling field
experiments to coexistence theory. Ecology 95:726–736
Goldammer JG (ed) (1992) Tropical forests in transition: ecology of natural and anthropogenic
disturbance processes. Birhauser-Verlag, Basel
Goldammer JG (1993) Fire management. In: Pancel L (ed) Tropical forest handbook, vol 2.
Springer-Verlag, Berlin, pp 1221–1265
Goma-Tchimbakala J  (2009) Carbon and nitrogen storage in soil aggregates from differ-
ent Terminalia superba age plantations and natural forest in Kouilou, Congo. Int J  Soil Sci
4:104–113
Gond V, Fayolle A, Pennec A et al (2013) Vegetation structure and greenness in Central Africa
from Modis multi-temporal data. Phil Trans R Soc B Biol Sci 368:20120309
González A, Loreau M (2009) The causes and consequences of compensatory dynamics in eco-
logical communities. Annu Rev Ecol Evol Syst 40:393–414
González OJ, Zak DR (1996) Composition and structure of topical dry forest of St. Lucia, West
Indies: the influence of edaphic properties and disturbance. Biotropica 28:618–626
Gordon JE, Hawthorne WD, Reyes-García A et al (2004) Assessing landscapes: a case study of
tree and shrub diversity in the seasonally dry tropical forests of Oaxaca, Mexico and southern
Honduras. Biol Conserv 117:429–442
Gotsch SG, Geiger EL, Franco AC et al (2010) Allocation to leaf area and sapwood area affects
water relations of co-occurring savanna and forest trees. Oecologia 163:291–301
Goulden ML, Wofsy SC, Harden JW et al (1998) Sensitivity of boreal forest carbon balance to soil
thaw. Science 279:214–217
Gove AD, Majer JD, Rico-Gray V (2005) Methods for conservation outside of formal reserve
systems: the case of ants in the seasonally dry tropics of Veracruz, Mexico. Biol Conserv
126:328–338
Gower ST, Reich PB, Son Y (1993) Canopy dynamics and aboveground production of five tree
species with different leaf longevities. Tree Physiol 12:327–345
Graham E, Mulkey S, Kitajima K et al (2003) Cloud cover limits net CO2 uptake and growth of a
rainforest tree during tropical rainy seasons. Proc Natl Acad Sci USA 100:572–576
Grainger A (1996) An evaluation of the FAO tropical forest resource assessment, 1990. Geogr
J 162:73–79
Greene DF, Johnson EA (1993) Seed mass and dispersal capacity in wind-dispersed diaspores.
Oikos 67:69–74
Greve P, Orlowsky B, Mueller B et al (2014) Global assessment of trends in wetting and drying
over land. Nat Geosci 7:716–721
Griffiths E, Birch HF (1961) Microbiological changes in freshly moistened soil. Nature
189:424–424
Grime JP (1997) Biodiversity and ecosystem function, the debate deepens. Science 277:1260–1261
Grime JP, Crick JC, Rincon JE (1986) The ecological significance of plasticity. In: Jennings DH,
Trewavas AJ (eds) Plasticity in plants, Proceedings of the society for experimental biology,
40th symposium. The Society for Experimental Biology, Cambridge, pp 5–29
Griscom HP, Ashton MS (2011) Restoration of dry tropical forests in Central America: a review of
pattern and process. For Ecol Manag 261:1564–1579
Griscom HP, Griscom BW, Ashton MS (2009) Forest regeneration from pasture in the dry tropics
of Panama: effects of cattle, exotic grass, and forested riparia. Restor Ecol 17:117–126
212 References

Gritti ES, Cassignat C, Flores O et al (2010) Simulated effects of a seasonal precipitation change
on the vegetation in tropical Africa. Clim Past 6:169–178
Griz LMS, Machado ICS (2001) Fruiting phenology and seed dispersal syndromes in caatinga, a
tropical dry forest in the Northeast of Brazil. J Trop Ecol 17:303–321
Grogan J, Peña-Claros M, Günter S (2011) Managing natural populations of big-leaf mahogany.
In: Günter S, Weber M, Stimm B et al (eds) Silviculture in the tropics. Springer-Verlag, Berlin,
pp 227–235
Gryj EO (1990) Dispersion de frutos del arbusto Erythroxylum havanense Jacq. En Chamela,
Jalisco. Tesis de Licenciatura, Facultad de Ciencias, Universidad Nacional Autonoma de
Mexico, Mexico
Guan K, Wood EF, Medvigy D et al (2014) Terrestrial hydrological controls on land surface phe-
nology of African savannas and woodlands. J Geophys Res Biogeosci 119:2013JG002572
Guevara SA (1986) Plant species availability and regeneration in a Mexican tropical rain forest.
Ph.D. dissertation, Uppsala University, Uppsala, Sweden
Guo Q, Brown JH, Valone TJ et al (2000) Constraints of seed size on plant distribution and distur-
bance. Ecology 81:2149–2155
Gupta VVSR, Germida JJ (1988) Distribution of microbial biomass and its activity in different soil
aggregate size classes as affected by cultivation. Soil Biol Biochem 20:777–786
Gutierrez V, Keijzer M-N (2015) Funding forest landscape restoration using a business-centered
approach: an NGO’s perspective. Unasylva 66:99–105
Hammond DS, Brown VK (1995) Seed size of woody plants in relation to disturbance, dispersal,
soil type in wet neotropical forests. Ecology 76:2544–2561
Han Q, Luo G, Li C et al (2013) Modeling the grazing effect on dry grassland carbon cycling with
Biome-BGC model. Ecol Complex 17:149–157
Hansen MC, Potavov PV, Moore R et al (2013) High-resolution global maps of 21st century cover
change. Science 342:850
Hanski I, Kouki J, Halkka A (1993) Three explanations of the positive relationship between distri-
bution and abundance of species. In: Ricklefs RE, Schluter D (eds) Species diversity in ecologi-
cal communities. University of Chicago Press, Illinois, pp 108–116
Harikant, Ghildiyal MC (1982) Working plan, Renukoot Forest Division, South Circle, Uttar
Pradesh from 1982–83 to 1991–92, Working Plan Circle 2. Uttar Pradesh Forest Department,
Nainital
Harper JL, Benton RA (1966) The behaviour of seeds in soil. Part 2. The germination of seeds on
the surface of a water supply substrate. J Ecol 54:151–166
Harrington RA, Fownes JM, Vitousek PM (2001) Production and resource use efficiencies in N-
and P-limited tropical forest: a comparison of responses to long-term fertilization. Ecosystems
4:646–657
Harrison-Kirk T, Beare M, Meenken E et al (2014) Soil organic matter and texture affect responses
to dry/wet cycles: changes in soil organic matter fractions and relationships with C and N min-
eralization. Soil Biol Biochem 74:50–60
Hättenschwiler S, Tiunov AV, Scheu S (2005) Biodiversity and litter decomposition in terrestrial
ecosystems. Annu Rev Ecol Evol Syst 36:191–218
Hayden B, Greene DF, Quesada M (2010) A field experiment to determine the effect of dry-season
precipitation on annual ring formation and leaf phenology in a seasonally dry tropical forest.
J Trop Ecol 26:237–242
Hebber R, Sashidhar VR, Uma Shaanker R et al (1993) Dispersal mode of species influences the
trypsin inhibitor levels in fruits. Naturwissenshaften 80:519–521
Hendrix PF, Parmelee RW, Crossley DA Jr et al (1986) Detritus food webs and conventional and
no-tillage agroecosystems. Bioscience 36:374–380
Heneghan L, Coleman DC, Zou X Jr et al (1998) Soil microarthropod community structure and
litter decomposition dynamics: a study of tropical and temperate sites. Appl Soil Ecol 9:33–38
Herrera CM (1998) Long-term dynamics of mediterranean frugivorous birds and fleshy fruits: a
12-year study. Ecol Monogr 68:511–538
References 213

Hester AJ, Mitchell FJG, Kirby KJ (1996) Effects of season and intensity of sheep grazing on tree
regeneration in a British upland woodland. For Ecol Manag 88:99–106
Hewitt N (1998) Seed size and shade tolerance: a comparative analysis of North American temper-
ate trees. Oecologia 114:432–440
Heywood VH, Stuart SN (1992) Species extinctions in tropical forests. In: Whitmore TC, Sayer JA
(eds) Tropical deforestation and species extinction. Kluwer Academic Publishers, Dordrecht,
pp 91–118
Higgins SI, Keretetse M, February EC (2015) Feedback of trees on nitrogen mineralization to
restrict the advance of trees in C4 savannahs. Biol Lett 11:20150572
Hikosaka K, Hirose T (1998) Leaf and canopy photosynthesis of C3 plants at elevated CO2 in rela-
tion to optimal partitioning of nitrogen among photosynthetic components: theoretical predic-
tion. Ecol Model 106:247–259
Hill JL, Curran PJ (2001) Species composition in fragmented forests: conservation implications of
changing forest area. Appl Geogr 21:157–174
Hill JL, Curran PJ (2003) Area, shape and isolation of tropical forest fragments: effects on tree
species diversity and implications for conservation. J Biogeogr 30:1391–1403
Hillebrand H, Matthiessen B (2009) Biodiversity in a complex world: consolidation and progress
in functional biodiversity research. Ecol Lett 12:1–15
Hinckley TM, Richter H, Schulte PJ (1991) Water relations. In: Raghavendra AS (ed) Physiology
of trees. Wiley, New York, pp 137–162
Hiremath AJ, Sundaram B (2005) The fire-Lantana cycle hypothesis in Indian forests. Conserv
Soc 3:26–42
Hirota M, Holmgren M, Van Nes EH et al (2011) Global resilience of tropical forest and savanna
to critical transitions. Science 334:232–235
Hobbie SE (1992) Effects of plant species on nutrient cycling. Trends Ecol Evol 7:336–339
Hobbs RJ, Arico S, Aronson J et al (2006) Novel ecosystems: theoretical and management aspects
of the new ecological world order. Glob Ecol Biogeogr 15:1–7
Hobbs RJ, Higgs ES, Hall CM (eds) (2013) Novel ecosystems: intervening in the new ecological
world order. Willey-Blackwell, West Sussex
Hoekstra JM, Boucher TM, Ricketts TH, Roberts C (2005) Confronting a biome crisis: global
disparities of habitat loss and protection. Ecol Lett 8:23–29
Hoffmann WA, Franco AC, Moreira MZ et al (2005) Specific leaf area explains differences in leaf
traits between congeneric savanna and forest trees. Funct Ecol 19:932–940
Högberg P (1992) Root symbioses of trees in African dry tropical forests. J Veg Sci 3:393–400
Högberg P, Alexander IJ (1995) Roles of root symbioses in African woodland and forest: evidence
from 15N abundance and foliar analysis. J Ecol 83:217–224
Holdridge LR (1967) Life zone ecology. Tropical Science Center, San Jose
Holl KD (1999) Factors limiting tropical rain forest regeneration in abandoned pasture: seed rain,
seed germination, microclimate, and soil. Biotropica 31:229–242
Hooper ER, Condit R, Legendre P (2002) Responses of 20 native tree species to reforestation
strategies for abandoned farmland in Panama. Ecol Appl 12:1626–1641
Hooper E, Legendre P, Condit R (2005) Barriers to forest regeneration of deforested and aban-
doned land in Panama. J Appl Ecol 42:1165–1174
Hoorens B, Stroetenga M, Aerts R (2010) Litter mixture interactions at the level of plant functional
types are additive. Ecosystems 13:90–98
Houghton RA, Hall F, Goetz SJ (2009) Importance of biomass in the global carbon cycle.
J Geophys Res 114:G00E03
Howe HF, Smallwood J (1982) Ecology of seed dispersal. Annu Rev Ecol Syst 13:201–228
Huang W, Pohjonen V, Johansson S et  al (2003) Species diversity, forest structure and species
composition in Tanzanian tropical forests. For Ecol Manag 173:11–24
Huang QQ, Wu JM, Bai YY et al (2009) Identifying the most noxious invasive plants in China: role
of geographical origin, life form and means of introduction. Biodivers Conserv 18:305–316
214 References

Huante P, Rincón E, Gavito M (1988) Root system analysis of seedlings of seven tree species from
a tropical dry forest in Mexico. Trees Struct Funct 6:77–82
Huante P, Rincon E, Allen EB (1993) Effect of vesicular arbuscular mycorrhizae on seedling growth
of four tree species from the tropical deciduous forest in Mexico. Mycorrhiza 2:141–145
Huante P, Rincón E, Chapin FS III (1995) Responses to phosphorous of contrasting succession
tree-seedling species from the tropical deciduous forest of Mexico. Funct Ecol 9:760–766
Hubbell SP (2001) The unified neutral theory of biodiversity and biogeography. Princeton
University Press, Princeton
Hubbell SP, Foster RB, O’brien ST et al (1999) Light gap disturbances, recruitment limitation, and
tree diversity in a neotropical forest. Science 283:554–557
Hulshof CM, Stegen JC, Swenson NG et al (2011) Interannual variability of growth and reproduc-
tion in Bursera simaruba: the role of allometry and resource variability. Ecology 93:180–190
Hunt HW, Elliott ET, Walter DE (1989) Inferring trophic transfers from pulse-dynamics in detrital
food webs. Plant Soil 115:247–259
Hutyra LR, Munger JW, Saleska SR et al (2007) Seasonal controls on the exchange of carbon and
water in an Amazonian rain forest. J Geophys Res 112:G03008
Huxman T, Snyder K, Tissue D et al (2004) Precipitation pulses and carbon fluxes in semiarid and
arid ecosystems. Oecologia 141:254–268
Idso SB, Jackson RD, Reginato RJ (1978) Extending the “degree-day” concept of plant phenologi-
cal development to include water stress effects. Ecology 59:431–433
Imbert D, Portecop J (2008) Hurricane disturbance and forest resilience: assessing structural vs.
functional changes in a Caribbean dry forest. For Ecol Manag 255:3494–3501
Imbert D, Labbe P, Rousteau A (1996) Hurricane damage and forest structure in Guadeloupe,
French West Indies. J Trop Ecol 12:663–680
IPCC (2013) Climate change 2013: the physical science basis. In: Stocker TF, Qin D, Plattner G-K
et al (eds) Contribution of working group I to the fifth assessment report of the Intergovernmental
Panel on Climate Change. Cambridge University Press, Cambridge, pp 1029–1136
Ishida A, Diloksumpun S, Ladpala P et al (2006) Contrasting seasonal leaf habits of canopy trees
between tropical dry-deciduous and evergreen forests in Thailand. Tree Physiol 26:643–656
Jackson RB, Caldwell MM (1993) The scale of nutrient heterogeneity around individual plants and
its quantification with geostatistics. Ecology 74:612–614
Jackson RB, Manwaring JH, Caldwell MM (1990) Rapid physiological adjustment of roots to
localized soil enrichment. Nature 344:58–60
Janzen DH (1967) Synchronization of sexual reproduction of trees within the dry season in central
America. Evolution 21:620–637
Janzen DH (1970) Herbivores and the number of tree species in tropical forests. Am Nat
104:501–528
Janzen DH (1981) Patterns of herbivory in tropical deciduous forest. Biotropica 13:271–282
Janzen DH (ed) (1983) Costa Rican natural history. University of Chicago Press, Chicago, p 816
Janzen DH (1988a) Guanacaste National Park: tropical ecological and biocultural restoration.
In: Cairns J  Jr (ed) Rehabilitating damaged ecosystems, vol 11. CRC Press, Boca Raton,
pp 143–192
Janzen DH (1988b) Tropical dry forest: the most endangered major tropical ecosystem. In: Wilson
EO (ed) Biodiversity. National Academy Press, Washington, DC, pp 130–137
Janzen DH (2002) Tropical dry forest: area de conservación Guanacaste, northwestern Costa Rica.
In: Perrow M, Davy AJ (eds) Handbook of ecological restoration, vol II. Cambridge University
Press, Cambridge, pp 559–583
Jaramillo V, Martinez-yriznar A, Sanford R (2011) Primary productivity and biogeochemistry of
seasonally dry tropical forests. In: Dirzo R, Young H, Mooney H et al (eds) Seasonally dry
tropical forests. Island Press, Washington, DC, pp 109–128
Jarvis PG (1995) Scaling processes and problems. Plant Cell Environ 18:1079–1089
Jarvis P, Rey A, Petsikos C et  al (2007) Drying and wetting of Mediterranean soils stimulates
decomposition and carbon dioxide emission: the “Birch effect”. Tree Physiol 27:929–940
References 215

Jenerette G, Chatterjee A (2012) Soil metabolic pulses: water substrate and biological regulation.
Ecology 93:959–966
Jenkinson DS, Ladd JN (1981) Microbial biomass in soil: measurement and turnover. In: Paul EA,
Ladd JN (eds) Soil biochemistry, vol 5. Dekker, New York, pp 415–471
Jenkinson DS, Powlson DS (1976) Effect of biocidal treatments on metabolism in soil-V. A method
for measuring soil biomass. Soil Biol Biochem 8:209–213
Jha CS, Singh JS (1990) Compositions and dynamics of dry tropical forest in relation to soil tex-
ture. J Veg Sci 1:609–614
Jha PB, Kashyap AK, Singh JS (1996a) Effect of fertilizer and organic matter inputs on nitrifier
populations and N-mineralization rates in dry tropical region, India. Appl Soil Ecol 4:231–241
Jha PB, Singh JS, Kashyap AK (1996b) Dynamics of viable nitrifier community and nutrient
availability in dry tropical forest habitat as affected by cultivation and soil texture. Plant Soil
180:277–285
Jha CS, Goparaju L, Tripathi A et al (2005) Forest fragmentation and its impact on species diver-
sity: an analysis using remote sensing and GIS. Biodivers Conserv 14:1681–1698
Jobbágy EG, Jackson RB (2000) The vertical distribution of soil organic carbon and its relation to
climate and vegetation. Ecol Appl 10:423–436
Johnson RW, Tothill JC (1985) Definition and broad geographic outline of savanna lands. In: Tothill
JC, Mott JJ (eds) Ecology and management of the world’s savannas. Australian Academy of
Sciences, Canberra, pp 1–13
Johnston MH (1992) Soil vegetation relationships in Tabonuco forest community in the Luquillo
Mountains of Puerto Rico. J Trop Ecol 8:253–263
Joly AB (1970) Conheça a vegetação brasileira, EDUSP e Polí-gono. São Paulo, Brazil
Jones FA, Comita LS (2010) Density-dependent pre-dispersal seed predation and fruit set in a
tropical tree. Oikos 119:1841–1847
Jones RH, Zaharitz RR, Dixon MP et al (1994) Woody plant regeneration in four flooded forests.
Ecol Monogr 64:435–367
Jordan CF (1985) Nutrient cycling in tropical forest ecosystems: principles and their application in
conservation and management John. Wiley, Chichester
Jose S, Allen SC, Nair PKR (2008) Tree-crop interactions: lessons from temperate alley-cropping
systems. In: Batish DR, Kohli RK, Jose S et al (eds) Ecological basis of agroforestry. CRC
Press, Boca Raton, pp 15–36
Joshi B, Singh SP, Rawat YS et  al (2001) Facilitative effect of Coriaria nepalensis on species
diversity and growth of herbs on severely eroded hill slopes. Curr Sci 80:678–682
Juan-Baeza I, Martínez-Garza C, del-Val E (2015) Recovering more than tree cover: herbivores
and herbivory in a restored tropical dry forest. PLoS One 10:e0128583
Justiniano MJ, Fredericksen TS (2000) Phenology of tree species in Bolivian dry forests. Biotropica
32:276–281
Kammesheidt L (1999) Forest recovery by root suckers and above-ground sprouts after slash-and-
burn agriculture, fire and logging in Paraguay and Venezuela. J Trop Ecol 15:143–157
Keeling RF, Walker SJ, Piper SC et  al (2014) Scripps CO2 program. Scripps Institution of
Oceanography (SIO), University of California, La Jolla
Keitt TH (2008) Coherent ecological dynamics induced by large-scale disturbance. Nature
454:331–334
Kellner JR, Clark DB, Hubbell SP (2009) Pervasive canopy dynamics produce short-term stability
in a tropical rain forest landscape. Ecol Lett 12:155–164
Kelly CK, Bowler MG (2005) A new application of storage dynamics: differential sensitivity, dif-
fuse competition, and temporal niches. Ecology 86:1012–1022
Kemp PR, Reynolds JF, Virginia R et al (2003) Decomposition of leaf and root litter of Chihuahuan
desert shrubs: effect of three years of summer drought. J Arid Environ 53:21–39
Kemper WD, Rosenau RC (1986) Aggregate stability and size distribution. In: Klute A (ed) Methods
of soil analysis, part 1, 2nd edn. American Society of Agronomy, Madison, pp 425–442
216 References

Kemper J, Cowling RM, Richardson DM (1999) Fragmentation of South African Renosterveld


shrublands, effects on plant community structure and conservation implications. Biol Conserv
90:103–111
Kennard DK, Gould K, Putz FE et al (2002) Effect of disturbance intensity on regeneration mecha-
nisms in a tropical dry forest. For Ecol Manag 162:197–208
Khurana E, Singh JS (2000) Influence of seed size on seedling growth of Albizia procera under
different soil water levels. Ann Bot 86:1185–1192
Khurana E, Singh JS (2001a) Ecology of seed and seedling growth for conservation and restoration
of tropical dry forest: a review. Environ Conserv 28:39–52
Khurana E, Singh JS (2001b) Ecology of tree seed and seedlings: implications for tropical forest
conservation and restoration. Curr Sci 80:748–757
Khurana E, Singh JS (2004a) Germination and seedling growth of five tree species from tropical
dry forest in relation to water stress: impact of seed size. J Trop Ecol 20:385–396
Khurana E, Singh JS (2004b) Impact of elevated N inputs on seedling growth of five dry tropical
tree species as affected by life history traits. Can J Bot 82:158–167
Khurana E, Singh JS (2004c) Response of five dry tropical tree seedlings to elevated CO2: impact
of seed size and successional status. New For 27:139–157
Khurana E, Sagar R, Singh JS (2006) Seed size: a key trait determining species distribution and
diversity of dry tropical forest in northern India. Acta Oecol 29:196–204
Killingbeck KT (1996) Nutrients in senesced leaves: keys to the search for potential resorption and
resorption proficiency. Ecology 77:1716–1727
Kitajima K, Mulkey SS, Wright SJ (1997) Seasonal leaf phenotypes in the canopy of a tropical dry
forest: photosynthetic characteristics and associated traits. Oecologia 109:490–498
Kitao M, Lei TT, Koike T et al (2007) Interaction of drought and elevated CO2 concentration on
photosynthetic down-regulation and susceptibility to photoinhibition in Japanese white birch
seedlings grown with limited N availability. Tree Physiol 27:727–735
Kitaoka S, Matsuki S, Kitao M et al (2016) The photosynthetic response of four seral deciduous
broad-leaved tree seedlings grown under elevated CO2 concentrations. J Agri Meteor 72:43–49
Knapp AK, Fay PA, Blair JM et al (2002) Rainfall variability, carbon cycling, and plant species
diversity in a mesic grassland. Science 298:2202–2205
Kobe RK, Lepczyk CA, Iyer M (2005) Resorption efficiency decreases with increasing green leaf
nutrients in a global data set. Ecology 86:2780–2792
Kochummen KM, Lafrankie JV, Manokaran N (1990) Floristic composition of Pasoh Forest
Reserve, a lowland rain forest in Malaysia. J Trop For Sci 3:1–13
Koelmeyer KO (1960) The periodicity of leaf change and flowering in the principal forest com-
munities of Ceylon. Pts. 1–11. Ceylon For 4:157–189. 308–362
Koul VK, Bhardwaj SD, Kaushal AN (1995) Effect of N and P application on nutrient uptake and
biomass production in Bauhinia variegata Linn. seedlings. Ind For 121:14–22
Krasilnikov P, Gutiérrez-Castorena MC, Ahrens RJ et  al (2013) The soils of Mexico. Springer,
New York
Kushwaha CP, Tripathi SK, Singh KP (2015) Diversity of leaf deciduousness in important trees of
dry tropical forest, India. Biodivers Trop Ecosyst 2015:177–189
Lado-Monserrat L, Lull C, Bautista I et al (2014) Soil moisture increment as a controlling variable
of the ‘Birch effect’ interactions with the pre-wetting soil moisture and litter addition. Plant
Soil 379:21–34
Lal CB (1990) Dry matter and nutrient dynamics in leaves of certain tropical woody species. PhD
thesis, Department of Botany, Banaras Hindu University, Varanasi, India
Lal CB, Annapurna C, Raghubanshi AS et  al (2001a) Foliar demand and resource economy of
nutrients in dry tropical forest species. J Veg Sci 12:5–14
Lal CB, Annapurna C, Raghubanshi AS et al (2001b) Effect of leaf habit and soil type on nutri-
ent resorption and conservation in woody species of a dry tropical environment. Can J  Bot
79:1066–1075
References 217

Lamb D, Erskine PD, Parrotta JA (2005) Restoration of degraded tropical forest landscapes.
Science 310:1628–1632
Lambers H, Chapin FS III, Pons TL (1998) Plant physiological ecology. Springer-Verlag, New York
Lasky JR, Uriarte M, Muscarella R (2016) Synchrony, compensatory dynamics, and the functional
trait basis of phenological diversity in a tropical dry forest tree community: effects of rainfall
seasonality. Environ Res Lett 11:115003
Lathbridge G, Davidson MS (1983) Root-associated nitrogen-fixing bacteria and their role in the
nitrogen nutrition of wheat estimated by 15N isotope dilution. Soil Biol Biochem 15:375–376
Laurance WF (1999) Reflections on the tropical deforestation crisis. Biol Conserv 91:109–117
Laurance WF (2008) Theory meets reality: how habitat fragmentation research has transcended
island biogeographic theory. Biol Conserv 141:1731–1744
Laurance WF, Yensen E (1991) Predicting the impacts of edge effects in fragmented habitats. Biol
Conserv 55:77–92
Laurance WF, Useche DC, Rendeiro J et al (2012) Averting biodiversity collapse in tropical forest
protected areas. Nature 489(7415):290–294
Lavelle P, Blanchart E, Martin A et al (1993) A hierarchical model for decomposition in terrestrial
ecosystems: application to soils of the humid tropics. Biotropica 25:130–150
Lavorel S, McIntyre S, Landsberg J  et  al (1997) Plant functional classifications: from general
groups to specific groups based on response to disturbance. Trends Ecol Evol 12:474–478
Lavorel S, Díaz S, Cornelissen JHC et al (2007) Plant functional types: are we getting any closer to
the Holy Grail? In: Canadell JG, Pitelka LF, Pataki D (eds) Terrestrial ecosystems in a changing
world. Springer-Verlag, Berlin, pp 149–165
Leakey ADB, Ainsworth EA, Bernacchi CJ et  al (2009) Elevated CO2 effects on plant carbon,
nitrogen, and water relations: six important lessons from FACE. J Exp Bot 60:2859–2876
LeBauer DS, Treseder KK (2008) Nitrogen limitation of net primary production in terrestrial eco-
systems is globally distributed. Ecology 89:371–379
Lechowicz MJ (1995) Seasonality of flowering and fruiting in temperate forest trees. Can J Bot
73:175–182
Lehmann J, Kleber M (2015) The contentious nature of soil organic matter. Nature 528:60–68
Lehmann CER, Parr CL (2016) Tropical grassy biomes: linking ecology, human use and conserva-
tion. Philos Trans R Soc B 371:20160329
Leishman MR, Westoby M (1994a) The role of seed size in seedling establishment in dry soil
conditions—experimental evidence from semiarid species. J Ecol 82:249–258
Leishman MR, Westoby M (1994b) The role of large seed size in shaded conditions: experimental
evidence. Funct Ecol 8:205–214
Leiva MJ, Fernandez-Ales R (2003) Post-dispersive losses of acorns from Mediterranean savan-
nah-like forests and shrublands. For Ecol Manag 176:265–271
Lerdau M, Whitbeck J, Holbrook NM (1991) Tropical deciduous forest: death of a biome. Trends
Ecol Evol 6:201–202
Levis C, Clement CR, ter Steege H et al (2017) Forest conservation: humans’ handprints. Science
355(6324):466–467
Lewis RJ, Bannar-Martin KH (2012) The impact of cyclone Fanele on a tropical dry forest in
Madagascar. Biotropica 44:135–140
Li J, Ren Z, Zhou Z (2006) Ecosystem services and their values: a case study in the Qinba moun-
tains of China. Ecol Res 21:597–604
Li FL, Bao WK, Wu N (2009) Effects of water stress on growth, dry matter allocation and water-
use efficiency of a leguminous species, Sophora davidii. Agrofor Syst 77:193–201
Lieberman D (1982) Seasonallity and phenology in a dry tropical forest in Ghana. J  Ecol
70:791–806
Lieberman D, Li M (1992) Seedling recruitment patterns in a tropical dry forest in Ghana. J Veg
Sci 3:375–382
Lieberman D, Lieberman M (1984) The causes and consequences of synchronous flushing in a dry
tropical forest. Biotropica 16:193–201
218 References

Lima-Ribeiro MDS (2008) Edge effects on vegetation and population structure in Cerradao frag-
ments of southwest Goias, Brazil. Acta Bot Bras 22:535–545
Lin TC, Hamburg SP, Lin KC et al (2011) Typhoon disturbance and forest dynamics: lessons from
a Northwest Pacific subtropical forest. Ecosystems 14:127–143
Linares-Palomino R (2006) Phytogeography and floristics of seasonally dry tropical forests in Peru.
In: Pennington RT, Ratter JA, Lewis GP (eds) Neotropical savannas and seasonally dry forests:
plant biodiversity, biogeography and conservation. CRC Press, Boca Raton, pp 121–158
Linares-Palomino R, Oliveira-Filho AT, Pennington RT (2010) Neotropical seasonally dry forests:
diversity, endemism and biogeography of woody plants. In: Dirzo R, Mooney H, Ceballos G
et al (eds) Seasonally dry tropical forests: biology and conservation. Island Press, Washington,
DC
Liski J, Nissinen A, Erhard M et al (2003) Climatic effect on litter decomposition from arctic tun-
dra to tropical rainforest. Glob Chang Biol 9:575–584
Livingston GP, Vitousek PM, Matson PA (1988) Nitrous oxide flux and nitrogen transformations
across a landscape gradient in Amazonia. J Geophys Res 93(D3):1593–1599
Lloyd J, Farquhar GD (1996) The CO2 dependence of photosynthesis, plant growth responses
to elevated atmospheric CO2 concentrations and their interaction with soil nutrient status.
I. General principles and forest ecosystems. Funct Ecol 10:4–32
Lobo JA, Quesada M, Stoner KE et  al (2003) Factors affecting phenological patterns of
Bombacaceous trees in seasonal forests in Costa Rica and Mexico. Am J Bot 90:1054–1063
Lohbeck M, Lebrija-Trejos E, Martínez-Ramos M et al (2015a) Functional trait strategies of trees
in dry and wet tropical forests are similar but differ in their consequences for succession. PLoS
One 10:e0123741
Lohbeck M, Poorter L, Martínez-Ramos M et al (2015b) Biomass is the main driver of changes in
ecosystem process rates during tropical forest succession. Ecology 96:1242–1252
Lomolino MV (2001) The species area relationship: new challenges for an old pattern. Prog Phys
Geogr 25:1–21
Lorenz K, Lal R, Preston CM et al (2007) Strengthening the soil organic carbon pool by increasing
contributions from recalcitrant aliphatic bio(marc) molecules. Geoderma 142:1–10
Lott EJ, Atkinson TH (2002) Biodiversidad y fitogeografía de Chamela Cuixmala, Jalisco. In:
Noguera F, Vega JH, García Aldrete AN et al (eds) Historia Natural de Chamela. Instituto de
Biología, UNAM, México City, pp 83–97
Lott EJ, Atkinson TH (2006) Mexican and Central American seasonally dry tropical forests:
Chamela-Cuixmala, Jalisco, as a focal point for comparison. In: Pennington RT, Ratter JA,
Lewis GP (eds) Neotropical savannas and seasonally dry forests: plant biodiversity, biogeogra-
phy and conservation. CRC Press, Boca Raton, pp 121–158
Lovejoy TE (1980) A projection of species extinctions, The global 2000 report to the President:
Entering the 21st century. Council on the Environmental Quality and the Department of State.
Government Printing Office, Washington, DC, pp 328–331
Lovejoy TE, Bierregaard RO, Rankin JM et al (1983) Ecological dynamics of tropical forest frag-
ments. In: Slutton SL, Whitmore TC, Chadwick AC (eds) Tropical rainforests: ecology and the
management. Blackwell Scientific Publications, Oxford, pp 377–384
Lovejoy TE, Bierregaard RO, Rylands AB et  al (1986) Edge and other effects of isolation on
Amazon forest fragments. In: Soulé ME (ed) Conservation biology: the science of scarcity and
diversity. Sinauer Associates, Sunderland, pp 257–285
Lugo AE (2008) Visible and invisible effects of hurricanes on forest ecosystems: an international
review. Austr Ecol 33:368–398
Lugo A, Erickson H (2017) Novelty and its ecological implications to dry forest functioning and
conservation. Forests 8:161. https://doi.org/10.3390/f8050161
Lugo AE, Helmer E (2004) Emerging forests on abandoned land: Puerto Rico’s new forests. For
Ecol Manag 190:145–161
Lugo AE, Murphy PG (1986) Nutrient dynamics of a Puerto Rican sub-tropical dry forest. J Trop
Ecol 2:55–72
References 219

Lugo AE, Gonzalez-Liboy JA, Cintron B et al (1978) Structure, productivity and transpiration of a
subtropical dry forest in Puerto Rico. Biotropica 10:278–291
Lugo AE, Medina E, Trejo-Torres JC et al (2006) Botanical and ecological basis for the resilience
of Antillean dry forests. In: Pennington RT, Ratter JA, Lewis GP (eds) Neotropical savan-
nas and seasonally dry forests: plant biodiversity, biogeography and conservation. CRC Press,
Boca Raton, pp 359–382
Lull HW (1959) Soil compaction on forest and range lands, Miscellaneous publication 769.
U.S. Department of Agriculture, Washington, DC
Luo Y, Field CB, Mooney HA (1994) Predicting responses of photosynthesis and root fraction
to elevated [CO2] a: interactions among carbon, nitrogen and growth. Plant Cell Environ
17:1195–1204
Lupascu M, Welker JM, Seibt U et al (2014) High Arctic wetting reduces permafrost carbon feed-
backs to climate warming. Nat Clim Chang 4:51–55
Luttge U (1997) Physiological ecology of tropical plants. Springer-Verlag, Berlin, p 384
Luyssaert S, Inglima I, Jung M et al (2007) CO2 balance of boreal, temperate, and tropical forests
derived from a global data-base. Glob Chang Biol 13:2509–2537
Maass JM (1995) Tropical deciduous forest conversion to pasture and agriculture. In: Bullock
SH, Mooney HA, Medina E (eds) Seasonally dry tropical forests. Cambridge University Press,
Cambridge, pp 399–422
Maass JM, Burgos A (2011) Water dynamics at the ecosystem level in seasonally dry tropical
forests. In: Dirzo R, Young HS, Mooney HA et al (eds) Seasonally dry tropical forests, ecology
and conservation. Island Press, Washington, DC, pp 141–166
Maass JM, Balvanera P, Castillo A et al (2005) Ecosystem services of tropical dry forests: Insights
from long-term ecological and social research on the Pacific Coast of Mexico. Ecol Soc 10:1–23
Mac Arthur RH, Wilson EO (1967) The theory of island biogeography. Princeton University Press,
Princeton, p 203
Macedo MO, Resende AS, Garcia PC et al (2008) Changes in soil C and N stocks and nutrient
dynamics 13 years after recovery of degraded land using leguminous nitrogen fixing trees. For
Ecol Manag 255:1516–1524
Mahapatra AK, Tewari DD (2005) Importance of non-timber forest products in the economic valu-
ation of dry deciduous forests of India. For Pol Econom 7:455–467
Malaisse F (1974) Phenology of the Zambezian woodland area with emphasis on the miombo
ecosystem. In: Lieth H (ed) Phenology and seasonality modelling. Springer-Verlag, New York,
pp 269–286
Malcolm JR (1994) Edge effects in central Amazonian forest fragments. Ecology 75:2438–2445
Malcolm JR (2001) Extending models of edge effects to diverse landscape configurations, with a
test case from the Neotropics. In: Bierregaard RO Jr, Gascon C, Lovejoy TE et al (eds) Lessons
from Amazonia: the ecology and conservation of a fragmented forest. Yale University Press,
New Haven, pp 346–357
Malcolm JR, Valenta K, Lehman SM (2017) Edge effects in tropical dry forests of Madagascar:
additivity or synergy? Landsc Ecol 32:1–15
Malhi Y, Wright J (2004) Spatial patterns and recent trends in the climate of tropical rainforest
regions. Trans R Soc Lond Ser B Biol Sci 359:311–329
Malhi Y, Baldocchi DD, Jarvis PG (1999) The carbon balance of tropical, temperate and boreal
forests. Plant Cell Environ 22:715–740
Malhi Y, Baker TR, Phillips OL et al (2004) The above-ground coarse wood productivity of 104
Neotropical forest plots. Glob Chang Biol 10:563–591
Malhi Y, Wood D, Baker T et al (2006) The regional variation of aboveground live biomass in old-
growth Amazonian forests. Glob Chang Biol 12:1107–1138
Malhi Y, Roberts JT, Betts RA et  al (2008) Climate change, deforestation and the fate of the
Amazon. Science 319:169–172
Malhi Y, Doughty CE, Goldsmith GR et al (2015) The linkages between photosynthesis, produc-
tivity, growth and biomass in lowland Amazonian forests. Glob Chang Biol 21:2283–2295
220 References

Maloney ED, Camargo SJ, Chang E et al (2013) North American climate in CMIP5 experiments:
part III: assessment of twenty-first-century projections. J Climatol 27:2230–2270
Marcano-Vega H, Aide TM, Baez D (2002) Forest regeneration in abandoned coffee plantations
and pastures in the Cordillera Central of Puerto Rico. Plant Ecol 161:75–87
Marín-Spiotta E, Sharma S (2013) Carbon storage in successional and plantation forest soils: a
tropical analysis. Glob Ecol Biogeogr 22:105–117
Markesteijn L, Poorter L, Bongers F (2007) Light-dependent leaf trait variation in 43 tropical dry
forest tree species. Am J Bot 94:515–525
Markesteijn L, Iraipi J, Bongers F et al (2010) Seasonal variation in soil and plant water potentials
in a Bolivian tropical moist and dry forest. J Trop Ecol 26:497–508
Markesteijn L, Poorter L, Bongers F et al (2011) Hydraulics and life history of tropical dry forest
tree species: coordination of species’ drought and shade tolerance. New Phytol 191:480–495
Martijena ND, Bullock SH (1994) Monospecific dominance of a tropical deciduous forest in
Mexico. J Biogeogr 21:63–74
Martin CE, Loeschen VS, Borchert R (1994) Photosynthesis and leaf longevity in trees of a tropi-
cal deciduous forest in Costa Rica. Photosynthetica 30:341–351
Martinez LJ, Zinck JA (2004) Temporal variation of soil compaction and deterioration of soil qual-
ity in pasture areas of Colombian Amazonia. Soil Tillage Res 75:3–17
Martínez-Yrízar A (1995) Biomass distribution and primary productivity of tropical dry forests. In:
Bullock SH, Mooney HA (eds) Seasonally dry tropical forests. Cambridge University Press,
Cambridge, pp 326–345
Martínez-Yrízar A, Sarukhán J (1990) Litterfall patterns in a tropical deciduous forest in Mexico
over a five-year period. J Trop Ecol 6:433–444
Martínez-Yrízar A, Sarukhán J (1993) Cambios estacionales del mantillo en el suelo de un bosque
caducifolio y uno subcaducifolio en Chamela, Jalisco, México. Acta Bot Mex 21:1–6
Martinuzzi S, Lugo AE, Brandeis TJ et al (2013) Geographic distribution and level of novelty of
Puerto Rican forests. In: Hobbs RJ, Higgs ES, Hall C (eds) Novel ecosystems: intervening in
the new ecological world order. Wiley, Oxford, pp 81–87
Marumoto T, Kai H, Yoshida T et al (1977) Chemical fractions of organic nitrogen in acid hydro-
lysates given from microbial cells and their cell wall substances and characterization of decom-
posable soil organic nitrogen due to drying. Soil Sci Plant Nutr 23:125–134
Maximov NA (1929) The plant in relation to water. George Allen and Unwin, London
Maxwell SL, Fuller RA, Brooks TM et al (2016) The ravages of guns, nets and bulldozers. Nature
536:143–145
Mayaux P, Holmgren P, Achard F et  al (2005) Tropical forest cover change in the 1990s and
options for future monitoring. Phil Trans R Soc B 360:373–384
Maza-Villalobos S, Poorter L, Martínez-Ramos M (2013) Effects of ENSO and temporal rainfall
variation on the dynamics of successional communities in old-field succession of a tropical dry
forest. PLoS One 8:e82040
McClaugherty CA, Aber JD, Melillo JM (1982) The role of fine roots in the organic matter and
nitrogen budgets of two forest ecosystems. Ecology 63:1481–1490
McConnaughay KDM, Bazzaz FA (1987) The relationship between gap size and performance of
several colonizing annuals. Ecology 68:411–416
McConnaughay KDM, Coleman JS (1999) Biomass allocation in plants: ontogeny or optimally?
A test along three resource gradients. Ecology 80:2581–2593
McCulloh K, Meinzer F, Sperry J et al (2011) Comparative hydraulic architecture of tropical tree
species representing a range of successional stages and wood density. Oecologia 167:27–37
McGroddy ME, Daufresne T, Hedin L (2004) Scaling C:N:P stoichiometry in forests worldwide:
implications of terrestrial redfield-type ratios. Ecology 85:2390–2401
McLaren KP, McDonald MA (2005) Seasonal patterns of flowering and fruiting in a dry tropical
forest in Jamaica. Biotropica 37:584–590
McWethy DB, Hansen AJ, Verschuyl JP (2009) Edge effects for songbirds vary with forest produc-
tivity. For Ecol Manag 257:665–678
References 221

Mediavilla S, Escudero A (2003) Stomatal responses to drought at a Mediterranean site: a compar-


ative study of co-occurring woody species differing in leaf longevity. Tree Physiol 23:987–996
Medina E (1980) Ecology of tropical American savannas: an ecophysiological approach. In: Morns
DR (ed) Human ecology in savanna environments. Academic, London, pp 297–319
Medina E (1984) Nutrient balance and physiological processes at the leaf level. In: Medina E,
Mooney HA, Vazque-Yanes C (eds) Physiological ecology of plants of the wet tropics. Dr W
Junk Publications, The Hague
Medina E (2005) Diversity of life forms of higher plants in neotropical dry forest. In: Bullock
SH, Mooney HA, Medina E (eds) Seasonally dry tropical forest. Cambridge University Press,
Cambridge, MA, pp 221–224
Medina E, Francisco M (1994) Photosynthesis and water relations of savanna tree species differing
in leaf phenology. Tree Physiol 14:1367–1381
Meentemeyer V (1978) Macroclimate and lignin control of litter decomposition rates. Ecology
59:465–472
Meentemeyer V (1984) The geography of organic decomposition rates. Ann Am Assoc Geogr
74:551–560
Meir P, Pennington RT (2011) Climatic change and seasonally dry tropical forests. In: Dirzo R,
Young HS, Mooney HA et al (eds) Seasonally dry tropical forests: ecology and conservation.
Island Press, Washington, DC, pp 279–299
Meli P (2003) Restauración ecológica de bosques tropicales. Veinte años de investigación aca-
démica. Interciencia 28:581–589
Mencuccini M, Comstock J (1997) Vulnerability to cavitation in populations of two desert spe-
cies, Hymenoclea salsola and Ambrosia dumosa, from different climatic regions. J Exp Bot
48:1323–1334
Méndez-Alonzo R, Paz H, Zuluaga RC et al (2012) Coordinated evolution of leaf and stem eco-
nomics in tropical dry forest trees. Ecology 93:2397–2406
Méndez-Barroso LA, Vivoni ER (2010) Observed shifts in land surface conditions during the
North American monsoon: implications for a vegetation–rainfall feedback mechanism. J Arid
Environ 74:549–555
Méndez-Barroso LA, Vivoni ER, Watts CJ et al (2009) Seasonal and interannual relation between
precipitation, surface soil moisture and vegetation dynamics in the North American monsoon
region. J Hydrol 377:59–70
Mendivelso HA, Camarero JJ, Gutiérrez E et  al (2014) Time-dependent effects of climate and
drought on tree growth in a Neotropical dry forest: short-term tolerance vs. long-term sensitiv-
ity. Agric For Meteor 188:13–23
Milberg P, Lamont BB (1997) Seed/cotyledon size and nutrient content play a major role in early
performance of species on nutrient-poor soil. New Phytol 137:665–672
Miles L, Newton AC, DeFries RS et  al (2006) A global overview of the conservation status of
tropical dry forests. J Biogeogr 33:491–505
Miller PM, Kauffman JB (1998) Effects of slash and burn agriculture on species abundance and
composition of a tropical deciduous forest. For Ecol Manag 103:191–201
Miranda A (2002) Diversidad, historia natural, ecología y conservación de los mamíferos de
Chamela. In: Noguera FA, Vera Rivera JH, García Aldrete AN et  al (eds) Historia Natural
de Chamela. Instituto de Biología, Universidad Nacional Autónoma de México, México,
pp 359–377
Misra R (1983) Indian savannas. In: Bourliere F (ed) Tropical savannas. Elsevier, Amsterdam,
pp 151–166
MoEF (1999) National forestry action plan. Ministry of environment and forests. Government of
India, New Delhi
Mohamed-Yasseen Y, Barringer SA, Splittstoesser WE et al (1994) The role of seed coats in seed
viability. Bot Rev 60:426–439
Mokany K, Raison RJ, Prokushkin AS (2006) Critical analysis of root:shoot ratios in terrestrial
biomes. Glob Chang Biol 12:84–96
222 References

Monasterio M, Sarmiento G (1976) Phenological strategies of plant species in the tropical savanna
in the semi-deciduous forest of the Venezuelan Llanos. J Biogeogr 3:325–356
Montaño NM, Garcia-Oliva F, Jaramillo VJ (2007) Dissolved organic carbon affects soil microbial
activity and nitrogen dynamics in a Mexican tropical deciduous forest. Plant Soil 295:265–277
Montpellier Panel (2013) Sustainable intensification: a new paradigm for African agriculture.
Agriculture for Impact, London
Mooney HA, Bullock SH, Medina E (1995) Introduction. In: Bullock SH, Mooney HA, Medina E
(eds) Seasonally dry tropical forests. Cambridge University Press, Cambridge, pp 1–8
Moore JM, Wein RW (1977) Viable seed population by soil depth and potential site recolonization
after disturbance. Can J Bot 55:2408–2412
Moore TR, Trofymow JA, Prescott CE et al (2006) Patterns of carbon, nitrogen and phosphorus
dynamics in decomposing foliar litter in Canadian forests. Ecosystems 9:46–62
Moorhead DL, Currie WS, Rasttetter EB et al (1999) Climate and litter quality controls on decom-
position: an analysis of modeling approaches. Glob Biogeochem Cycles 13:575–589
Moreira F, Viedma O, Arianoutsou M et  al (2011) Landscape–wildfire interactions in southern
Europe: implications for landscape management. J Environ Manag 92:2389–2402
Moretto AS, Distel RA, Didone NG (2001) Decomposition and nutrient dynamic of leaf litter and
roots from palatable and unpalatable grasses in semi-arid grassland. Appl Soil Ecol 18:31–37
Motzer T, Munz N, Küppers M et al (2005) Stomatal conductance, transpiration and sap flow of
tropical montane rain forest trees in the southern Ecuadorian Andes. Tree Physiol 25:1283–1129
Moyano F, Vasilyeva N, Bouckaert L et  al (2012) The moisture response of soil heterotrophic
respiration: interaction with soil properties. Biogeosciences 9:1173–1182
Moyano F, Manzoni S, Chenu C (2013) Responses of soil heterotrophic respiration to moisture
availability: an exploration of processes and models. Soil Biol Biochem 59:72–85
Mucunguzi P, Oryem-Origa H (1996) Effects of heat and fire on the germination of Acacia siebe-
riana D.C. and Acacia gerrardii Benth. in Uganda. J Trop Ecol 12:1–10
Muick PC, Bartolome JW (1987) Factors associated with oak regeneration in California. In:
Plumb TR, Pillsbury NH (eds) Proceedings of the symposium on multiple-use management of
California’s hardwood resources. USDA Forest Service General Technical Report PSW-100,
Berkeley, pp 89–91
Murali KS, Sukumar R (1993) Leaf flushing phenology and herbivory in a tropical dry deciduous
forest, southern India. Oecologia 94:114–119
Murali KS, Sukumar R (1994) Reproductive phenology of a tropical dry forest in Mudumalai,
Southern India. J Ecol 82:759–767
Murcia C (1995) Edge effects in fragmented forests: implications for conservation. Trends Ecol
Evol 10:58–62
Murphy PG, Lugo AE (1986) Ecology of tropical dry forest. Annu Rev Ecol Evol Syst 17:67–88
Murphy PG, Lugo AE (1995) Dry forests of Central America and the Caribbean. In: Mooney
HA, Bullock SH (eds) Seasonally tropical forests. Cambridge University Press, Cambridge,
pp 9–34
Murphy BP, Andersen AN, Parr CL (2016) The underestimated biodiversity of tropical grassy
biomes. Philos Trans R Soc B 371:20150319
Muscarella R, Uriarte M (2016) Do community-weighted mean functional traits reflect optimal
strategies? Proc R Soc Lond B 283:20152434
Myers N (1988) Threatened biota’s, ‘Hot spots’ in tropical forests. Environmentalist 8:187–208
Nakano H, Makino A, Mae T (1997) The effect of elevated partial pressures of CO2 on the relation-
ship between photosynthetic capacity and N content in rice leaves. Plant Physiol 115:191–198
Nayar MP (1997) Changing patterns of the Indian flora. Bull Bot Surv Ind 19:145–154
Neelin JD, Münnich M, Su H et al (2006) Tropical drying trends in global warming models and
observations. Proc Natl Acad Sci USA 103:6110–6115
Nepstad DC, Uhl C, Pereira CA et al (1996) A comparative study of tree establishment in aban-
doned pasture and mature forest of eastern Amazonia. Oikos 76:25–39
References 223

Newmark WD, Jenkins CN, Pimm SL et al (2017) Targeted habitat restoration can reduce extinc-
tion rates in fragmented forests. Proc Natl Acad Sci USA 114:9635–9640
Newton AC, Del Castillo RF, Echeverria C et al (2012) Forest landscape restoration in the drylands
of Latin America. Ecol Soc 17:21
Niinemets Ü (1999) Components of leaf dry mass per area-thickness and density-alter leaf photo-
synthetic capacity in reverse directions in woody plants. New Phytol 144:35–47
Niinemets Ü, Tenhunen JD (1997) A model separating leaf structural and physiological effects
on carbon gain along light gradients for the shade-tolerant species Acer saccharum. Plant Cell
Environ 20:845–866
Niinemets Ü, Díaz-Espejo A, Flexas J et al (2009) Role of mesophyll diffusion conductance in
constraining potential photosynthetic productivity in the field. J Exp Bot 60:2249–2270
Norris J, Allen R, Amato E et al (2016) Evidence for climate change in the satellite cloud record.
Nature 536:72–75
Nye PH, Greenland DJ (1964) Changes in the soil after clearing tropical forests. Plant Soil
21:101–112
O’Connor TG (1996) Individual, population and community response of woody plants to browsing
in Africa savannas. Bull Grassl Soc S Africa 7:14–18
Oades JM (1984) Soil organic matter and structural stability: mechanisms and implications for
management. Plant Soil 76:319–337
Olesen JM, Bascompte J, Elberling H et al (2008) Temporal dynamics in a pollination network.
Ecology 89:1573–1582
Olivares E, Medina E (1992) Water and nutrient relations of woody perennials from tropical dry
forests. J Veg Sci 3:383–392
Olson DM, Dinerstein E, Abell R et al (2000) The global 200: a representation approach to con-
serving the Earth’s distinctive ecoregions, Conservation science program. World Wildlife
Fund-US, Washington, DC
Olson DM, Dinerstein E, Wikramanayake ED et al (2001) Terrestrial ecoregions of the world: a
new map of life on earth. Bioscience 51:933–938
Ong CK, Deans JD, Wilson J et al (1999) Exploring belowground complementarity in agroforestry
using sap flow and root fractal techniques. Agrofor Syst 44:87–103
Oni O, Bada SO (1992) Effects of seed size on seedling vigour in idigbo (Terminalia ivorensis).
J Trop For Sci 4:215–224
Onishi N (2016) Africa’s charcoal economy is cooking. The trees are paying. Available online:
https://www.nytimes.com/2016/06/26/world/africa/africas-charcoal-economy-is-cooking-
thetrees-are-paying.html?_r=0
Onoda Y, Hikosaka K, Hirose T (2005) Seasonal change in the balance between capacities of RuBP
carboxylation and RuBP regeneration affects CO2 response of photosynthesis in Polygonum
cuspidatum. J Exp Bot 56:755–763
Opler PA, Frankie GW, Baker HG (1976) Rainfall as a factor in the synchronization, release and
timing of anthesis by tropical trees and shrubs. J Biogeogr 3:231–236
Opler PA, Frankie GW, Baker HG (1980) Comparative phonological studies of Treelet and shrub
species in tropical wet and dry forests in the woodlands of Costa Rica. J Ecol 68:167–188
Ouédraogo D-Y, Mortier F, Gourlet-Fleury S et al (2013) Slow-growing species cope best with
drought: evidence from long-term measurements in a tropical semideciduous moist forest of
Central Africa. J Ecol 101:1459–1470
Ouédraogo O, Bondé L, Boussim JI et al (2015) Caught in a human disturbance trap: Responses of
tropical savanna trees to increasing land-use pressure. For Ecol Manag 354:68–76
Ouédraogo D-Y, Fayolle A, Gourlet-Fleury S et  al (2016) The determinants of tropical for-
est deciduousness: disentangling the effects of rainfall and geology in central Africa. J Ecol
104:924–935
Pallardy SG (2008) Nitrogen metabolism. In: Kramer PJ, Kozlowski TT (eds) Physiology of
woody plants, 3rd edn. Academic, New York, pp 233–254
224 References

Palma B, Vogt G, Neville P (1995) Endogenous factors that limit seed germination of Acacia sen-
egal Willd. Phyton (Buenos Aires) 57:97–102
Palmiotto PA, Davies SJ, Vogt KA et al (2004) Soil-related habitat specialization in dipterocarp
rain forest tree species in Borneo. J Ecol 92:609–623
Pan Y, Birdsey RA, Fang J et al (2011) A large and persistent carbon sink in the world’s forests.
Science 333:988–993
Pandey SK, Shukla RP (1999) Plant diversity and community patterns along the disturbance gradi-
ent in plantation forests of sal (Shorea robusta Gaertn.) Curr Sci 77:814–818
Pandey SK, Shukla RP (2001) Regeneration strategy and plant diversity status in degraded sal
forests. Curr Sci 81:95–102
Pandey CB, Singh JS (1992a) Rainfall and grazing effects on net primary productivity in a tropical
savanna, India. Ecology 73:2007–2021
Pandey CB, Singh JS (1992b) Influence of rainfall and grazing on herbage dynamics in a season-
ally dry tropical savanna. Vegetatio 102:107–124
Pandey SK, Singh H, Singh JS (2014a) Contrasting leaf phenology of woody species of dry tropi-
cal forest. Plant Biosyst 148:655–665
Pandey SK, Singh H, Singh JS (2014b) Effect of environmental conditions on decomposition in
eight woody species of a dry tropical forest. Plant Biosyst 148:410–418
Paquette A, Messier C (2011) The effect of biodiversity on tree productivity: from temperate to
boreal forests. Glob Ecol Biogeogr 20:170–180
Parthasarathy N, Sethi P (1997) Trees and liana species diversity and population structure in a
tropical dry evergreen forest in south India. Trop Ecol 38:19–30
Pattanayak SK, Wunder S, Ferraro PJ (2010) Show me the money: do payments supply environ-
mental services in developing countries? Rev Environ Econ Policy 4(2):254–274
Pau S, Okin GS, Gillespie TW (2010) Asynchronous response of tropical forest leaf phenology to
seasonal and El Niño-driven drought. PLoS One 5:e11325
Pau S, Wolkovich EM, Cook BI et al (2013) Clouds and temperature drive dynamic changes in
tropical flower production. Nat Clim Chang 3:838–842
Pearce F (2015) The new wild: why invasive species will be nature’s salvation. Beacon Press,
Boston
Peng S, Piao S, Shen Z et al (2013) Precipitation amount, seasonality and frequency regulate car-
bon cycling of a semi-arid grassland ecosystem in Inner Mongolia, China: a modeling analysis.
Agr For Meteor 178–179:46–55
Pennington RT, Prado DA, Pendry C (2000) Neotropical seasonally dry forests and Pleistocene
vegetation changes. J Biogeogr 27:261–273
Pennington RT, Ratter JA, Lewis GP (2006a) An overview of the plant diversity, biogeography and
conservation of neotropical savannas and seasonally dry forests. In: Pennington RT, Ratter JA,
Lewis GP (eds) Neotropical savannas and seasonally dry forests: plant biodiversity, biogeogra-
phy and conservation. CRC Press, Boca Raton, pp 1–29
Pennington RT, Ratter JA, Lewis GP (eds) (2006b) Neotropical savannas and seasonally dry for-
ests: plant biodiversity, biogeography and conservation. CRC Press, Boca Raton
Pennington RT, Lavin M, Oliveira-Filho A (2009) Woody plant diversity, evolution, and ecology
in the tropics: perspectives from seasonally dry tropical forests. Annu Rev Ecol Evol Syst
40:437–457
Pereira HM, Ferrier S, Walters M et  al (2013) Essential biodiversity variables. Science
339(6117):277–278
Pérez-Harguindeguy N, Diaz S, Cornelissen JHC et al (2000) Chemistry and toughness predict
leaf litter decomposition rates over a wide spectrum of functional types and taxa in central
Argentina. Plant Soil 218:21–30
Pérez-Harguindeguy N, Díaz S, Garnier E et al (2013) New handbook for standardised measure-
ment of plant functional traits worldwide. Aust J Bot 61:167–234
Perrings C, Williamson M, Barbier EB et al (2002) Biological invasion risks and public good: an
economic perspective. Conserv Ecol 6:1
References 225

Perry DA, Amaranthus MP, Borchers JG et  al (1989) Bootstrapping in ecosystems. Bioscience
39:230–237
Persiani AM, Tosi S, Delfrate G et al (2011) High spots for diversity of soil and litter micro fungi
in Italy. Plant Biosyst 145:969–977
Pimm SL, Raven P (2000) Extinction by numbers. Nature 403:843–845
Pineda-García F, Paz H, Meinzer FC (2013) Drought resistance in early and late secondary succes-
sional species from a tropical dry forest: the interplay between xylem resistance to embolism,
sapwood water storage and leaf shedding. Plant Cell Environ 36:405–418
Pisani O, Hills KM, Courtier-Murias D et al (2014) Accumulation of aliphatic compounds in soil
with increasing mean annual temperature. Org Geochem 76:118–127
Plieninger T, Pulido FJ, Schaich H (2004) Effects of land-use and landscape structure on holm
oak recruitment and regeneration at farm level in Quercus ilex L. dehesas. J  Arid Environ
57:345–364
Poorter L, Bongers F (2006) Leaf traits are good predictors of plant performance across 53 rain
forest species. Ecology 87:1733–1743
Poorter H, Garnier E (1999) Ecological significance of inherent variation in relative growth rate
and its components. In: Pugnaire FI, Valladares F (eds) Handbook of functional plant ecology.
Marcel Dekker, New York, pp 81–120
Poorter L, Markesteijn L (2008) Seedling traits determine drought tolerance of tropical tree spe-
cies. Biotropica 40:321–331
Poorter H, Niinemets Ü, Poorter L et al (2009) Causes and consequences of variation in leaf mass
per area (LMA): a meta analysis. New Phytol 182:565–588
Posada JM, Schuur EAG (2011) Relationships among precipitation regime, nutrient availability,
and carbon turnover in tropical rain forests. Oecologia 165:783–795
Posada JM, Aide TM, Cavelier J (2000) Cattle and weedy shrubs as restoration tools of tropical
montane rainforest. Restor Ecol 8:370–379
Posada JM, Lechowicz MJ, Kitajima K (2009) Optimal photosynthetic use of light by tropical
tree crowns achieved by adjustment of individual leaf angles and nitrogen content. Ann Bot
103:795–805
Powell SJ, Prosser JI (1992) Inhibition of biofilm populations of Nitrosomonas europaea. Microb
Ecol 24:43–50
Powers JS, Tiffin P (2010) Plant functional type classifications in tropical dry forests in Costa Rica:
leaf habit versus taxonomic approaches. Funct Ecol 24:927–936
Powers JS, Becknell JM, Irving J  et  al (2009a) Diversity and structure of regenerating tropical
dry forests in Costa Rica: geographic patterns and environmental divers. For Ecol Manag
258:959–970
Powers JS, Montgomery RA, Adair EC et al (2009b) Decomposition in tropical forests: a pan-
tropical study of the effects of litter type, litter placement and mesofaunal exclusion across a
precipitation gradient. J Ecol 97:801–811
Powers JS, Corre MD, Twine TE et al (2011) Geography bias of field observations of soil carbon
stocks with tropical land-use changes precludes spatial extrapolations. Proc Natl Acad Sci USA
108:6318–6322
Powers JS, Becklund KK, Gei MG et al (2015) Nutrient addition effects on tropical dry forests: a
mini-review from microbial to ecosystem scales. Front Earth Sci 3:34
Prado-Junior JA, Schiavini I, Vale VS et al (2016) Conservative species drive biomass productivity
in tropical dry forests. J Ecol 104:817–827
Prasad N, Hegde M (1986) Phenology and seasonality in the tropical dry deciduous forest of
Bandipur, South India. Proc Ind Acad Sci (Plant Sci) 96:121–133
Prasad KG, Rawat VRS (1992) Fertilizer use efficiency of different tree species for higher biomass
production. Ind For 118:265–270
Prasad KG, Rawat VRS (1994) Fertilizer response of Eucalyptus tereticornis seedlings. Ind For
120:699–710
226 References

Prasad KG, Gupta GN, Mohan S et al (1984) Fertilization in Eucalyptus grandis on severely trun-
cated soil III: nutrient uptake. Ind For 110:1033–1048
Prasad R, Date GP, Jalil P (1988) Observation on storage and germination of seeds of Lagerstroemia
parviflora (Roxb.) Vaniki Sandesh 12:2–6
Prentice IС, Farquhar GD, Fasham MJR et al (2001) The carbon cycle and atmospheric carbon
dioxide. In: Houghton JT, Ding Y, Griggs DJ et al (eds) Climate change 2001: the scientific
basis, Contribution of Working Group I to the Third Assessment Report of the Intergovernmental
Panel on Climate Change. Cambridge University Press, Cambridge, pp 183–237
Preston KA, Cornwell WK, DeNoyer JL (2006) Wood density and vessel traits as distinct corre-
lates of ecological strategy in 51 California coast range angiosperms. New Phytol 170:807–818
Pretty J (2011) Sustainable intensification in Africa. Int J Agric Sustain 9:3–4
Prior LD, Eamus D, Bowman DMJS (2003) Leaf attributes in the seasonally dry tropics: a com-
parison of four habitats in northern Australia. Funct Ecol 17:504–515
Prior LD, Eamus D, Bowman DMJS (2004) Tree growth rates in north Australian savanna habitats:
seasonal patterns and correlations with leaf attributes. Aust J Bot 52:303–314
Prosser JI (1989) Autotrophic nitrification in bacteria. Adv Microb Physiol 30:125–181
Pugnaire F, Chapin FS (1993) Controls over nutrient resorption from leaves of evergreen
Mediterranean species. Ecology 74:124–129
Putz FE, Coley PD, Lu K et al (1983) Uprooting and snapping of trees: structural determinants and
ecological consequences. Can J For Res 13:1011–1020
Quesada M, Aguilar R, Rosas F et  al (2009) Human impacts on pollination, reproduction and
breeding systems in tropical forest plants. In: Dirzo R, Mooney H, Ceballos G (eds) Tropical
dry forests. Island Press, New York, pp 173–194
Quesada CA, Phillips OL, Schwarz M et al (2012) Basin-wide variations in Amazon forest struc-
ture and function are mediated by both soils and climate. Biogeosciences 9:2203–2246
Quézel P, Médail F (2003) Ecologie et Biogéographie des forêts du bassinméditerranéen. Elsevier,
Amsterdam
Quigley MF, Platt WJ (2003) Composition and structure of seasonally deciduous forests in the
Americas. Ecol Monogr 73:87–106
Rachmawali I, Effendi M, Sinaga M (1996) Growth and development performance of multipur-
pose tree species under the effect of fertilizer. Bull Penelitian Kehutanan-Kupang 1:43–51
Raghubanshi AS (1992) Effect of topography on selected soil properties and nitrogen mineraliza-
tion in a dry tropical forest. Soil Biol Biochem 24:145–150
Raghubanshi AS, Tripathi A (2009) Effect of disturbance, habitat fragmentation and alien invasive
plants on floral diversity in dry tropical forests of Vindhyan highland: a review. Trop Ecol
50:57–69
Raherison SM, Grouzis M (2005) Plant biomass, nutrient concentration and nutrient storage in a
tropical dry forest in the south–west of Madagascar. Plant Ecol 180:33–45
Raich JW (1983) Effects of forest conversion on the carbon budget of a tropical soil. Biotropica
15:177–184
Ramjohn IA, Murphy PG, Burton TM et al (2012) Survival and rebound of Antillean dry forests:
role of forest fragments. For Ecol Manag 284:124–132
Rao GR (1985) Effect of different levels of nitrogen and phosphorus on growth and chemical
composition of Celtis australis Linn. and Acacia mollissima Wild. M.Sc. thesis, Dr Y.S. Parmar
University of Horticulture and Forestry, Solan, India
Rao MR, Nair PKK, Ong CK (1998) Biological interactions in tropical agroforestry systems.
Agrofor Syst 38:3–50
Rasse DP, Rumpel C, Dignac MD (2005) Is soil carbon mostly root carbon? Mechanisms for a
specific stabilisation. Plant Soil 269:341–356
Rathcke B, Lacey EP (1985) Phenological patterns of terrestrial plants. Annu Rev Ecol Syst
16:179–214
Ratnam J, Sankaran M, Hanan NP et  al (2008) Nutrient resorption patterns of plant functional
groups in a tropical savanna: variation and functional significance. Oecologia 157:141–151
References 227

Ratnam J, Tomlinson KW, Rasquinha DN et al (2016) Savannahs of Asia: antiquity, biogeography,
and an uncertain future. Phil Trans R Soc Lond B 371:20150305
Ratter JA, Ribeiro JF, Bridgewater S (1997) The Brazilian cerrado vegetation and threats to its
biodiversity. Ann Bot 80:223–230
Raven PH (1987) The scope of plant conservation problem worldwide. In: Bramwell D, Hamann
O, Heywood V et  al (eds) Botanic gardens and the world conservation strategy. Academic,
London, pp 19–29
Ray GJ, Brown BJ (1995) Restoring Caribbean dry forests: evaluation of tree propagation tech-
niques. Restor Ecol 3:86–94
Read L, Lawrence D (2003) Litter nutrient dynamics during succession in dry tropical forests of
the Yucatan: regional and seasonal effects. Ecosystems 6:747–761
Rees M (1995) Community structure in sand dune annuals: is seed size weight a key quantity?
J Ecol 83:857–863
Reich PB (1993) Reconciling apparent discrepancies among studies relating life-span, structure
and function of leaves in contrasting plant life forms and climates: “the blind men and the
elephant retold”. Funct Ecol 7:721–725
Reich PB (1995) Phenology of tropical forests: patterns, causes, and consequences. Can J  Bot
73:164–174
Reich PB, Borchert R (1982) Phenology and ecophysiology of the tropical tree Tabebuia neochry-
santha (Bignoniaceae). Ecology 63:294–299
Reich PB, Borchert R (1984) Water stress and tree phenology in a tropical dry forest in the low-
lands of Costa Rica. J Ecol 72:61–74
Reich PB, Borchert R (1988) Changes with leaf age in stomatal function and water status of several
tropical tree species. Biotropica 20:60–69
Reich PB, Uhl C, Walters MB et al (1991) Leaf lifespan as a determinant of leaf structure and func-
tion among 23 tree species in Amazonian forest communities. Oecologia 86:16–24
Reich PB, Walters MB, Ellsworth DS (1992) Leaf lifespan in relation to leaf, plant, and stand
characteristics among diverse ecosystems. Ecol Monogr 62:365–392
Reich PB, Walters MB, Ellsworth DS (1997) From tropics and tundra: global convergence in plant
functioning. Proc Natl Acad Sci USA 94:13730–13734
Reich PB, Walters MB, Ellsworth DS et al (1998a) Relationships of leaf dark respiration to leaf
nitrogen, specific leaf area and leaf life-span: a test across biomes and functional groups.
Oecologia 114:471–482
Reich PB, Tjoelker MG, Walters MB et al (1998b) Close association of RGR, leaf and root mor-
phology, seed mass and shade tolerance in seedlings of nine boreal tree species grown in high
and low light. Funct Ecol 12:327–328
Reich PB, Ellsworth DS, Walters MB et al (1999) Generality of leaf trait relationships: a test across
six biomes. Ecology 80:1955–1969
Reichmann LG, Sala OE, Peters DP (2013) Precipitation legacies in desert grassland primary
production occur through previous-year tiller density. Ecology 94(2):435–443
Reichstein M, Bahn M, Ciais P et  al (2013) Climate extremes and the carbon cycle. Nature
500:287–295
Reid WV (1992) How many species will there be? In: Whitmore TC, Sayer JA (eds) Tropical
deforestation and species extinction. Chapman & Hall, London, pp 55–74
Reid WV, Miller KR (1989) Keeping options alive, the scientific basis for conserving biodiversity.
World Resources Institute, Washington, DC
Rentería LY, Jaramillo VJ (2011) Rainfall drives leaf traits and leaf nutrient resorption in a tropical
dry forest in Mexico. Oecologia 165:201–211
Rentería LY, Jaramillo VJ, Martínez-Yrízar A et al (2005) Nitrogen and phosphorus resorption in
trees of a Mexican tropical dry forest. Trees Struct Funct 19:431–441
Reyes-García C, Andrade JL, Simá JL et al (2012) Sapwood to heartwood ratio affects whole-tree
water use in dry forest legume and non-legume trees. Trees 26:1317–1330
228 References

Reynolds J, Kemp P, Ogle K et al (2004) Modifying the ‘pulse–reserve’ paradigm for deserts of
North America: precipitation pulses, soil water, and plant responses. Oecologia 141:194–210
Richards PW (1996) The tropical rain forest: an ecological study, 2nd edn. Cambridge University
Press, Cambridge
Richards JH, Caldwell MM (1987) Hydraulic lift: substantial nocturnal water transport between
soil layers by Artemesia tridentata roots. Oecologia 73:486–489
Richardson SJ, Peltzer DA, Allen RB et al (2005) Resorption proficiency along a chronosequence:
responses among communities and within species. Ecology 86:20–25
Ries L, Fletcher RJ, Battin J et al (2004) Ecological responses to habitat edges: mechanisms, mod-
els, and variability explained. Annu Rev Ecol Evol Syst 35:491–522
Rincón E, Huante P (1993) Growth responses of tropical deciduous tree seedlings to contrasting
light conditions. Trees Struct Funct 7:202–207
Roa-Fuentes LL, Campo J, Parra V (2012) Plant biomass allocation across a precipitation gradient:
an approach to seasonally dry tropical forest at Yucatan, Mexico. Ecosystems 15:1234–1244
Roa-Fuentes LL, Hidalgo C, Etchevers JD et al (2013) The effects of precipitation regime on soil
carbon pools on the Yucatan Peninsula. J Trop Ecol 29:463–466
Robinson D, Rorison IH (1983) A comparison of the responses of Lolium perenne L., Holcus
lanatus L., and Deschampsia flexuosa (L.) Trin. to a localized supply of nitrogen. New Phytol
94:263–273
Robles-Morua A, Che D, Mayer AS et al (2015) Hydrologic assessment of proposed reservoirs
in the Sonora River Basin, Mexico, under historical and future climate scenarios. Hydrol Sci
J 60:50–66
Rodriguez-Echeverria S, Perez-Fernandez MA (2003) Soil fertility and herb facilitation by Retama
sphaerocarpa. J Veg Sci 14:807–814
Rohr T, Manzoni S, Feng X et al (2013) Effect of rainfall seasonality on carbon storage in tropical
dry ecosystems. J Geophys Res 118:1156–1167
Rossatto DR, Hoffmann WA, Franco AC (2009) Differences in growth patterns between co-occur-
ring forest and savanna trees affect the forest–savanna boundary. Funct Ecol 23:689–698
Roupsard O, Ferhi A, Granier A et al (1999) Reverse phenology and dry-season water uptake by
Faidherbia albida (Del.) A. Chev. in an agro-forestry parkland of Sudanese West Africa. Funct
Ecol 13:460–472
Roy S, Singh JS (1994) Consequences of habitat heterogeneity for availability of nutrients in a dry
tropical forest. J Ecol 82:503–509
Roy S, Singh JS (1995) Seasonal and spatial dynamics of plant-available N and P pools and
N-mineralization in relation to fine roots in a dry tropical forest habitat. Soil Biol Biochem
27:33–40
Rozendaal DMA, Chazdon RL (2015) Demographic drivers of tree biomass change during sec-
ondary succession in northeastern Costa Rica. Ecol Appl 25:506–516
Rozendaal DMA, Hurtado VH, Poorter L (2006) Plasticity in leaf traits of 38 tropical tree species
in response to light: relationships with light demand and adult stature. Funct Ecol 20:207–216
Rudel TK (2005) Tropical forests: regional paths of destruction and regeneration in the late twen-
tieth century. Columbia University Press, New York
Rudel TK (2017) The dynamics of deforestation in the wet and dry tropics: a comparison with
policy implications. Forests 8:108
Rudel TK, Lugo MP, Zichal H (2000) When fields revert to forest: development and spontaneous
reforestation in post-war Puerto Rico. Prof Geogr 52:386–397
Rundel PW, Becker PF (1987) Cambios estacionales en las relaciones hidricias y en la fenologi
vegetativa de plantas del estrata bajo del bosque tropical de la Isla de Barro Colarado, Panama.
Rev Biol Trop l35:71–84
Rundel PW, Boonpragob K (1995) Dry forest ecosystems of Thailand. In: Bullock SH, Mooney
HA, Medina E (eds) Seasonally dry tropical forests. Cambridge University Press, Cambridge,
pp 93–123
References 229

Ryan DE, Bormann FH (1982) Nutrient resorption in northern hardwood forests. Bioscience
32:29–32
Ryser P, Urbas P (2000) Ecological significance of leaf life span among Central European grass
species. Oikos 91:41–50
Rzedowski J (1991) El endemismo en laflora fanerogamica mexicana: una apreciacion analitica
preliminar. Acta Bot Mex 15:47–64
Saatchi S, Houghton R, Santos Alvala Dos R et al (2007) Distribution of aboveground live biomass
in the Amazon basin. Glob Chang Biol 13:816–837
Saatchi SS, Harris NL, Brown S et al (2011) Benchmark map of forest carbon stocks in tropical
regions across three continents. Proc Natl Acad Sci USA 108:9899–9904
Sabey BR, Frederick LR, Bartholomew WB (1959) The formation of nitrate from ammonium-
nitrogen in soils. III. Influence of temperature and initial population of nitrifying organisms on
the maximum rate and delay period. Soil Sci Soc Am Proc 23:462–465
Sagar R, Singh JS (2003) Predominant phenotypic traits of disturbed tropical dry deciduous forest
vegetation in northern India. Comm Ecol 4:63–71
Sagar R, Singh JS (2004) Local plant species depletion in tropical dry deciduous forest of northern
India. Environ Conserv 31:1–8
Sagar R, Singh JS (2005) Structure, diversity and regeneration of tropical dry deciduous forest of
northern India. Biodivers Conserv 14:935–959
Sagar R, Singh JS (2006) Tree density, basal area and species diversity in a disturbed dry tropical
forest of northern India: implications for conservation. Environ Conserv 33:256–262
Sagar R, Raghubanshi AS, Singh JS (2003) Tree species composition, dispersion and diversity
along a disturbance gradient in a dry tropical forest region of India. For Ecol Manag 186:61–71
Sagar R, Raghubanshi AS, Singh JS (2008a) Comparison of community composition and species
diversity of understorey and overstorey tree species in a dry tropical forest of northern India.
J Environ Manag 88:1037–1046
Sagar R, Singh A, Singh JS (2008b) Differential effect of woody plant canopies on species compo-
sition and diversity of ground vegetation: a case study. Trop Ecol 49:189–197
Saha S, Howe HF (2003) Species composition and fire in a dry deciduous forest. Ecology
84:3118–3123
Saha PK, Bhattacharya A, Ganguly SN (1992) Problems with regard to the loss of seed viability of
Shorea robusta Gaertn. F. Indian For 118:70–76
Saitta A, Bernicchia A, Gorjón SP et al (2011) Biodiversity of wood-decay fungi in Italy. Plant
Biosyst 145:958–968
Sakai S (2001) Phenological diversity in tropical forests. Popul Ecol 43:77–86
Sala OE, Parton WJ, Joyce LA et al (1988) Primary production of the central grassland region of
the United States. Ecology 69:40–45
Sala OE, Chapin FS III, Armesto JJ et al (2000) Global biodiversity scenarios for the year 2100.
Science 287:1770–1774
Saldaña-Acosta A, Meave JA, Paz H et al (2008) Variation of functional traits in trees from a bio-
geographically complex Mexican cloud forest. Acta Oecol 34:111–121
Sampaio AB, Scariot A (2011) Edge effect on tree diversity, composition and structure in a decidu-
ous dry forest in central Brazil. Rev Árvore 35:1121–1134
Sampaio E, Salcedo IH, Kauffman JB (1993) Effect of different fire severities on coppicing of
caatinga vegetation in Serra Talhada, Pe, Brazil. Biotropica 25:452–460
Sanches L, Valentini CMA, Júnior OBP et  al (2008) Seasonal and interannual litter dynamics
of a tropical semideciduous forest of the southern Amazon Basin, Brazil. J  Geophys Res
113:G04007
Sanchez PA (1976) Properties and management of soils in the tropics. Wiley, New York
Sánchez-Azofeifa GA, Kalacska M, Quesada M et al (2005) Need for integrated research for a
sustainable future in tropical dry forests. Conserv Biol 19:1–2
Sandquist DR, Cordell S (2007) Functional diversity of carbon-gain, water-use, and leaf-allocation
traits in trees of a threatened lowland dry forest in Hawaii. Am J Bot 94:1459–1469
230 References

Santa Regina I, Rico M, Rapp M et al (1997) Seasonal variation in nutrient concentration in leaves
and branches of Quercus pyrenaica. J Veg Sci 8:651–654
Santantonio D, Herman RK, Overton WS (1977) Root biomass studies in forest ecosystems.
Pedobiologia 17:1–31
Santiago LS (2003) Leaf traits of canopy trees on a precipitation gradient in Panama: integrating
plant physiological ecology and ecosystem science. PhD dissertation, University of Florida,
Florida
Santiago LS, Kitajima K, Wright SJ et al (2004) Coordinated changes in photosynthesis, water
relations and leaf nutritional traits of canopy trees along a precipitation gradient in lowland
tropical forest. Oecologia 139:495–502
Saxena AK, Rao OP, Singh BP (1995) Effect of shade on seedling growth of Dalbergia sissoo,
Acacia catechu and Casuarina equisetifolia. Ann For 3:152–157
Sayer EJ (2006) Using experimental manipulation to assess the roles of leaf litter in the function-
ing of forest ecosystems. Biol Rev 81:1–31
Saynes V, Hidalgo C, Etchevers JD et al (2005) Soil C and N dynamics in primary and secondary
seasonally dry tropical forests in Mexico. Appl Soil Ecol 29:282–289
Scariot A (1999) Forest fragmentation effects on palm diversity in central Amazonia. J  Ecol
87:66–76
Schimel DS (1995) Terrestrial ecosystems and the carbon cycle. Glob Chang Biol 1:77–91
Schimel J, Balser T, Wallenstein M (2007) Microbial stress response physiology and its implica-
tions for ecosystem function. Ecology 88:1386–1394
Schlesinger WH, Andrews JA (2000) Soil respiration and the global carbon cycle. Biogeochemistry
48:7–20
Schliemann SA, Bockheim JG (2011) Methods for studying treefall gaps: a review. For Ecol
Manag 261:1143–1151
Schmidt EL (1982) Nitrification in soil. In: Stevenson FJ (ed) Nitrogen in agricultural soils.
American Society of Agronomy, Madison, pp 253–288
Schmidt MWI, Torn MS, Abiven S et al (2011) Persistence of soil organic matter as an ecosystem
property. Nature 478:49–56
Schnitzer SA, Carson WP (2001) Treefall gaps and the maintenance of species diversity in a tropi-
cal forest. Ecology 82:913–919
Schöngart J, Piedade MTF, Ludwigshausen S et al (2002) Phenology and stem-growth periodicity
of the tree species in Amazonian floodplain forests. J Trop Ecol 18:581–597
Schrire BD, Lavin M, Lewis GP (2005a) Global distribution patterns of the Leguminosae: insights
from recent phylogenies. Biol Skr 55:375–422
Schrire BD, Lewis GP, Lavin M (2005b) Biogeography of the Leguminosae. In: Lewis GP, Schrire
BD, Lock MD et al (eds) Legumes of the world. Royal Botanical Garden, Kew, pp 21–54
Schulze ED, Gebauer G, Ziegler H et al (1991) Estimates of nitrogen fixation by trees on an aridity
gradient. Oecologia 88:451–455
Schuur EAG (2001) The effect of water on decomposition dynamics in mesic to wet Hawaiian
montane forests. Ecosystems 4:259–273
Schuur EAG (2003) Productivity and global climate revisited: the sensitivity of tropical forest
growth to precipitation. Ecology 84:1165–1170
Schuur EAG, Matson PA (2001) Net primary productivity and nutrient cycling across a mesic to
wet precipitation gradient in Hawaiian montane forest. Oecologia 128:431–442
Scifres CJ (1974) Salient aspects of huisache seed germination. Southwest Nat 18:383–392
Scott RL, Jenerette GD, Potts DL et al (2009) Effects of seasonal drought on net carbon dioxide
exchange from a woody-plant-encroached semiarid grassland. J Geophys Res 114:G04004
Seastedt TR, Crossley DA Jr, Meentemeyer V et al (1983) A two-year study of leaf litter decom-
position as related to macroclimatic factors and microarthropod abundance in the southern
Appalachians. Holarct Ecol 6:11–16
Seghieri J, Floret C, Pontanier P (1995) Plant phenology in relation to water availability: herba-
ceous and woody species in the savannas of northern Cameroon. J Trop Ecol 11:237–254
References 231

Seghieri J, Vescovo A, Padel K et  al (2009) Relationships between climate, soil moisture and
phenology of the woody cover in two sites located along the West African latitudinal gradient.
J Hydrol 375:78–89
Segura G, Balvanera P, Durán E et al (2003) Tree community structure and stem mortality along a
water availability gradient in a Mexican tropical dry forest. Plant Ecol 169:259–271
Selwyn MA, Parthasarathy N (2006) Reproductive traits and phenology of plants in tropical dry
evergreen forest on the coromandel coast of India. Biodivers Conserv 15:3207–3234
Semb G, Robinson JBD (1969) The natural nitrogen flush in different arable soils and climates in
East Africa. East Afr Agric For J 34:350–370
Semenova GV, van der Maarel E (2000) Plant functional types–a strategic perspective. J Veg Sci
11:917–922
Sharma GP, Raghubanshi AS (2006) Tree population structure, regeneration and expected future
composition at different levels of Lantana camara L. invasion in the Vindhyan tropical dry
deciduous forest of India. Lyonia 11:25–37
Sharma GP, Raghubanshi AS (2007) Effect of Lantana camara L. cover on local depletion of
tree population in the Vindhyan tropical dry deciduous forest of India. Appl Ecol Environ Res
5:109–121
Sharma GP, Singh JS, Raghubanshi AS (2005a) Plant invasions: emerging trends and future impli-
cations. Curr Sci 88:726–734
Sharma GP, Raghubanshi AS, Singh JS (2005b) Lantana invasion: an overview. Weed Biol Manag
5:157–165
Sheil D (1999) Tropical forest diversity, environmental change and species augmentation: after
intermediate disturbance hypothesis. J Veg Sci 10:851–860
Silver WL, Miya RK (2001) Global patterns in root decomposition: comparisons of climate and
litter quality. Oecologia 129:407–419
Simberloff D (1986) Are we on the verge of a mass extinction in tropical rainforests? In: Elliott DK
(ed) Dynamics of extinction. Wiley, New York, pp 165–180
Simberloff D, Parker IM, Windle PM (2005) Introduced species policy, management and future
research needs. Front Ecol Environ 3:12–20
Singh JS (2002) The biodiversity crisis: a multifaceted review. Curr Sci 82:638–647
Singh JS, Chaturvedi RK (2017) Diversity of ecosystem types in India: a review. Proc Ind Natl Sci
Acad – INSA 83(3):569–594
Singh KP, Kushwaha CP (2005) Emerging paradigms of tree phenology in dry tropics. Curr Sci
89:964–975
Singh KP, Kushwaha CP (2006) Diversity of flowering and fruiting phenology of trees in a tropical
deciduous forest in India. Ann Bot 97:265–276
Singh KP, Misra R (1979) Structure and functioning of natural, modified and silvicultural eco-
systems, Eastern Uttar Pradesh, Final Technical Report (1975–1978), MAB research project.
Banaras Hindu University, Varanasi, p 160
Singh J, Mukherjee IN (1986) A study of soil Pauropoda and Symphyla from different fields at
Varanasi. J Soil Biol Ecol 6:104–108
Singh L, Singh JS (1991a) Species structure, dry matter dynamics and carbon flux of a dry tropical
forest in India. Ann Bot 68:263–273
Singh L, Singh JS (1991b) Storage and flux of nutrients in a dry tropical forest in India. Ann Bot
68:275–284
Singh JS, Singh VK (1992a) Phenology of seasonally dry tropical forest. Curr Sci 63:684–688
Singh VP, Singh JS (1992b) Energetics and environmental costs of agriculture in a dry tropical
region of India. Environ Manag 16:495–503
Singh H, Singh KP (1993a) Effect of residue placement and chemical fertilizer on soil microbial
biomass under tropical dryland cultivation. Biol Fertil Soils 16:275–281
Singh L, Singh JS (1993b) Importance of short-lived components of a dry tropical forest for bio-
mass production and nutrient cycling. J Veg Sci 4:681–686
232 References

Singh S, Singh JS (1995) Microbial biomass associated with water stable aggregates in forest,
savanna and cropland soils of a seasonally dry tropical region, India. Soil Biol Biochem
27:1027–1033
Singh S, Singh JS (1996) Water-stable aggregates and associated organic matter in forest, savanna,
and cropland soils of a seasonally dry tropical region, India. Biol Fertil Soils 22:76–82
Singh JS, Singh KD (2011) Silviculture of dry deciduous forests, India. In: Günter S, Weber M,
Stimm B et al (eds) Silviculture in the tropics. Springer-Verlag, Berlin, pp 273–283
Singh JS, Singh KP, Yadava PS (1979) Ecosystem synthesis. In: Coupland RT (ed) Grassland
ecosystems of the world: analysis of grasslands and their uses, International biological pro-
gramme, vol 18. Cambridge University Press, London, pp 231–239
Singh JS, Rawat YS, Chaturvedi OP (1984) Replacement of oak forest with pine in the Himalaya
affects the nitrogen cycle. Nature 311:54–56
Singh JS, Hanxi Y, Sajise PE (1985) Structural and functional aspects of Indian and southeast
Asian savanna ecosystems. In: Tothill JC, Mott JJ (eds) Ecology and management of world’s
savannas. The Australian Academy of Sciences, Canberra, p 384
Singh JS, Raghubashi AS, Singh RS et al (1989) Microbial biomass acts as a source of plant nutri-
ents in dry tropical forest and savanna. Nature 338:499–500
Singh KP, Singh JS (1988) Certain structural and functional aspects of dry tropical forest and
savanna. Int J Ecol Environ Sci 14:31–45
Singh RS, Srivastava SC, Raghubanshi AS et  al (1991a) Microbial C, N and P in dry tropical
savanna: effects of burning and grazing. J Appl Ecol 28:869–878
Singh JS, Singh L, Pandey CB (1991b) Savannization of dry tropical forest increases carbon flux
relative to storage. Curr Sci 61:477–480
Singh L, Singh KP, Singh JS (1992) Biomass, productivity and nutrient cycling in four contrast-
ing forest ecosystems of India. In: Singh KP, Singh JS (eds) Tropical ecosystems: ecology and
management. Wiley Eastern Limited, New Delhi, pp 415–430
Singh KP, Shukla AN, Singh JS (2010) State-level inventory of invasive alien plants, their source
regions and use potential. Curr Sci 99:107–114
Sivasupiramanium S, Kasaeng AK, Shelton HM (1988) Effect of nitrogen and lime on growth of
Leucaena leucocephala. Aust J Exp Agri 26:23–29
Skoglund J (1992) The role of seed banks in vegetation dynamics and restoration of dry tropical
ecosystems. J Veg Sci 3:357–360
Slocum MG, Horvitz CC (2000) Seed arrival under different genera of trees in a neotropical pas-
ture. Plant Ecol 149:51–62
Snow DW (1971) Evolutionary aspect of fruit eating by birds. Ibis 113:194–202
Sobrado MA (1986) Aspects of tissue water relations and seasonal changes in leaf water poten-
tial components of evergreen and deciduous species coexisting in tropical forests. Oecologia
68:413–416
Sobrado MA (1991) Cost-benefit relationships in deciduous and evergreen leaves of tropical dry
forest species. Funct Ecol 5:608–616
Sobrado MA (1993) Trade-off between water transport efficiency and leaf life-span in a tropical
dry forest. Oecologia 96:19–23
Sobrado MA (1997) Embolism vulnerability in drought-deciduous and evergreen species of a
tropical dry forest. Acta Oecol 18:383–391
Sobrado MA, Cuenca G (1979) Aspectos del uso de agua de especies deciduas y siempreverdes en
un bosque seco tropical de Venezuela. Acta Cient Venez 30:302–308
Sorensen LH (1974) Rate of decomposition of organic matter in soil as influenced by repeated
air-drying-rewetting and repeated additions of organic matter. Soil Biol Biochem 6:287–292
Soriano D, Orozco-Segovia A, Márquez-Guzmán J et al (2011) Seed reserve composition in 19
tree species of a tropical deciduous forest in Mexico and its relationship to seed germination
and seedling growth. Ann Bot 107:939–951
Sparks JP, Black RA (1999) Regulation of water loss in populations of Populus trichocarpa: the
role of stomatal control in preventing xylem cavitation. Tree Physiol 19:453–459
References 233

Sperry JS (2003) Evolution of water transport and xylem structure. Int J Plant Sci 164:S115–S127
Sprent JI (1987) The ecology of the nitrogen cycle. Cambridge University Press, New York
Srimathi P, Rai RSV, Surendran C (1991) Studies on the effect of seed coat colour and seed size on
seed germination in Acacia mellifera. Ind J For 14:5–7
Srivastava SC (1989) Variation in microbial C. N and P in soil of selected terrestrial ecosystems of
dry tropical forest environment. Ph.D. thesis, Department of Botany, Banaras Hindu University,
Varanasi, India
Srivastava SC, Singh JS (1988) Carbon and phosphorus in the soil biomass of some tropical soils
of India. Soil Biol Biochem 20:743–747
Srivastava SC, Singh JS (1989) Effect of cultivation on microbial biomass C and N of dry tropical
forest soil. Biol Fertil Soils 8:343–348
Srivastava SC, Singh JS (1991) Microbial C, N and P in dry tropical forest Soils: effects of alter-
nate land-uses and nutrient flux. Soil Biol Biochem 23:117–124
Srivastava SC, Jha AK, Singh JS (1989) Changes with time in soil biomass C, N and P of mine
spoils in a dry tropical environment. Can J Soil Sci 49:849–855
Staaf H (1982) Plant nutrient changes in beech leaves during senescence as influenced by site
characteristics. Oecologia Plant 3:161–170
Staaf H, Berg B (1981) Plant litter input to soil. In: Clark FE, Rosswall T (eds) Terrestrial nitro-
gen cycles, Ecological bulletin 33. Swedish Natural Science Research Council, Stockholm,
pp 147–162
Standiford R, McDougald N, Phillips R et al (1991) South Sierra oak regeneration weak in sapling
stage. Calif Agric 45:12–14
Stark JM, Firestone MK (1995) Mechanisms for soil moisture effects on activity of nitrifying
bacteria. Appl Environ Microbiol 61:218–221
Staver AC, Archibald S, Levin SA (2011) The global extent and determinants of savanna and forest
as alternative biome states. Science 334:230–232
Steffens T, Lehman SM (2016) Factors determining Microcebus abundance in a fragmented
landscape. In: Lehman SM, Radespiel U, Zimmermann E (eds) Dwarf and mouse lemurs
of Madagascar: biology, behavior, and conservation biogeography of the Cheirogaleidae.
Cambridge University Press, Cambridge, pp 477–497
Steininger MK, Tucker CJ, Ersts P et al (2001) Clearance and fragmentation of tropical deciduous
forest in the Tierras Bajas, Santa Cruz, Bolivia. Conserv Biol 15:856–866
Stephan HH, Sylvain C, Sandra B et al (2011) Leaf traits and decomposition in tropical rainforests:
revisiting some commonly held views and towards a new hypothesis. New Phytol 189:950–965
Sterck F, Markesteijn L, Schieving F et al (2011) Functional traits determine trade-offs and niches
in a tropical forest community. Proc Natl Acad Sci USA 108:20627–20632
Stevenson FJ (1986) Cycles of soil carbon, nitrogen, phosphorus, sulfur micronutrients. Willey,
New York
Stiles FG (1977) Coadapted competitors: the flowering seasons of hummingbird-pollinated plants
in a tropical forest. Science 198:1177–1178
Stitt M (1991) Rising CO2 levels and their potential significance for carbon flow in photosynthetic
cells. Plant Cell Environ 14:741–762
Stroo HF, Jenks EM (1982) Enzymatic activities and respiration in mine soil. Soil Sci Soc Am
J 46:548–553
Struhsaker TT (1997) Ecology of an African rain forest: logging in Kibale and the conflict between
conservation and exploitation. University Press of Florida, Gainesville
Suding KN, Miller AE, Bechtold H et al (2006) The consequence of species loss on ecosystem
nitrogen cycling depends on community compensation. Oecologia 149:141–149
Sumina OI (1994) Plant communities on anthropogenically disturbed sites on Chukotka Peninsula,
Russia. J Veg Sci 5:885–896
Sunderland T, Apgaua D, Baldauf C et al (2015) Global dry forests: a prologue. Int For Rev 17:1–9
234 References

Suresh HS, Dattaraja HS, Sukumar R (2010) Relationship between annual rainfall and tree mortal-
ity in a tropical dry forest: results of a 19-year study at Mudumalai, southern India. For Ecol
Manag 259:762–769
Swaine MD (1992) Characteristics of dry forests in West Africa and the influence of fire. J Veg
Sci 3:365–374
Swaine MD, Lieberman D, Hall JB (1990) Structure and dynamics of a tropical dry forest in
Ghana. Vegetatio 88:31–51
Swamy PS, Sundarapandian SM, Chandrasekar P et  al (2000) Plant species diversity and tree
population structure of a humid tropical forest in Tamil Nadu, India. Biodivers Conserv
9:1643–1669
Swiecki TJ, Bernhardt EA, Drake C (1993) Factors affecting blue oak sapling recruitment
and regeneration. A report to the California Department of Forestry and Fire Protection,
Sacramento, CA, p 131
Swift MJ, Heal OW, Anderson JM (1979) Decomposition in terrestrial ecosystems. Blackwell
Science Publication, Oxford
Tateno R, Takeda H (2003) Forest structure and tree species distribution in relation to topography-
mediated heterogeneity of soil nitrogen and light at the forest floor. Ecol Res 18:559–571
Teketay D (1994) Germination ecology of two endemic multipurpose species of Erythrina from
Ethiopia. For Ecol Manag 65:81–87
Teketay D (1996) Germination ecology of twelve indigenous and eight exotic multipurpose legu-
minous species from Ethiopia. For Ecol Manag 80:209–223
Teketay D, Granstrom A (1997) Germination ecology of forest species from the highlands of
Ethiopia. J Trop Ecol 14:793–803
Thapliyal RC, Connor KF (1997) Effects of accelerated ageing on viability, leachate exudation and
fatty acid content of Dalbergia sissoo Roxb. seeds. Seed Sci Technol 25:311–319
Tilman D (1987a) On the meaning of competition and the mechanisms of competitive superiority.
Funct Ecol 1:304–315
Tilman D (1987b) Secondary succession and the pattern of plant dominance along experimental
nitrogen gradients. Ecol Monogr 57:2179–2191
Tisdall JM, Oades JM (1982) Organic matter and water-stable aggregates in soils. J  Soil Sci
33:141–163
Toledo M, Poorter L, Pena-Claros M et al (2011) Climate is a stronger driver of tree and forest
growth rates than soil and disturbance. J Ecol 99:254–264
Top N, Mizoue N, Ito S et al (2009) Effects of population density on forest structure and species
richness and diversity of trees in Kampong Thom Province, Cambodia. Biodivers Conserv
18:717–738
Tosi JA Jr, Voertman RF (1964) Some environmental factors in the economic development of the
tropics. Econ Geogr 40:189–205
Trejo I (2005) Análisis de la diversidad de la selva baja caducifolia en México. In: Halffter G,
Soberón J, Koleff P et al (eds) Sobre Diversidad Biológica: el Significado de las Diversidades
Alfa, Beta y Gamma. M3M: Monografías Tercer Milenio, vol 4. Sociedad Entomológica
Aragonesa, Zaragoza, pp 111–122
Trejo I, Dirzo R (2000) Deforestation of seasonally dry tropical forest: a national and local analysis
in Mexico. Biol Conserv 94:133–142
Trejo I, Dirzo R (2002) Floristic diversity of Mexican seasonally dry tropical forests. Biodivers
Conserv 11:2063–2084
Troup RS (1921) The silviculture of Indian trees, vol I–III. Clarendon Press, Oxford
Turner IM, Corlett RT (1996) The conservation value of small, isolated fragments of lowland tropi-
cal rain forest. Trends Ecol Evol 8:330–333
Turton SM, Duff GA (1992) Light environments and floristic composition across an open forest-
rainforest boundary in northeastern Queensland. Aust J Ecol 17:415–423
Tyree MT, Davis SD, Cochard H (1994) Biophysical perspectives of xylem evolution: is there a
tradeoff of hydraulic efficiency for vulnerability to dysfunction? IAWA J 115:335–360
References 235

Tyrell JG, Coe MJ (1974) The rainfall regime of Tsavo National Park, Kenya and its potential
phenological significance. J Biogeogr 1:187–192
Uhl C, Buschbacher R, Serrao EAS (1988) Abandoned pastures in eastern Amazonia. I. Patterns of
plant succession. J Ecol 76:663–681
UNEP (2001) India: state of the environment – 2001. United Nations Environment Programme,
Regional Resource Centre for Asia and the Pacific, UNEP RRC.AP
Unger S, Máguas C, Pereira J et al (2010) The influence of precipitation pulses on soil respiration-
assessing the ‘Birch effect’ by stable carbon isotopes. Soil Biol Biochem 42:1800–1810
Upadhyay MD, Srivastava SCN (1980) Working plan, Obra Forest Division, South Circle,
Uttar Pradesh from 1980–1981 to 1989–1990, Working plan circle 2. Uttar Pradesh Forest
Department, Nainital
Uriarte M, Canham CD, Thompson J et al (2009) Natural disturbance and human land use as deter-
minants of tropical forest dynamics: results from a forest simulator. Ecol Monogr 79:423–443
Valencia R, Balslev H, Pazy Mino CG (1994) High tree alpha-diversity in Amazonian Ecuador.
Biodivers Conserv 3:21–28
Valentini R, Matteucci G, Dolman AJ et al (2000) Respiration as the main determinant of carbon
balance in European forests. Nature 404:861–865
Van Bloem S, Lugo A, Murphy P (2006) Structural response of Caribbean dry forests to hurricane
winds: a case study from Guánica Forest, Puerto Rico. J Biogeogr 33:517–523
van Heerwaarden LM, Toet S, Aerts R (2003) Current measures of nutrient resorption efficiency
lead to a substantial underestimation of real resorption efficiency: facts and solutions. Oikos
101:664–669
van Schaik CP, Terborgh JW, Wright SJ (1993) The phenology of tropical forests: adaptive signifi-
cance and consequences for primary consumers. Annu Rev Ecol Syst 24:353–377
Van Veen JA, Paul EA (1981) Organic carbon dynamics in grassland soils. 1. Background informa-
tion and computer simulation. Can J Soil Sci 61:185–201
van Veen JA, Ladd JN, Frissel MJ (1984) Modelling C and N turnover through the microbial bio-
mass in soil. Plant Soil 76:257–274
Vargas-Rodriguez YL, Vázquez-García JA, Williamson GB (2005) Environmental correlates of
tree and seedling-sapling distributions in a Mexican tropical dry forest. Plant Ecol 180:117–134
Vasconcelos SS, Zarin DJ, Silva MB et al (2007) Leaf decomposition in a dry season irrigation
experiment in Eastern Amazonian forest regrowth. Biotropica 35:593–600
Vasseur DA, Fox JW, Gonzalez A et al (2014) Synchronous dynamics of zooplankton competitors
prevail in temperate lake ecosystems. Proc R Soc B Biol Sci 281:20140633
Venable DF, Brown JS (1988) The selective interactions of dispersal, dormancy and seed size as
adaptations for reducing risk in variable environments. Am Nat 131:360–384
Vendramini F, Díaz S, Gurvich DE et al (2002) Leaf traits as indicators of resource-use strategy in
floras with succulent species. New Phytol 154:147–157
Verduzco VS, Garatuza-Payán J, Yépez EA et al (2015) Variations of net ecosystem production due
to seasonal precipitation differences in a tropical dry forest of northwest Mexico. J Geophys
Res Biogeosci 120:2081–2094
Vesk PA, Westoby M (2004) Sprouting ability across diverse disturbances and vegetation types
worldwide. J Ecol 92:310–320
Vetaas OR (1992) Micro-site effects of tree and shrubs in a dry savanna. J Veg Sci 3:337–344
Vicente-Serrano SM, Gouveia C, Camarero JJ et al (2013) Response of vegetation to drought tim-
escales across global land biomes. Proc Natl Acad Sci USA 110:52–57
Vico GT, Sally EM, Stefano M et al (2014) Climatic, ecophysiological, and phonological controls
on plant ecohydrological strategies in seasonally dry ecosystems. Ecohydrology 8:660–681
Vieira DLM, Scariot A (2006) Principles of natural regeneration of tropical dry forests for restora-
tion. Restor Ecol 14:11–20
Vieira J, Rossi S, Campelo F et al (2013) Seasonal and daily cycles of stem radial variation of
Pinus pinaster in a drought-prone environment. Agric For Meteorol 180:173–181
236 References

Villar R, Merino J (2001) Comparison of leaf construction costs in woody species with differing
leaf life-spans in contrasting ecosystems. New Phytol 151:213–226
Visser S, Griffiths CL, Parkinson D (1983) Effect of surface mining on the microbiology of a prai-
rie site in Alberta, Canada. Can J Soil Sci 63:177–183
Visser ME, Caro SP, van Oers K et al (2010) Phenology, seasonal timing and circannual rhythms:
towards a unified framework. Philos Trans R Soc B 365:3113–3127
Vitousek PM (1984) Litterfall, nutrient cycling and nutrient limitation in tropical forests. Ecology
65:285–298
Vitousek PM (1998) Foliar and litter nutrients, nutrient resorption, and decomposition in Hawaiian
Metrosideros polymorpha. Ecosystems 1:401–407
Vitousek PM, Matson PA (1982) Causes of delayed nitrate production in two Indiana forests. For
Sci 31:122–131
Vitousek PM, Matson PA (1988) Nitrogen transformations in a range of tropical forest soils. Soil
Biol Biochem 20:361–367
Vitousek PM, Porder S, Houlton BZ et al (2010) Terrestrial phosphorus limitation: mechanisms,
implications, and nitrogen-phosphorus interactions. Ecol Appl 20:5–15
Vogel CS, Curtis PS, Thomas RB (1997) Growth and nitrogen accretion of dinitrogen fixing Alnus
glutinosa (L.) Gaertn. under elevated carbon dioxide. Plant Ecol 130:63–70
Vogt KA, Grier CC, Meier CE (1982) Mycorrhizal role in net primary production and nutrient
cycling in Abies amabilis ecosystems in Western Washington. Ecology 63:370–380
Vogt KA, Grier CC, Vogt DJ (1986) Production, turnover and nutrient dynamics of above- and
below-ground detritus of world forests. Adv Ecol Res 15:303–377
Von Willert DJ, Eller BM, Werger MJ et  al (1990) Desert succulents and their life strategies.
Vegetatio 90:133–143
Von Willert DJ, Eller BM, Werger MJA et al (1992) Life strategies of succulents in deserts with
special reference to the Namib Desert. Cambridge University Press, Cambridge
Wagner HH, Edwards PJ (2001) Quantifying habitat specificity to assess the contribution of a
patch to species richness at a landscape scale. Landsc Ecol 16:121–131
Wali MK, Evrendilek F, West TO et al (1999) Assessing terrestrial ecosystem sustainability: use-
fulness of regional carbon and nitrogen models. Nat Res 35:21–33
Walter H (1971) Ecology of tropical and subtropical vegetation. Oliver and Boyd, Edinburgh
Walters MB, Reich PB (1996) Are shade tolerance survival and growth linked? Low light and
nitrogen effects on hardwood seedlings. Ecology 77:841–853
Wang L, Manzoni S, Ravi S et al (2015) Dynamic interactions of ecohydrological and biogeo-
chemical processes in water-limited systems. Ecosphere 6:133
Wareing PF, Patrick J  (1975) Source sink relation and partition of assimilation in the plant.
In: Cooper JP (ed) Photosynthesis and productivity in different environments. Cambridge
University Press, Cambridge, pp 481–499
Waring BG, Powers JS (2016) Unravelling the mechanisms underlying pulse dynamics of soil
respiration in tropical dry forests. Environ Res Lett 11:105005
Waring B, Adams R, Branco S et al (2016) Scale-dependent variation in nitrogen cycling and soil
fungal communities along gradients of forest composition and age in regenerating tropical dry
forests. New Phytol 209:845–854
Warren JM, Jensen AM, Medlyn BE et  al (2015) Carbon dioxide stimulation of photosynthe-
sis in Liquidambar styraciflua is not sustained during a 12-year field experiment. AoB Plants
7:plu074
Watanabe M, Watanabe Y, Kitaoka S et al (2011) Growth and photosynthetic traits of hybrid larch
F1 (Larix gmelinii var. japonica x L. kaempferi) under elevated CO2 concentration with low
nutrient availability. Tree Physiol 31:965–975
Watson SW, Valos FW, Waterbury JB (1981) The family nitrobacteraceae. In: Starr MP, Stolp H,
Truper HG et al (eds) The prokaryotes. Springer-Verlag, Berlin, pp 1005–1022
Weiher E, van der Werf A, Thompson K et al (1999) Challenging Theophrastus: a common core
list of plant traits for functional ecology. J Veg Sci 10:609–620
References 237

Weisberg PJ, Bugmann H (2003) Forest dynamics and ungulate herbivory: from leaf to landscape.
For Ecol Manag 181:1–12
Weisberg PJ, Bonavia F, Bugmann H (2005) Modeling the interacting effects of browsing and
shading on mountain forest tree regeneration (Picea abies). Ecol Model 185:213–230
Weltzin JF, Coughenour MB (1990) Savanna tree influence on understorey vegetation and soil
nutrients in north-western Kenya. J Veg Sci 1:325–334
Weltzin JF, Loik ME, Schwinning S et al (2003) Assessing the response of terrestrial ecosystems
to potential changes in precipitation. Bioscience 53:941–952
Werner P (1984) Changes in soil properties during tropical wet forest succession in Costa Rica.
Biotropica 16:43–50
Werner FA, Gradstein SR (2009) Diversity of dry forest epiphytes along a gradient of human dis-
turbance in the tropical Andes. J Veg Sci 20:59–68
Westoby M (1998) A leaf-height-seed (LHS) plant ecology strategy scheme. Plant Soil 199:213–227
Westoby M, Wright IJ (2006) Land-plant ecology on the basis of functional traits. Trends Ecol
Evol 21:261–268
Westoby M, Jurado E, Leishman M (1992) Comparative evolutionary ecology of seed size. Trends
Ecol Evol 7:368–372
Wheelwright NT (1985) Competition for dispersers, and the timing of flowering and fruiting in a
guild of tropical trees. Oikos 4:65–77
Whigham DF, Dickinson MB, Brokaw NVT (1999) Tropical forest treefalls and windstorms. In:
Walker L (ed) Ecosystems of disturbed ground. Elsevier, Amsterdam, pp 223–252
Whigham DF, Olmsted I, Cabrera Cano E et  al (2003) Impacts of hurricanes on the forests of
Quintana Roo, Yucatan Peninsula, Mexico. In: Gomez-Pompa A, Allen MF, Fedick SL et al
(eds) The lowland maya area: three millennia at the human-wildland interface. The Haworth
Press, New York, pp 193–213
White PS (1979) Pattern, process and natural disturbance in vegetation. Bot Rev 45:229–299
Whitmore TC (1975) Tropical rain forests of the Far East. Clarendon Press, Oxford
Wickens GE (1969) A study of Acacia albida Del. (Mimosoideae). Kew Bull 23:181–202
Wieder RK, Wright SJ (1995) Tropical forest litter dynamics and dry season irrigation on Barro
Colorado Island, Panama. Ecology 76:1971–1979
Wieder WR, Cleveland CC, Townsend AR (2009) Controls over litter decomposition in wet tropi-
cal forests. Ecology 90:3333–3341
Wieder WR, Cleveland CC, Smith WK et al (2015) Future productivity and carbon storage limited
by terrestrial nutrient availability. Nat Geosci 8:441–445
Wijdeven SMJ, Kuzee ME (2000) Seed availability as a limiting factor in forest recovery processes
in Costa Rica. Restor Ecol 8:414–424
Wikander T (1984) Mechanismos de dispersion de diasporas de una selva decidua en Venezuela.
Biotropica 16:276–283
Williams CA, Hanan N, Scholes RJ et al (2009) Complexity in water and carbon dioxide fluxes
following rain pulses in an African savanna. Oecologia 161:469–480
Williamson M (1996) Biological invasion. Chapman and Hall, London
Willson MF, Crome FHJ (1989) Patterns of seed rain at the edge of a tropical Queensland rain-
forest. J Trop Ecol 5:301–308
Wilson PJ, Thompson K, Hodgson JG (1999) Specific leaf area and leaf dry matter content as
alternative predictors of plant strategies. New Phytol 143:155–162
Wolkovich EM, Cook BI, Davies TJ (2014a) Progress towards an interdisciplinary science of
plant phenology: building predictions across space, time and species diversity. New Phytol
201:1156–1162
Wolkovich EM, Cook BI, McLauchlan KK et al (2014b) Temporal ecology in the Anthropocene.
Ecol Lett 17:1365–1379
Worbes M (1995) How to measure growth dynamics in tropical trees – a review. IAWA J 16:337–351
Worbes M, Blanchart S, Fichtler E (2013) Relations between water balance, wood traits and phe-
nological behavior of tree species from a tropical dry forest in Costa Rica–a multifactorial
study. Tree Physiol 33:527–536
238 References

Wright SJ (1991) Seasonal drought and the phenology of understory shrubs in a tropical moist
forest. Ecology 72:1643–1657
Wright SJ, Cornejo FH (1990) Seasonal drought and leaf fall in a tropical forest. Ecology
71:1165–1175
Wright RA, Wein RW, Dancik BP (1992) Population differentiation in seedling root size between
adjacent stands of jack pine. For Sci 38:777–785
Wright IJ, Westoby M (2003) Nutrient concentration, resorption and lifespan: leaf traits of
Australian sclerophyll species. Funct Ecol 17:10–19
Wright IJ, Reich PB, Westoby M et al (2004a) The world-wide leaf economics spectrum. Nature
428:821–827
Wright IJ, Reich PB, Cornelissen JHC et al (2004b) Assessing the generality of global leaf trait
relationships. New Phytol 166:485–496
Wright IJ, Ackerly DD, Bongers F et al (2007) Relationships among ecologically important dimen-
sions of plant trait variation in seven Neotropical forests. Ann Bot 99:1003–1015
Wunderle JM Jr (1997) The role of animal seed dispersal in accelerating native forest regeneration
on degraded tropical lands. For Ecol Manag 99:223–236
Xiang SR, Doyle A, Holden P et al (2008) Drying and rewetting effects on C and N mineralization
and microbial activity in surface and subsurface California grassland soils. Soil Biol Biochem
40:2281–2289
Xuluc-Tolosa FJ, Vester HFM, Ramírez-Marcial N et al (2003) Leaf litter decomposition of tree
species in three successional phases of tropical dry secondary forest in Campeche, Mexico. For
Ecol Manag 174:401–412
Yan L, Chen S, Xia J et al (2014) Precipitation regime shift enhanced the rain pulse effect on soil
respiration in a semi-arid steppe. PLoS One 9:e104217
Yang S, Albert R, Carlo TA (2013) Transience and constancy of interactions in a plant-frugivore
network. Ecosphere 4:147
Yavitt J, Wright J (2001) Drought and irrigation effects on fine root dynamics in a tropical moist
forest, Panama. Biotropica 33:421–434
Yépez E, Scott R, Cable W et al (2007) Intraseasonal variation in water and carbon dioxide flux
components in a semiarid riparian woodland. Ecosystems 10:1100–1115
Yuan Z, Chen HYH (2009a) Global trends in senesced-leaf nitrogen and phosphorus. Glob Ecol
Biogeogr 18:532–542
Yuan Z, Chen HYH (2009b) Global-scale patterns of nutrient resorption associated with latitude,
temperature and precipitation. Glob Ecol Biogeogr 18:11–18
Zak DR, Pregitzer KS, Curtis PS et al (2000) Atmospheric CO2, soil N availability and allocation
of biomass and nitrogen by Populus tremuloides. Ecol Appl 10:34–46
Zanetti S, Hartwig UA, Luschner A et al (1996) Stimulation of symbiotic N atmospheric pCO2 in
a grassland ecosystem. Plant Physiol 112:575–583
Zhang YJ, Meinzer FC, Hao GY et al (2009) Size-dependent mortality in a Neotropical savanna
tree: the role of height-related adjustments in hydraulic architecture and carbon allocation.
Plant Cell Environ 32:1456–1466
Zheng S, Fu C, Xu X et al (2002) The Asian nitrogen cycle case study. Ambio 31:79–87
Zida D, Sawadogo L, Tigabu M et al (2007) Dynamics of sapling population in savanna woodlands
of Burkina Faso subjected to grazing, early fire and selective tree cutting for a decade. For Ecol
Manag 243:102–115
Zimmerman JK, Pascarella JB, Aide TM (2000) Barriers to forest regeneration in an abandoned
pasture in Puerto Rico. Restor Ecol 8:350–360
Zimmerman B, Elsenbeer H, Moraes JM (2006) The influence of land-use changes on soil hydrau-
lic properties: implications for runoff generation. For Ecol Manag 222:29–38
Zimmerman JK, Wright SJ, Calderón O et al (2007) Flowering and fruiting phenologies of sea-
sonal and aseasonal Neotropical forests: the role of annual changes in irradiance. J Trop Ecol
23:231–251
Zobel K, Zobel M, Rosen E (1994) An experimental test of diversity maintenance mechanisms,
by a species removal experiment in a species rich wooded meadow. Folia Geobot Phytotaxon
29:449–457

Anda mungkin juga menyukai