Sei sulla pagina 1di 285

Sustainable Development and Biodiversity 5

N. Parthasarathy Editor

Biodiversity
of Lianas

123
Sustainable Development and Biodiversity

Volume 5

Series editor
Kishan Gopal Ramawat
M.L. Sukhadia University
Botany Department
Udaipur, Rajasthan, India
This book series provides complete, comprehensive and broad subject based reviews
about existing biodiversity of different habitats and conservation strategies in the
framework of different technologies, ecosystem diversity, and genetic diversity.
The ways by which these resources are used with sustainable management and
replenishment are also dealt with. The topics of interest include but are not restricted
only to sustainable development of various ecosystems and conservation of hotspots,
traditional methods and role of local people, threatened and endangered species,
global climate change and effect on biodiversity, invasive species, impact of various
activities on biodiversity, biodiversity conservation in sustaining livelihoods and
reducing poverty, and technologies available and required. The books in this series
will be useful to botanists, environmentalists, marine biologists, policy makers,
conservationists, and NGOs working for environment protection.

More information about this series at http://www.springer.com/series/11920


N. Parthasarathy
Editor

Biodiversity of Lianas
Editor
N. Parthasarathy
Ecology and Environmental Sciences
Pondicherry University
Puducherry, India

ISSN 2352-474X ISSN 2352-4758 (electronic)


Sustainable Development and Biodiversity
ISBN 978-3-319-14591-4 ISBN 978-3-319-14592-1 (eBook)
DOI 10.1007/978-3-319-14592-1

Library of Congress Control Number: 2015932961

Springer Cham Heidelberg New York Dordrecht London


© Springer International Publishing Switzerland 2015
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.

Printed on acid-free paper

Springer International Publishing AG Switzerland is part of Springer Science+Business Media (www.


springer.com)
Preface

The scope of biodiversity is certainly vast. The concern on global environment


has been increasing due to ever increasing deforestation rate and its consequent
biodiversity loss. With burgeoning population, the conservation of biological
diversity of our planet earth, sustainable use of bio-resources and their wiser
management become crucial and a challenging task as well. Modernization,
industrialization and further a complex array of environmental factors exert their
effect on natural resources and their use patterns. Bio-resources constitute an
important link in ecosystem food chain and their sustainable development needs
to be addressed effectively both from the present use context and to conserve
them for future generation.
Forests, particularly in the tropics, are known for high biodiversity, and among
plants, of the various life-forms (trees, shrubs, climbers, herbs etc.), the lianas or the
woody climbers form one of the fascinating groups, yet remain relatively under-
researched, as compared to trees which provide structural framework for all other
life-forms in a forest community. Lianas occur in all forests, but particularly diverse
in tropical forests, and their density being high in dry tropics. After the classical
work of Darwin (1867), deVries (1880) and Schenck (1892), in the recent two
decades, studies on lianas have gained momentum and some notable publications
on lianas have come out (Putz and Mooney 1991; Bongers et al. 2005; Acevedo–
Rodriguez 2005; Schnitzer et al. 2015) besides individual research papers and few
chapters on lianas included in books dealing with tropical forest ecology (Richards
1996; Whitmore 1991; Ghazoul and Sheil 2010 to mention a few).
Lianas play an important role in forest functioning, contribute substantially to
forest aboveground biomass and render several ecosystem services (rewarding leaf,
flower and fruit resources) to various faunal communities, and provide several
goods of economic importance utilized by humans (from ropes, fruits to medicine).
That being the importance of lianas in forest ecosystem, it calls for greater attention
in sustainable resource use and wiser management of biodiversity and conservation
for ecosystem well-being and human welfare.
This book is a compilation of research contribution on liana diversity and ecol-
ogy comprising research articles (incorporating reviews and recent trends in the

v
vi Preface

subject), drawn from various parts of the world – from Canada, North America,
South America, Africa, Europe, China, South and Southeast Asia and Australia.
Many chapters address liana diversity in the context of geographical, climate and
various other environmental attributes and finally emphasize biodiversity conserva-
tion, sustainable use of bioresources and future direction of research. The publica-
tion of this book aptly comes at a time when the world is witnessing increase in
liana density, especially in the tropics, which is expected to impact forest structure
and functioning, and in turn the bio-resources management in years to come.
I sincerely hope that this book on liana diversity and ecology will be useful to
many readers – students, researchers, forest managers and conservation scientists. I
gratefully acknowledge Professor K. G. Ramawat for the invitation to edit this book
and the encouragement and suggestions provided at various stages in completing
this work. I thank Springer International Publishing AG, Switzerland, for publish-
ing this book. I thank all the chapter contributors of this book and all the reviewers
who helped in the peer-review process of the chapters included in this book. I thank
all the authorities of Pondicherry University for the facilities.

Puducherry, India N. Parthasarathy


Contents

1 General Introduction .............................................................................. 1


N. Parthasarathy and P. Vivek
2 Patterns of Liana Abundance, Diversity and Distribution
in Temperate Forests ............................................................................... 7
Bruce P. Allen
3 Geographical, Taxonomical and Ecological Aspects
of Lianas in Subtropical Forests of Argentina ...................................... 17
Agustina Malizia, Paula I. Campanello,
Mariana Villagra, and Sergio Ceballos
4 Liana Effects on Carbon Storage and Uptake in Mature
and Secondary Tropical Forests............................................................. 43
Sandra M. Durán and G.A. Sánchez-Azofeifa
5 Diversity and Distribution of Climbing Plants
in Eurasia and North Africa................................................................... 57
Liang Hu and Mingguang Li
6 Liana Assemblages in Tropical Forests of Africa
and Southeast Asia: Diversity, Abundance, and Management ........... 81
Patrick Addo-Fordjour and Zakaria B. Rahmad
7 Diversity of Lianas in Eastern Himalayas
and North-Eastern India ........................................................................ 99
S.K. Barik, D. Adhikari, A. Chettri, and P.P. Singh
8 Biodiversity of Lianas and Their Functional Traits
in Tropical Forests of Peninsular India ................................................. 123
N. Parthasarathy, P. Vivek, C. Muthumperumal,
S. Muthuramkumar, and N. Ayyappan

vii
viii Contents

9 The Contribution of Lianas to Forest Ecology,


Diversity, and Dynamics ......................................................................... 149
Stefan A. Schnitzer
10 Liana Diversity and Their Ecosystem Services in Tropical
Dry Evergreen Forest on the Coromandel Coast of India................... 161
N. Parthasarathy, P. Vivek, and K. Anil
11 A Review of Biotechnological Approaches to Conservation
and Sustainable Utilization of Medicinal Lianas in India ................... 179
Shaily Goyal, Varsha Sharma, and Kishan Gopal Ramawat
12 Biological Invasion of Vines, Their Impacts and Management........... 211
SM. Sundarapandian, C. Muthumperumal, and K. Subashree
13 Liana Diversity and the Future of Tropical Forests ............................. 255
Mason Campbell, Ainhoa Magrach, and William F. Laurance

Index ................................................................................................................. 275


Contributors

Patrick Addo-Fordjour Department of Theoretical and Applied Biology, College


of Science, Kwame Nkrumah University of Science and Technology (KNUST),
Kumasi, Ghana
School of Biological Sciences, Universiti Sains Malaysia, Penang, Malaysia
D. Adhikari Department of Botany, Centre for Advanced Studies in Botany, North-
Eastern Hill University, Shillong, India
Bruce P. Allen NH DRED-FL (New Hampshire Division of Resources and
Economic Development, Forests and Lands), Springfield, NH, USA
K. Anil Department of Ecology and Environmental Sciences, Pondicherry
University, Puducherry, India
N. Ayyappan French Institute of Pondicherry, Puducherry, India
S.K. Barik Department of Botany, Centre for Advanced Studies in Botany, North-
Eastern Hill University, Shillong, India
Paula I. Campanello CONICET, Instituto de Biología Subtropical, Facultad de
Ciencias Forestales, Universidad Nacional de Misiones, Puerto Iguazú, Misiones,
Argentina
Mason Campbell Centre for Tropical Environmental and Sustainability Science
(TESS) and College of Marine and Environmental Sciences, James Cook University,
Cairns, QLD, Australia
Sergio Ceballos CONICET, Instituto de Ecología Regional, Facultad de Ciencias
Naturales e IML, Universidad Nacional de Tucumán, Yerba Buena, Tucumán,
Argentina
A. Chettri Department of Botany, Sikkim University, Gangtok, Sikkim, India
Sandra M. Durán Department of Earth and Atmospheric Sciences, University of
Alberta, Edmonton, Canada

ix
x Contributors

Shaily Goyal Laboratory of Bio-Molecular Technology, Department of Botany,


M. L. Sukhadia University, Udaipur, Rajasthan, India
Erie, PA, USA
Liang Hu School of Geography and Planning, Sun Yat-sen University, Guangzhou,
China
William F. Laurance Centre for Tropical Environmental and Sustainability
Science (TESS) and College of Marine and Environmental Sciences, James Cook
University, Cairns, QLD, Australia
Mingguang Li State Key Laboratory of Biocontrol, Sun Yat-sen University,
Guangzhou, China
Ainhoa Magrach Ecosystem Management Group, Institute of Terrestrial
Ecosystems, ETH Zurich, Zurich, Switzerland
Agustina Malizia CONICET, Instituto de Ecología Regional, Facultad de Ciencias
Naturales e IML, Universidad Nacional de Tucumán, Yerba Buena, Argentina
C. Muthumperumal School of Biological Sciences, Department of Plant Sciences,
Madurai Kamaraj University, Madurai, Tamil Nadu, India
S. Muthuramkumar V.H.N.S.N. College, Virudhunagar, Tamil Nadu, India
N. Parthasarathy Department of Ecology and Environmental Sciences,
Pondicherry University, Puducherry, India
Zakaria B. Rahmad Department of Theoretical and Applied Biology, College of
Science, Kwame Nkrumah University of Science and Technology (KNUST),
Kumasi, Ghana
The Centre for Marine and Coastal Studies (CEMACS), Universiti Sains Malaysia,
Penang, Malaysia
Kishan Gopal Ramawat M.L. Sukhadia University, Botany Department, Udaipur,
Rajasthan, India
G.A. Sánchez-Azofeifa Department of Earth and Atmospheric Sciences,
University of Alberta, Edmonton, Canada
Stefan A. Schnitzer Department of Biological Sciences, Marquette University,
Milwaukee, WI, USA
Smithsonian Tropical Research Institute, Apartado, Balboa, Republic of Panama
Varsha Sharma Laboratory of Bio-Molecular Technology, Department of Botany,
M. L. Sukhadia University, Udaipur, Rajasthan, India
Department of Biochemical Engineering & Biotechnology, Indian Institute of
Technology (IIT), New Delhi, India
Contributors xi

P.P. Singh Department of Botany, Centre for Advanced Studies in Botany, North-
Eastern Hill University, Shillong, India
K. Subashree Department of Ecology and Environmental Sciences, Pondicherry
University, Puducherry, India
SM. Sundarapandian Department of Ecology and Environmental Sciences,
Pondicherry University, Puducherry, India
Mariana Villagra CONICET, Instituto de Biología Subtropical, Facultad de
Ciencias Forestales, Universidad Nacional de Misiones, Puerto Iguazú, Misiones,
Argentina
P. Vivek Department of Ecology and Environmental Sciences, Pondicherry
University, Puducherry, India
Chapter 1
General Introduction

N. Parthasarathy and P. Vivek

Lianas are woody climbers rooted in soil and are incapable of autonomous vertical
growth above a certain height and rely on external support to reach forest canopy
(Wyka et al. 2013). Lianas have long attracted the interest of botanists because of
the peculiar characteristics associated with this life-form including their climbing
mechanisms (Darwin 1865; Isnard and Silk 2009), biomechanical properties (Rowe
et al. 2004), anatomical modifications (Bowling and Vaughn 2009), extreme stem
hydraulic capacities (Gartner et al. 1990; Ewers et al. 1991) and their extraordinary
developmental plasticity (Lee and Richards 1991). Generally, lianas are abundant in
tropical forests with a high taxonomic diversity (Schnitzer and Bongers 2002;
Mascaro et al. 2004). Although less diverse, they are present in great abundance in
many temperate forests as well (Givnish and Vermeij 1976; Putz 1984). Lianas con-
stitute 19–30 % of species diversity in tropical forests (Jongkind and Hawthrone
2005), 9.6–19 % in subtropical forests (Cai and Song 2000) and 5.6–7 % in temper-
ate forests (Gentry 1991a). The diversity and abundance of lianas are expected to be
explained by abiotic factors such as water, temperature, soil and community struc-
tural attributes (Balfour and Bond 1993; Schnitzer and Bongers 2002; Schnitzer
et al. 2005). The high diversity of lianas in tropical forests has been assigned to
diversity in microhabitats (DeWalt et al. 2006) and the availability of a wide array
of dimensions, shapes and morphological features of trees that provide support to
them (Clark and Clark 1990). Liana research has been stimulated, especially during
the last two decades owing to their increasing presence and dominance in disturbed
vegetation and discoveries of the multidimensional role they play in forest dynam-
ics (Schnitzer et al. 2012; Yong et al. 2012). Lianas exhibit a diversity in functional
types and ecological strategies (DeWalt et al. 2000; Gerwing 2004; Yuan et al. 2009)
and although they have a similar growth form, they do differ in their functional traits

N. Parthasarathy (*) • P. Vivek


Department of Ecology and Environmental Sciences, Pondicherry University,
Puducherry 605014, India
e-mail: parthapu@yahoo.com

© Springer International Publishing Switzerland 2015 1


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_1
2 N. Parthasarathy and P. Vivek

including climbing mechanism (Putz 1984; Putz and Holbrook 1991), light require-
ments (Baars et al. 1998; Gianoli et al. 2010; Putz 1984), flower size and diaspore
type (Bullock 1995; Cai et al. 2009; Gentry 1991b).
Lianas play a major role in tropical forest dynamics, structuring and functioning
by competing with trees for both aboveground and belowground resources (Schnitzer
and Bongers 2002), resulting in reduced recruitment, regeneration, growth, fecun-
dity and survival of trees (Schnitzer et al. 2000; Toledo-Aceves and Swaine 2008;
Ingwell et al. 2010; Schnitzer and Carson 2010) which have enormous consequences
for tropical forest diversity and community composition, as well as ecosystem-level
dynamics such as carbon, nutrient and water sequestration. Since lianas rely on trees
for their physical support to reach canopy, they save on the heavy investments made
by their hosts for the production and maintenance of supporting tissue (Ewers and
Fisher 1991; Caballe 1993). Instead they invest more resource in vertical growth,
leaf production and sexual reproduction. Moreover, lianas typically have a high
canopy to stem ratio, which allows them to deploy a large canopy of leaves above
the host tree, thus competing aggressively with their hosts. In addition, lianas may
act as transformer species (Richardson et al. 2000), particularly when they become
invasive. The natural and man-made canopy disturbances in rainforests pave the way
to invasion of exotic vines (Floyd 1989), which may restrict both existing canopy
vegetation as well as ground layer (Groves and Willis 1999; Greenberg et al. 2001;
Kriticos et al. 2003; Timmins and Reid 2000). For example, species such as Mikania
micrantha tend to form dense ground cover carpets, suppressing native flora.
Apart from the deleterious effect of lianas, their presence can also have an impor-
tant role in ecosystem processes, structural diversity, and habitat heterogeneity of
tropical forests. In tropical forests, lianas add both physical structure and resources
to the forest. Lianas greatly contribute to canopy closure after tree fall and help
stabilize the microclimate underneath (Schnitzer and Bongers 2002), giving shade
tolerant species a chance to establish. Lianas provide valuable habitat and connec-
tions among tree canopies that provide a pathway for arboreal animals to transverse
the tree tops (Emmons and Gentry 1983; Schnitzer and Carson 2001). Their flowers,
and fruits provide nutrients to a diverse group of vertebrate and invertebrate fauna
(Schnitzer and Bongers 2002; Martins 2009) and their leaves are used as oviposit
hosts for lepidopterans and as a foliar resource for diverse forest insects (Odegaard
2000). Martins (2009) found that lianas constituted 27 % of the diet for the brown
howling monkey (Aouatta guariba) and 33 % of the diet for the Southern muriqui
(Brachyteles arachnoides). Tra Bi et al. (2005) found that 29 climbers were used by
indigenous and rural communities in Cote d’ Ivoire, Africa, for their edible parts and
derivatives. Millions of people worldwide are fed by agricultural vine crops or vines
cultivated from the forest. Many common foods depend on the cultivation and trans-
port of vine products such as the grape (Vitis spp.), sweet potato (Ipomoea batatas),
yams (Dioscorea spp.), black pepper (Piper nigrum), vanilla (Vanilla planifolia),
and melons, squash, and cucumber. Passion fruit (Passiflora edulis) and chayote
(Sechium edule) are commercially valuable foods in tropical countries (Phillips
1991). However, the over exploitation of these economically and culturally important
1 General Introduction 3

species threatens their survival. For example, the rattan palms and Heteropsis roots
are becoming sparse in Cote d’Ivoire, west central Sumatra, and other areas of
Southeast Asia due to over exploitation (Peters et al. 2007). Greater research into
the use and sustainable harvest of climber products can be useful in the foundation
of management tools to protect those species from extinction. In recent times, the
role of biotechnology in conservation of such species is also emphasized (Arora
et al. 2010; Sharma et al. 2011) so as to meet the growing requirements.
One of the most prominent structural changes occurring now in tropical forests
is the increase in liana abundance and biomass (Schnitzer and Bongers 2011). The
leading hypotheses to explain increasing liana abundance include increasing forest
disturbance, increasing duration and severity of seasonal drought, and elevated
atmospheric CO2 (Schnitzer and Bongers 2011). Increasing forest disturbance
would favor lianas relative to trees by creating more edge and gap habitat, where
lianas proliferate (Putz 1984; Schnitzer et al. 2000; Schnitzer and Carson 2010).
Stronger seasonal drought may benefit lianas because they suffer less water stress
and grow more than trees during dry periods (Schnitzer 2005; Cai et al. 2009). It has
been reported that lianas comprise up to 40 % of the woody stems in tropics, thereby
contributing substantially to the forest leaf area and biomass (Gerwing and Farias
2000; Chave 2001). Currently it is known that lianas can contribute up to 30 % of
the total aboveground biomass in liana-dense tropical forests (Schnitzer and Bongers
2011). The increase in liana abundance and biomass poses serious consequences for
tropical forest dynamics and functioning. For example, lianas may lower whole-
forest carbon storage by reducing tree growth and increasing tree mortality, espe-
cially for shade-tolerant trees with high wood density. Since liana stems contain far
less carbon than trees, lianas do not compensate for the tree biomass that they dis-
place (e.g., van der Heijden and Phillips 2009; Schnitzer and Bongers 2011).
Tropical forests are enriched with half of the earth’s terrestrial species and contrib-
ute more than a third of global terrestrial carbon stocks as well as nearly a third of
terrestrial net primary productivity (Dixon et al. 1994; Field et al. 1998; Wright
2010). Thus, any alteration to tropical forests has important potential ramifications
for species diversity, productivity and the global carbon cycle. Thus, considering the
extent of liana diversity, their role in forest functioning under the current era of cli-
mate change and progressive deforestation rates, this book is prepared.
This book includes original research articles, case studies and reviews (regional
and global) on biodiversity, ecology and phytogeography and conservation of lianas
from temperate, sub-tropical and tropical forests. Further, it provides an insight into
the patterns of liana diversity, distribution, the role of lianas in structuring forest
community, and functional ecology (carbon uptake, ecosystem services, dynamics
and invasion), biotechnological tool for conservation of lianas and finally summa-
rizes the significance and the need for conservation of lianas in the changing global
environmental scenario. This book is unique as it covers wide array of topics on this
subject covering contributions from all continents and will serve as a valuable refer-
ence material for students, researchers and forest managers who are concerned with
biodiversity, forest ecology and sustainable development of forest resources.
4 N. Parthasarathy and P. Vivek

References

Arora J, Goyal S, Ramawat KG (2010) Enhanced stilbene production in cell cultures of Cayratia
trifolia through co-treatment with abiotic and biotic elicitors and sucrose. In Vitro Cell Dev
Biol Plant 46:430–436
Baars R, Kelly D, Sparrow AD (1998) Liana distribution within native forest remnants in two
regions of the South Island, New Zealand. New Zeal J Ecol 22:71–85
Balfour DA, Bond WJ (1993) Factors limiting climber distribution and abundance in a southern
African forest. J Ecol 11:93–99
Bowling AJ, Vaughn KC (2009) Gelatinous fibers are widespread in coiling tendrils and twining
vines. Am J Bot 96:719–727
Bullock SH (1995) Plant reproduction in Neotropical dry forests. In: Bullock SH, Mooney HA,
Medina E (eds) Seasonally dry tropical forests. Cambridge University Press, Cambridge,
pp 277–303
Caballe G (1993) Liana structure, function and selection: a comparative study of xylem cylinders
of tropical rainforest species in Africa and America. Bot J Linn Soc 113:41–60
Cai YL, Song YC (2000) Diversity of vines in subtropical zone of East China. J Wuhan Bot Res
18:390–396
Cai ZQ, Schnitzer SA, Bongers F (2009) Liana communities in three tropical forest types in
Xishuangbanna, South-West China. J Trop For Sci 21:252–264
Chave J (2001) Estimation of biomass in a Neotropical forest of French Guiana: spatial and tem-
poral variability. J Trop Ecol 17:79–96
Clark DB, Clark DA (1990) Distribution and effects of tree growth of lianas and woody hemiepi-
phytes in a Costa Rican tropical wet forest. J Trop Ecol 6:321–331
Darwin C (1865) On the movements and habits of climbing plants. Bot J Linn Soc 9:1–118
DeWalt SJ, Schnitzer SA, Denslow JS (2000) Density and diversity of lianas along a chronose-
quence in central Panamanian lowland forest. J Trop Ecol 16:1–19
DeWalt SJ, Ickles K, Nilus R et al (2006) Liana habitat association and community structure in a
Bornean lowland tropical forest. Plant Ecol 186:203–216
Dixon RK, Solomon AM, Brown S et al (1994) Carbon pools and flux of global forest ecosystems.
Science 263:185–190
Emmons LH, Gentry AH (1983) Tropical forest structure and the distribution of gliding and
prehensile-tailed vertebrates. Am Nat 121:513–524
Ewers FW, Fisher JB (1991) Why vines have narrow stems: histological trends in Bauhinia
(Fabaceae). Oecologia 88:233–237
Ewers FW, Fisher JB, Fichtner K (1991) Water flux and xylem structure in vines. In: Putz FE,
Mooney HA (eds) The biology of vines. Cambridge University Press, Cambridge, pp
127–160
Field CB, Behrenfeld MJ, Randerson JT et al (1998) Primary production of the biosphere: integrat-
ing terrestrial and oceanic components. Science 281:237–240
Floyd AG (1989) The vine weeds of coastal rainforests. In: Proceedings of the 5th biennial noxious
plants conference. New South Wales Department of Agriculture and Fisheries, Sydney,
pp 1109–1115
Gartner BL, Bullock SH, Mooney HA et al (1990) Water transport properties of vine and tree stems
in a tropical deciduous forest. Am J Bot 77:742–749
Gentry AG (1991a) The distribution and evolution of climbing plants. In: Putz FE, Mooney HA
(eds) The biology of vines. Cambridge University Press, Cambridge, pp 73–97
Gentry AH (1991b) Breeding and dispersal systems of lianas. In: Putz FE, Mooney HA (eds) The
biology of vines. Cambridge University Press, Cambridge, pp 393–426
Gerwing JJ (2004) Life history diversity among six species of canopy lianas in an old-growth for-
est of the eastern Brazilian Amazon. For Ecol Manage 190:57–72
Gerwing JJ, Farias DL (2000) Integrating liana abundance and forest stature into an estimate of
total aboveground biomass for an eastern Amazonian forest. J Trop Ecol 16:327–335
1 General Introduction 5

Gianoli E, Saldana A, Jimenez-Castillo M et al (2010) Distribution and abundance of vines along


the light gradient in a southern temperate rain forest. J Veg Sci 21:66–73
Givnish TJ, Vermeij GJ (1976) Sizes and shapes of liane leaves. Am Nat 110:743–778
Greenberg CH, Smith LM, Levey DJ (2001) Fruit fate, seed germination and growth of an invasive
vine – an experimental test of ‘sit and wait’ strategy. Biol Invasions 3:363–372
Groves RH, Willis AJ (1999) Environmental weeds and loss of native plant biodiversity: some
Australian examples. Aust J Environ Manage 6:164–171
Ingwell LL, Joseph Wright S, Becklund KK et al (2010) The impact of lianas on 10 years of tree
growth and mortality on Barro Colorado Island, Panama. J Ecol 98:879–887
Isnard S, Silk WK (2009) Moving with climbing plants from Charles Darwin’s time into the 21st
century. Am J Bot 96:1205–1221
Jongkind CCH, Hawthorne WD (2005) A botanical synopsis of lianes and other forest climbers.
In: Bongers F, Parren MPE, Traore D (eds) Forest climbing plants of West Africa: diversity,
ecology and management. CABI Publishing, Oxfordshire, pp 19–39
Kriticos DJ, Sutherst RW, Brown JR, Adkins SW, Maywald GF (2003) Climate change and biotic
invasions: a case history of a tropical woody vine. Biol Invasions 5:145–165
Lee DW, Richards JH (1991) Heteroblastic development in vines. In: Putz FE, Mooney HA (eds)
The biology of vines. Cambridge University Press, Cambridge, pp 205–243
Martins MM (2009) Lianas as a food resource for brown howlers (Alouatta guariba) and southern
muriquis (Brachyteles arachnoides) in a forest fragment. Anim Biodivers Conserv 32:51–58
Mascaro J, Schnitzer SA, Carson WP (2004) Liana diversity, abundance, and mortality in a tropical
wet forest in Costa Rica. For Ecol Manage 190:3–14
Odegaard F (2000) The relative importance of trees versus lianas as host for phytophagous beetles
(Coleopteran) in tropical forest. J Biogeogr 27:283–296
Peters CM, Henderson A, Maung UM et al (2007) The rattan trade of Northern Myanmar: species,
supplies, and sustainability. Econ Bot 61:3–13
Phillips OL (1991) The ethnobotany and economic botany of tropical vines. In: Putz FE, Mooney
HA (eds) Biology of vines. Cambridge University Press, Cambridge, pp 427–476
Putz FE (1984) The natural history of Lianas on Barro Colorado Island, Panama. Ecology
65:1713–1724
Putz FE, Holbrook NM (1991) Biomechanical studies of vines. In: Putz FE, Mooney HA (eds) The
biology of vines. Cambridge University Press, Cambridge, pp 73–96
Richardson DM, Pyusek P, Rejmanek M, Barbour MG, Panetta FD, West CJ (2000) Naturalisation
and invasion of alien plants: concepts and definitions. Divers Distrib 6:93–107
Rowe NP, Isnard S, Speck T (2004) Diversity of mechanical architecture in climbing plants: an
evolutionary perspective. J Plant Growth Regul 23:108–128
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
Schnitzer SA, Bongers F (2011) Increasing liana abundance and biomass in tropical forests:
emerging patterns and putative mechanisms. Ecol Lett 14:397–406
Schnitzer SA, Carson WP (2001) Treefall gaps and maintenance of species diversity in a tropical
forest. Ecology 82:913–919
Schnitzer SA, Carson WP (2010) Lianas suppress tree regeneration and diversity in Treefall gaps.
Ecol Lett 13:849–857
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical for-
est canopy gaps: evidence for an alternate pathway of gap-phase regeneration. J Ecol 88:655–666
Schnitzer SA, Kuzee ME, Bongers F (2005) Disentangling above- and below-ground competition
between lianas and trees in a tropical forest. J Ecol 93:1115–1125
Schnitzer SA, Mangan SA, Dalling JW et al (2012) Liana abundance, diversity, and distribution on
Barro Colorado Island, Panama. PLoS One 7:e52114
Sharma V, Goyal S, Ramawat KG (2011) Increased puerarin biosynthesis during in vitro shoot
formation in Pueraria tuberosa grown in Growtek bioreactor with aeration. Physiol Mol Biol
Plants 17:87–92
6 N. Parthasarathy and P. Vivek

Timmins SM, Reid V (2000) Climbing asparagus, Asparagus scandens Thunb.: a South African in
your forest patch. Austral Ecol 25:533–538
Toledo-Aceves T, Swaine MD (2008) Effects of lianas on tree regeneration in gaps and forest
understorey in a tropical forest in Ghana. J Veg Sci 19:717–728
Tra Bi FH, Kouame FN, Traore D (2005) Utilisation of climbers in two forest reserves in west Cote
d’Ivoire. In: Bongers F, Parren MPE, Traore D (eds) Forest climbing plants of West Africa:
diversity, ecology, and management. CABI Publishing, Wallingford, pp 167–182
Van der Heijden GMF, Phillips OL (2009) Environmental effects on Neotropical liana species rich-
ness. J Biogeogr 36:1561–1572
Wright SJ (2010) The future of tropical forests. Ann NY Acad Sci 1195:1–27
Wyka TP, Oleksyn J, Karolewski P et al (2013) Phenotypic correlates of the lianescent growth
form: a review. Ann Bot 112:1667–1681
Yong T, Kitching RL, Cao M (2012) Lianas as structural parasites: a reevaluation. Chin Sci Bull
57:307–312
Yuan CM, Liu WY, Tang CQ et al (2009) Species composition, diversity and abundance of lianas
in different secondary and primary forests in a subtropical mountainous area, SW China. Ecol
Res 24:1361–1370
Chapter 2
Patterns of Liana Abundance, Diversity
and Distribution in Temperate Forests

Bruce P. Allen

Abstract Lianas are a growing part of temperate forests that are responding to
environmental changes that give lianas a competitive advantage. Shifts in climactic
factors like growing season, precipitation, CO2, and disturbance frequency (both
natural and anthropogenic) are affecting woody vine communities. Long-term stud-
ies of forest communities as well as dendrochronology provide insights into how
communities are changing through time.

2.1 Introduction

Lianas, or large woody vines, are important features of temperate forest that are
changing through time. Though historically ignored in studies of temperate forest
ecology and dynamics, lianas represent a growing influence on forest structure,
dynamics and regeneration. The growing interest in lianas in both temperate and
tropical forests has provided important developments in our understanding of their
role in changing forests that are confronted by shifts in climate, disturbance and
atmospheric chemistry. Within temperate regions, liana species richness increases
with decreasing latitude and precipitation. The tools used to identify shifts in liana
communities have grown to include long-term forest plots and dendrochronology.
Liana density in temperate forests appears to be increasing, though studies of com-
munities at the colder end of their distribution may not be responding to the same
limits on growth. Lianas in temperate forests are changing and our understanding of
how and why is expanding as well.

B.P. Allen (*)


NH DRED-FL (New Hampshire Division of Resources and Economic Development,
Forests and Lands), 191 Four Corners Rd, Springfield, NH 03284, USA
e-mail: Bruce.P.Allen@gmail.com

© Springer International Publishing Switzerland 2015 7


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_2
8 B.P. Allen

2.2 Patterns of Liana Abundance and Distribution


in Temperate Forests

Temperate forests, like tropical forests, are experiencing dramatic structural changes
in the form of increases in liana abundance and biomass. In 1991, Teramura et al.
(1991) noted that liana abundance had shifted dramatically in North America during
the twentieth century but their focus was on invasive exotic vines like Lonicera
japonica Thunb. (Japanese honeysuckle), Celastrus orbiculata Thunb. (Asian bit-
tersweet) and Pueraria lobata McNab (kudzu) (Meeker 1987; Patterson 1973;
Wechsler 1977; Sasek and Strain 1991). Invasive lianas continue to increase in den-
sity in temperate forest (Fike and Niering 1999; Horvitz and Koop 2001; Gallagher
et al. 2010; Ladwig and Meiners 2010b) but increases in liana density are no longer
limited to exotic lianas (Allen et al. 2007). Changes continued into the late twentieth
century and early twenty-first century in the form of increases in liana density, pro-
portion of stems and basal area of native species (Allen et al. 2007; Allen 2007).
Early in the twenty-first century the focus shifted to changes in both native and
exotic liana density in temperate forests and tropical forests (see Schnitzer and
Bongers 2011). Increases in native liana density were first noted in tropical forests
of Central and South America (Philips and Gentry 1994; Philips et al. 2002; Benitex-
Malvido and Martinez-Ramos 2003; Wright et al. 2004; Korner 2006; Chave et al.
2008; Foster et al. 2008; Rutishauser 2011; Laurance et al. 2013). Two studies of
tropical forests in Africa were not consistent with the pattern of increasing tropical
liana density (Caballe and Martin 2004; Ewango 2010). Like tropical forests, pat-
terns of changes in liana density were mixed in temperate forests. Increasing liana
density was reported in Germany (Dierschke 2005), New Jersey, USA (Ladwig and
Meiners 2010a, b) and South Carolina, USA (Allen 2007; Allen et al. 2007). In
contrast, Londré and Schnitzer (2006) found no changes in liana density in
Wisconsin, USA.
Liana stem density increases in temperate forests have been documented in three
locations. In Germany, Dierschke (2005) documented dramatic increases in Hedera
helix (English ivy) density and cover over a 24 year period – 1980–2004. In North
America, Allen et al. (2007) found linear increases in large (≥4.5 cm dbh) liana
density, proportion of vines and liana basal area in undisturbed second growth
floodplain forests in South Carolina, USA between 1979 and 2001. In old growth
floodplain forests of the Congaree National Park subjected to both moderate and
severe hurricane damage, liana densities (stem ≥ 2.5 cm dbh) nearly doubled regard-
less of disturbance level 16 years post-hurricane. Similarly, Ladwig and Meiner
(2010b) found that the per cent cover of two liana species (Vitis sp. and Celastrus
orbiculata) was increasing over the initial 50 years of forest development in New
Jersey. In contrast to the above studies, Londré and Schnitzer (2006) found that
liana density had not increased over 45 years (1960–2005) in Wisconsin, USA. This
study examined all climbing liana stems and found that liana density decreased with
distance from the forest edge. The limited number of studies either supporting or
refuting liana increases reflect the fact that lianas have been ignored in most
long-term studies in temperate forests historically. One pattern that emerges from
2 Patterns of Liana Abundance, Diversity and Distribution in Temperate Forests 9

this evidence is a latitudinal shift: liana populations near their northern end of their
range, like Wisconsin, USA, may be limited by factors other than water use effi-
ciency or even light availability. Lianas may be limited by the number of frost-free
days or the vulnerability of large vessel elements to extremely cold winter tempera-
tures or other factors associated with the northern end of their range (Schnitzer and
Bongers 2002).
Another factor affecting the detection of changes in liana density or abundance
is the proportion of the size-class distribution examined. In addition, shifts in size
class distribution my interact with sampling protocol. Some of the detected shifts in
liana density may well be a function of shifts in liana size class distribution from
negative exponential to modal as disturbed forests age or simply maintaining the
distribution shape but growing into larger size classes. It may also be a function of
portion of the distribution you are looking at and the minimum stem diameter
studied. If you look at all stems larger than 1 cm diameter, you might get a very
different conclusion than if you looked at only stems larger than 4.5 cm. Increases
in the number of large lianas associated with older forests are not a new phenome-
non (Philips et al. 2002) but they have been observed in temperate forests
(Allen et al. 2007).

2.3 Possible Mechanisms for Increasing Liana


Abundance in Temperate Forests

Putative mechanism for increases in liana abundance include increases in evopo-


transpirational demands (Schnitzer 2005; Cai et al. 2009), shifts in natural forest
disturbance rate (Putz 1984; Allen et al. 1997, 2005; DeWalt et al. 2000; Schnitzer
et al. 2000), changing land use patterns, elevated surface level CO2 (Granados and
Korner 2002; Mohan et al. 2006; Zotz et al. 2006; Ziska et al. 2007) as well as shifts
in climate in the form of longer growing seasons (Anderson-Teixeira et al. 2013).
Increases in liana abundance not associated with disturbance require a re-examination
for forest dynamic models that suggest that lianas are most abundant in young forests
following disturbance (DeWalt et al. 2000; Ladwig and Meiners 2010b).
Increases in evopotranspirational demand would favor lianas over other life
forms if lianas have dry season growth advantages due to more efficient soil water
use and stronger stomatal control during seasonal drought that allows them grow
while trees are dormant (Schnitzer 2005; DeWalt et al. 2010). Swaine and Grace
(2007) found that liana species richness increased as mean annual rainfall decreased
going from wet to dry tropical forests in Ghana. DeWalt et al. (2010) conducted a
meta-analysis on pan-tropical data set across a wide range of mean annual precipita-
tion (860–7,250 mm/year) and dry season length (0–7 months) and found that liana
density and basal area decreased with increasing precipitation and increased with
increasing dry season length. Schnitzer and Bongers (2011) suggest that the combi-
nation of seasonal advantages and increasing evopotranspirational demand through
time could explain increases in liana abundance and biomass. The evidence for
similar patterns in temperate forests is weaker where liana species richness appears
10 B.P. Allen

to decrease as rainfall decreases (Teramura et al. 1991). The high leaf area to stem
area ratios reported for lianas might counteract this advantage.
Increases in the frequency of natural disturbance may explain shifts in liana
abundance by expanding areas that have been recently disturbed. Liana responses to
disturbance have been widely reported in temperate forests (Lutz 1943; Allen et al.
1997, 2005; Londré and Schnitzer 2006; Allen et al. 2010). If the frequency of natu-
ral disturbance is increasing through time, this could explain increases in liana den-
sity. Historical changes in climatic patterns have increased the rate of disturbance in
temperate forests in Canada, Europe, and Russia (Kurz et al. 1995; Dale et al. 2001).
This would not explain increases in liana density in undisturbed forests (Phillips
et al. 2002; Wright et al. 2004; Allen et al. 2007).
Human disturbance to natural communities in the form of development, forest
management, and changes in land use, are influencing liana communities. If rates are
changing (increasing) through time, they could cause increases in liana density.
Londré and Schnitzer (2006) reported that liana densities were four times more abun-
dant at the forest edge, ~250 stems ha−1, when compared with the forest interior in
Wisconsin, USA and that liana density decreased with distance from the forest edge.
This suggests that human activities that tend to increase the proportion of forest edge
to forest interior would likely cause increases in liana densities in temperate forests.
A number of studies have demonstrated that increases in carbon dioxide level
give lianas a competitive advantage over other plant life-forms in temperate forests
(Belote et al. 2003; Hattenschwiler and Korner 2003; Mohan et al. 2006; Zotz et al.
2006; Ziska et al. 2007). Belote et al. (2003) found that the invasive liana, Lonicera
japonica (Japanese honeysuckle) consistently accumulated greater biomass under
elevated CO2 when compared to ambient levels in controlled exposure studies at
Oak Ridge National Laboratory, Tennessee, USA. Hattenschwiler and Korner
(2003) found that Hedera helix (English ivy) seedlings had a large linear growth
response to elevated CO2 and postulated that it could result in increasing coloniza-
tion of trees in forests. Zotz et al. (2006) also studied Hedera helix and found that
when exposed to elevated CO2 concentrations, growth and biomass production was
stimulated particularly in low light environments. In a controlled exposure study at
Duke’s Free Air CO2 Enrichment (FACE) study in North Carolina, USA, Mohan
et al. (2006) reported that Toxicodendron radicans (poison ivy) responded to ele-
vated CO2 levels in an intact forest ecosystem by stimulating photosynthetic rate,
water use efficiency, growth, and causing it to produce more allergenic form of
urushiol. Ziska et al. (2007) examined the effect of historic, current, and projected
future atmospheric CO2 concentrations on poison ivy growth in a controlled expo-
sure study. This study found significant increases in leaf area, leaf and stem weight,
and rhizome length relative to plants grown at historic CO2 concentrations. These
patterns of CO2-stimulated growth were similar to controlled exposure studies for
three tropical lianas (Granados and Korner 2002).
Elevated CO2 may benefit lianas more than other plant life-forms due to
interaction between improved water use efficiency and the high leaf area to stem
area ratios associated with lianas (Schnitzer and Bongers 2011). This would also
improve liana drought tolerance. If water transport is the limiting factor on liana
2 Patterns of Liana Abundance, Diversity and Distribution in Temperate Forests 11

growth, improved water use efficiency would increase growth. This may explain
differential patterns of liana stem density in temperate forests. Lianas at the north-
ern end of their range in North America may be limited other factors like tempera-
ture and the length of growing season rather than soil moisture.

2.4 Patterns of Liana Diversity in Temperate Forests

Temperate liana communities differ from tropical communities primarily in species


richness and to a lesser degree in stem density and climbing mechanism, particu-
larly in lower latitude temperate regions (Schnitzer and Bongers 2002; Durigon
et al. 2014). Liana abundance and species richness decrease with increasing lati-
tude, more rapidly than other life-forms (Schnitzer and Bongers 2002). Within the
temperate forest, liana density and diversity decreases with distance from the equa-
tor, number of frost free days and moisture (Teramura et al. 1991). In North America,
liana community composition changed dramatically in the twentieth century, where
invasive exotics (Celastrus orbiculata, Lonicera japonica and to a lesser extent
Pueraria lobata (Willd. Owhi)) rapidly invaded temperate forests (Patterson 1973;
Wechsler 1977; Sase 1985; McNab and Meeker 1987; Teramura et al. 1991).
If we shift our focus from broad regional trends in temperate floodplain ecosys-
tems to floodplain forests in South Carolina, USA, we can look at specific trends in
liana communities. Three types of long-term data provide insights into how liana
communities are changing through time. Evidence from long-term permanent plot
studies of liana communities in two floodplain forests in South Carolina provides
community dynamics information. In second-growth floodplain forests of the
Savannah River system that have been minimally disturbed in the last 70 years
(Odum, personnel communication), five 1-ha plots established in 1979 and moni-
tored for 22 years provide insights in long-term trends in second growth forests. In
old-growth floodplain forests of the Congaree National Park, liana density changes
have been followed over 16 years in six 1-ha plots after Hurricane Hugo disturbed
the forests in 1989. One quarter of the liana stems present when the hurricane hit
were killed primarily by loss or severe damage to their host tree (Allen et al. 1997).
Liana mortality rates (5–6 %/year) were approximately double that of canopy and
sub canopy trees for 12 years post-hurricane (Allen et al. 2005). Despite the high
mortality rate, liana stem density increased from ~ 140 stems/ha to >200 stems/ha in
bottomland hardwood sites.
South Carolina is a southern temperate area where the lianas most common in
floodplains are not subtropical as some have stated. In Congaree National Park, the
flora includes 28 woody taxa of which 17 species grow larger than 2.5 cm dbh and
would be considered lianas. The liana species that dominated floodplains are
Toxicodendron radicans (poison ivy), Vitus spp. (grapes), Campsis radicans (trum-
pet creeper), and Parthenocissus quinquefolia (Virginia creeper), a distinctly tem-
perate species mix. The 28 woody vine species present in Congaree National Park
include four exotic species, seven species which exceed 10 cm dbh and 17 species
that exceed 2.5 cm dbh (Gaddy and Nelson 2006).
12 B.P. Allen

In the Savannah river system, liana density (stems ≥ 4.5 cm) increased from a
single stem in 1979 to 10 stems/ha in 2001 in minimally disturbed plots (Allen et al.
2007). Stem density, proportion of lianas, and liana basal area increased linearly
(R2 = 0.99) over the 22 years studied, while tree density declined over the same
period.
In the old-growth floodplain forests of Congaree National Park, six 1-ha plots
were followed from 1989, immediately following Hurricane Hugo to 2006. This
study examined stems 2.5 cm dbh and larger. Liana densities declined 41 % in heav-
ily damaged areas, but only 7 % in low damage areas (Allen et al. 2005). Within 12
years liana density increased in areas of both high (60 %) and low damage (25 %)
to 212–214 stems/ha.
Further supporting evidence for changes in temperate liana communities can be
found by examining long-term diameter growth patterns. Dendrochronologic tech-
niques frequently used with trees to examine long-term growth and longevity pat-
terns have been applied to three temperate liana species (Allen 2007; Heuze et al.
2008; Allen et al. 2010). Allen (2007) examined 100 radial cores from Toxicodendron
radicans (poison ivy) and Campsis radicans (trumpet creeper) and found different
long-term radial growth patterns. Poison ivy radial growth increased linearly from
the 1960s to 2000s in decadal average growth in Congaree National Park, South
Carolina. This was not merely a case of older vines growing slower, when growth
was examined by vine age, radial growth did not decrease with liana age (Allen
et al. 2010). In contrast, trumpet creeper radial growth was more than double that of
poison ivy initially but it did not increase through time. Unlike poison ivy, trumpet
creeper’s radial growth decreased with stem age (Allen et al. 2010).
Dendrochronology also provides insights into how long vine stems live. Heuzé
et al. (2008) cored 118 English ivy vines (Hedera helix) in three floodplain sites in
France and found stems up to 66 years old. Allen (2007) found poison ivy stems up
to 58 years old and trumpet creeper stems up to 38 years old, though trumpet creeper
stems larger than 14 cm dbh, sometimes much larger (up 40 cm dbh), suffered from
heart rot preventing accurate aging. The age of a stem may not reflect the age of the
individual, as many lianas appear to be clonal, meaning the roots systems may be
much older than the individual stem. Lianas have long been known to be strongly
clumped (Caballé 1984; Putz 1984; Schnitzler and Heuzé 2006). The primary evi-
dence for this clonal nature is provided by stem mapping studies that show clumps
of individual species (Allen et al. 1997, 2005).

2.5 Conclusion

Temperate liana communities continue to change through time, but the changes go
beyond increases in invasive lianas noted in the twentieth century (Teramura et al.
1991). Our understanding of lianas and liana ecology continues to grow in leaps and
bounds as research interest increases. Lianas represent an important and growing
influence on temperate forests that appear to be responding to environmental changes.
2 Patterns of Liana Abundance, Diversity and Distribution in Temperate Forests 13

Why would lianas gain a competitive advantage over other plant life-forms?
What unique characteristics do woody vines have that other forms of plants do not?
First, lianas are not self-supporting; by using trees (and other forms of support) they
reduce the need to allocate resources to support structure. This allows lianas to have
higher leaf area to stem area ratios than trees (Putz 1984; Gerwing and Farias 2000;
Gehring et al. 2004). Gerwing and Farias (2000) found that liana leaf mass to stem
mass ratios were 4–5 times higher than trees in eastern Amazonian forests. This
substantially higher leaf area requires proportionally greater water transport to sup-
port the leaves. If leaf area is limited by a liana’s ability to transport water, factors
that increase their water use efficiency will increase growth.
So why would lianas increase in density, proportion of stems, and biomass in
relation to other life-forms in forests? The idea that forests reach a stable climax
through a successional process requires a stable environment through time, stable
atmospheric chemistry, stable disturbance regime, and a stable suite of species. Is
there any reason to expect any of these factors to be stable? Most have changed
dramatically over the last 100 years.

References

Allen BP (2007) Vegetation dynamics and response to disturbance in floodplain forest ecosystems
with a focus on lianas. Ph.D. Dissertation, Ohio State University, Columbus. 242p
Allen BP, Pauley EF, Sharitz RR (1997) Hurricane impacts on liana populations in an old-growth
southeastern bottomland forest. J Torrey Bot Soc 124:34–42
Allen BP, Sharitz RR, Goebel PC (2005) Twelve years post-hurricane liana dynamics in an old-
growth southeastern floodplain forest. For Ecol Manage 218:259–269
Allen BP, Sharitz RR, Goebel PC (2007) Are lianas increasing in importance in temperate flood-
plain forests in the southeastern United States? For Ecol Manag 242:17–23
Allen BP, Goebel PC, Sharitz RR (2010) Long-term effects of wind disturbance on the old-growth
forests and lianas of the Congaree National Park. Final report, USDI NPS CA #5000-03-5040,
42p
Anderson-Teixeira KJ, Miller AD, Mohan JE, Hudiburg TW, Duval BD, DeLucia EH (2013)
Altered dynamics of forest recovery under a changing climate. Glob Chang Biol
19:2001–2021
Belote RT, Weltzen JF, Norby RJ (2003) Response of an understory plant community to elevate
CO2 depends on differential responses of dominant invasive species and is mediated by soil
water availability. New Phytol 161:827–835
Benitez-Malvido J, Martinez-Ramos M (2003) Impact of forest fragmentation on understory plant
species richness in Amazonia. Conserv Biol 17:389–400
Caballé G (1984) Essaisur la dynamique des peuplements de lianesligneusesd’uneforet du Nord-
Est du Gabon. Rev Ecol (Terre Vie) 39:3–35
Caballé G, Martin A (2004) Thirteen years of change in trees and lianas in a Gabonese rainforest.
Plant Ecol 152:167–173
Cai ZQ, Schnitzer SA, Bongers F (2009) Season difference in leaf-level physiology give lianas a
competitive advantage over trees in tropical seasonal forest. Oecologia 161:25–33
Chave J, Olivier J, Bongers F, Chatelet P, Forget PM, van der Meer P, Norden N, Riera B, Charles-
Dominique P (2008) Aboveground biomass and productivity in rain forest of eastern South
America. J Trop Ecol 24:355–366
14 B.P. Allen

Dale VH, Joyce LA, McNulty S, Neilson RP, Ayres MP, Flannigan MD, Hanson PJ, Irland LC,
Lugo AE, Peterson CJ, Simberloff D, Swanson FJ, Stocks BJ, Wotton BM (2001) Climate
change and forest disturbance: climate change can affect forests by altering the frequency,
intensity, duration, and timing of fire, drought, introduced species, insects and pathogen out-
breaks, hurricanes, windstorms, ice storms or landslides. Bioscience 51:723–734
DeWalt SJ, Schnitzer SA, Denslow JS (2000) Density and diversity of lianas along a chronose-
quence in a central Panamanian lowland forest. J Trop Ecol 16:139–151
DeWalt SJ, Schnitzer SA, Chave J, Bongers F, Burnham RJ, Cai ZQ, Chuyong G, Clark DB,
Ewango CEN, Gerwing JJ, Gortaire E, Hart T, Ibara-Manriquez G, Ickes K, Kenfack D, Macia
MJ, Makana JR, Martiez-Ramos M, Mascaro J, Moses S, Muller-Landau HC, Parren MPE,
Parthasarathy N, Perez-Salicrup DR, Putz FE, Romero-Saltos H, Tomas D (2010) Annual rain-
fall and seasonality predict pan-tropical patterns of liana density and basal area. Biotropica
42:309–317
Dierschke H (2005) Laurophyllisaion – aucheineErscheinungimnordlichenMitteleuropa?
ZuraktuellenAusbreitung von Hederaheliz in sommergrunenLaubwandern. BerReingTuxenGes
17:151–168
Durigon J, Miotto STS, Gianoli E (2014) Distribution and traits of climbing plants in subtropical
and temperate South America. J Veg Sci. doi:10.1111/jvs.12197
Ewango CEN (2010) The liana assemblage of Congolian rainforest: diversity, structure and func-
tion. Ph.D. Dissertation, Wageningen University Wageningen
Fike J, Niering WA (1999) Four decades of old field vegetation development and the role Celastrus
orbiculata in the northeastern United States. J Veg Sci 10:483–492
Foster JR, Townsend PA, Zganjar CE (2008) Spatial and temporal patterns of gap dominance by
low-canopy lianas detected using EO-1 Hyperion and Landsat Thematic Mapper. Remote Sens
Environ 112:2104–2117
Gaddy LL, Nelson JB (2006) The vascular flora of the Congaree National Park, South Carolina.
National Park Service. Southeast Coast Network. Atlanta, GA, 42p
Gallagher RV, Hughes L, Leishman MR, Wilson PD (2010) Predicted impact of exotic vines on the
endangered ecological community under future climate change. Biol Invasions 12:4049–4063
Gehring C, Park S, Denich M (2004) Liana allometric biomass equations for Amazonian primary
and secondary forest. For Ecol Manag 195:69–83
Gerwing JJ, Frias DL (2000) Integrating liana abundance and forest stature into an estimate of total
aboveground biomass for an eastern Amazonian forest. J Trop Ecol 16:327–335
Granados J, Korner C (2002) In deep shade, elevated CO2 increases the vigor of tropical climbing
plants. Glob Chang Biol 8:1109–1117
Hattenschwiler S, Korner C (2003) Does elevated CO2 facilitate naturalization of non-indigenous
Prunus laurocerasus in Swiss temperate forests. Funct Ecol 17:778–785
Heuzé P, Dupouey JL, Schnitzler A (2008) Radial growth response of Hedera helix to hydrological
changes and climatic variability in the Rhine floodplain. River Res Appl 5:393–404
Horvitz CC, Koop A (2001) Removal of nonnative vines and post-hurricane recruitment in tropical
hardwood forests of Florida. Biotropica 33:268–281
Korner C (2006) Forests, biodiversity and CO2: surprises are certain. Biologist 53:82–90
Kurz WA, Apps MJ, Stocks BJ, Volney WJA (1995) Global climate change: disturbance regimes
and biospheric feedbacks of temperate and boreal forests. In: Woodwell GM, Mackenzie FT
(eds) Biotic feedbacks in the global climatic system. Oxford University Press, New York,
pp 119–133
Ladwig I, Meiners S (2010a) Liana host preference and implication for deciduous forest regenera-
tion. J Torrey Bot Soc 137:103–112
Ladwig I, Meiners S (2010b) Spatiotemporal dynamics of lianas during the 50 years of succession
to temperate forests. Ecology 91:671–680
Laurance WF, Andrade AS, Magrach A, Camargo JC, Valsko JJ, Campbell M, Fearnside PM,
Edwards W, Lovejoy TE, Laurance SG (2013) Long-term changes in liana abundance and
forest dynamics in undisturbed Amazonian forests. Ecology 95:1604–1611
2 Patterns of Liana Abundance, Diversity and Distribution in Temperate Forests 15

Londré RA, Schnitzer SA (2006) The distribution of lianas and their change in abundance in tem-
perate forests over the past 45 years. Ecology 87:2973–2978
Lutz HJ (1943) Injuries to trees caused by Celastrus and Vitis. Bull Torrey Bot Soc 70:436–439
McNab WH, Meeker M (1987) Oriental bittersweet: a growing threat to hardwood silviculture in
the Appalachians. North J Appl For 4:174–177
Mohan JE, Ziska LH, Schlesinger WH, Thomas RB, Sicher RC, George K, Clark JS (2006)
Biomass and toxicity responses of poison ivy (Toxicodendron radicans) to elevated atmo-
spheric CO2. Proc Natl Acad Sci 103:9086–9089
Patterson DT (1973) The ecology of oriental bittersweet, Celastrus orbiculatus, a weedy introduce
ornamental vine. Ph.D. Dissertation, Duke University, Durham
Phillips OL, Gentry AH (1994) Increasing turnover through time in tropical forests. Science
263:954–958
Phillips OL, Martinez RV, Arroyo L, Baker TR, Killeen T, Lewis SL, Malhi Y, Mendoza AM, Neill
D, Vargas PN, Alexiades M, Cerón C, De Fiore A, Erwin T, Jardim A, Palacios W, Saldias M,
Vinceti B (2002) Increasing dominance of large lianas in Amazonian forests. Nature
418:770–774
Putz FE (1984) The natural history of lianas on Barro Colorado Island, Panama. Ecology
65:1713–1724
Rutishauser SE (2011) Increasing liana abundance and biomass in tropical forests: testing mecha-
nistic explanations. M.S. thesis, University of Wisconsin – Milwaukee, Milwaukee
Sasek TW (1985) Implications of atmospheric carbon dioxide enrichment for the physiological
ecology and distribution of two introduced woody vines, Puerarialobata Ohwi (kuduzu) and
Lonicera japonica Thumb. (Japanese honeysuckle). Dissertation, Duke University, Durham
Sasek TW, Strain BR (1991) Effects of CO2 enrichment on growth and morphology of a native and
an introduce honeysuckle vine. Am J Bot 78:69–75
Schnitzer SA (2005) A mechanistic explanation of global patterns of liana abundance and distribu-
tion. Am Nat 166:262–276
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
Schnitzer SA, Bongers F (2011) Increasing liana abundance and biomass in tropical forests:
emerging patterns and putative mechanisms. Ecol Lett 14:397–406
Schnitzer A, Heuzé P (2006) Ivy (Hedera helix L.) dynamics in riverine forests: effects of river
regulation and forest disturbance. For Ecol Manage 236:12–17
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical
forest canopy gaps: evidence for an alternative pathway to gap-phase regeneration. J Ecol
88:655–666
Swaine MD, Grace J (2007) Lianas may be favored by low rainfall: evidence from Ghana. Plant
Ecol 192:271–276
Teramura AH, Gold WG, Forseth IN (1991) The biology of vines. In: Putz FE, Mooney HA (eds)
Physiological ecology of mesic, temperate woody vines. Cambridge University Press,
Cambridge, MA, pp 245–285
Wechsler NR (1977) Growth and physiological characteristics of kudzu, Pueraria loabata (Willd.)
Ohwi, in relation to its competitive success. Thesis, University of Georgia, Athens
Wright SJ, Calderon O, Hernandez A, Paton S (2004) Are lianas increasing in importance in tropi-
cal forests? A 17 year record from Panama. Ecology 85:484–489
Ziska LH, Sicher RC, George K, Mohan JE (2007) Rising atmospheric carbon dioxide and poten-
tial impacts on the growth and toxicity of poison ivy (Toxicodendron radicans). Weed Sci
55:288–292
Zotz G, Cueni N, Korner C (2006) In situ growth stimulation of a temperate zone liana (Hedera
helix) in elevated CO2. Funct Ecol 20:763–769
Chapter 3
Geographical, Taxonomical and Ecological
Aspects of Lianas in Subtropical Forests
of Argentina

Agustina Malizia, Paula I. Campanello, Mariana Villagra,


and Sergio Ceballos

Abstract Lianas are more diverse and typically more abundant in tropical than
temperate forests, with subtropical forests being intermediate. In this chapter, we
analyze geographical, taxonomical and ecological patterns of lianas in subtropical
forests of Northern Argentina, including Mountain Forests (MF), Atlantic Forests
(AF); and Dry and Humid Chaco Forests (DCh and HCh, respectively). A total of
184 woody species of climbing plants were recognized in all four subtropical for-
ests, with 35 species exclusive to MF, 38 exclusive to AF, while DCh and HCh had
2 and 8 exclusive species, respectively. In MF most liana species belonged to
Sapindaceae and Bignoniaceae (16 % each), followed by Malpighiaceae (11 %) and
Apocynaceae (10 %). In AF most liana species belonged to Bignoniaceae (21 %)
followed by Apocynaceae (12 %), Fabaceae (11 %), Malpighiaceae (11 %) and
Sapindaceae (10 %). Considering all liana species together, the most common
climbing mechanisms included tendrils and twiners. The highest liana density was
observed in the semideciduous Atlantic Forest, followed by the deciduous Humid
Chaco Forest and the semideciduous Montane Forest. The semideciduous Atlantic
Forest has also relatively high liana species richness as compared to other subtropi-
cal forests, followed by semideciduous MF. Besides geographical location and for-
est disturbances, little is known about how lianas respond to other environmental
factors that drive patterns of liana density and diversity in these subtropical forests.

Keywords Atlantic Forest • Chaco • Disturbance • Subtropical Forest • Woody


climbers • Yungas

A. Malizia (*) • S. Ceballos


CONICET, Instituto de Ecología Regional, Facultad de Ciencias Naturales e IML,
Universidad Nacional de Tucumán, Casilla de Correo 34, Yerba Buena,
Tucumán 4107, Argentina
e-mail: agustinamalizia@yahoo.com
P.I. Campanello • M. Villagra
CONICET, Instituto de Biología Subtropical, Facultad de Ciencias Forestales, Universidad
Nacional de Misiones, Bertoni 85, Puerto Iguazú, Misiones 3370, Argentina

© Springer International Publishing Switzerland 2015 17


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_3
18 A. Malizia et al.

3.1 Introduction

Lianas, the woody climbing plants, have been intensively studied in the tropics
(Putz 1984; Gentry 1991; Phillips et al. 2002; Schnitzer 2005; Schnitzer and
Bongers 2011). It is well known that they are an important component of tropical
forests where they constitute up to 32 % of the stems and up to 35 % of the woody
species diversity (for stems ≥ 1 cm dbh) (Gentry 1991; DeWalt and Chave 2004;
Schnitzer et al. 2012). Further, lianas are more diverse and typically more abundant
in tropical than temperate forests, with subtropical forests being intermediate
(Gentry 1991; Schnitzer and Bongers 2002; Schnitzer 2005). Recently DeWalt et al.
(2014) reported that subtropical montane forest of Argentina and Australia support
relatively high liana density but low diversity. Even the liana-dense subtropical
forests tend to have only a fraction of the diversity found in tropical forests
(DeWalt et al. 2014).
The abundance and diversity of lianas, like many other plant growth forms, are
governed by various factors including altitude, latitude, climate and edaphic gra-
dients which lead to geographical differences among and within regions (DeWalt
et al. 2014; Durigon et al. 2014; Alves et al. 2012). Understanding the processes
responsible for the abundance and distribution of organisms is a central goal in
biology, as well as revealing their patterns across scales. In this chapter, we
describe geographical, taxonomical and ecological patterns of lianas in subtropi-
cal forests of Northern Argentina, including Mountain Forests (MF), Atlantic
Forests (AF), Dry and Humid Chaco Forests (DCh and HCh, respectively). We
review geographical and taxonomic patterns at a larger scale and ecological aspects
at a smaller scale through case studies in the subtropical region of Argentina. Even
when knowledge on lianas is still incomplete in subtropical forests of Argentina,
studies carried out especially in the last two decades have shown their important
role in contributing to woody species richness (Meyer 1963; Giusti et al. 1995;
Killeen et al. 1998; Ayarde et al. 1999; Ayarde 2005), forest structure (Killeen
et al. 1998) and in the ecology of these forests (Lorea and Brassiolo 2007; Lorea
et al. 2008; Malizia 2003; Malizia and Grau 2006, 2008; Malizia et al. 2009, 2010;
Campanello et al. 2007a, 2009, 2012). The scientific nomenclature was updated
according to the Darwinion IRIS database (http://www.darwin.edu.ar/Proyectos/
FloraArgentina/Especies.asp).

3.2 Geographical and Taxonomical Perspective

For this section, we considered four subtropical forest types of northern Argentina,
the northwest region which covers the subtropical Mountain Forest, including
deciduous and semideciduous forests; the dry forests of Chaco and the northeast
region which includes deciduous forests of the Humid Chaco and semideciduous
forests of the Atlantic Forest (Fig. 3.1).
3 Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical… 19

Fig. 3.1 Subtropical forests types of northern Argentina, including Mountain Forests, Atlantic
Forests, and Dry and Humid Chaco Forests. Circles correspond to study sites

The subtropical Mountain forests (MF), also known as Yungas, extend along
discontinuous mountain ranges in north-western Argentina, form the border with
Bolivia (22°LS), where they continue to the north, down to Catamarca province
(29°LS) going through Salta, Jujuy and Tucumán provinces (Fig. 3.1). They repre-
sent a forest belt of 700 km extent and 50 km wide, with an altitudinal range between
400 and 3,000 m asl, covering an area of 4 million ha. These forests include the
three altitudinal belts: the deciduous piedmont belt (400–700 m asl), the semide-
ciduous lower montane forest (700–1,500 m asl) and semideciduous upper montane
forest belt (15,00–3,000 m asl) which corresponds to cloud forests. Yungas consti-
tute the southernmost extension distribution of Neotropical montane forests (Brown
et al. 2001).
The semideciduous Atlantic Forest is a subtropical forest distributed in the
extreme north-eastern Argentina, eastern Paraguay and southern Brazil. In Argentina
it is also known as “selva paranaense” (Cabrera 1971). It harbours the highest
20 A. Malizia et al.

biodiversity of the country (Burkart et al. 1999) and extends throughout the prov-
ince of Misiones, which includes the largest remaining area of continuous semide-
ciduous Atlantic Forest (10,000 km2) (Fig. 3.1). In Brazil and Paraguay, in contrast
to Argentina, this forest is highly fragmented and occupies less than 6 % of the
original cover mainly due to conversion to agriculture (Galindo-Leal and Gusmão-
Câmara 2003). The Atlantic Forest is considered one of the hotspots of biodiversity
in the world and is at high risk of extinction (Myers et al. 2000).
In Argentina, the Chaco region includes the Dry and Humid Chaco Forests
(Fig. 3.1). The Dry Chaco Forest (DCh) is one of the largest dry forest biomes of
Argentina. It is a warm region extending between Yungas and Humid Chaco Forest
and includes part of Salta, Formosa, Chaco, Santiago del Estero, Catamarca and La
Rioja provinces (Fig. 3.1). Its typical habitats are dominated by matorral and thorny
vegetation. It is the most unexplored and least documented region in the country.
The Humid Chaco Forest (HCh) covers between 16 and 19 million ha and occupies
approximately the eastern half of the provinces of Chaco and Formosa and northern
part of Santa Fe (Fig. 3.1). It is an extremely flat plain with a soft slope to the east.
The predominant landscape is fluvial and fluvio-lacustrine type that drained in
Paraguay and Parana rivers (Morello 2012). The forest vegetation is discontinuous
with grasslands and savannas (Cabrera 1971; Burkart et al. 1999).
Both the deciduous piedmont belt of the Mountain Forest and semideciduous
Atlantic Forest could be considered Seasonally Dry Neotropical Forests (SDTFs)
and are also known as the Piedmont and Misiones nucleus, respectively (Pennington
et al. 2009). The distribution of SDTF in South America forms an arc with the ends
positioned at the Caatinga domain of north-eastern Brazil and the Caribbean coast
of Colombia and Venezuela and a long curved route connecting the ends through the
seasonal forests of the Atlantic Forest domain (i.e., Misiones nucleus), the patches
of seasonal forests of the cerrado domain, and the seasonal forests of the Andean
piedmont, inter-Andean valleys, Pacific coast and Caribbean coast (Oliveira-Filho
et al. 2006). These SDTFs share many genera, and tree species such as Astronium
urundeuva (Anacardiaceae), Enterolobium contortisiliquum (Fabaceae), Ruprechtia
laxiflora (Polygonaceae), Peltophorum dubium (Fabaceae), Aspidosperma polyneu-
ron (Apocynaceae), Cordia americana (Boraginaceae), Cordia trichotoma
(Boraginaceae), Gleditsia amorphoides (Fabaceae), Cedrela fissilis (Meliaceae),
Pisonia zapallo (Nictaginaceae), Phyllostylon rhamnoides (Ulmaceae), Luehea
divaricata (Tiliaceae). In the semideciduous Atlantic Forest, although without rain-
fall seasonality, many tree species are deciduous or brevideciduous. Particularly,
species shared with the Piedmont nucleus and other SDTFs are deciduous during
winter, while tree species of Amazonian lineage or link to the Atlantic moist forest
(many Lauraceae and Sapotaceae) are evergreen (Pennington et al. 2009).
Furthermore, the SDTFs have been postulated as a phytogeographical domain in
South America based on tree species distribution (Prado 2000). In Fig. 3.2 the dis-
tribution of six liana species occurring in subtropical Argentina is shown, two spe-
cies conform to the SDTF arc distribution (Fig. 3.2a, b), while the others are
exclusive to one of the subtropical forests in the country (Fig. 3.2c–f).
Fig. 3.2 Geographic distribution for six liana species in South America according to Global
Biodiversity Information Facility databases. Note this widespread distribution of Tanaecium selloi
(a) and Vigna caracalla (b); distribution of Forsteronia glabrescens in the MF, HCh and AF of
Argentina (c), Vernonia fulta in MF (d); Fridericia truncata in the piedmont of the MF and DCh
(e), and Dicella nucifera in the semideciduous AF (f)
22 A. Malizia et al.

An extensive and detailed herbarium compilation was made by Ayarde (2005)


for subtropical MF. In the other three subtropical forests we compiled species based
on the Darwinion database and on our own records in the case of the AF. For the
Chaco region (including both DCh and HCh Forests) we also used the Darwinion
database and the revisions of Vogt (2011, 2012a, b). The distribution of all the spe-
cies was double checked at the Darwinion and the Global Biodiversity Information
Facility databases (GBIF; www.gbif.org).
A total of 184 woody species of climbing habit were registered in all four sub-
tropical forests of Argentina (Appendix). In the Mountain Forests, more than 98 %
of the species were concentrated in the piedmont and lower montane forest belt
while a small proportion (9 %) reached the upper forest belt of the mountain (i.e.
upper montane forest). Within the latitudinal gradient of Yungas, the number of spe-
cies decreased sharply from north (90 %) to south (45 %) (Malizia et al. 2009). The
total of 97 species in the MF of Argentina belong to 23 botanical families, with most
liana species belonging to the Sapindaceae and Bignoniaceae (16 % each), followed
by Malpighiaceae (11 %), Apocynaceae (10 %), Fabaceae (7 %), Vitaceae (6 %) and
Asteraceae (6 %) (Fig. 3.3). In the AF, 28 plant families were found, with most liana
species belonging to the Bignoniaceae (21 %), followed by Apocynaceae (12 %),
Fabaceae (11 %), Malpighiaceae (11 %) and Sapindaceae (10 %). In HCh the spe-
cies-rich families include Bignoniaceae (19 %), Sapindaceae (15 %) and
Malpighiaceae (14 %). In DCh three families comprised more than 50 % of the

Fig. 3.3 Plant families with most liana species in the four subtropical forests of Argentina:
Mountain Forest (a), Dry Chaco Forest (b), Humid Chaco Forest (c) and Atlantic Forest (d)
3 Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical… 23

species present and they include the Bignoniaceae (25 %), Malpighiaceae (17 %)
and Apocynaceae (13 %) (Fig. 3.3).
Considering all liana species together, the most common climbing mechanisms
used were tendrils (AF −56 species; HCh −47 species; MF −44 species; DCh −12
species) and twining (AF −39 species; HCh −30 species; MF −24 species; DCh −10
species). Tendril climbers are common in the Bignoniaceae, Sapindaceae and
Fabaceae, while species of the other most conspicuous families viz., Apocyaneae
and Malpighiaceae are usually twiners. Several species showed more than one
climbing mechanism, like scramble and spines (AF −7 species; MF −7 species;
HCh −6 species) and scramble and twining (MF = 6 species; HCh = 4 species; AF = 3
species). In MF three lianas species presented no special organ to climb but they
were capable of climbing by the deformation of all or part of their aerial system, like
stems, branches, petioles/petiolules or leaves (Caballé and Martin 2001).
Of the total 97 species encountered in MF, 35 species were exclusive of the
region while 39 species were shared with the AF and 47 species with the HCh
(Fig. 3.4). The MF shared also 19 species with the DCh. A total of 24 species were
found in the DCh, 19 of which were also present in the HCh. The HCh shared 65
species with the AF in Argentina, which had 38 exclusive species. The DCh and the
HCh had 2 and 8 exclusive species, respectively. In general, dry forests are poor in
woody climbers (Gentry 1991), which may explain the low number of species in the
DCh. But these forests have many herbaceous climbers that are present in the other
subtropical forests of Argentina and Brazil (Durigon et al. 2014). Just six species
were shared by all the subtropical forests: Forsteronia pubescens, Dolichandra
cynanchoides. Dolichandra unguis-cati, Fridericia dichotoma, Janusia guaranitica
and Cissus verticillata. The liana D. unguis-cati is a widespread species in the AF
not only of Argentina but also in Brazil (Durigon and Waechter 2011). The Venn
diagram of Fig. 3.4 supports the idea that the Chaco region, particularly the DCh
forests, does not belong to the SDTF domain (Pennington et al. 2009). In fact, we
could checked that some liana species present in both Dry and Humid Chaco forests
in Argentina also occur in the Brazilian cerrado.

Fig. 3.4 Venn diagram


showing the number of
shared and exclusive species
in the subtropical forests of
Argentina (MF mountain
forests, DCh Dry Chaco
Forests, HCh Humid Chaco
forests, AF Atlantic forests)
(Software Venny developed
by Oliveros (2007) was used
for this)
24 A. Malizia et al.

3.3 Ecological Perspective

Studies on ecological aspects of lianas in subtropical forest of Argentina have been


published in the last few years (Lorea and Brassiolo 2007; Lorea et al. 2008; Malizia
2003; Malizia and Grau 2006, 2008; Malizia et al. 2009, 2010; Campanello et al.
2007a, 2009, 2012). For the Mountain Forests of northwest Argentina, we reviewed
Malizia and Grau (2006, 2008), and Malizia et al. (2010), who carried out studies in
a contiguous 6-ha permanent plot of mature semideciduous forest characteristic of
the lower montane forest belt (i.e. at 1,000 m elevation; Brown et al. 2001). These
forests harbour an average of 23 tree species ha−1 (≥10 cm diameter) including
deciduous and evergreen species. The tree canopy (15–20 m high) is dominated by
Blepharocalyx salicifolius (Myrtaceae), Cinnamomum porphyrium (Lauraceae) and
Pisonia zapallo (Nyctaginaceae) (Malizia et al. 2010). For the Humid Chaco Forest
of northeast Argentina we reviewed Lorea (2006), Lorea and Brassiolo (2007),
Lorea et al. (2008) who established 13 plots of 250 m2 in the forest type that corre-
sponds to Monte Alto (Morello and Adámoli 1974), where the predominant vegeta-
tion is of deciduous type with the presence of numerous bromeliads, grasses and, to
a lesser extent, cacti. These forests are discontinuous stands dominated by Schinopsis
balansae (Anacardiaceae) and Aspidosperma quebracho-blanco (Apocynaceae)
that emerge like islands in depressed areas surrounded by grasslands (Morello and
Adámoli 1974). Many other species, mostly deciduous, such as Cordia americana,
Gleditsia amorphoides, Pisonia zapallo, Phyllostylon rhamnoides, Peltophorum
dubium, Ruprechtia laxiflora are also present in the Humid Chaco Forest (Giménez
and Moglia 2003). Some of these species are shared with the semideciduous Atlantic
Forest region. The forest was selectively logged 80 years ago mostly for S. balansae
exploitation. For the Dry Chaco Forest of Argentina there are no ecological studies.
Finally, for the Atlantic Forests of northeast Argentina we revised Campanello et al.
(2007a, 2012), whose studies were carried out in 6 ha permanent plots of native
subtropical semideciduous forest that was subjected to selective logging of isolated
commercial trees 40–50 years ago. Tree canopy is 25–30 m high and some domi-
nant trees include Balfourodendron riedelianum (Rutaceae), Nectandra megapota-
mica (Lauraceae), Bastardiopsis densiflora (Malvaceae), Cedrela fissilis, Cordia
americana, Cordia trichotoma and Lonchocarpus leucanthus (Fabaceae).
For all three forests sites there exists quantitative information on liana density
(Table 3.1). The highest liana density was observed in the semideciduous Atlantic
Forest (Campanello et al. 2007a), followed by the deciduous Humid Chaco Forest
(Lorea et al. 2008) and the semideciduous Montane Forest (Malizia et al. 2010).
Recently, De Walt et al. (2014) found that the number of months per year
with ≤ 100 mm year−1 (DSL- dry season length) is a factor determining liana density
worldwide. In the case of the Atlantic Forest, there is a high liana density compared
to tropical humid forests and subtropical dry forest (DeWalt et al. 2014; GLD –
global liana database). It is possible that more recent disturbance in the forest stud-
ied increased liana density compared to the old-growth forest. However, even when
the Atlantic Forest in Argentina has no dry season, it is geographically and
taxonomically linked to seasonal dry forests (Pennington et al. 2009). Dry spells of
3

Table 3.1 Density of liana stems ≥ 2 cm and ≥ 2.5 cm in all three subtropical regions of northern Argentina and values found for tropical forest in South
America. Density is the mean number of liana stems, either expressed per ha and per 0.1 ha at a given site. Richness is the number of species with stems ≥ 2.5 cm
per 0.1 ha. Fisher’s alpha was calculated for liana diversity taking into account stems ≥ 2.5 cm. Altitude – is mean altitude (m above sea level). MAP – mean
annual precipitation (mm year−1). MAT = mean annual temperature (°C). DSL – number of months per year with ≤ 100 mm rainfall
Density Density Richness
Sampled (stems ≥ 2 cm (stems ≥ 2.5 cm (species Fisher’s Altitude MAP (mm
Subtropical region area (ha) ha−1) 0.1 ha−1) 0.1 ha−1) Alpha (masl) year−1) MAT (°C) DSL (months)
Semideciduous 6 474 32 10 2.1 1,000 1,300 18 5
Montain foresta
Deciduous Humid 0.325 638 34 2 0.7 75 1,170 21 5
Chaco forestb
Semideciduous 0.680 711 50 17 8.6 250 2,000 21 0
Atlantic forestc
Tropical forests in 0.20–14 180–1,414 – – 7.6–21.8 70–620 1,500–3,000 – 0–6
South Americad
a
Malizia et al. (2010)
b
Lorea et al. (2008)
c
Campanello et al. (2007a)
d
De Walt et al. (2014)
Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical…
25
26 A. Malizia et al.

several weeks occur any time of the year (Campanello et al. 2007b). The air satura-
tion deficit is very high at midday particularly during the summer reaching values of
4.5 kPa. Thus, physiological traits related to regulation of water use and water bal-
ance are important adaptations to the environment in the semideciduous AF (Oliva
Carrasco et al. 2015). These forests contain deciduous, brevi-deciduous and ever-
green species with deciduous ones accounting for 25–50 % of the tree species (Leite
and Klein 1990). Owing to these reasons a high liana density could be expected in
the subtropical forests. Studies on species specific adaptations of liana species to the
environment would help to explain its patterns of distribution.
The semideciduous Atlantic Forest also has relatively high liana richness and
diversity compared to other subtropical forests (Table 3.1), and is comparable to
tropical seasonal forests in Bolivia (De Walt et al. 2014). According to Gentry
(1991), the highest richness and abundance of lianas occur in equatorial forests
with values of approximately 30 species and 70 individuals per 0.1 ha, and
decreases toward higher latitudes. De Walt et al. (2014) did not find definitive trend
between mean annual precipitation (MAP) and dry season length (DSL) and liana
diversity (measured through Fisher’s alpha). The confluence of two phytogeo-
graphical domains in the semideciduous Atlantic Forest (Prado 2000; Pennington
et al. 2009) would possibly contribute to a high liana diversity compared to other
subtropical forests.
A comparison of liana species composition and abundance in the three studied
forests (Fig. 3.5) reveals the differences among the three liana communities. The
semideciduous Atlantic forest is the most diverse forest, while the Humid Chaco
Forest was clearly dominated by the species Forsteronia glabrescens (83 % of
the stems measured) (Lorea et al. 2008). The shapes of the rank-abundance plots
are quite different. The curve of the semideciduous Montane Forest reveals a high
incidence of locally abundant species, which is typical of communities with high
dispersal limitation (Hubbell 1979). Phylogenetic studies support the idea that
SDTF is a highly dispersal-limited biome, with isolated units sharing a small pro-
portion of the total species (Pennington et al. 2006, 2009). In the case of the semi-
deciduous Atlantic Forest, the slope is less pronounced with many species having
intermediate and low abundances (this is most clearly visualized when stems larger
than 1 cm are taken into account, probably because of the small sample size). It is
possible that a mixture of floristic elements in this case is shaping a curve more typi-
cal of rain forests, in which the most abundant species coexist with low abundant
and rare species (Hubbell 1979). The most abundant liana species in the Atlantic
Forest were Adenocalymna marginatum, Acacia velutina, Fridericia mutabilis,
Seguieria aculeata, Adenocalymna paulistarum and Tanaecium selloi. Most of
these species have either tendrils or are scramblers and belong to the Bignoniaceae
and Fabaceae families, respectively. In this forest F. glabrescens, although present,
was not dominant. In the Montane Forest the liana community is in an intermediate
situation. The most abundant species in this forest include Cissus striata, Chamissoa
altissima and Celtis iguanaea. In the subtropical Montane and the Atlantic forests
of Argentina, at the local scale, the 18 % and 17 % of the species accounted for the
55 % and 60 % of the stems, similar to patterns of species dominance in other
Neotropical forests (Pitman et al. 2001). At a broader scale, species oligarchies (i.e.,
3 Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical… 27

1 Semideciduous Montane Forest


Deciduous Chaco Forest
Semideciduous Atlantic Forest
Relative abundance
0.1

0.01

0.001

0.0001
0 10 20 30 40 50
Species rank in abundance

Fig. 3.5 Rank-abundance plots for liana species for stems ≥ 2 cm in the subtropical forests of
Argentina (■, ▲, ●). In the case of the Atlantic Forest the distribution of lianas ≥ 1 cm is also
shown (○)

small group of species locally abundant and dominant in different type of habitats
in a biogeographical region) of trees and liana species were found to dominate
entire ecosystems such as the cerrado savannas in Brazil and different types of for-
ests in Brazil, Ecuador, Peru, Panamá and Costa Rica (Bridgewater et al. 2004;
Burnham 2002; De Walt et al. 2000; Mascaro et al. 2004; Pitman et al. 2001).
However, when considering tree species of SDTFs, no oligarchy was found to domi-
nate all the different nuclei (Pennington et al. 2009). Dominant patterns of a few
species were also found in other semideciduous Atlantic Forests in Brazil. For
example, Hora and Soares (2002) reported that the species Mansoa difficilis (also
found in Argentina) had a relative abundance of 19.5 % and was the dominant spe-
cies which agrees with the presence of oligarchies at small scales.
Lianas depend on trees for support, thus they are considered interstitial organ-
isms (sensu Houston 1995). Malizia and Grau (2006) reported that the percentage
of tree colonization (trees ≥ 10 cm dbh with at least one liana) was 65 % while Lorea
and Brassiolo (2007) reported 70 % of trees were liana-infested. These values of
tree colonization are comparable or even higher than various other tropical forests
where total tree occupancy by lianas (considering trees ≥10 cm dbh) ranged between
40 % and 63 % (Boom and Mori 1982; Clark and Clark 1990; Perez-Salicrup and
de Meijere 2005; Putz 1983; Talley et al. 1996, but see Pérez Salicrup and Sork
2001; Reddy and Parthasarathy 2006). Furthermore, disturbance, in general,
enhances liana abundance in tropical forests (Schnitzer and Bongers 2002). Forests
of the semideciduous Atlantic Forest clearly respond to disturbances. Trees in semi-
deciduous Atlantic Forests with low logging intensity have relatively low liana colo-
nization in their crowns (i.e., more than 50 % of the trees do not have any lianas and
just 2 % of the trees carried lianas covering most of their crown), while more heavily
28 A. Malizia et al.

logged forests have few trees without lianas and more than 28 % of trees carried
lianas covering most of their crown (Campanello et al. 2009). Trees may offer a
variety of niches to lianas, and the differences in their morphological and physiolog-
ical features could lead to associations between lianas and trees, either at species or
life-form levels. Alternatively, trees may serve as ecologically neutral support struc-
tures, colonized by the lianas that happen to occur close to them, and liana commu-
nities may be mostly structured by other factors largely unrelated to tree composition.
The relative contribution of tree-related niches to the assemblage of lianas may be
important in understanding the ecology of lianas (Malizia and Grau 2006).

3.4 Importance of Lianas with an Appropriate Conservation


Context and Sustainable Utilization

Tropical and subtropical forests, particularly in the Neotropics, are experiencing large-
scale structural changes; the most outstanding may be the increase in liana abundance
and biomass (Schnitzer and Bongers 2011). As lianas rely on trees for support much
of their influence on forests is likely due to their interactions with trees. They typically
have a high canopy to stem ratio, which allows them to deploy a large canopy of leaves
above those of the host tree, thus competing aggressively with their hosts and likely
having a huge effect on tree diversity, recruitment, growth and survival, which, in turn,
can alter tree community composition, carbon storage and carbon, nutrient and water
fluxes (Schnitzer and Bongers 2011). Several mechanisms have been proposed to
explain this pattern, including increasing evapotranspirative demand (i.e. decreasing
rainfall and increasing seasonality and temperature), increasing forest disturbance and
turnover, changes in land use and fragmentation, elevated atmospheric CO2 and
increasing nitrogen deposition. Each of these mechanisms probably contributes to the
observed patterns of increasing liana abundance and biomass, and the mechanisms are
likely to be interrelated and synergistic (Schnitzer and Bongers 2011).
Many subtropical forests are subject to timber extraction. Unplanned selective
logging by untrained and poorly supervised crews is still the most common method
used for timber extraction in tropical and subtropical forests. Natural gap dynamics
are profoundly altered by this timber extraction method because the removal of a
few trees per unit area causes substantial and avoidable damage to the vegetation
adjacent to the harvested trees, including trees that could be used in successive har-
vests (Campanello et al. 2009). When large-gaps are created the structure and
dynamics of these forests may be altered, either by increasing tree diversity
(Vandermeer et al. 2000; Molino and Sabatier 2001) or by enhancing the prolifera-
tion of lianas and other light-demanding fast growing species (Whitmore 1990; Putz
1991). For example, lianas respond and grow fast under these environmental condi-
tions (i.e. increases in solar radiation) and climb suitable host trees covering most of
the upper canopy and they may also act as climbing supports facilitating to other
lianas the access to the upper canopy (Campanello et al. 2007a). Thus, lianas may
inhibit the gap-phase regeneration process by inhibiting tree seedling regeneration
3 Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical… 29

and growth (Putz 1984). Similarly, when small gap opening occurs, lianas may
inhibit or delay these successional processes and the recovery of the forest may be
stalled or delayed (Schnitzer et al. 2000). In addition, the spatial context of gap
density (i.e. gaps surrounded by other gaps that occur at a similar time) is particu-
larly important for lianas composition and diversity, probably by increasing propa-
gule input and the availability of small trellises for support (Malizia and Grau 2008).
In logged forests, the abundance of lianas tends to increase inexorably over time
due, in part, to the increasing rate of large gap formation, in addition to the opening
of logging roads and log landings; and, in part due to the facilitation processes gen-
erated by themselves where lianas cause even more disturbance by increasing the
probability of falling branches and trees that in turn favors their growth and devel-
opment. This positive feedback process between the disturbance and the abundance
of lianas could help explain the increase in basal area of lianas and tree-liana loads
in the forests of northeast Argentina which actually experience selective logging
(Campanello et al. 2009) and in other forests where similar disturbances occur. In
the semideciduous Atlantic Forest, lianas were cut regardless of their diameters and,
after 10 years, substantial effect of this silvicultural treatment applied was observed
to enhance tree regeneration (Campanello et al. 2012) suggesting the effectiveness
of the treatment. Therefore, this should be performed more frequently if increased
timber productivity of forests is desired. However, frequent liana cutting could
excessively reduce the abundance of some species which in turn could affect eco-
system-level processes (i.e. several lianas share pollinators and fruit dispersers with
trees, and provide food and shelter for animals).
Besides disturbances, the increased in basal area of lianas and tree-liana loads
registered in the semideciduous Atlantic Forest may also be the result of elevated
atmospheric CO2 and increasing nitrogen deposition. For example, atmospheric N
deposition could further promote the growth of lianas than trees, due to the response
of lianas to added fertilizer even when radiation is limiting under the canopy
(Hättenschwiler 2002). According to models of nitrogen deposition worldwide, the
Atlantic Forest, one of the hotspots of biodiversity on the planet, would be currently
receiving high rates of atmospheric N deposition (between 5 and 10 kg N ha−1 yr−1)
which could be increased to 20 or 40 kg N ha−1 yr−1 in the next 40 years (Bobbink
et al. 2010; Phoenix et al. 2006). Studies to evaluate the specific responses of liana
vs. tree species to these and other ongoing environmental changes (i.e. increasing
seasonality, land use change) are needed in order to explain patterns of liana
distribution and change in the Atlantic Forest as well as in the other subtropical
forests. Unravelling the main drivers of change is difficult for several reasons,
including the paucity of environmental and historical records in most forest loca-
tions, short observation windows, temporal co-variation among potential drivers,
and delayed or non-linear responses (Clark 2007). Thus, to untangling the relative
contribution of these potential factors is a current challenge that would help to
explain the patterns of liana distribution in the subtropical forests of Argentina.

Acknowledgments Karina Buzza from SIGA PROYUNGAS made the figure of the study area.
30

Appendix

Liana species recorded in subtropical forests of northern Argentina, listed by plant family, climbing mechanism, and forest type
Forest type
Liana species Climbing mechanism UM LM P DCh HCh AF
Amaranthaceae
Alternanthera scandens Herzog Scrambler ●
Chamissoa altissima (Jacq.) Kunth Scrambler ● ● ● ●
Hebanthe eriantha (Poir.) Pedersen Scrambler ●
Hebanthe occidentalis (R.E. Fr.) Borsch & Pedersen Scrambler ● ●
Apocynaceae
Condylocarpon isthmicum (Vell.) A.DC. Twiner ●
Fischeria stellata (Vell.) E.Fourn. Twiner ●
Forsteronia glabrescens Mull. Arg. Twiner ● ● ●
Forsteronia pubescens A. DC. Twiner ● ● ● ●
Forsteronia refracta Müll. Arg. Twiner ●
Forsteronia thyrsoidea (Vell.) Müll. Arg. Twiner ● ● ●
Jobinia lindbergii E. Fourn. Twiner ●
Macropharynx meyeri (C. Ezcurra) Xifreda Twiner ●
Macroscepis aurea E. Fourn. Twiner ● ● ●
Mandevilla pentladiana (A. DC.) Woodson Twiner ●
Marsdenia macrophylla (Humb. & Bonpl. ex Schult) E. Fourn. Twiner ●
Mesechites trifidus (Jacq.) Müll. Arg. Twiner ● ●
Oxypetalum erianthum Decne. Twiner ● ● ●
Peltastes peltatus (Vell.) Woodson Twiner ● ●
Prestonia coalita (Vell.) Woodson Twiner ● ●
Prestonia cyaniphylla (Rusby) Woodson Twiner ● ●
A. Malizia et al.
3

Prestonia quinquangularis (Jacq.) Spreng. Twiner ● ● ●


Prestonia riedelli (Mull. Arg.) Markgr Twiner ● ● ● ●
Aristolochiaceae
Aristolochia esperanzae Kuntze Tendril ●
Aristolochia gibertii Hook. Tendril ●
Aristolochia odoratissima L. Tendril ●
Aristolochia triangularis Cham. Tendril ● ●
Euglypha rojasiana Chodat & Hassl. Tendril ●
Asteraceae
Mikania glomerata Spreng. Twiner ● ●
Mikania variifolia Hieron. Twiner ● ●
Mutisia acuminata Ruiz et Pav. var paucijuga (Griseb.) Cabrera Tendril ●
Mutisia saltensis Cabrera Tendril ●
Piptocarpha sellowii (Sch. Bip.) Baker Scrambler ● ●
Pseudogynoxys cabrerae H. Rob. et Cuatrec. Twiner ● ● ● ●
Salmea scandens (L.) DC. Scrambler ● ●
Vernonia fulta Griseb. Scrambler ● ●
Wedelia saltensis Cabrera Scrambler ● ●
Bignoniaceae
Adenocalymma marginatum (Cham.) DC. marginatum Tendril ● ●
Adenocalymma paulistarum Bureau & K. Schum. Tendril ●
Adenocalymma scansile Miers Tendril ●
Amphilophium carolinae (Lindl.) L.G. Lohmann Tendril ● ● ●
Amphilophium crucigerum (L.) L.G. Lohmann Tendril ● ● ●
Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical…

Amphilophium paniculatum (L.) Kunth Tendril ● ●


Amphilophium pannosum (DC.) Bureau & K. Schum. Tendril ●
Amphilophium sandwithii Fabris Tendril ● ●

(continued)
31
32

Forest type
Liana species Climbing mechanism UM LM P DCh HCh AF
Bignonia binata Thumb. Tendril ● ●
Bignonia callistegioides Cham. Tendril ● ● ●
Bignonia decora (S. Moore) L.G. Lohmann Tendril ●
Bignonia sciuripabulum (K. Schum.) L.G. Lohmann Tendril ●
Cuspidaria convoluta (Vell.) A.H. Gentry Tendril ● ●
Dolichandra chodatii (Hassl.) L.G. Lohmann Tendril ● ●
Dolichandra cynanchoides Cham. Tendril ● ● ● ● ●
Dolichandra quadrivalvis (Jacq.) L.G. Lohmann Tendril ● ● ●
Dolichandra uncata (Andrews) L.G. Lohmann Tendril ● ●
Dolichandra unguis-cati (L.) L.G. Lohmann Tendril ● ● ● ● ●
Fridericia candicans (Rich.) L.G. Lohmann Tendril ●
Fridericia caudigera (S. Moore) L.G. Lohmann Tendril ● ●
Fridericia chica (Bonpl.) L.G. Lohmann Tendril ●
Fridericia dichotoma (S. Moore) L.G. Lohmann Tendril ● ● ● ●
Fridericia florida (DC.) L.G. Lohmann Tendril ● ●
Fridericia samydoides (Cham.) L.G. Lohmann Tendril ● ●
Fridericia triplinervia (DC.) L.G. Lohmann Tendril ● ●
Fridericia truncata (Sprague) L.G. Lohmann Tendril ● ●
Mansoa difficilis (Cham.) Bureau & K. Schum. Tendril ● ●
Pyrostegia venusta (Ker Gawl.) Miers Tendril ● ●
Tanaecium cyrtathum (Mart. ex DC.) Bureau et K. Schum. Tendril ● ●
Tanaecium mutabile (Bureau & K. Schum.) L.G. Lohmann Tendril ● ●
Tanaecium selloi (Spreng.) L.G. Lohmann Tendril ● ● ●
Tourretia lappacea (L’Her) Willd. Tendril ● ●
Tynanthus micranthus Corr. Méllo ex K. Schum. Tendril ●
A. Malizia et al.
3

Cactaceae
Pereskia aculeata Mill. Twiner ● ●
Celtidaceae
Celtis iguanaea (Jacq.) Sarg Scrambler/Spines ● ● ● ●
Cucurbitaceae
Siolmatra brasiliensis (Cogn.) Baill. var. pubescens (Griseb.) Cogn. Tendril ● ● ●
Dilleniaceae
Tetracera oblongata DC. Twiner ●
Euphorbiaceae
Dalechampia stipulacea Müll. Arg. Twiner ● ●
Fabaceae
Acacia nitidifolia Speg. Scrambler ●
Acacia parviceps (Speg.) Burkart Scrambler/Spines ● ● ●
Acacia tucumanensis Griseb. Scrambler/Spines ● ● ● ●
Acacia velutina DC. Scrambler/Spines ●
Bauhinia microstachya (Raddi) J.F. Macbr. Tendril ●
Canavalia piperi Killip et J.F. Macbr. Twiner ●
Clitoria cordobensis Burkart Twiner ●
Cratylia intermedia (Hassl.) L.P. Queiroz & R. Monteiro Tendril ●
Dalbergia frutescens (Vell.) Britton Tendril ● ●
Dioclea violacea Mart. ex Benth. Tendril ● ●
Mimosa bimucronata (DC.) Kuntze Tendril ● ●
Mimosa sensibilis Griseb. Scrambler/Spines ● ●
Mimosa velloziana Mart. Scrambler/Spines ● ● ●
Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical…

Mucuna sloanei Fawc. & Rendle Twiner ●


Nissolia fruticosa Jacq. Twiner ● ● ● ●
Vigna caracalla (L.) Verdc. Tendril ● ●

(continued)
33
34

Forest type
Liana species Climbing mechanism UM LM P DCh HCh AF
Hippocrateaceae
Elachyptera micrantha (Cambess.) AC. Sm. Twiner ●
Hippocratea volubilis L. Twiner ● ●
Pristimera andina Miers Twiner ● ●
Liliaceae
Herreria bonplandii Lecomte Scrambler/Twiner ● ● ●
Herreria montevidensis Klotzsch ex Griseb. Twiner ● ●
Logoniaceae
Strychnos brasiliensis (Spreng.) Mart. Scrambler ●
Malphighiaceae
Alicia anisopetala (A. Juss.) W.R. Anderson Twiner ●
Banisteriopsis caapi (Griseb.) C.V. Morton Twiner ● ● ●
Banisteriopsis lutea (Griseb.) Cuatrec. Twiner ● ●
Banisteriopsis muricata (Cav.) Cuatrec. Scrambler/Twiner ● ● ●
Callaeum psilophyllum (A. Juss.) D.M. Johnson Scrambler/Twiner ● ●
Dicella nucifera Chodat Twiner ● ●
Heladena multiflora (Hook. & Arn.) Nied. Twiner ●
Heteropterys amplexicaulis Morong Twiner ●
Heteropterys argyrophaea A. Juss. Twiner ● ●
Heteropterys bicolor A. Juss. Twiner ●
Heteropterys cochleosperma A. Juss. Scrambler/Twiner ● ●
Heteropterys dumetorum (Griseb.) Nied. Scrambler/Twiner ● ● ● ●
Heteropterys intermedia (A. Juss.) Griseb. Twiner ●
Heteropterys mollis (Nied.) Nied. Scrambler ●
Heteropterys schulziana W.R. Anderson Scrambler/Twiner ●
A. Malizia et al.
3

Heteropterys sylvatica A. Juss Scrambler ● ●


Hiraea fagifolia (DC.) A. Juss. Twiner ● ●
Janusia guaranitica (St. Hil.) A. Juss. Twiner ● ● ● ● ●
Mascagnia brevifolia Griseb. Twiner ● ●
Mascagnia divaricata (Kunth) Nied. Twiner ● ●
Niedenzuella sericea A. Juss. W.R. Anderson Twiner ●
Stigmaphyllon bonariense (Hook. & Arn.) C.E. Anderson Twiner ● ●
Stigmaphyllon jatrophifolium A. Juss. Twiner ● ●
Menispermaceae
Odontocarya acuparata Miers Twiner ● ●
Odontocarya asarifolia Barneby Twiner ● ● ●
Nyctaginaceae
Pisionella arborescens (Lag. Et Rodr.) Standl Scrambler ● ●
Pisonia aculeata L. Scrambler ● ●
Passifloraceae
Passiflora alata Curtis Tendril ●
Passiflora edulis Sims Tendril ●
Passiflora elegans Mast. Tendril ●
Phytolaccaceae
Seguieria aculeata Jacq. Scrambler/Spines ● ● ●
Trichostigma octandrum (L.) H. Walter Scrambler/Spines ● ● ●
Polygonaceae
Muehlenbeckia sagittifolia (Ortega) Meisn. Twiner ● ● ● ●
Ranunculaceae
Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical…

Clematis bonariensis Juss. Ex DC. Twiner ●


Clematis campestris A. St.-Hil. Twiner ●
Clematis dioica L. Twiner ●

(continued)
35
36

Forest type
Liana species Climbing mechanism UM LM P DCh HCh AF
Clematis haenkeana C. Presl. No special organs ● ●
Clematis montevidensis Spreng. No special organs ● ● ● ●
Rhamnaceae
Gouania latifolia Reissek. Tendril ●
Gouania lupuloides (L.) Urb. Tendril ● ● ●
Gouania polygama (Jacq.) Urb. Tendril ● ● ●
Gouania ulmifolia Hook. & Arn. Tendril ●
Rosaceae
Rubus imperialis Cham et Schltdl. Spines ● ● ●
Rubiaceae
Chiococca alba (L.) C.L. Hitchc Scrambler ● ●
Sapindaceae
Paullinia elegans Cambess Tendril ● ●
Paullinia meliaefolia Juss. Tendril ●
Paullinia pinnata L. Tendril ● ●
Serjania ampelopsis Planch. et Linden Tendril ●
Serjania caracasana (Jacq.) Willd. Tendril ● ● ● ●
Serjania confertiflora Radlk. Tendril ●
Serjania foveata Griseb. Tendril ● ● ●
Serjania fuscifolia Radlk. Tendril ● ●
Serjania glabrata Kunth Tendril ● ● ● ●
Serjania glutinosa Radlk. Tendril ● ●
Serjania hebecarpa Benth. Tendril ● ● ● ●
Serjania laruotteana Cambess Tendril ● ●
Serjania longistipula Radlk Tendril ●
A. Malizia et al.
3

Serjania marginata Casar. Tendril ● ● ●


Serjania meridionalis Cambess Tendril ● ● ● ●
Serjania perulacea Radlk. Tendril ● ● ●
Serjania sufferruginea Radlk. Tendril ● ●
Serjania tripleuria Ferrucci Tendril ● ● ●
Thinouia mucronata Radlk. Tendril ● ● ● ●
Urvillea chacoensis Hunz. Tendril ● ●
Urvillea laevis Radlk. Tendril ●
Urvillea ulmacea Kunth Tendril ● ● ●
Urvillea uniloba Radlk. Tendril ●
Smilacaceae
Smilax assumptionis A.DC. Tendril ●
Smilax campestris Griseb. Tendril ● ● ● ●
Smilax cognata Kunth Tendril ●
Smilax fluminensis Steud. Tendril ●
Smilax pilcomayensis Guagl. & Gattuso Tendril ●
Solanaceae
Salpichroa scandens Dammer Scrambler ●
Salpichroa tristis Miers var. tristis Scrambler ●
Solanum calileguae Cabrera Scrambler ●
Solanum hirtellum (Spreng.) Hassl. Scrambler ●
Solanum ipomoeoides Chodat No special organs ● ● ●
Sterculiaceae
Byttneria catalpifolia Jacq. Twiner ●
Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical…

Byttneria filipes Mart. ex K. Schum Twiner ● ●


Byttneria oranensis Cristóbal Twiner ● ●

(continued)
37
38

Forest type
Liana species Climbing mechanism UM LM P DCh HCh AF
Violaceae
Anchietea pyrifolia Hallier Twiner ● ●
Vitaceae
Cissus palmata Poir. Tendril ● ● ●
Cissus simsiana Schult. et Schult. fil. Tendril ●
Cissus striata Ruiz et Pav. Tendril ● ● ●
Cissus subrhomboidea (Baker) Planch. Tendril ●
Cissus sulcicaulis Planch Tendril ● ●
Cissus tweediana (Baker) Griseb. Tendril ● ● ●
Cissus verticillata (L.) Nicolson et C.E. Jarvis Tendril ● ● ● ● ●

UM upper montane forest, LM lower montane forest, P piedmont, DCh Dry Chaco Forest, HCh Humid Chaco forest, AF Atlantic forest
A. Malizia et al.
3 Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical… 39

References

Alves LF, Assis MA, van Melis J, Barros ALS, Vieira SA, Martins FR, Martinelli LA, Joly CA
(2012) Variation in liana abundance and biomass along an elevational gradient in the tropical
Atlantic Forest (Brazil). Ecol Res 27:323–332
Ayarde HR (2005) Vegetación lianescente de las áreas montanas del noroeste de Argentina. Lilloa
42:95–128
Ayarde HR, Boero C, Moris M, Slanis A, González JA (1999) Flora y Vegetación de Tariquía. In:
González JA, Scrocchi GJ, Lavilla EO (eds) Relevamiento de la biodiversidad de la Reserva
Nacional de Flora y Fauna Tariquía (Tarija, Bolivia). Fundación Miguel Lillo, Tucumán,
pp 29–62
Bobbink R, Hicks WK, Galloway JN, Spranger T, Alkemade R, Ashmore M, Bustamante M,
Cinderby S, Davidson E, Dentener F, Emmett B, Erisman JW, Fenn M, Gillian F, Nordin A,
Pardo L, De Vries W (2010) Global assessment of nitrogen deposition effects on terrestrial
plant diversity: a synthesis. Ecol Appl 20:30–59
Boom BM, Mori SA (1982) Falsification of two hypotheses on liana exclusion from tropical trees
possessing buttresses and smooth bark. Bull Torrey Bot Club 109:447–450
Bridgewater S, Ratter JA, Ribeiro JF (2004) Biogeographic patterns, β-diversity and dominance in
the cerrado biome of Brazil. Biodiv Conserv 13:2295–2318
Brown AD, Grau HR, Malizia LR, Grau A (2001) Argentina. In: Kappelle M, Brown AD (eds)
Bosques nublados del Neotrópico. INBIO, Santo Domingo de Heredia, pp 623–658
Burkart R, Bárbaro NO, Sánchez RO, Gómez DA (1999) Eco-regiones de la Argentina.
Administración de Parques Nacionales, Buenos Aires
Burnham RJ (2002) Dominance, diversity and distribution of lianas in Yasuní, Ecuador: who is on
top? J Trop Ecol 18:845–864
Caballé G, Martin A (2001) Thirteen years of change in trees and lianas in a Gabonese rain forest.
Plant Ecol 152:167–173
Cabrera AL (1971) Fitogeografía de la República Argentina. Bol Soc Argent Bot 14:1–42
Campanello PI, Garibaldi JF, Gatti MG, Goldstein G (2007a) Lianas in a subtropical Atlantic
Forest: host preference and tree growth. For Ecol Manage 242:250–259
Campanello PI, Gatti MG, Ares A, Montti L, Goldstein G (2007b) Tree regeneration and microcli-
mate in a liana and bamboo-dominated semideciduous Atlantic forest. For Ecol Manage
252:108–117
Campanello PI, Montti L, Mac Donagh P, Goldstein G (2009) Reduced impact logging and post-
harvesting forest management in the Atlantic Forest: Alternative approaches to enhance canopy
tree growth and regeneration. In: Grossberg SP (ed) Forest management. Nova, New York,
pp 39–59
Campanello PI, Villagra M, Garibaldi JF, Ritter LJ, Araujo JJ, Goldstein G (2012) Liana abun-
dance, tree crown infestation, and tree regeneration ten years after liana cutting in a subtropical
forest. For Ecol Manage 284:213–221
Clark DA (2007) Detecting tropical forests’ responses to global climatic and atmospheric change:
current challenges and a way forward. Biotropica 39:4–19
Clark DB, Clark DA (1990) Distribution and effects on tree growth of lianas and woody hemi-
epiphytes in a Costa Rican tropical wet forest. J Trop Ecol 6:321–331
DeWalt SJ, Chave J (2004) Structure and biomass of four lowland Neotropical forests. Biotropica
36:7–19
DeWalt SJ, Schnitzer SA, Denslow JS (2000) Density and diversity of lianas along a chronose-
quence in a central Panamanian lowland forest. J Trop Ecol 16:1–19
DeWalt SJ, Schnitzer SA, Alves LF, Bongers F, Burnham RJ, Cai Z, Carson WP, Chave J, Chuyong
GB, Costa FRC, Ewango CEN, Gallagher RV, Gerwing JJ, Gortaire Amezcua E, Hart T,
Ibarra-Manríquez G, Ickes K, Kenfack D, Letcher S, Macía MC, Makana JR, Malizia A,
Martínez-Ramos M, Mascaro J, Muthumperumal C, Muthuramkumar S, Nogueira A, Parren
MPE, Parthasarathy N, Pérez-Salicrup DR, Putz FE, Romero-Saltos H, Sridhar Reddy M,
40 A. Malizia et al.

Nsanyi Sainge M, Thomas D, van Melis J (2014) Biogeographical patterns in liana abundance
and diversity. In: Schnitzer S, Burnham R, Putz J, Bongers F (eds) Ecology of lianas. Wiley-
Blackwell, Oxford (in press)
Durigon J, Waechter JL (2011) Floristic composition and biogeographic relations of a subtropical
assemblage of climbing plants. Biodiv Conserv 20:1027–1044
Durigon J, Miotto STS, Gianoli E (2014) Distribution and traits of climbing plants in subtropical
and temperate South America. J Veg Sci 25:1484–1492
Galindo-Leal C, Gusmão-Câmara I (2003) The Atlantic forest of South America: biodiversity sta-
tus, threats and outlook. Island Press, Washington, DC
Gentry AH (1991) The distribution and evolution of climbing plants. In: Putz FE, Mooney HA
(eds) The biology of Vines. Cambridge University Press, Cambridge, pp 3–49
Giménez AM, Moglia JG (2003) Árboles del Chaco Argentino: guía para el reconocimiento den-
drocronológico. Editorial El Liberal, Santiago del Estero
Giusti L, Slanis A, Aceñolaza P (1995) Fitosociología de los bosques de aliso (Alnus acuminata
HBK. ssp. acuminata) de Tucumán (Argentina). Lilloa 38:93–120
Hättenschwiler S (2002) Liana seedling growth in response to fertilization in a Neotropical forest
understorey. Basic Appl Ecol 3:135–143
Hora RC, Soares JJ (2002) Estrutura fitossociológica da comunidade de lianas em uma floresta
estacional semidecidual na Fazenda Canchim, São Carlos, SP. Rev Bras Bot 25:323–329
Houston M (1995) Biological diversity: the coexistence of species in changing landscapes.
Cambridge University Press, Cambridge
Hubbell SP (1979) Tree dispersion, abundance, and diversity in a tropical dry forest. Science
203:1299–1309
Killen TJ, Jardim A, Mamani F, Rojas N (1998) Diversity, composition and structure of tropical
semidecidua forest in the Chiquitanía region of Santa Cruz, Bolivia. J Trop Ecol 14:803–827
Leite PF, Klein RM (1990) Vegetação. In: IBGE (ed) Geografia do Brasil: Região Sul. Instituto
Brasileiro de Geografia e Estatística, Rio de Janeiro, pp 113–150
Lorea L (2006) Lianas en bosques del Chaco húmedo. Descripción de su participación en la estruc-
tura del bosque. Thesis. Universidad Nacional de Santiago del Estero
Lorea L, Brassiolo MM (2007) Establecimiento de lianas sobre los árboles de un bosque del Chaco
Húmedo Argentino. Rev Forest Venez 51:47–55
Lorea L, Brassiolo MM, Gomez C (2008) Abundancia y diversidad de lianas en un bosque del
Chaco húmedo argentino. Quebracho 16:41–50
Malizia A (2003) Host tree preference of vascular epiphytes and climbers in a subtropical montane
cloud forest of Northwest Argentina. Selbyana 24:196–205
Malizia A, Grau HR (2006) Liana – host tree associations in a subtropical montane forest of north-
western Argentina. J Trop Ecol 22:331–339
Malizia A, Grau HR (2008) Landscape context and microenvironment influences on liana com-
munities within treefall gaps. J Veg Sci 19:597–604
Malizia A, Ayarde H, Sasal Y (2009) Ecología y diversidad de lianas en la selva pedemontana de
las yungas australes. In: Brown AD, Blendinger P, Lomáscolo T, García Bes P (eds) Ecología,
historia natural y conservación de la Selva Pedemontana de las Yungas Australes. Ediciones del
Subtrópico, Tucumán, pp 75–104
Malizia A, Grau HR, Lichstein JW (2010) Soil phosphorus and disturbance control liana commu-
nities in a subtropical montane forest. J Veg Sci 21:551–560
Mascaro J, Schnitzer SA, Carson WP (2004) Liana diversity, abundance, and mortality in a tropical
wet forest in Costa Rica. For Ecol Manage 190:3–14
Meyer T (1963) Estudios sobre la Selva Tucumana. La Selva de Mirtáceas de las Pavas. Opera
Lilloana 10:1–144
Molino JF, Sabatier D (2001) Tree diversity in tropical rain forests: a validation of the intermediate
disturbance hypothesis. Science 294(5547):1702–1704
Morello J (2012) Ecoregión Chaco Húmedo. In: Morello J, Matteucci SD, Rodriguez AF, Silva ME
(eds) Ecoregiones y complejos ecosistémicos argentinos. Ediciones FADU, Buenos Aires,
pp 205–233
3 Geographical, Taxonomical and Ecological Aspects of Lianas in Subtropical… 41

Morello J, Adámoli J (1974) Las grandes unidades de vegetación y ambiente del Chaco Argentino.
Vegetación y Ambiente de la Provincia del Chaco. INTA, Argentina
Myers N, Mittermeier RA, Mittermeier CG, Fonseca GAB, Kent J (2000) Biodiversity hotspots for
conservation priorities. Nature 403:853–858
Oliva Carrasco L, Bucci SJ, Di Francescantonio D, Lezcano OA, Campanello PI, Scholz FG,
Rodríguez S, Madanes N, Cristiano PM, Hao GY, Holbrook NM, Goldstein G (2015) Water
storage dynamics in the main stem of subtropical tree species differing in wood density, growth
rate, and life history traits. Tree Physiol. doi:10.1093/treephys/tpu087
Oliveira-Filho AT, Jarenkow JA, Rodal MJN (2006) Floristic relationships of seasonally dry for-
ests of eastern South America based on tree species distribution patterns. In: Pennington RT,
Lewis GP, Ratter JA (eds) Neotropical savannas and dry forests: plant diversity, biogeography
and conservation. Taylor & Francis/CRC Press, Oxford, pp 159–192
Oliveros JC (2007) VENNY. An interactive tool for comparing lists with Venn Diagrams. http://
bioinfogp.cnb.csic.es/tools/venny/index.html
Pennington RT, Richardson JA, Lavin M (2006) Insights into the historical construction of species-
rich biomes from dated plant phylogenies, phylogenetic community structure and neutral eco-
logical theory. New Phytol 172:605–616
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
Pérez Salicrup DR, Sork VL (2001) Lianas and trees in a liana forest of Amazonian Bolivia.
Biotropica 33:34–47
Perez-Salicrup DR, de Meijere W (2005) Number of lianas per tree and number of trees climbed
by lianas at Los Tuxtlas, Mexico. Biotropica 37:153–156
Phillips OL, Vásquez Martínez R, Arroyo L, Baker TR, Killeen TJ, Lewis SL, Malhi Y, Monteagudo
Mendoza A, Neill D, Núñez Vargas P, Alexiades M, Cerón C, Di Fiore A, Erwin T, Jardim A,
Palacios W, Saldias M, Vinceti B (2002) Increasing dominance of large lianas in Amazonian
forests. Nature 418:770–774
Phoenix GK, Hicks WK, Cinderby S, Kuylenstierna JCI, Stock WD, Dentener FJ, Giller KE,
Austin AT, Lefroy RDB, Gimeno BS, Ashmore MR, Ineson P (2006) Atmospheric nitrogen
deposition in world biodiversity hotspots: the need for a greater global perspective in assessing
N deposition impacts. Glob Change Biol 12:470–476
Pitman NCA, Terborgh JW, Silman MR, Nuñez PV, Neil DA, Cerón CE, Palacios WA, Aulestia M
(2001) Dominance and distribution of tree species in upper Amazonian terra firme forests.
Ecology 82:2101–2117
Prado DE (2000) Seasonally dry forests of tropical South America: from forgotten ecosystems to
a new phytogeographic unit. Edinb J Bot 57:437–461
Putz FE (1983) Liana biomass and leaf area of a “tierra firme” forest in the Rio Negro basin,
Venezuela. Biotropica 15:185–189
Putz FE (1984) The natural history of lianas on Barro Colorado Island, Panama. Ecology
65:1713–1724
Putz FE (1991) Silvicultural effects on lianas. In: Putz FE, Mooney HA (eds) The biology of vines.
Cambridge University Press, New York, pp 493–501
Reddy MS, Parthasarathy N (2006) Liana diversity and distribution on host tress in four inland
tropical dry evergreen forests of peninsular India. Trop Ecol 47:103–116
Schnitzer SA (2005) A mechanistic explanation for global patterns of liana abundance and distri-
bution. Am Nat 166:262–276
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
Schnitzer SA, Bongers F (2011) Increasing liana abundance and biomass in tropical forests:
emerging patterns and putative mechanisms. Ecol Lett 14:397–406
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical
forest canopy gaps: evidence for an alternative pathway of gap-phase regeneration. J Ecol
88:655–666
42 A. Malizia et al.

Schnitzer SA, Mangan SA, Dalling JW, Baldeck CA, Hubbell SP, Ledo A, Muller-Landau H,
Tobin MF, Aguilar S, Brassfield D, Hernandez A, Lao S, Perez R, Valdes O, Rutishauser Yorke
S (2012) Liana abundance, diversity, and distribution on Barro Colorado Island, Panama.
PLoS ONE 7:e52114. doi:10.1371/journal.pone.0052114
Talley SM, Setzer WN, Jackes BR (1996) Host associations of two adventitious root climbing
vines in a north Queensland tropical rain forest. Biotropica 28:356–366
Vandermeer J, de la Cerda IG, Boucher D, Perfecto I, Ruiz J (2000) Hurricane disturbance and
tropical tree species diversity. Science 290:788–791
Vogt CV (2011) Composición de la flora vascular del Chaco Boreal, Paraguay I. Pteridophyta y
Monocotyledonae. Steviana 3:13–47
Vogt CV (2012a) Composición de la flora vascular del Chaco Boreal, Paraguay II. Dycotiledonae:
Achantaceae-Fabaceae. Steviana 4:65–116
Vogt CV (2012b) Composición de la flora vascular del Chaco Boreal, Paraguay III. Dycotiledonae:
Gesneriaceae-Zygophyllaceae. Steviana 5:5–40
Whitmore TC (1990) An introduction to tropical rain forests. Clarendon, Oxford
Chapter 4
Liana Effects on Carbon Storage and Uptake
in Mature and Secondary Tropical Forests

Sandra M. Durán and G.A. Sánchez-Azofeifa

Abstract Lianas are a key structural component of tropical forests, where they
represent approximately 25 % of woody plant species. Lianas reduce tree growth,
inhibit tree regeneration and increase tree mortality. Thus, lianas are able to reduce
carbon stored as tree biomass. Infestation rates on trees by lianas are stronger in
shade-tolerant species with high wood density, which store more carbon than fast-
growing species. Therefore, lianas may promote shifts in species composition and
threaten tree carbon storage capacity of tropical forests. Lianas have shown consis-
tent increases in density and biomass in tropical regions in the last decade, which
may have profound consequences for forest dynamics. In this chapter, we review
available evidence of liana effects on carbon cycling in mature and secondary tropi-
cal forests. Secondary forests now cover larger areas than mature forests, but their
role in carbon cycling is unclear. Lianas are more prevalent in early stages of suc-
cession, and could have disproportionate effects on carbon uptake in secondary for-
ests. Current knowledge indicates that lianas could reduce carbon stocks by up to
50 % and reduce carbon increment by 10 % in mature tropical forests. In secondary
forests, evidence is quite limited; but one study found that lianas reduce 9–18 % of
carbon accumulation in treefall gaps. Changes in composition by lianas are not yet
supported by literature. We identify research needs required to improve predictions
of how tropical carbon sinks will respond to liana increases.

Keywords Liana increases • Global change • Tropical forests • Carbon cycling •


Secondary forests

S.M. Durán (*) • G.A. Sánchez-Azofeifa


Department of Earth and Atmospheric Sciences, University of Alberta,
Edmonton, AB T6G 2E3, Canada
e-mail: sduran@ualberta.ca

© Springer International Publishing Switzerland 2015 43


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_4
44 S.M. Durán and G.A. Sánchez-Azofeifa

4.1 Introduction

Tropical forests store over 30 % of the global carbon and account for 32 % of the
global primary productivity (Malhi 2012). Thus, any alteration in tropical forests
can have strong impacts in the global carbon cycle and ultimately in the global cli-
mate. Growing evidence suggests that tropical forests are experiencing major struc-
tural changes as a consequence of forest fragmentation and logging (Schnitzer and
Bongers 2011). Some of these changes include increases in rainfall seasonality, air
temperature, atmospheric CO2 and liana abundance and biomass (Lewis et al. 2009).
Liana abundance relative to trees in tropical forests has increased by 1.7–4.6 % per
year (Phillips et al. 2002). Nonetheless, most research in tropical forest dynamics
and variation in tropical forest carbon stocks continue to focus on the role of climate
(Stegen et al. 2011).
Lianas are a key structural component of tropical forests, they represent on aver-
age 20–45 % of the woody biomass in mature tropical forests and 25 % of all woody
species (Schnitzer and Bongers 2002). Lianas depend on trees for physical support
in order to reach the forest canopy and, as a result can be detrimental for host trees
by creating mechanical stresses (Pérez-Salicrup and Barker 2000), competing
for above and belowground resources (Chen et al. 2008), reducing tree growth
(Schnitzer 2005; van der Heijden and Phillips 2009), fecundity (Wright et al. 2005),
survival and recruitment (Schnitzer and Carson 2010).
Lianas have been considered light-loving plants, because they respond positively
to disturbance and are more prevalent in areas of secondary forest succession (Paul
and Yavitt 2011). Nonetheless, liana communities have been described primarily in
disturbed areas or mature forests, with little research on lianas during secondary forest
succession (Letcher and Chazdon 2009). Secondary forest succession is defined as the
woody vegetation that re-grows after complete forest clearance for pasture, agricul-
ture or other human activities such as clear-cutting or timber extraction (Chazdon
et al. 2007). Secondary forests increasingly dominate tropical regions, and currently
occupy more area than mature forests (Laurance 2010), thus, tropical forest succes-
sion constitutes a central topic in forest ecology (Chazdon et al. 2007). Secondary
forests show fast rates of aboveground production, especially during early stages of
succession, and can sequester up to 1.4 Pg C year−1, ameliorating raising levels of
atmospheric CO2 (Hughes et al. 1999). Expansion of secondary forests, however, may
promote an increment in liana abundance and biomass, since they provide an optimal
balance of tree host availability and high light (Schnitzer and Bongers 2011).
Here we review current evidence on the effects of lianas on carbon cycling in
mature and secondary forests in the tropics, specifically, we focus on aboveg-
round carbon storage and net primary productivity. We compiled information
available through published literature and in online databases. Most data found
provided values of forest carbon as plant biomass, thus we converted these values
to carbon pools by assuming that carbon accounts for 47 % of woody biomass
(Hughes et al. 1999). Although, net primary productivity includes both annual
changes in plant growth and litterfall production, we discuss them separately as
lianas have an important contribution to litterfall production; though can cause
4 Liana Effects on Carbon Storage and Uptake in Mature and Secondary Tropical… 45

reductions in woody productivity (e.g. tree growth). Throughout the review, we


provide information on the contribution of lianas to carbon cycling in order to
illustrate whether liana proliferation in tropical forests could compensate for
reduction in carbon pools.

4.2 Liana Effects on Carbon Cycling in Mature Forests

4.2.1 Aboveground Carbon Storage

Estimations of the current carbon sink in tropical areas are still under debate.
According to Pan et al. (2011), tropical forests sequester 1.1 Pg C year−1 but it
could be less (Wright 2013). Moreover, carbon sinks in tropical forests have
already undergone a decline from 1.5 Pg C year−1 to 1.1 Pg C year−1 from 1990
to 2007, and future projections still hold 10–20 % uncertainty (Pan et al. 2011).
This uncertainty is due to in part to the unknown response of tropical forests to
greater CO2 concentrations. Liana proliferation in tropical forests has been asso-
ciated with raising CO2, but the overall effects on net carbon balance are not yet
understood. This limited knowledge is threatened for the lack long-term assess-
ment of forest dynamics, and the lack of liana censuses in most inventories in
tropical regions.
To date only two reports provided confirmation of the negative effect of liana
density on biomass carbon in mature tropical forests. One of them examined the
relationship between the number of lianas (2.5 cm diameter) and tree carbon stor-
age (>10 cm diameter) in 145 locations worldwide, and estimated that lianas could
reduce aboveground carbon storage by up to 50 % (Durán and Gianoli 2013). A
more recent study in 36 sites in the Amazon, also found a negative effect of liana
(2 cm diameter) density on biomass carbon, with lianas explaining 18 % variation
in biomass of trees > 10 cm diameter (Laurance et al. 2014). The loss in tree bio-
mass, however, may not necessarily be compensated by a buildup in liana biomass.
Liana stems generally constitute less than 10 % of the aboveground carbon storage
in mature tropical forests (DeWalt and Chave 2004; Durán and Gianoli 2013) due to
their relatively slender stems and low wood density (Laurance et al. 1997; DeWalt
and Chave 2004). In central Amazon, total aboveground biomass of trees declined
after forest fragmentation by 36.1 Mg ha−1, while liana biomass raised by only
0.46 Mg ha−1 (Laurance et al. 1997).
Reductions in aboveground carbon storage due to liana density may be due in
part to greater rates of tree mortality. The probability of mortality of trees infested
by lianas is two to three times greater than for liana-free trees (Ingwell et al. 2010).
Moreover, the susceptibility of trees infested by lianas is higher in shade-tolerant
and slow-growing trees, which in general show higher basal areas and high wood
density (van der Heijden et al. 2008). Thus, lianas may be promoting directional
changes in species composition towards more fast-growing species with low wood
density that store less carbon (van der Heijden et al. 2013).
46 S.M. Durán and G.A. Sánchez-Azofeifa

Currently, censuses from permanent plots in tropical forests have indicated a


growth in tree carbon storage of 0.46 Mg C ha−1 year−1 due to an increment in
recruitment rates (Baker et al. 2004). Nonetheless, it is still unknown whether this
increment is accompanied by changes in community composition, which could cut-
back the carbon storage of tropical forests in the long term (Phillips et al. 2002).
Baker et al. (2004) found greater biomass carbon in lowland forests in the Amazon
from 1983 to 2001 due to greater recruitment rates. During the same period,
decreases in mean wood density of tree species were also registered, indicating a
greater number of fast-growing tree species (van der Heijden et al. 2013). Whether
these changes are influenced by liana density has not been determined yet. A simu-
lation analysis of the effects of different biodiversity scenarios on carbon storage in
a tropical forest, found that liana-induced shifts in species composition towards
fast-growing trees could lead to reductions in the carbon storage capacity by 34 %
(Bunker et al. 2005). This estimate could be quite conservative, since it excludes
smaller lianas (<10 cm diameter), which not only are more abundant than lianas
greater than 10 cm, but also account for more than 50 % liana biomass in mature
forest (Phillips et al. 2005).
An important unanswered question is to determine whether rates of change in
carbon stocks are driven by rates of change in liana density. To answer this ques-
tion, we obtained data on the rates of change in tree carbon storage and liana den-
sity (>10 cm diameter) per year, estimated from permanent plots censused every
2–5 years during 25 years in 26 mature forest plots (Phillips et al. 2002). We found
that changes in aboveground carbon storage per year decreased with annual incre-
ments in liana density per hectare (Fig. 4.1). Although, it is unknown whether

Fig. 4.1 Relationship between changes in liana density and changes in aboveground carbon
(AGC) storage in 26 mature forests in the Amazon region (Data source: Phillips et al. (2002) and
Baker et al. (2004))
4 Liana Effects on Carbon Storage and Uptake in Mature and Secondary Tropical… 47

extrinsic mortality events such as droughts have influenced reductions in carbon


stocks (Laurance et al. 2014), this result clearly suggests that rates of change in
liana density have the potential to diminish rates of change in carbon stocks, espe-
cially if lianas further increase in the future (Schnitzer and Bongers 2011). More
research on the simultaneous effects of climate and lianas on forest dynamics
are needed to disentangle the environmental correlates of carbon stocks in mature
tropical forests.

4.2.2 Aboveground Woody Productivity

It is well established that liana infestation reduces tree growth rates in tropical for-
ests (Clark et al. 1990; Ingwell et al. 2010; van der Heijden and Phillips 2009; van
der Heijden et al. 2013). Competition for below and aboveground resources appears
as important mechanisms explaining these reductions. In humid forests, lianas
deploy leaves on the canopy competing intensely with trees for above and below-
ground resources. Aboveground, high liana loads in the canopy reduce light avail-
ability and incoming solar radiation, which can lower photosynthetic rates and
carbon uptake (Graham et al. 2003). Reductions in light availability can also pro-
vide cover for seed and seedling predators decreasing the reproductive output of tree
species (Schnitzer et al. 2000). Belowground, lianas and trees can compete for
nutrients or soil resources (Chen et al. 2008). In seasonal forests, lianas appear to
have a competitive advantage over trees due to their efficient vascular system, which
allow lianas to tap water during seasonal drought while their tree competitors are
dormant (Schnitzer 2005). The broad overlap of life history strategies of lianas and
trees as seedlings constitute another example of their competitive interactions
(Gilbert et al. 2006).
Despite the cumulative knowledge of liana impacts on tree growth, there is lim-
ited information on how tree growth reductions translate into declines in forest car-
bon. A study in the Peruvian Amazon used data on tree growth rates, local
environmental conditions, and liana competition for aboveground resources to
quantify changes in carbon uptake (van der Heijden and Phillips 2009). The results
indicated that liana-induced reductions of tree growth rates diminish tree carbon
uptake by 0.25 Mg C ha−1 year−1, which correspond to 10 % reduction in tree carbon
increment in this mature forest (van der Heijden and Phillips 2009). Tree growth
rates in this forest averaged 2.70 Mg C ha−1 year−1, while liana biomass growth
was only 0.09 Mg C ha−1 year−1, which represents 3.3 % of total stem production
(van der Heijden et al. 2013). Thus, carbon uptake by lianas is not able to compen-
sate for reductions in tree carbon uptake.
Understanding liana-tree interactions may provide a more accurate assessment
of the effects of lianas on carbon uptake (Ingwell et al. 2010). Turnover rates of both
lianas and trees are increasing in tropical forests (Phillips et al. 2004, 2005).
Permanent plots in the Amazon region showed that tree turnover rates have
48 S.M. Durán and G.A. Sánchez-Azofeifa

augmented on average by 2 % per year in a period of 25 years (Phillips et al. 2004),


with steeper turnover rates for lianas (Phillips et al. 2002). Greater turnover rates of
trees and lianas are presumably the result of greater concentration of CO2 in the
atmosphere (Phillips et al. 2004). Carbon dioxide enrichment may intensify photo-
synthesis and accelerate forest productivity and plant growth (Körner 2004).
Nonetheless, greater forest productivity does not necessarily translate in greater car-
bon sinks in the long term. Accelerated growth intensifies plant competition, which
can lead to rapid tree mortality and recruitment as well as faster tree senescence
(Laurance et al. 2014). Moreover, enrichment of carbon dioxide might fertilize lia-
nas to a greater extent than trees leading to further liana increases (Körner 2004).
Recruitment and mortality rates of lianas appear to be three times greater than those
reported by trees (Phillips et al. 2005). In addition, turnover rates of lianas are posi-
tively associated with high tree turnover rates rather than influenced by changes in
climatic conditions or soil factors (Dalling et al. 2012; Laurance et al. 2014). Since
greater liana infestations may accelerate tree mortality rates, the detrimental effects
of lianas on trees constitute a positive feedback of liana dynamics on stand produc-
tivity (Phillips et al. 2005; Ingwell et al. 2010). Long-term data on liana and tree
dynamics are imperative to provide more accurate calculations of the net losses and
gains of forest carbon.

4.2.3 Primary Productivity: Litterfall Production

Liana contribution to litterfall production is predicted to be high since lianas allo-


cate few resources to a self-supporting system, and rather assign more resources to
leaf productivity in the canopy (Schnitzer and Bongers 2002). Most research in lit-
terfall production concentrate on total litterfall production and discriminating by
leaf and total litterfall, but few attempts exist to estimate contributions of litterfall
by different life-form types (e.g., trees and lianas). Information about the contribu-
tion of lianas to aboveground net primary productivity is quite limited as well, as
most detailed measurements are provided for trees and their components (e.g.,
branch, stem, and leaves). In a lowland forest, liana contribution to aboveground
primary productivity (ANPP) averaged 1.32 Mg C ha−1 year−1, which corresponds to
14.8 % of the total ANPP (van der Heijden et al. 2013). Aboveground primary pro-
ductivity across six tropical forests worldwide showed that leaf litterfall production
of lianas averaged 2.15 Mg C ha−1 (from 0.8 to 3.1 Mg C ha−1), which represents
23.5 % (11–38 %) of total litterfall in mature tropical forests (Hladik 1974;
Burghouts et al. 1994; Wright et al. 2004; Pragasan and Parthasarathy 2005; Chave
et al. 2008; Da hora et al. 2008). Since liana density and biomass are becoming
higher in mature tropical forest, the overall contribution of lianas to ANPP is prob-
ably greater as well. Therefore, total contribution of trees to forest canopy produc-
tivity may be diminished as well, but overall effects in ANPP are still unknown (van
der Heijden et al. 2013).
4 Liana Effects on Carbon Storage and Uptake in Mature and Secondary Tropical… 49

4.3 Liana Effects on Carbon Cycling in Mature Forests

Research on carbon dynamics in secondary forests has traditionally been focused on


evaluating recovery rates of structural characteristics (e.g., basal area, stem density,
plant growth), and carbon pools (above and belowground), and estimating the time
it would take for secondary forests to reach similar values to those found in mature
forests (Hughes et al. 1999; Chazdon et al. 2007). In general, this research has
revealed that secondary forests have the potential to accumulate carbon pools simi-
lar to those in mature forests, with the rate and pattern of this recovery extremely
affected by the severity and duration of previous land uses (Read and Lawrence
2003; Letcher and Chazdon 2009). Still little is known about what other factors
besides land use have the potential to accelerate or slow down recovery rates in
these regenerating forests.
Lianas may turn dominants in disturbed vegetation or following forest fragmen-
tation (Gehring et al. 2004). Secondary stands favored liana abundance by providing
both high light availability and abundant small trees that act as trellises (Schnitzer
and Bongers 2002). In treefall gaps the high dominance of lianas inhibits tree
growth, regeneration, and suppresses the density of shade-tolerant trees by obstruct-
ing light penetration (Schnitzer et al. 2000). Early in the successional recovery of
forest after disturbance in mature and secondary forests, lianas can form dense
stands, often referred as tangles, which can persist for long periods and alter the
pathway of forest recovery to one stalled by liana abundance (Uhl et al. 1988;
Buschbacher et al. 1988; Hegarty 1991; Schnitzer et al. 2000; Paul and Yavitt 2011).
Thus, lianas are able to arrest forest succession, negatively affect the development
of tree species (Schnitzer et al. 2000) and even change the rate of carbon accumula-
tion in regenerating forests (Schnitzer et al. 2014). Several studies have demon-
strated that lianas are more abundant in secondary than mature forests in tropical
regions, but few have examined changes of liana abundance and biomass during
succession (DeWalt et al. 2000; Letcher and Chazdon 2009; Madeira et al. 2009).

4.3.1 Aboveground Carbon Storage

The effects of lianas on carbon storage in secondary forests are still unknown.
Evidence on the recovery of carbon pools provides some insights in the relative
contribution of lianas to forest carbon. The overall contribution of lianas (>10 cm
diameter) to carbon stocks in secondary forests is less than 10 %, while tree contri-
bution varies from 60 % to 94 % (Table 4.1). Comparisons of liana biomass across
stand ages are mixed, with one study displaying significant increases during succes-
sion (Letcher and Chazdon 2009), while others showing no variation in liana bio-
mass with forest age (DeWalt et al. 2000; Feldpaush et al. 2005). Tree carbon
storage (>10 cm diameter) in secondary forests accounted for 60–95 % of total
carbon pools depending of land use history (Table 4.1). Tree carbon storage and
50 S.M. Durán and G.A. Sánchez-Azofeifa

Table 4.1 Contribution of trees and lianas to aboveground carbon storage in secondary tropical
forests. Rate of recovery refers to the time it would take for carbon stocks to reach mature levels
Stand Recovery
age Trees Lianas Previous land use rate
Forest Mg C Mg C
typea (years) ha−1 % ha−1 % (years) Source
Moist 0.5–50 105.9 75–85 1.6 1 Grazing, crops 73 years Hughes et al.
(1999)
Moist 20–70 − – 3.3 – Grazing, crops DeWalt et al.
(2000)
Dry 7–13 12.3 94 1.4 6 Logging, burning Restom and
Nepstad (2001)
Dry 2–25 16.1 92 1.3 5 Logging, crops, 65–120 Read and
development Lawrence
(2003)
Moist 0–14 23.2 90 1.8 2 Grazing Feldpausch
et al. (2005)
Moist 4–22 20.6 92 0.6 3 Grazing, crops, >60 Sierra et al.
mining (2007)
Dry 9–82 34.4 61 1.6 5 Grazing 35 Cifuentes-Jara
Moist 0.4–40 23.4 65 2.5 7 Grazing 80 (2008)
Wet 0.5–60 40.2 71 1.9 3 Grazing 108–124
Dry 9–50 25.5 85 2.1 6 Grazing, crops Madeira et al.
(2009)
a
Forest type follows Chave et al. (2005)

basal area in secondary forests accumulate with age, with older stands showing
greater values and sometimes attaining similar values to mature forests after the first
35 years of regeneration (Cifuentes-Jara 2008).
It is unknown how long lianas can persist with dominance strong enough to
change the regeneration process (Paul and Yavitt 2011). Some have found liana
density to increase up until 20 years after disturbance, and then decline (DeWalt
et al. 2000; Letcher and Chazdon 2009). The decline in liana density appears to be
associated with increases in canopy height, and declines in tree-host availability
during succession (Putz 1984; Letcher and Chazdon 2009). Lianas that fail to reach
the canopy early in succession have lower chances of doing (Letcher and Chazdon
2009). As the canopy closes, light availability is reduced and tree diameter is
increased, thus it becomes difficult for lianas to gain the vertical growth necessary
to compete with other plants (Letcher and Chazdon 2009; Paul and Yavitt 2011).
Consequently, the role of lianas for secondary forests may be more important in
early stages of forest regeneration. Different studies have shown that the relative
contribution of lianas (>10 cm diameter) to total carbon stocks could reach up to
8 % in young stands (<20 years), while in old stands (>40 years), liana contribution
is reduced by almost half to values lower than 4 % (Read and Lawrence 2003;
Feldpaush et al. 2005; Cifuentes-Jara 2008). It is still undetermined whether these
4 Liana Effects on Carbon Storage and Uptake in Mature and Secondary Tropical… 51

changes in liana biomass have an impact on tree dynamics, and if so the potential
consequences for carbon gain.

4.3.2 Aboveground Carbon Accumulation

In general accumulation of tree biomass during succession occurs very rapidly


(Chazdon et al. 2007), while lianas show relatively slow recovery of biomass during
succession due to slower growth rates (Letcher and Chazdon 2009). Information on
the role of lianas in carbon accumulation in secondary forests comes entirely from
small-scale disturbances such as treefall gaps (Dupuy and Chazdon 2006; Schnitzer
and Carson 2010). Dupuy and Chazdon (2006) evaluated the effect of removal veg-
etation in secondary forests on the recruitment, mortality and density of seedlings of
lianas and trees over 2.5 years. They found that recruitment of tree saplings was
positively affected by light availability, but was unrelated to recruitment of liana
saplings, which have lower numbers in advance regeneration. Tree seedlings expe-
rienced high mortality, probably related to greater competition to herbaceous spe-
cies rather than lianas. Similar to what it has been found in chronosequences, found
that density of liana saplings declined in advance regeneration, and had lower densi-
ties than tree saplings (Dupuy and Chazdon 2006).
Another study experimentally demonstrated the effects of lianas on biomass
carbon accumulation in treefall gaps. Schnitzer et al. (2014) quantified rates of
tree growth and mortality during 8 years in treefall gaps with and without lianas.
They found that lianas substantially decreased tree carbon accumulation by 4.2–
8.4 % through reductions mainly in tree growth. Liana growth only contributed
24 % of the tree biomass accumulation they displaced. Reductions of biomass
carbon increment in treefall gaps depended on initial tree biomass though, with
lower declines of biomass accumulation in gaps with low initial tree biomass
(Schnitzer et al. 2014).
Although, canopy gaps provide an essential mechanism for regeneration of lia-
nas and fast-growing trees (Schnitzer and Carson 2010), they are relatively small
and infrequent in secondary tropical forests compared to mature forests (Chazdon
et al. 2007). In addition, gap closure and dynamics may occur at a faster rate than
mature forests, since gap size and canopy height are smaller, and woody growth and
plant density are greater (Paul and Yavitt 2011). Certainly, more comparable studies
in stands of different ages are needed.

4.3.3 Litterfall Production

Lianas devote a large proportion of their energy to leaf production, and have a
higher ratio of leaf mass to basal area (Paul and Yavitt 2011). Moreover, lianas
have shorter leaf life-spans than trees (Hegarty 1991); hence the proportion of
52 S.M. Durán and G.A. Sánchez-Azofeifa

leaf litter of lianas in secondary forests may be higher than in mature forests
(DeWalt et al. 2000). Unfortunately, assessments of litterfall production in sec-
ondary forests are scarce. To our knowledge only two studies have compared
litterfall production of lianas across stands of different ages. Buschbacher et al.
(1988) evaluated litterfall production in abandoned pastures in the Amazon and
found that liana leaf litterfall production varied from 0.1 to 0.5 Mg C ha−1, with
lower values in abandoned pastures that were grazed for less than 5 years, while
greater values found in pastures previously grazed for more than 10 years. More
recent data in a deciduous secondary forests showed that contribution of lianas to
litterfall was greater in intermediate stages of regeneration (25–35 years) com-
pared to younger or older stands. These differences were related to the greatest
abundance of lianas in intermediate stages (Kalácska et al. 2005). Together these
two studies also support previous findings of densities of lianas declining with
the advance of secondary succession.

4.4 Conclusions

This review provides evidence that lianas can have negative effects in carbon
stocks and sequestration in tropical forests. Although some mechanisms by which
lianas could impact carbon pools are well known (e.g., reduction of tree growth
and increases in tree mortality), more research is needed to determine whether the
negative effects of lianas are consistent across tropical areas. Liana density is
positively associated with rainfall seasonality and evapotranspirative demand
(Schnitzer and Bongers, 2011), with greater abundance of lianas in seasonal than
unseasonal forests (Schnitzer 2005). Thus, the role of lianas for carbon sequestra-
tion may be more important in seasonal forests rather than rainforests where most
research are currently being carried out. It is also essential to estimate liana lit-
terfall production and liana growth rates in order to assess whether reductions of
carbon stocks and accumulation due to lianas can be compensated by carbon gain
and uptake through litterfall production and increases in liana biomass. This is
particularly important in secondary forests, where liana contribution to primary
productivity could be even greater to what it has been reported for mature forests,
since lianas are more abundant in secondary forests, and have shorter leaf life
span and higher turnover rates. Priorities for future research in mature forests
include determining whether increases in tree turnover rates and reductions in
wood density are caused by increases in liana density. In secondary forests, exam-
ining long-term changes in trees and lianas are urgently required for a general
understanding of the contribution of this ecosystem to global carbon cycling.
Ultimately, estimates of liana-induced changes need to be incorporated to global
circulation models to predict whether tropical forests in the future will act as car-
bon sinks (e.g. greater carbon stocks over time) or carbon sources (e.g. lower
carbon stocks over time).
4 Liana Effects on Carbon Storage and Uptake in Mature and Secondary Tropical… 53

References

Baker TR, Phillips OL, Malhi Y et al (2004) Increasing biomass in Amazonian forest plots. Philos
Trans R Soc Lond B Biol Sci 359:353–365. doi:10.1098/rstb.2003.1422
Bunker D, DeClerck F, Bradford J (2005) Species loss and aboveground carbon storage in a tropi-
cal forest. Science 310:1029–1031
Burghouts TBA, Campbell EJF, Kolderman PJ (1994) Effects of tree species heterogeneity on leaf
fall in primary and logged dipterocarp forest in the UluSegama Forest Reserve, Sabah,
Malaysia. J Trop Ecol 10:1. doi:10.1017/S0266467400007677
Buschbacher R, Uhl C, Serrao E (1988) Abandoned pastures in eastern Amazonia II. Nutrient
stocks in the soil and vegetation. J Ecol 76:682–699
Chave J, Andalo C, Brown S et al (2005) Tree allometry and improved estimation of carbon stocks
and balance in tropical forests. Oecologia 145:87–99. doi:10.1007/s00442-005-0100-x
Chave J, Olivier J, Bongers F et al (2008) Above-ground biomass and productivity in a rain forest
of eastern South America. J Trop Ecol 24:355–366. doi:10.1017/S0266467408005075
Chazdon RL, Letcher SG, van Breugel M et al (2007) Rates of change in tree communities of
secondary Neotropical forests following major disturbances. Philos Trans R Soc Lond B Biol
Sci 362:273–289. doi:10.1098/rstb.2006.1990
Chen YJ, Bongers F, Cao KF, Cai ZQ (2008) Above-and below-ground competition in high and
low irradiance: tree seedling responses to a competing liana Byttneria grandifolia. J Trop Ecol
24:517–524. doi:http://dx.doi.org/10.1017/S0266467408005233
Cifuentes-Jara M (2008) Aboveground biomass and ecosystem carbon pools in tropical secondary
forests growing in six life zones of Costa Rica. Dissertation, Oregon State University, Corvallis,
OR, USA
Clark DB, Clark DA, Studies T et al (1990) Distribution and effects on tree growth of lianas and
woody hemiepiphytes in a Costa Rican tropical wet forest. J Trop Ecol 6:321–331
Dalling JW, Schnitzer SA, Baldeck C et al (2012) Resource-based habitat associations in a
Neotropical liana community. J Ecol 100:1174–1182. doi:10.1111/j.1365-2745.2012.01989.x
DeWalt S, Chave J (2004) Structure and biomass of four lowland Neotropical forests. Biotropica
36:7–19
Dewalt SJ, Schnitzer SA, Denslow JS (2000) Density and diversity of lianas along a chronose-
quence in a central Panamanian lowland forest. J Trop Ecol 16:1–19. doi:10.1017/
S0266467400001231
Dupuy JM, Chazdon RL (2006) Effects of vegetation cover on seedling and sapling dynamics
in secondary tropical wet forests in Costa Rica. J Trop Ecol 22:65. doi:10.1017/
S0266467405002890
Durán SM, Gianoli E (2013) Carbon stocks in tropical forests decrease with liana density. Biol Lett
9(4):20130301. doi:10.1098/rsbl.2013.0301
Feldpausch TR, Riha SJ, Fernandes ECM, Wandelli EV (2005) Development of forest structure
and leaf area in secondary forests regenerating on abandoned pastures in Central Amazônia.
Earth Interact 9:1–22. doi:10.1175/EI140.1
Gehring C, Park S, Denich M (2004) Liana allometric biomass equations for Amazonian primary
and secondary forest. For Ecol Manage 195:69–83. doi:10.1016/j.foreco.2004.02.054
Gilbert B, Wright SJ, Muller-Landau HC et al (2006) Life history trade-offs in tropical trees and
lianas. Ecology 87:1281–1288. doi:10.1890/0012-9658(2006)87[1281:LHTITT]2.0.CO;2
Graham EA, Mulkey SS, 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 U S A 100:572–576.
doi:10.1073/pnas.0133045100
Hegarty EE (1991) Leaf litter production by lianes and trees in a sub-tropical Australian rain forest.
J Trop Ecol 7:201. doi:10.1017/S0266467400005356
Hladik A (1974) Phenology of leaf production in rain forest of Gabon: distribution and composi-
tion of food for folivores. C R Acad Sci 278:2527–2530
54 S.M. Durán and G.A. Sánchez-Azofeifa

da Hora RC, Primavesi O, Soares JJ (2008) Contribuição das folhas de lianas na produção de
serapilheira em um fragmento de floresta estacional semidecidual em São Carlos, SP. Rev Bras
Botânica 31:277–285. doi:10.1590/S0100-84042008000200010
Hughes RF, Kauffman JB, Jaramillo VJ (1999) Biomass, carbon, and nutrient dynamics of second-
ary forests in a humid tropical region of Mexico. Ecology 80:1892. doi:10.2307/176667
Ingwell LL, Joseph Wright S, Becklund KK et al (2010) The impact of lianas on 10 years of
tree growth and mortality on Barro Colorado Island, Panama. J Ecol 98:879–887.
doi:10.1111/j.1365-2745.2010.01676.x
Kalácska M, Calvo-Alvarado JC, Sánchez-Azofeifa GA (2005) Calibration and assessment of sea-
sonal changes in leaf area index of a tropical dry forest in different stages of succession. Tree
Physiol 25:733–744. doi:10.1093/treephys/25.6.733
Körner C (2004) Through enhanced tree dynamics carbon dioxide enrichment may cause tropical
forests to lose carbon. Philos Trans R Soc Lond B Biol Sci 359:493–498. doi:10.1098/
rstb.2003.1429
Laurance WF (2010) Habitat destruction: death by a thousand cuts. In: Sodhi NS, Ehrlich PR (eds)
Conservation biology for all. Oxford University Press, Oxford, pp 73–87
Laurance WF, Laurance SG, Ferreira LV, Rankin-de Merona JM, Gascon C, Lovejoy TE (1997)
Biomass collapse in Amazonian forest fragments. Science 278:1117–1118. doi:10.1126/
science.278.5340.1117
Laurance W, Andrade A, Magrach A (2014) Long-term changes in liana abundance and forest
dynamics in undisturbed Amazonian forests. Ecology 95:1604–1611
Letcher SG, Chazdon RL (2009) Lianas and self-supporting plants during tropical forest succes-
sion. For Ecol Manag 257:2150–2156. doi:10.1016/j.foreco.2009.02.028
Lewis SL, Lloyd J, Sitch S et al (2009) Changing ecology of tropical forests: evidence and drivers.
Annu Rev Ecol Evol Syst 40:529–549. doi:10.1146/annurev.ecolsys.39.110707.173345
Madeira BG, Espírito-Santo MM, Neto SD et al (2009) Changes in tree and liana communities
along a successional gradient in a tropical dry forest in south-eastern Brazil. Plant Ecol
201:291–304. doi:10.1007/s11258-009-9580-9
Malhi Y (2012) The productivity, metabolism and carbon cycle of tropical forest vegetation. J Ecol
100:65–75. doi:10.1111/j.1365-2745.2011.01916.x
Pan Y, Birdsey RA, Fang J et al (2011) A large and persistent carbon sink in the world’s forests.
Science 333:988–993. doi:10.1126/science.1201609
Paul GS, Yavitt JB (2011) Tropical vine growth and the effects on forest succession: a review of
the ecology and management of tropical climbing plants. Bot Rev 77:11–30. doi:10.1007/
s12229-010-9059-3
Pérez-Salicrup DR, Barker MG (2000) Effect of liana cutting on water potential and growth of
adult Sennamultijuga (Caesalpinioideae) trees in a Bolivian tropical forest. Oecologia 124:469–
475. doi:10.1007/PL00008872
Phillips O, Martínez R, Arroyo L, Baker T (2002) Increasing dominance of large lianas in
Amazonian forests. Nature 418:770–774
Phillips OL, Baker TR, Arroyo L et al (2004) Pattern and process in Amazon tree turnover, 1976–
2001. Philos Trans R Soc Lond B Biol Sci 359:381–407. doi:10.1098/rstb.2003.1438
Phillips OL, Vásquez Martínez R, Monteagudo Mendoza A et al (2005) Large lianas as hyperdy-
namic elements of the tropical forest canopy. Ecology 86:1250–1258. doi:10.1890/04-1446
Pragasan LA, Parthasarathy N (2005) Litter production in tropical dry evergreen forests of south
India in relation to season, plant life-forms and physiognomic groups. Curr Sci 88:1255–1263
Putz FE (1984) The natural history of Lianas on Barro Colorado Island, Panama. Ecology 65:1713.
doi:10.2307/1937767
Read L, Lawrence D (2003) Recovery of biomass following shifting cultivation in dry tropical
forests of the Yucatan. Ecol Appl 13:85–97
Restom TG, Nepstad DC (2001) Contribution of vines to the evapotranspiration of a secondary
forest in eastern Amazonia. Plant and Soil 236:153–163
Schnitzer SA (2005) A mechanistic explanation for global patterns of liana abundance and distri-
bution. Am Nat 166:262–276. doi:10.1086/431250
4 Liana Effects on Carbon Storage and Uptake in Mature and Secondary Tropical… 55

Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
Schnitzer SA, Bongers F (2011) Increasing liana abundance and biomass in tropical forests: emerging
patterns and putative mechanisms. Ecol Lett 14:397–406. doi:10.1111/j.1461-0248.2011.01590.x
Schnitzer SA, Carson WP (2010) Lianas suppress tree regeneration and diversity in treefall gaps.
Ecol Lett 13:849–857. doi:10.1111/j.1461-0248.2010.01480.x
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical
forest canopy gaps: evidence for an alternative pathway of gap-phase regeneration. J Ecol
88:655–666. doi:10.1046/j.1365-2745.2000.00489.x
Schnitzer SA, van der Heijden GMF, Mascaro J, Carson WP (2014) Lianas in gaps reduce carbon
accumulation in a tropical forest. Ecol 140515081851004. doi:10.1890/13-1718.1
Sierra CA, del Valle JI, Orrego SA et al (2007) Total carbon stocks in a tropical forest landscape of
the Porce region, Colombia. For Ecol Manage 243:299–309. doi:10.1016/j.foreco.2007.03.026
Stegen JC, Swenson NG, Enquist BJ, White EP, Phillips OL, Jørgensen PM, Weiser MD,
Monteagudo-Mendoza A, Núñez-Vargas P (2011) Variation in above-ground forest biomass
across broad climatic gradients. Glob Ecol Biogeogr 20:744–754.
doi:10.1111/j.1466-8238.2010.00645.x
Uhl C, Buschbacher R, Serrao E (1988) Abandoned pastures in eastern Amazonia. I. Patterns of
plant succession. J Ecol 76:663–681
van der Heijden GMF, Phillips OL (2009) Liana infestation impacts tree growth in a lowland tropi-
cal moist forest. Biogeosciences 6:2217–2226. doi:10.5194/bg-6-2217-2009
van der Heijden GMF, Healey JR, Phillips OL (2008) Infestation of trees by lianas in a tropical
forest in Amazonian Peru. J Veg Sci 19:747–756. doi:10.3170/2008-8-18459
van der Heijden GMF, Schnitzer SA, Powers JS, Phillips OL (2013) Liana impacts on carbon
cycling, storage and sequestration in tropical forests. Biotropica 45:682–692. doi:10.1111/
btp.12060
Wright SJ (2013) The carbon sink in intact tropical forests. Glob Chang Biol 19:337–339.
doi:10.1111/gcb.12052
Wright SJ, Calderón O, Hernandéz A, Paton S (2004) Are lianas increasing in importance in tropi-
cal forests? A 17-year record from panama. Ecology 85:484–489. doi:10.1890/02-0757
Wright SJ, Jaramillo MA, Pavon J, Condit R, Hubbell SP, Foster RB (2005) Reproductive size
thresholds in tropical trees: variation among individuals, species and forests. J Trop Ecol
21:307–315. doi:http://dx.doi.org/10.1017/S0266467405002294
Chapter 5
Diversity and Distribution of Climbing Plants
in Eurasia and North Africa

Liang Hu and Mingguang Li

Abstract A total of 12,382 climbers from 143 families and 1,415 genera are
recorded in the Old World, including 57 families with climbers only in the Old
World (e.g. Actinidiaceae). As for Eurasia and North Africa, 6,659 climbers were
documented to be native, belonging to 101 families and 809 genera. About 30.8 %
of them are herbaceous and 69.2 % are somewhat woody. Only one family, but 285
genera (35.2 %) and 5,283 species (79.3 %) are climber-endemic to Eurasia and
North Africa. Leguminosae and Apocynaceae are the two largest families with
climbers, followed by Convolvulaceae, Vitaceae, Cucurbitaceae and Rubiaceae.
The most climber-abundant families in Eurasia and North Africa are similar to Sub-
Saharan Africa, except Ranunculaceae, Rosaceae and Arecaceae. The Bignoniaceae,
Araceae and Compositae are less prevalent in Eurasia and North Africa compared
with the Americas, while Ranunculaceae, Rosaceae, Arecaceae and Annonaceae
stand out. The study area was divided into 19 regions and 211 districts. Diversity
and geographical distribution of climbing plants in these regions and districts are
discussed. South-east Asia, South China, South Asia and the Himalayas are the top
four climber-abundant regions and each have more than 1,000 climbers and signifi-
cantly higher than the rest regions. The proportion of climbing plants in the sper-
matophyte flora declined from 20 % in Malay Peninsula to less than 2 % in Sahara,
Asian plateaus and the northern Eurasia.

Keywords Climbers • Lianas • Vines • Europe • Asia • Africa • West Malay


Archipelago

L. Hu (*)
School of Geography and Planning, Sun Yat-sen University,
Guangzhou 510275, China
e-mail: huliang_hy@163.com
M. Li
State Key Laboratory of Biocontrol, Sun Yat-sen University,
Guangzhou 510275, China
e-mail: lsslmg@mail.sysu.edu.cn

© Springer International Publishing Switzerland 2015 57


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_5
58 L. Hu and M. Li

5.1 Introduction

The importance of climbers in forest ecosystems has been increasingly emphasized


with our more in-depth understanding on them. The climbing device have been well
developed in climbers, as successful strategies to compete with self-supporting
plants for light, space and the chance of survival as found in both ferns and sper-
matophytes, particularly in angiosperms (Gentry 1991). This special group of plants
has attracted the interest of many researchers since Darwin’s time. However, our
knowledge on climbers is still lagging behind self-supporting plants. Even now,
there are too many puzzles left on their function, ecology, evolution and even the
most basic question: how many kinds of them and where they are distributed?
At present, the large-scale researches of liana diversity and geographical distri-
bution are mainly from the Americas and Africa. Gentry (1991) described the
climber composition of the Americas. According to him, 9,216 climbing plants
belonging to 97 families exist in the Americas and at least 133 families include at
least a few climbers in the world. Jongkind and Hawthorne (2005) has attached a
checklist of 746 climbers found in west tropical Africa.
There are many reports on liana composition at local (eg. Yan and Qi 2007;
Muthumperumal and Parthasarathy 2009) or regional (eg. Yan et al. 2006; Hu 2011)
scales in Asia, although research in large-scale is scarce. Recently, Yan (2009) listed
784 climbers in Central China. A total of 3,073 spermatophyte species are listed as
climbers in China, accounting for 11.3 % of the flora of China (Hu et al. 2010a).
Obviously, these studies are limited to the administrative regions and have limited
significance for the understanding of the geographical distribution of climbers. In
this study, climbers of Eurasia and North Africa and the adjacent areas are all sur-
veyed, including most part of Indomalayan Realm and the whole Palaearctic and
Afrotropical Realms.

5.2 Methods

5.2.1 Study Areas

The focus study area is Eurasia (excluding West Malay Archipelago) and North
Africa (including Sahara region). We divide this area into 19 regions and 211 dis-
tricts (Fig. 5.1). The ranges of these regions and the number of districts included (in
brackets) are as follows:
1. SE. Asia (12 districts): including Andaman & Nicobar, Cambodia, Laos, Malay
Peninsula, Myanmar, Thailand, Vietnam and the southernmost of China
(S. Yunnan, S. Guangxi, S. Guangdong, Hainan, Xisha Islands and Dongsha
Islands).
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 59

Fig. 5.1 Regions and districts of Eurasia and North Africa

2. S. China (20 districts): from the east side of the Tibetan Plateau to Taiwan and
the Ryukyu Islands.
3. S. Asia (15 districts): including the South Asian subcontinent except the
Himalayas.
4. Himalaya (6 districts): including Kashmir, NW. India, Nepal, Bhutan, Sikkim,
Darjiling, and the southern slope of the Himalayas of India and China.
5. Tibetan Plateau (12 districts): most part of Qinghai and Tibet Provinces of
China.
6. Iranian Plateau (5 districts): including Pakistan, Iran, Afghanistan and the
S. Caucasus.
7. W. Asia (10 districts): Yemen (excluding Socotra), Oman, Saudi Arabia, UAE,
Qatar, Kuwait, Iraq, Jordan and Sinai Peninsula.
8. Mediterranean (19 districts): consists of Iberian Peninsula, Italian Peninsula,
Balkan Peninsula, French Mediterranean, Israel, Cyprus and North African
countries in the Mediterranean region. Turkey and Syria are not sub-divided
and they are classified into this region.
9. Sahara (10 districts): including Western Sahara, the northern part of Mauritania,
Mali, Niger, Chad and Sudan, and the southern part of Algeria, Libya and
Egypt.
10. E. Asia (20 districts): including Northeast China, North China, Korean
Peninsula, Japanese Islands, Kuril Islands, Sakhalin, Amur, Primorsky and
South Khabardvskiy.
60 L. Hu and M. Li

11. Mongolian Plateau (19 districts): including Mongolia and Inner Mongolia
Province, China.
12. Tianshan-Pamir-Kunlun (11 districts): Tajikistan, Kyrgyzstan, Xinjiang and
NE. Gansu provinces of China.
13. C. Asia (3 districts): Turkmenistan, Uzbekistan and Kazakhstan.
14. S. Siberia (9 districts): Nine districts in the south part of Siberia are
included in.
15. E. Europe (12 districts): Belarus, Bulgaria, Czech, Hungary, Poland, Romania,
Slovakia, Ukraine, Estonia, Latvia, Lithuania, south and central part of
European Russia.
16. W. Europe (7 districts): Austria, Belgium, Germany, Swiss, France, Great
Britain and Ireland.
17. Russian Far East (4 districts): Chukotka, Magadan, Kamchatka and North
Khabardvskiy.
18. N. Siberia (11 districts): Eleven districts in the north part of Siberia are
included.
19. N. Europe (6 districts): Norway, Sweden, Finland, Denmark and the north part
of European Russia.
Ranges of other areas or regions mentioned in this study are as follows:
1. Americas (or the New World): including North America and South America.
2. East Africa: including Burundi, Djibouti, Eritrea, Ethiopia, Kenya, Rwanda,
Socotra, Somalia, Tanzania, Uganda and sub-Saharan Sudan.
3. Oceania: including East Malay Archipelago, Australia and the Polynesian
Triangle.
4. Old World: including Eurasia & North Africa, Sub-Saharan Africa and West
Malay Archipelago.
5. Sub-Saharan Africa: including the continent of Africa that lies south of the
Sahara Desert, Madagascar, Mauritius and Seychelles.
6. West Africa: including Benin, Burkina Faso, Gambia, Ghana, Guinea, Guinea-
Bissau, Ivory Coast, Liberia, Nigeria, Senegal, Sierra Leone, Togo and sub-
Saharan Chad, Mali, Mauritania and Niger.

5.2.2 Data Collection

The definitions of terms about climbing plants are chaotic. This problem has also
been discussed by Putz and Mooney (1991) in the preface of the book itself. The
words “lianas”, “lianes” and “vines” are used for climbing plants (Parsons 2005),
while some others define lianas as woody climbers (e.g. Schnitzer and Bongers
2002) and vines as herbaceous ones (e.g. Hu et al. 2010b). In this study, we adopted
the division of climbers into lianas (woody climbers, including half-woody
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 61

climbers and scandent shrubs) and vines (herbaceous climbers) based on stem
characteristics described in literature. The scope of climbing plants, followed
Darwin (1865) and Gentry (1991), and includes twiners (e.g. Ipomoea), tendril-
climbers (e.g. Vitis), hook-climbers (e.g. Rubus), adhesive-climbers (e.g. Pothos)
and scandent or scrambling shrubs (e.g. Zizyphus). Creepers and herbaceous epi-
phytes are usually excluded. Parasitic climbers (e.g. Cassytha) and epiphytes
climbing upward with adhesive roots and elongated stems (e.g. Hoya) are included.
Lentibulariaceae (some species of this family have twining inflorescences) is also
excluded in this paper.
Data have been collected from vast literature including standard floras, check-
lists, keys, monographs, and a large number of botanical journals. Only wild native
spermatophyte species have been taken into account and analyzed, while natural-
ized, introduced and cultivated plants have been excluded. As of paper into the
draft, at least 450 kinds/series, over 1,800 volumes of references have been used as
data sources.
Exploring climbing plants in such a vast geographic range is a tough task,
although the criteria were carried out in advance, we still encounter various troubles
when we decide whether a species is a liana or vine, or not even a climber. The big-
gest reason being lack of clear definition of climbers for what we call as well as the
difference between woody and herbaceous climbers. Therefore we also depend
partly on our field information. For example, Periploca sepium is recorded as shrub
in Flora of China, but trailing or arching plants in some earlier data. However, our
field survey finds that the proper description should be “arching shrub, rarely twin-
ing” (Fig. 5.2).

Fig. 5.2 Periploca sepium with twining stem (Photo by Liang HU)
62 L. Hu and M. Li

5.2.3 Data Analysis

In each district, all records are supported by at least one reference as proof of status
(endemic, native or alien) and habit (liana or vine). All species names have been
checked in The Plant List website (The Plant List 2013) and synonyms have been
amended to accepted name. Genera and species of flowering plants are presented in
families following the updated classification of Angiosperm Phylogeny Group
(APG 2009). In this study, we coined the compound word “climber-endemic” to
indicate a taxon (family/genus/species) which contains climbing plant life-form
only in specific region or district. For example, there are eight species in Cladrastis
(Leguminosae) and they are distributed in Asia and North America, only Cladrastis
scandens endemic to S. China is a liana while the rest are non-climbing trees or
shrubs. Thus we describe Cladrastis as a genus climber-endemic to S. Asia.
The division of regions and districts inevitably affected by administrative bound-
aries. Flora size also makes comparison difficult between adjacent regions. Thus,
the proportion of climbing plants in the spermatophyte flora (CPF) has been calcu-
lated following Hu et al. (2010b). A higher CPF indicates higher relative climber
diversity while ignoring the influence of flora size.
We also comparatively analyzed the difference between the Old World and the
Americas based on checklist of families with climbers described by Gentry (1991).
Necessary revisions have been made before comparing. For example, climber diver-
sity in Asclepiadaceae is submerged with Apocynaceae, while Tiliaceae is merged
into Malvaceae and thus it is no longer climber-endemic to the Old World.

5.3 Diversity of Climbers

A total of 169 seed plant families contain at least one climber in the world (or 170
if Lentibulariaceae is also included), higher than Gentry’s suggested 133 families.
The increase of our estimation attributed to both supplements of new data and fam-
ily rearrangement.
Thirty-six of the 169 families are climber-endemic to the Old World (Table 5.1).
Twenty-one families have climbers in both the Old World and Oceania but not in the
Americas (e.g. Flagellariaceae, Nepenthaceae, Rutaceae, Sabiaceae and
Santalaceae). Three families (Berberidopsidaceae, Onagraceae and Violaceae) have
climbers in both Oceania and the Americas but not in the Old World.
Climbing plant abundance in Oceania is not so rich as other parts of the world.
However, the proportion of endemic climbers in these areas are very high. Ninety
(85.7 %) of the 105 climbers in Hawaii, 30 (54.5 %) of the 55 climbers in New
Zealand and 71 (47.0 %) of the 151 climbers in Fiji are endemic. According to our
uncompleted list of Oceanian climbers, the total number of this area will be over
3,000 and at least two-thirds of the climbers are endemic considering characteristics
of their floras. Accordingly, we estimate that the climbing seed plants of the world
may exceed 25,000 species.
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 63

Table 5.1 Families climber-endemic to the Old World, the Americas or Oceania
The Old World (36) The Americas (18) Oceania (5)
Achariaceae Lophopyxidaceae Alstroemeriaceae Aphanopetalaceae
Actinidiaceae Meliaceae Bromeliaceae Austrobaileyaceae
Ancistrocladaceae Montiniaceae Cactaceae Petermanniaceae
Aquifoliaceae Ochnaceae Calceolariaceae Rhipogonaceae
Balsaminaceae Pandaceae Calophyllaceae Xanthorrhoeaceae
Barbeuiaceae Papaveraceae Caricaceae
Brassicaceae Paulowniaceae Cleomaceae
Chrysobalanaceae Pedaliaceae Clusiaceae
Crassulaceae Pentadiplandraceae Cyclanthaceae
Dioncophyllaceae Peraceae Loasaceae
Ebenaceae Peridiscaceae Marcgraviaceae
Ephedraceae Portulacaceae Polemoniaceae
Gelsemiaceae Resedaceae Schlegeliaceae
Gerrardinaceae Salicaceae Siparunaceae
Hamamelidaceae Salvadoraceae Staphyleaceae
Hypericaceae Sapotaceae Symplocaceae
Iteaceae Stilbaceae Trigoniaceae
Lecythidaceae Talinaceae Tropaeolaceae

Table 5.2 Climbing plant diversity in the Old World


Proportion (%)
Area Family Genera Species Vine Liana
Eurasia & N. Africa 101 809 6,659 30.8 69.2
West Malay Archipelago 96 500 2,343 16.1 83.9
Sub-Saharan Africa 122 825 4,843 30.1 69.9
Total 143 1,415 12,382 27.9 72.1

5.3.1 Old World Climbers

The Old World has abundant climbing plants. A total of 12,382 species from 143
families and 1,415 genera are recorded as climbers (Table 5.2). Only 38 species
(0.3 %) from two families (Gnetaceae and Ephedraceae) are gymnosperms and
the rest are angiosperms. About 27.9 % species are herbaceous vines and the rest
are lianas.
Leguminosae (1,517 species) and Apocynaceae (including Asclepiadaceae,
1,355 species) are the two largest families with climbers in the Old World, followed
by Vitaceae (621), Convolvulaceae (593), Rubiaceae (589) and Cucurbitaceae
(552), four other families with more than 500 climber species (Appendix) each. The
top six families account for 42 % (5,227 species) of the total climbers in the Old
World. Thirty-six families have more than 100 climber species each and account for
84.5 % of the total climber species of in the Old World. Twenty-four families have
64 L. Hu and M. Li

only one climber species, nine families have only two climbers, and additional
twenty six families have less than ten climber species each.
However, our figure on climbers of the Old World may still be less than the actual
value, as data in some districts and taxa are not all satisfactory. For example, the
number of climbers in Cambodia and Sumatra (327 and 427 climbers have been
documented for now, respectively) are far underestimated and many records are
expected to be report in future, which may include a lot of endemic species. In addi-
tion, some species of Convolvulaceae and Apocynaceae of Africa (especially
Central Africa) are difficult to classify as a climber or not.

5.3.2 Eurasian and North African Climbers

The corresponding figures for climbers of Eurasia and North Africa are 6,659 spe-
cies, 809 genera and 101 families. A total of 285 genera (35.2 %) and 5,283 species
(79.3 %) are climber-endemic to this area. Leguminosae and Apocynaceae are still
the two largest families with climbers in Eurasia and North Africa, followed by
Convolvulaceae, Vitaceae, Cucurbitaceae and Rubiaceae. A total of 16 families with
more than 100 climber species each account for 67.6 % of the total climber flora of
this area. The top 15 largest families are listed in Table 5.3 and the last one remain-
ing is Smilacaceae with 107 tendrillar climbers.
The most climber-abundant families in Eurasia and North Africa are similar to Sub-
Saharan Africa. Seven of the top ten climber-abundant families are the same, except
for Ranunculaceae, Rosaceae and Arecaceae which have only a few climbers in

Table 5.3 Top 15 climber-abundant families of Eurasia & N. Africa, Sub-Saharan Africa and the
Americas (Data of the Americas cited and recalculated from Gentry 1991)
Eurasia & N. Africa Sub-Saharan Africa The Americas
Family Species Family Species Family Species
Leguminosae 915 Leguminosae 570 Apocynaceae 1,350
Apocynaceae 740 Apocynaceae 537 Convolvulaceae 750
Convolvulaceae 395 Vitaceae 287 Leguminosae 720
Vitaceae 327 Rubiaceae 285 Compositae 470
Cucurbitaceae 275 Cucurbitaceae 264 Araceae 400
Rubiaceae 261 Convolvulaceae 217 Bignoniaceae 400
Ranunculaceae 201 Annonaceae 208 Sapindaceae 400
Rosaceae 199 Celastraceae 174 Malpighiaceae 400
Annonaceae 194 Combretaceae 141 Passifloraceae 360
Arecaceae 185 Compositae 133 Cucurbitaceae 311
Celastraceae 174 Euphorbiaceae 119 Ericaceae 300
Menispermaceae 153 Menispermaceae 118 Rubiaceae 220
Piperaceae 151 Malvaceae 117 Dioscoreaceae 200
Dioscoreaceae 117 Passifloraceae 99 Aristolochiaceae 180
Lamiaceae 109 Malpighiaceae 90 Euphorbiaceae 170
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 65

Table 5.4 Climbing plant diversity (genera/species) of some families in different geographical
areas (Data of the Americas cited and recalculated from Gentry 1991)
Eurasia & West Malay Sub-Saharan The
Family N. Africa Archipelago Africa Americas
Ranunculaceae 6/201 2/13 2/25 1/10
Rosaceae 6/199 3/31 3/27 2/10
Annonaceae 24/194 13/60 24/208 2/6
Arecaceae 7/185 8/185 4/13 2/65
Araceae 12/91 8/50 5/12 7/~400
Compositae 22/71 14/22 32/133 23/470
Actinidiaceae 3/64 2/4 – –
Nepenthaceae 1/18 1/88 1/3 –
Bignoniaceae 5/8 3/4 6/6 53/~400
Crassulariaceae 1/2 – – –

Sub-Saharan Africa (Table 5.3). However, the number of climber-endemic families in


Eurasia and North Africa is less than Sub-Saharan Africa. Crassulariaceae is the only
family climber-endemic to Eurasia and North Africa, while 30 families have climbers
only in Sub-Saharan Africa. Eurasia and North Africa is also the distribution center of
Actinidiaceae, only four lianas of this family are spread to West Malay Archipelago.
The three largest families with climbers are the same in Eurasia and North Africa
and the Americas, however, six of the top ten families are different (Table 5.3). In addi-
tion to these families with climber absent from each other (Table 5.2), Bignoniaceae,
Araceae and Compositae are less prevalent, while Ranunculaceae, Rosaceae, Arecaceae
and Annonaceae stand out in Eurasia and North Africa (Table 5.4).

5.4 Distribution of Climbers in Eurasia and North Africa

In general, SE. Asia, S. China, S. Asia and Himalaya are the top four climber-
abundant regions, each has more than 1,000 climbers and the proportion of lianas is
significantly higher than herbaceous vines. Lianas are much lacking in plateaus,
deserts, and temperate regions of Eurasia and North Africa but herbaceous vines
occur there (Table 5.5).

5.4.1 Distribution of Climbers in Regions


5.4.1.1 SE. Asia

Dominant taxa: Leguminosae (535), Apocynaceae (481), Vitaceae (221),


Convolvulaceae (220) and Rubiaceae (178) are the five largest families with
climbers, and account for 37.8 % of the total climbers. Calamus (109), Piper
(104), Dioscorea (87), Tetrastigma (84) and Bauhinia (80) are the five largest
66 L. Hu and M. Li

Table 5.5 Climbing plant composition of 19 regions in Eurasia and North Africa and in adjacent
areas
Proportion (%)
Region Family Genera Species Vines Lianas
Eurasia & N. Africa
SE. Asia 97 649 4,329 22.6 77.4
S. China 83 373 2,288 32.2 67.8
S. Asia 88 473 1,775 28.7 71.3
Himalaya 77 333 1,090 33.1 66.9
E. Asia 41 113 369 53.1 46.9
Iranian Plateau 36 109 328 66.1 33.9
W. Asia 37 99 306 71.2 28.8
Mediterranean 25 53 296 84.1 15.9
Tibetan Plateau 32 63 161 55.3 44.7
Sahara 29 66 136 72.0 28.0
E. Europe 18 29 124 79.8 20.2
W. Europe 17 27 106 81.9 18.1
Tianshan-Pamir-Kunlun 15 23 103 85.4 14.6
Mongolian Plateau 21 35 89 77.5 22.5
C. Asia 12 23 88 90.9 9.1
S. Siberia 9 16 54 100.0 0.0
N. Europe 12 16 39 92.1 7.9
N. Siberia 5 7 20 100.0 0.0
Russian Far East 5 8 15 93.3 6.7
Adjacent areas
W. Malay Archipelago 96 500 2,343 16.1 83.9
W. Africa 86 431 1,514 26.0 74.0
E. Africa 95 478 1,692 39.6 60.4
Madagascar Islands 71 256 1,025 22.7 77.3
Canada & Alaska 19 30 66 66.7 33.3

genera with climbers. This region is climber-diverse center for many families
and genera. Besides the families and genera mentioned above, the Annonaceae
(174) and Jasminum (71) are prominent.
Endemic taxa: 81 genera (12.5 %) and 1,758 species (40.6 %) are climber-endemic
to this region. However, we are not sure of these figures because our data in
districts of SE. Asia and W. Malay Archipelago are not so satisfactory as districts
of other regions, especially in Cambodia and Sumatra.
CPFs: All CPFs of districts in SE. Asia are higher than 15 %, and decline from Malay
Peninsula to Myanmar (Table 5.6, Fig. 5.3). Overall, CPFs of SE. Asia are lower
than those of W. Tropical Africa and higher than those of S. Asia and E. Tropical
Africa (Table 5.6). Interestingly, tropical forests in SE. Asia harbored more climb-
ers, but have lower CPFs compared with tropical West Africa (Table 5.6). Lianas
are absolutely dominant in SE. Asia and account for 77.4 % of the total.
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 67

Table 5.6 Climbing plants diversity and their proportion in the spermatophyte flora (CPF) in
tropical districts of Asia and Africa
Region/district Family Genus Species CPF (%)
SE. Asia
Malay Peninsula 81 369 1,420 20.0
Thailand 88 405 1,517 18.7
Vietnam 88 412 1,563 17.3
Andaman and Nicobar 58 167 319 16.3
S. Asia
Tamil Nadu 67 243 621 15.2
Srilanka 66 228 454 15.1
Kerela 61 225 568 14.4
W. Tropical Africa
S. Ivory Coast 70 294 804 23.5
Liberia 67 277 771 23.2
Senegal 58 207 439 22.1
Sierra Leone 63 209 446 21.6
E. Tropical Africa
Uganda 74 309 752 16.7
Kenya 81 347 948 15.0
Ethiopia and Eritrea 66 268 671 12.2
Somalia 48 159 340 11.0

Fig. 5.3 Climbing plant proportion in the spermatophyte flora (CPF) of districts in SE. Asia,
S. Asia, S. China, Himalaya and Tibetan Plateau
68 L. Hu and M. Li

Table 5.7 Top 10 districts with high proportion of endemic climbing plants
Endemic
Districts Family Genera Species Climber %
Malay Peninsula 81 369 1,420 411 28.9
Thailand 88 405 1,517 257 16.9
Sri Lanka 66 228 454 75 16.5
Vietnam 88 412 1,563 231 14.8
Taiwan 61 180 410 56 13.7
Andaman and Nicobar 58 167 319 42 13.2
Myanmar 87 384 1,200 148 12.3
Hainan 74 260 651 70 10.8
Israel 23 42 134 14 10.4
Iberia 14 21 92 9 9.8

Notes: Tropical forests of SE. Asia are rich in climbing plants. It is the most climber-
abundant region in Eurasia and North Africa. It also has the highest endemic
climber proportion. About 34.3 % climbers in SE. Asia are shared with S. China,
23.9 % shared with W. Malay Archipelago, 21.6 % shared with S. Asia and
16.8 % shared with the Himalayas. Climber-endemic proportion in islands seems
to be higher than neighboring districts in the continent. Four of the top 10 dis-
tricts with high proportion of endemic climbing plants are islands (Table 5.7).

5.4.1.2 S. China

Dominant taxa: Leguminosae (272), Apocynaceae (241), Cucurbitaceae (134),


Ranunculaceae (123) and Vitaceae (122) are the five largest families with climb-
ers, and account for 39.0 % of the total climbers distributed in this region.
Clematis (103), Rubus (96), Smilax (60), Actinidia (57) and Dioscorea (51) are
the five largest genera with climbers.
Endemic taxa: Ten genera are climber-endemic to this region: Monimopetalum (1
climber/1 species in genus), Heteroplexis (1/3), Sinobaijiania (3/3), Decumaria
(1/1), Deutzia (1/60), Archakebia (1/1), Cladrastis (1/8), Bredia (1/15),
Veronicastrum (1/20), Oreocnide (1/18). However, according to Flora of China,
one species of the Sinobaijiania is also distributed in N. Laos and the only liana
of Oreocnide is also distributed in N. Vietnam. In addition, 583 species (25.5 %)
are climber-endemic to this region.
CPFs: CPF ranges from 16.6 % to 5.6 % (Fig. 5.3). Lianas are still dominant in
S. China and account for 67.8 %.
Notes: Evergreen broadleaf forests are well developed in S. China and climbing
plants in this region are also very rich. For example, climber diversity of
Rosaceae, Ranunculaceae, Actinidiaceae, Lardizabalaceae, Schisandraceae and
Campanulaceae are much higher than the adjacent regions, while Smilax,
Dioscorea, Aristolochia are second only to SE. Asia.
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 69

5.4.1.3 S. Asia

Dominant taxa: Leguminosae (279), Apocynaceae (203), Convolvulaceae (144),


Cucurbitaceae (82) and Vitaceae (79) are the five largest families with climbers,
and account for 44.3 % of the total climbers distributed in this region. Ipomoea
(45), Piper (45), Argyreia (44), Jasminum (36) and Calamus (35) are the five
largest genera with climbers.
Endemic taxa: Although the S. Asia forests harbored fewer climbers than S. China,
it has higher climber-endemic proportions. Overall, 47 genera (9.9 %) and 614
species (34.6 %) are climber-endemic to this region. However, only Baeolepis
(1/1) and Decalepis (1/1) are truly endemic genera. The others are climber-
endemic and mostly contain some scandent shrubs.
CPFs: CPFs gradually decrease northward and westward in this region, and range from
15.2 % to 9.3 % (Fig. 5.3). Lianas are dominant in S. Asia and account for 71.3 %.
Notes: About 52.7 % climbers in S. Asia are shared with SE. Asia, 39.4 % shared
with Himalaya and 28.8 % shared with S. China. Climber-endemic proportion of
Sri Lanka is up to 16.5 %, only next to Malay Peninsula and Thailand (Table 5.7).

5.4.1.4 Himalaya

Dominant taxa: Leguminosae (163), Apocynaceae (113), Convolvulaceae (68),


Cucurbitaceae (61) and Vitaceae (49) are the five largest families with climbers
and account for 41.3 % of the total climbers distributed in this region. Clematis
(41), Rubus (34), Piper (27), Dioscorea (27) and Jasminum (20) are the five larg-
est genera with climbers.
Endemic taxa: Thirteen genera (3.9 %) are climber-endemic to Himalayan moun-
tains. However, only Edgaria (1/1) and Indofevillea (1/1) are truly endemic gen-
era. The others are climber-endemic and mostly have several scandent shrubs. In
addition, 147 species (13.5 %) are climber-endemic to this region.
CPFs: CPF declined from 13.3 % in SE. Himalaya to only 3.6 % in Kashmir
(Fig. 5.3). Lianas account for 66.9 % of the total climbers.
Notes: Climbing plant composition of Himalaya is strongly affected by adjacent
regions. The proportion of climbers shared with these regions are 66.1 % for
SE. Asia, 63.7 % for S. Asia, 49.5 % for S. China and 14.8 % for Iranian plateau.
Only 9.5 % of climbers in Himalaya are shared with Tibetan Plateau.

5.4.1.5 Tibetan Plateau

Dominant taxa: Ranunculaceae (27), Leguminosae (26), Apocynaceae (14),


Rubiaceae (11) and Vitaceae (9) are the five largest families with climbers and
account for 54.0 % of the total climbers distributed in this region. Clematis (24),
Vicia (13), Cynanchum (7), Galium (6), Cuscuta (5), Dioscorea (5) and Jasminum
(5) are the seven largest genera with climbers.
70 L. Hu and M. Li

Endemic taxa: Only four species are climber-endemic to Tibetan Plateau (Aconitum
longilobum, Clematis zandaensis, Crawfurdia nyingchiensis, Euonymus
tibeticus).
CPFs: CPF is up to 4.8 % in southeastern district (Nyingchi) which is adjacent to
Himalayas, while the hinterlands lacks climbers (Fig. 5.3). Only one liana
(Clematis tangutica) has been recorded in W. Nagqu district recently (Wu 2008)
and no climbers have been found in W. Xigaze and East Ali districts. Vines are
dominant in Tibetan Plateau and account for 55.3 %.
Notes: Climbers are lacking in Asian plateaus. Tibetan Plateau climbing plant flora
is most affected by S. China, 96.8 % genera and 88.8 % species in this region are
shared with S. China.

5.4.1.6 Iranian Plateau

Dominant taxa: Leguminosae (106), Convolvulaceae (44), Apocynaceae (30),


Cucurbitaceae (26) and Ranunculaceae (15) are the five largest families with
climbers and account for 67.4 % of the total climbers distributed in this region.
Vicia (41), Cuscuta (27), Lathyrus (23), Clematis (15) and Galium (10) are the
five largest genera with climbers.
Endemic taxa: Only 28 species (8.5 %, e.g. Hedera caucasigena, Dioscorea cauca-
sica) are climber-endemic to this region.
CPFs: CPF ranges from 4.5 % in Pakistan to 2.2 % in S. Caucasus (Fig. 5.4). Vines
account for 66.2 % of the total climbers in this region.

Fig. 5.4 Climbing plant proportion in the spermatophyte flora (CPF) of districts in Iranian Plateau,
C. Asia, W. Asia, Mediterranean, Sahara, W. Africa and E. Africa
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 71

Notes: This region is the second largest diversity center of Cuscuta (27) in Eurasia
and North Africa.

5.4.1.7 W. Asia

Dominant taxa: Leguminosae (92), Convolvulaceae (47), Apocynaceae (32),


Cucurbitaceae (24) and Rubiaceae (12) are the five largest families with climbers
and account for 67.2 % of the total climbers distributed in this region. Vicia (32),
Lathyrus (21), Cuscuta (16), Ipomoea (14) and Convolvulus (14) are the five
largest genera with climbers.
Endemic taxa: Only 25 species (8.1 %, e.g. Boerhaavia arabicus, Ceropegia foli-
osa) are climber-endemic to this region. The genus Tylophoropsis including two
species (vines), one endemic to W. Asia (T. heterophylla) and another distributed
in E. Africa.
CPFs: CPF ranges from 6.5 % in Yemen to 1.9 % in UAE (Fig. 5.4). Vines account
for 71.4 % of the total climbers in this W. Asia.
Notes: About 52.3 % climbers in this region are shared with E. Africa, while 45.1 %
shared with Mediterranean and 42.2 % shared with Iranian plateau.

5.4.1.8 Mediterranean

Dominant taxa: The dominant families and genera with climbers in Mediterranean
region are similar to that of W. Asia. Leguminosae (138), Convolvulaceae (44),
Apocynaceae (17), Rubiaceae (14) and Cucurbitaceae (12) are the five largest
families with climbers and account for 76.0 % of the total climbers distributed in
this region. Vicia (76), Lathyrus (55), Cuscuta (22), Convolvulus (17) and Galium
(9) are the five largest genera with climbers.
Endemic taxa: Despite the low diversity, Mediterranean has impressive endemic
proportions. One monotypic genus (Cyprinia gracilis, however, some references
treated it under genus Periploca) and 70 species (23.7 %) are climber-endemic to
this region. Vicia (16 endemic) and Lathyrus (14 endemic) are the two largest
genera with endemic climbers in this region.
CPFs: CPF ranges from 3.8 % in Sardegna Island to 1.3 % in Iberian Peninsula
(Fig. 5.4). About 84.1 % climbers are herbaceous.
Notes: This region is the climber diversity center of Vicia, Lathyrus and Convolvulus.

5.4.1.9 Sahara

Dominant taxa: In general, Convolvulaceae (34), Leguminosae (27), Cucurbitaceae


(17), Apocynaceae (14), Capparaceae (5) and are the five largest families with
climbers and account for 71.3 % of the total climbers distributed in this region.
Ipomoea (16), Vicia (10), Cuscuta (8), Lathyrus (6) and Convolvulus (5) are the
five largest genera with climbers.
72 L. Hu and M. Li

Endemic taxa: None.


CPFs: CPF ranges from 9.2 % in North Chad to 1.8 % in Algerian and Libyan
Sahara (Fig. 5.4). Vines account for 72.1 % of the total climbers in this region.
Notes: About 98.5 % genera and 80.1 % species in this region are shared with tropi-
cal Africa, while 68.2 % genera and 65.4 % species shared with W. Asia, and
40.9 % genera and 41.9 % species shared with Mediterranean. Only 40 climbers
have been found in N. Sahara (excludes Nile Valley) adjacent to Mediterranean,
while 98 climbers distributed in S. Sahara next to the Sahel and the tropical Africa.

5.4.1.10 E. Asia

Dominant taxa: Leguminosae (54), Ranunculaceae (52), Apocynaceae (39),


Vitaceae (25) and Rubiaceae (20) and are the five largest families with climbers
and account for 51.5 % of the total climbers distributed in this region. Clematis
(42), Cynanchum (21), Vicia (19), Vitis (12) and Dioscorea (11) are the five larg-
est genera with climbers.
Endemic taxa: Although the diversity in E. Asia is much less compared to S. China,
the endemic proportions are still more. No genus, but 67 species (18.2 %) are
climber-endemic to this region. Clematis (14 endemic) and Cynanchum (12
endemic) are the two largest genera with endemic climbers in this region.
CPFs: CPF ranges from 6.4 % (N. Shaanxi Province) to 2.3 % (Sahalin). Vines are
slightly more than lianas here and account for 53.1 % of the total climbers.
Notes: It is also the liana diversity center of Polygonaceae, including 16 climbers here.

5.4.1.11 Mongolian Plateau

Dominant taxa: Leguminosae (20), Ranunculaceae (14), Convolvulaceae (12),


Apocynaceae (8) and Rubiaceae (6) constitute the five largest families with
climbers and account for 67.4 % of the total climbers distributed in this region.
Clematis (13), Vicia (13), Cynanchum (6), Cuscuta (6), Lathyrus (5) and Galium
(5) are the six largest genera with climbers.
Endemic taxa: Only two vines are endemic to this region (Cynanchum gobicum and
Galium trifidum).
CPFs: CPF ranges from 3.4 % (C. Inner Mongolia) to 0.7 % (Khubsugul of
Mongolia). Vines account for 77.5 % of the total climbers in this region.

5.4.1.12 Tianshan-Pamir-Kunlun

Dominant taxa: Leguminosae (33), Convolvulaceae (33), Rubiaceae (7) and


Ranunculaceae (6) are the four largest families with climbers and account for
76.7 % of the total climbers distributed in this region. Cuscuta (28), Vicia (19),
Lathyrus (13) and Clematis (6) are the four largest genera with climbers.
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 73

Endemic taxa: Only nine species are climber-endemic to this region (Ampelopsis
tadshikistanica, Bryonia lappifolia, Clematis iliensis, Cuscuta callinema,
Cuscuta lophosepala, Cuscuta pamirica, Cuscuta syrtorum, Cuscuta tianschan-
ica and Rosa silverhjelmii).
CPFs: CPF ranges from 1.7 % (Tajikistan, Kyrgyzstan) to none (Junggar Basin).
Vines account for 85.4 % of the total.
Notes: This is the largest diversity center of Cuscuta (28) in Eurasia and North
Africa.

5.4.1.13 Central Asia

Dominant taxa: Leguminosae (34), Convolvulaceae (24), Rubiaceae (7),


Ranunculaceae (6) and Cucurbitaceae (6) are the five largest families with climb-
ers and account for 87.5 % of the total climbers distributed in this region. Cuscuta
(21), Vicia (19), Lathyrus (13), Galium (5), Clematis (4) and Asparagus (4) are
the six largest genera with climbers.
Endemic taxa: Only four climbers are endemic to Central Asia (Bryonia melano-
carpa, Clematis ispahanica, Cuscuta elpassiana and Cuscuta karatavica).
CPFs: CPF is 2.3 % in Turkmenistan, 1.6 % in Uzbekistan and 1.3 % in Kazakhstan.
Vines account for 90.9 % of the total climbers in C. Asia.

5.4.1.14 S. Siberia

Dominant taxa: Leguminosae (26) and Convolvulaceae (10) are the two largest
families with climbers and account for 66.7 % of the total climbers distributed in
this region. Vicia (19), Cuscuta (6) and Lathyrus (6) are the three largest genera
with climbers.
Endemic taxa: Only five vines are climber-endemic to this region (Asparagus pal-
lasii, Cicer songaricum, Vicia olchonensis, Vicia popovii and Vicia tsydenii).
However, according to Flora of China, C. songaricum is also distributed in
C. Asia and Tianshan-Pamir-Kunlun regions.
CPFs: CPF ranges from 2.2 % (W. Siberia of this region) to 0.9 % (E. Siberia). All
climbers in this region are herbaceous.

5.4.1.15 E. Europe

Dominant taxa: Leguminosae (57) and Convolvulaceae (23) are the two largest
families with climbers and account for 64.5 % of the total climbers distributed in
this region. Vicia (30), Lathyrus (23) and Cuscuta (14) are the three largest gen-
era with climbers.
Endemic taxa: Only three vines are endemic to this region (Cuscuta glabrior,
Vincetoxicum juzepczukii and Vincetoxicum rossicum).
74 L. Hu and M. Li

CPFs: CPF ranges from 2.1 % (Bulgaria, S. European Russia) to 1.2 % (Poland).
Vines account for 79.8 % of the total climbers.

5.4.1.16 W. Europe

Dominant taxa: The dominant families and genera with climbers in W. Europe are
similar to that of E. Europe. Leguminosae (55) and Convolvulaceae (16) are the
two largest families with climbers and account for 67.0 % of the total climbers
distributed in this region. Vicia (31), Lathyrus (20) and Cuscuta (9) are the three
largest genera with climbers.
Endemic taxa: Only one species is climber-endemic to this region (Rubus lejeunei),
described as shrub with long, climbing stems in Flora of Europe.
CPFs: CPF ranges from 2.3 % (France) to 1.6 % (Germany). Vines account for
81.9 % of the total climbers.

5.4.1.17 N. Eurasia (including Russian Far East, N. Siberia


and N. Europe)

Dominant taxa: Vicia (16) and Lathyrus (10) are the two largest genera with climb-
ers and account for 53.1 % of the total climbers and make Leguminosae the larg-
est family with climbers in this region.
Endemic taxa: No species are climber-endemic to any of the three regions in
N. Eurasia. However, Vicia macrantha is endemic to Russian Far East and east-
ern N. Siberia.
CPFs: CPF in those regions rarely exceed 2 %, and never exceed 1 % in the most
northern districts. All climbers in N. Siberia are herbaceous, while one subspe-
cies (Clematis alpina subsp. ochotensis) in Russian Far East and three species
(Clematis alpina, Lonicera periclymenum and Hedera helix) in N. Europe are
lianas.
Notes: The total of 49 climbers distributed in N. Eurasia is thus less than that of
Canada and Alaska (Table 5.5). Eight families and nine genera are shared by
N. Eurasia and Canada & Alaska. However, only four species (Calystegia sepium,
Galium aparine, Lathyrus japonicas and Lathyrus palustris) are distributed in
both the areas. Although vines of Vicia and Lathyrus have weak or developed
tendrils modified from end leaflets, they are rarely described as climbing plants
in those regions. Most genera and species in N. Eurasia are widespread. Nine
genera and six species are distributed in no less than 100 districts in Eurasia and
North Africa. Four more genera and 11 more species have been recorded in no
less than 50 districts.
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 75

5.4.2 Distribution of Genus and Species

At least four geographical centers are distinguished in Eurasia and North Africa at
generic level: (a) SE. Asia (and W. Malay Archipelago) - Many genera are climber
diverse in this area such as Calamus, Bauhinia, Derris, Ficus and Nepenthes.
(b) S. China. Most climber-abundant in Actinidia, Clematis, Hemsleya, Rubus,
Schisandra, Vitis etc. (c) Mediterranean. Geographical center of Vicia and Lathyrus
and Convolvulus vines and it is also abundant in Cuscuta. (d) S. Asia. No geograph-
ical center of climber-abundant genus is located in S. Asia. However, 47 genera with
one or a few climbers each are climber-endemic to this region, such as Baeolepis,
Decalepis and Oianthus.
The top ten widespread genera with climbers are belong to eight different fami-
lies (Table 5.8). Vicia is the most widespread genus with climbers in Eurasia and
North Africa. Mediterranean is the climber diversity center and this region is
climber-absent from 28 districts, including most districts of SE. Asia and some dis-
tricts of S. Asia, Tibetan Plateau, W. Asia and S. Sahara. Cuscuta is the second
widespread genus with climbers. Tianshan-Pamir-Kunlun and Iranian plateau are
the climber diversity centers and it is absent from most districts of N. Eurasia and
some districts of S. Sahara and Tibetan Plateau. Climbers of the top 10 widespread
genera are mostly herbaceous dominant, except for Clematis. S. China is the climber
diversity center of Clematis and 80 % climbers in Eurasia & N. Africa are somewhat
woody. Fallopia is the smallest widespread genus with only eight climbers (one
liana and seven vines) but distributed in 130 districts. F. convolvulus is the only truly
widespread species in Fallopia and the rest are mostly restricted in less than 30
districts except for F. dumetorum.

Table 5.8 Top 10 widespread climbers and genera with climbers in Eurasia and North Africa
Widespread genera Widespread species
Recorded Recorded
No. Genus districts Species districts
1 Vicia (Legu.) 182 Convolvulus arvensis (Conv.) 136
2 Cuscuta (Conv.) 179 Vicia cracca (Legu.) 122
3 Clematis (Ranu.) 167 Fallopia convolvulus (Poly.) 112
4 Galium (Rubi.) 164 Vicia hirsuta (Legu.) 109
5 Lathyrus (Legu.) 153 Cuscuta europaea (Conv.) 104
6 Convolvulus (Conv.) 140 Galium spurium (Rubi.) 100
7 Cynanchum (Apoc.) 138 Lathyrus pratensis (Legu.) 88
8 Fallopia (Poly.) 130 Lathyrus palustris (Legu.) 86
9 Rubia (Rubi.) 128 Calystegia sepium (Conv.) 83
10 Asparagus (Aspa.) 126 Rubia cordifolia (Rubi.) 83
76 L. Hu and M. Li

Table 5.9 Recorded districts of climbers and genera with climbers in Eurasia & N. Africa
Genus Species
Recorded districts Num. Proportion Num. Proportion
1 154 19.0 2,504 37.6
2 86 10.6 1,000 15.0
3 59 7.3 623 9.4
4 27 3.3 435 6.5
5 37 4.6 322 4.8
6 ~ 10 100 12.4 799 12.0
11 ~ 20 121 15.0 585 8.8
21 ~ 50 167 20.6 364 5.5
51 ~ 100 45 5.6 22 0.3
>100 13 1.6 5 0.1
Total 809 100 6,659 100

Most climbers in Eurasia and North Africa are steno-choric. About 52.6 %
species and 29.6 % genera are restricted to one or two districts (Table 5.9). The real
widespread climbers are very limited. Only 14.7 % species and 42.8 % genera have
been recorded in more than ten districts, and only 27 species (0.4 %) and 58 genera
(7.2 %) have been recorded in more than 50 districts. The top ten widespread
recorded climbers (all vines) are listed in Table 5.8.

5.5 Conclusion

Although climbing habits have been developed in 169 families, most climbers
belong to few families or genera. Apocynaceae and Leguminosae are the two largest
families with climbers and both have more than 2,000 climbers in the world.
Convolvulaceae is the third largest family and have more than 1,000 climbers.
About 25 % climbers in the world belong to the three largest families. Similar pat-
tern has also been discussed by Gentry (1991) at generic level.
Generally, both CPF and climber diversity in Eurasia and North Africa decrease
from low latitudes to high latitudes and decrease from coastal or humid areas to
inland or arid regions. Water and energy are suggested to be critical environmental
factors responsible for these diversity gradients (Molina-Freaner et al. 2004;
Schnitzer 2005), and both geographical and environmental gradients affected lianas
significantly but not vines (Hu et al. 2010b). Our data support these theories and
lianas are dominant in SE. Asia, S. Asia, S. China and Himalaya, while less impor-
tant in plateau, desert and northern districts of Eurasia and North Africa.
Climbing plants are important components of ecosystems and they make such an
appreciable contribution to global species richness. Their responses to geographical
and environmental factors as well as geographical and environmental patterns aris-
ing therefrom are both important for explanation of their historical and ecological
phytogeography.
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 77

Acknowledgements We thank Yuan Huang for assistance in database organization. We thank our
students M.J. SHU, L. LIN, J.Y. LIU, H.R. WANG, H.Y. DENG, Y. ZHANG, C.Y. GUO, Q. DAI,
Y.Y. GUAN for assistance with data collection. We thank J. Mugnier for data on Senegal flora,
C.S. Chang for assistance on Korea climbers, D. Podlech for assistance on Afghanistan climbers.
This research was financially supported by the National Natural Science Foundation of China
(Project 41101057).

Appendix

Climbing plant diversity of the Old World (Eurasia, Africa & West Malay Archipelago)
Num. Family Genus Species Num. Family Genus Species
1 Leguminosae 149 1,517 34 Icacinaceae 17 106
2 Apocynaceae 176 1,355 35 Dichapetalaceae 1 101
3 Vitaceae 11 621 36 Nepenthaceae 1 101
4 Convolvulaceae 41 593 37 Asparagaceae 8 96
5 Rubiaceae 88 589 38 Primulaceae 4 93
6 Cucurbitaceae 76 552 39 Orchidaceae 24 87
7 Annonaceae 43 426 40 Araliaceae 7 83
8 Celastraceae 32 359 41 Campanulaceae 6 82
9 Arecaceae 12 340 42 Rutaceae 12 76
10 Menispermaceae 50 297 43 Actinidiaceae 3 66
11 Rosaceae 7 239 44 Gentianaceae 6 64
12 Ranunculaceae 6 230 45 Urticaceae 20 60
13 Compositae 46 207 46 Caprifoliaceae 3 57
14 Dioscoreaceae 3 203 47 Solanaceae 5 49
15 Piperaceae 3 188 48 Amaranthaceae 13 43
16 Combretaceae 6 181 49 Gesneriaceae 6 42
17 Melastomataceae 27 179 50 Polygonaceae 10 38
18 Euphorbiaceae 29 164 51 Cannabaceae 2 38
19 Lamiaceae 27 161 52 Lardizabalaceae 6 38
20 Malvaceae 26 154 53 Pandanaceae 2 38
21 Malpighiaceae 17 151 54 Phyllanthaceae 11 37
22 Oleaceae 4 138 55 Schisandraceae 2 37
23 Passifloraceae 10 138 56 Linaceae 3 35
24 Araceae 14 131 57 Boraginaceae 7 33
25 Moraceae 8 131 58 Ericaceae 6 32
26 Capparaceae 10 127 59 Gnetaceae 1 32
27 Poaceae 43 118 60 Dilleniaceae 2 31
28 Connaraceae 14 115 61 Sabiaceae 1 31
29 Rhamnaceae 12 114 62 Sapindaceae 10 30
30 Aristolochiaceae 4 113 63 Geraniaceae 3 29
31 Smilacaceae 2 110 64 Hernandiaceae 1 27
32 Loganiaceae 3 109 65 Papaveraceae 6 24
33 Acanthaceae 29 108 66 Elaeagnaceae 1 24
(continued)
78 L. Hu and M. Li

Num. Family Genus Species Num. Family Genus Species


67 Anacardiaceae 10 23 106 Gelsemiaceae 2 3
68 Commelinaceae 10 23 107 Myrtaceae 2 3
69 Crassulaceae 6 22 108 Ochnaceae 2 3
70 Ancistrocladaceae 1 20 109 Oxalidaceae 1 3
71 Nyctaginaceae 3 20 110 Salvadoraceae 2 3
72 Stemonaceae 1 17 111 Adoxaceae 1 2
73 Hydrangeaceae 6 16 112 Cardiopteridaceae 1 2
74 Polygalaceae 3 16 113 Ebenaceae 1 2
75 Thymelaeaceae 5 16 114 Flagellariaceae 1 2
76 Bignoniaceae 10 15 115 Grossulariaceae 1 2
77 Marantaceae 4 15 116 Hypericaceae 1 2
78 Scrophulariaceae 6 14 117 Myristicaceae 2 2
79 Begoniaceae 1 12 118 Pittosporaceae 1 2
80 Opiliaceae 5 12 119 Zygophyllaceae 2 2
81 Olacaceae 4 11 120 Achariaceae 1 1
82 Santalaceae 5 11 121 Barbeuiaceae 1 1
83 Lythraceae 2 10 122 Gerrardinaceae 1 1
84 Paulowniaceae 2 10 123 Goodeniaceae 1 1
85 Caryophyllaceae 4 9 124 Hamamelidaceae 1 1
86 Chrysobalanaceae 1 9 125 Iteaceae 1 1
87 Orobanchaceae 6 9 126 Lecythidaceae 1 1
88 Plantaginaceae 5 9 127 Lophopyxidaceae 1 1
89 Apiaceae 3 8 128 Monimiaceae 1 1
90 Meliaceae 1 8 129 Montiniaceae 1 1
91 Burseraceae 1 7 130 Pandaceae 1 1
92 Lauraceae 2 7 131 Pedaliaceae 1 1
93 Ephedraceae 1 6 132 Pentadiplandraceae 1 1
94 Salicaceae 4 6 133 Pentaphylacaceae 1 1
95 Balsaminaceae 1 5 134 Peraceae 1 1
96 Basellaceae 1 5 135 Peridiscaceae 1 1
97 Cornaceae 1 5 136 Phytolaccaceae 1 1
98 Loranthaceae 5 5 137 Portulacaceae 1 1
99 Plumbaginaceae 1 5 138 Resedaceae 1 1
100 Colchicaceae 2 4 139 Rhizophoraceae 1 1
101 Cyperaceae 2 4 140 Sapotaceae 1 1
102 Talinaceae 2 4 141 Stilbaceae 1 1
103 Aquifoliaceae 1 3 142 Ulmaceae 1 1
104 Brassicaceae 2 3 143 Verbenaceae 1 1
105 Dioncophyllaceae 3 3 Total 1,415 12,382
5 Diversity and Distribution of Climbing Plants in Eurasia and North Africa 79

References

Darwin C (1865) On the movements and habits of climbing plants. J Linn Soc (Bor) 9:1–118
Gentry AH (1991) The distribution and evolution of climbing plants. In: Putz FE, Mooney HA
(eds) The biology of vines. Cambridge University Press, Cambridge, pp 3–42
Jongkind CCH, Hawthorne WD (2005) A botanical synopsis of the lianes and other forest climb-
ers. In: Bongers F, Parren MPE, Traoré D (eds) Forest climbing plants of West Africa: diversity,
ecology and management. CABI Publishing, Oxford, pp 19–39
Hu L (2011) Distribution and diversity of climbing plants in temperate East Asia. Biodivers Sci
19(5):567–573 (in Chinese with English Abstract)
Hu L, Li MG, Li Z (2010a) The diversity of climbing plants in the spermatophyte flora of China.
Biodivers Sci 18(2):198–207 (in Chinese with English Abstract)
Hu L, Li MG, Li Z (2010b) Geographical and environmental gradients of lianas and vines in
China. Glob Ecol Biogeogr 19:554–561
Molina-Freaner F, Gamez RC, Tinoco-Ojanguren C, Castellanos AE (2004) Vine species diversity
across environmental gradients in northwestern Mexico. Biodivers Conserv 13:1853–1874
Muthumperumal C, Parthasarathy N (2009) Angiosperms, climbing plants in tropical forests of
southern Eastern Ghats, Tamil Nadu, India. Check List 5(1):092–111
Parsons RF (2005) Desert vines: a comparison of Australia with other areas. J Biogeogr
32:121–126
Putz FE, Mooney HA (1991) The biology of vines. Cambridge University Press, Cambridge
Schnitzer SA (2005) A mechanistic explanation for global patterns of liana abundance and distri-
bution. Am Nat 166:262–276
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
The Angiosperm Phylogeny Group (2009) An update of the Angiosperm Phylogeny Group
classification for the orders and families of flowering plants: APG III. Bot J Linn Soc
161:105–121
The Plant List (2013) Version 1.1. Published on the internet. http://www.theplantlist.org/. Accessed
11 Sept 2014.
Wu YH (2008) The vascular plants and their eco-geographical distribution of the Qinghai-Tibetan
Plateau. Science Press, Beijing (in Chinese with English Abstract)
Yan LH (2009) Climbing plants of the Central China. Hunan Science & Technology Press,
Changsha (in Chinese with English Abstract)
Yan LH, Qi CJ (2007) Vine diversity of Huping Mountain in Hunan Province. Sci Silvae Sin
43(6):20–26 (in Chinese with English Abstract)
Yan LH, Qi CJ, Liu XX (2006) A study on the flora of the seed vines in Central China region. Bull
Bot Res 26(4):497–507 (in Chinese with English Abstract)
Chapter 6
Liana Assemblages in Tropical Forests
of Africa and Southeast Asia: Diversity,
Abundance, and Management

Patrick Addo-Fordjour and Zakaria B. Rahmad

Abstract Lianas form an important component of tropical forest ecosystems in


Africa and Southeast Asia, but there is scanty information on liana ecology in these
two eco-regions. Furthermore, there is dearth of information about intercontinental
comparison of liana ecology in the tropics. This chapter therefore, describes and
synthesises studies conducted in these two regions. The patterns of liana abundance
and diversity, and the factors that affect them in the two continents were described
and compared in the chapter. Additionally, the chapter describes various manage-
ment interventions used to control lianas in Africa and Southeast Asia, and their
consequences on liana assemblages and forest biodiversity. Tropical forests of
Africa tend to harbour higher liana diversity and abundance than Southeast Asian
forests. Liana assemblages in the two regions are related with a number of factors
including human disturbance, soil properties, rainfall and topography. Liana cutting
was the main silvicultural tool used in controlling lianas in Africa and Southeast
Asia. This silvicultural tool is integrated in many forest management systems in the
two regions. Although liana cutting reduces liana abundance, it adversely affects
liana diversity which could influence the overall forest biodiversity of treated
forests.

Keywords Liana cutting • Liana diversity and abundance • Soil properties


• Topographic factors • Human disturbance

P. Addo-Fordjour (*)
Department of Theoretical and Applied Biology, College of Science, Kwame Nkrumah
University of Science and Technology (KNUST), Kumasi, Ghana
School of Biological Sciences, Universiti Sains Malaysia, Penang, Malaysia
e-mail: paddykay77@yahoo.com; paddofordjour.cos@knust.edu.gh
Z.B. Rahmad
Department of Theoretical and Applied Biology, College of Science, Kwame Nkrumah
University of Science and Technology (KNUST), Kumasi, Ghana
The Centre for Marine and Coastal Studies (CEMACS), Universiti Sains Malaysia,
11800 Pulau Pinang, Penang, Malaysia

© Springer International Publishing Switzerland 2015 81


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_6
82 P. Addo-Fordjour and Z.B. Rahmad

6.1 Introduction

Lianas are woody climbing plants that are rooted in the soil but due to their weak
stems use trees and other plants as structural support to get to the forest canopy.
They grow very fast because their non-self supporting nature allows them to spend
little resources on structural support, therefore being able to allocate more resources
for rapid stem growth (cf. Schnitzer and Bongers 2002; Tang et al. 2012). Although
lianas are found in many ecosystems in the world, they are more abundant and
diverse in tropical forest ecosystems (Schnitzer and Bongers 2002; Bongers et al.
2005). In the tropics, lianas may be more abundant in disturbed areas of forests such
as edges (Laurance et al. 2001) and gaps (Schnitzer et al. 2004). They exhibit rapid
recruitment of many individuals into gaps following disturbance that enables them
to predominate in disturbed areas (Schnitzer et al. 2004). Lianas could have strong
ramifications in tropical forests, affecting forest dynamics and functioning (Schnitzer
and Bongers 2011). They form an important component of tropical forest physiog-
nomy (Tang et al. 2012), and are one of the major physiognomic characteristics
distinguishing tropical forests from temperate forests (cf. Schnitzer and Bongers
2002). Lianas compete effectively with seedlings and saplings (Schnitzer et al.
2005; Toledo-Aceves and Swaine 2008) and adult trees (Ingwell et al. 2010; Tobin
et al. 2012) which can exert significant influence on forest natural regeneration
processes.
Although liana studies have been reported from Africa and Southeast Asia, the
patterns of liana assemblages in these two regions are not clear. This is due to the
fact that studies conducted so far have been scanty, scattered and non-synthesised.
Consequently, this chapter provides a synthesis of published papers on liana ecol-
ogy in Africa and Southeast Asia with the view to determining the main patterns of
liana diversity and abundance, and the factors that influence the patterns in these
eco-regions. Furthermore, the chapter describes management of lianas in tropical
forests of the two regions. The chapter also identifies existing gaps in liana ecology
literature in the two eco-regions. This information will be valuable for future
research on lianas in these continents.

6.2 Patterns of Liana Diversity in Tropical Forests of Africa


and Southeast Asia

Although lianas form an integral component of many forest ecosystems worldwide,


they show higher diversity in tropical forests (Schnitzer and Bongers 2002). The
higher diversity of lianas in tropical forest ecosystems is a major feature that distin-
guishes tropical forests from temperate forests. Available data from Africa and
Southeast Asia indicates that lianas in these regions add substantially to the baseline
diversity of plants in their tropical forests. Studies from other parts of the tropics
revealed that lianas comprise of about 25 % of woody plant species richness of
forests (cf. Schnitzer and Bongers 2002; Bongers et al. 2005). However, studies
6 Liana Assemblages in Tropical Forests of Africa and Southeast Asia… 83

from some tropical forests in Africa and Southeast Asia show that some forests can
even possess higher liana species richness in relation to woody plant species rich-
ness. For instance, liana species composed of about 37–38 % of woody plant spe-
cies richness in tropical forests in Ghana (Addo-Fordjour et al. 2008; Swaine et al.
2005). Kouame (1998) reported that lianas formed 30 % of woody species in Forêt
Classée de Haut Sassandra, Ivory Coast (see Bongers et al. 2002). Later, Kuzee
(2002) reported of a much higher liana species composition (40 % of total woody
species) in a different forest in Ivory Coast (see Bongers et al. 2002). Generally,
there is not much information about the proportion of liana species among woody
plant species in tropical forests of Southeast Asia. Nevertheless, available informa-
tion shows that the proportion of liana species among woody plant species in tropi-
cal Southeast Asian forests is lower than the figures reported by most studies in
African tropical forests. For example, lianas constituted 17 % of woody plant spe-
cies in a Philippine lowland forest (Langenberger et al. 2006) and Ghollasimood
et al. (2011) recorded a lower proportion of liana species among woody plant spe-
cies (9 %) in Malaysia. But a study conducted in some Malaysian rainforests
reported of 30 % liana species composition among woody plant species (Appanah
et al. 1993). So far, this is the only study from Southeast Asia whose liana composi-
tion is somehow close to the percentages recorded in African forests. Appanah et al.
(1993) reported that rattans (woody climbing palms) constitute a significant feature
of Southeast Asian tropical forests but are relatively scarce in African tropical for-
ests. This finding is corroborated by a comparative study conducted between Ghana
and Malaysia in which many rattan species were recorded among a number of for-
ests in Malaysia but none was recorded in Ghanaian forests (Addo-Fordjour and
Rahmad unpublished data). Boonsermsuk et al. (2007) noted that rattans mostly
occur in forests with high rainfall and humidity. Based on this finding, it is hypoth-
esized that the higher rattan species richness in Southeast Asian forests relative to
African forests is attributed to higher rainfall and humidity that prevail in Southeast
Asia.
As in other parts of the world, lianas in Africa and Southeast Asia also occur in a
wide range of plant families. Nevertheless some families contain more liana species
than others. In a more extreme case, a few plant families consist of only climber
species (Bongers et al. 2005). For instance, Dilleniaceae, Convolvulaceae, Linaceae,
Cucurbitaceae and Dioscoreaceae found in many African and Southeast Asian tropi-
cal forests are made up of only climbers (Bongers et al. 2005; Hawthorne and
Jongkind 2006). The dominance of some liana families (in terms of species rich-
ness) is pantropical as they contribute high species numbers to liana flora across
different countries and continents in the tropics. However, there are a few liana
families whose dominance is limited to one continent or the other (see Appanah
et al. 1993). In Africa and Southeast Asia, Fabaceae and Apocynaceae occur among
the most species rich families in different forests (Putz and Chai 1987; DeWalt et al.
2006; Nurfazliza et al. 2012). Nonetheless, Celastraceae occurs as a dominant liana
family among liana communities in Africa but not in Southeast Asia (Senbeta et al.
2005; Addo-Fordjour et al. 2008, 2009a, b, 2013a, b), and Annonaceae, Connaraceae
and Loganiaceae exhibit high dominance among liana communities in Southeast
84 P. Addo-Fordjour and Z.B. Rahmad

Table 6.1 Types of climbing mechanisms used by liana communities within West African tropical
forests
Type of climber Mode of climbing
Stem twining climber Uses main stem
Stem tendril climber Uses tendrils on stem
Leaf tendril climber Uses tendril of leaf or tendril-like petiole of leaf
Hook climber Uses hooks located on stems or branches
Grappler climber Uses horizontal or recurved branches to grasp the host
Thorn climber Uses recurved thorns on stem
Branch twining climber Uses branches of stem as its main climbing organ. The main stem may
be used in addition to the branches
Root climber Uses adventitious root located on the stem
Leaning climber Leans on the host using the main stem

Asia but not in Africa (Putz and Chai 1987; DeWalt et al. 2006; Nurfazliza et al.
2012; Addo-Fordjour et al. 2012).
Usually, liana communities in tropical forests of Africa and Southeast Asia
employ several climbing mechanisms to ascend their host (Bongers et al. 2005;
Jongkind 2005; Addo-Fordjour et al. 2008). The more diverse the climbing mecha-
nisms of a particular liana community, the better its ability to climb different host
sizes (Nabe-Nielsen 2001), and therefore inhabit a wide range of habitats within
tropical forests. The type of climbing mechanism employed by liana species deter-
mines the height they can attain in a forest (Baars et al. 1998). In West Africa, many
climbing mechanisms have been identified among several liana communities in
various forest types (Jongkind 2005; Addo-Fordjour et al. 2008). These climbing
mechanisms can be grouped into nine main types (Table 6.1). Although there is rela-
tively fair knowledge about diversity of liana climbing mechanisms in West Africa,
information from other parts of Africa is either inadequate or non-existent. For
example, in Central Africa, only one study assessed the climbing mechanisms
employed by lianas in Ituri forest in Democratic Republic of Congo (Ewango 2010).
A total of four liana climbing mechanisms were reported among the liana communi-
ties in the Ituri forest. Interestingly, all the four climbing mechanisms identified in
the Ituri forest are common to West African liana communities, suggesting that liana
communities in different parts of Africa may not differ much in their mechanisms of
climbing. In Southeast Asia there is little information on liana climbing mecha-
nisms as only one study has reported on this subject matter. That study reported of
six types of climbing mechanisms used by lianas in climbing trees in primary and
secondary forests in Malaysia (Addo-Fordjour et al. 2012). All the types of climbing
mechanisms identified in Malaysia were also reported in Ghana (and other parts of
Africa), albeit the predominance of the climbing mechanisms differed between the
two regions (Fig. 6.1). Though different studies have recorded different types and
number of climbing mechanisms, it is intriguing to note that stem twining has been
the most predominant climbing mechanism in terms of species richness and abun-
dance in the two regions (Bongers et al. 2005; Addo-Fordjour et al. 2008, 2012).
6 Liana Assemblages in Tropical Forests of Africa and Southeast Asia… 85

1000

900
Ghana
800
Malaysia
700
Abundance

600

500

400

300

200

100

0
Stem Branch Hook Thorn Stem Leaning Root Leaf tendril
twining twining tendril
Climbing mechanism

Fig. 6.1 Comparison of liana climbing mechanism abundance in Ghana and Malaysia

This phenomenon is however, not peculiar to Africa and Southeast Asia, as it has
also been reported from other parts of the tropics (e.g. Parthasarathy et al. 2004).
Generally, the number of lianas with particular climbing mechanism may depend on
the successional stage and disturbance level of forest ecosystems (Baars et al. 1998;
cf. DeWalt et al. 2000; Schnitzer and Bongers 2002). Thus, the age, disturbance his-
tory and diameter distribution of trees of forests to a great extent, determine the type
of climbing mode that dominates particular tropical forests. For instance, hook and
tendril climbers are limited to small diameter trees compared to stem twiners that
are able to climb up to trees of 30 cm diameter (Pinard and Putz 1994). For this
reason, hook and tendril climbers may be more common in young secondary forests
where small diameter trees and dense vegetations are predominant (Baars et al.
1998; Addo-Fordjour et al. 2012).

6.3 Liana Abundance in Tropical Forests of Africa


and Southeast Asia

A number of studies indicate that tropical forests have high liana abundance even
though there is inadequate information concerning liana abundance dynamics in the
forests. Long term data on liana abundance dynamics is important for better under-
standing of the patterns of liana abundance changes in the tropics. Currently, there
is scanty information on this subject matter in Africa, while there is no information
at all in Southeast Asia. Thus, the two eco-regions differ with respect to the attention
86 P. Addo-Fordjour and Z.B. Rahmad

given to studies on long term liana abundance dynamics. In Africa, Caballé and
Martin (2001) and Ewango (2010) examined long term changes in liana abundance
and observed that lianas decreased in abundance by about 20 % and 33 %, respec-
tively, over a period of 13 years each. Interestingly, the above-mentioned pattern
from Africa is in sharp contrast with that reported from other parts of the tropics
(see Schnitzer and Bongers 2011; Laurance et al. 2014), indicating that factors gov-
erning liana dynamics in Africa may differ from those in other continents.
Lianas are highly abundant in a number of forests in different tropical countries
but there is virtually no information about intercontinental comparisons of liana
abundance in tropical forests. Differences in liana abundance between some conti-
nents in the tropics are clear cut (e.g. South America contains more lianas than
Asia), but liana abundance variations between other areas are not obvious due to
limited number of studies in those areas (e.g. Africa and Asia). Currently, there are
mixed results concerning which of the two regions, Africa and Southeast Asia, har-
bours higher liana abundance. Emmons and Gentry (1983) indicated that lianas are
more abundant in some African forests in comparison with forests in Malaysia,
Southeast Asia. A comparative analysis of different studies by Lü et al. (2009)
showed that the abundance of large diameter lianas (dbh > 10 cm) per hectare was
higher in African tropical forests than in Southeast Asian forests. Similarly, Gentry
(1991) found that the abundance of lianas with dbh ≥ 2.5 cm (110 stems per 0.1 ha)
was higher in African forests than Southeast Asian forests (70 stems per 0.1 ha) (cf.
Parren and Bongers 2001). Furthermore, a recent study conducted in Malaysia and
Ghana using the same methods and forest types indicated that liana abundance
(dbh ≥ 2 cm) in Ghanaian forests were about three times more abundant than in
Malaysian forests (Addo-Fordjour and Rahmad unpublished data). Despite the
overwhelming evidence demonstrating higher liana abundance in African forests
than Asian forests, Appanah and Putz (1984) recorded liana abundance for
dbh > 5 cm in a tropical forest in Pahang, Malaysia, which was higher than what was
reported by a similar study conducted in Cameroon, Africa (cf. Parren and Bongers
2001). The above stated data suggests that majority of the studies overwhelmingly
support a pattern of higher liana abundance in African forests than Southeast Asian
forests. This trend may be due to differences in human disturbance and amounts of
rainfall experienced by tropical forests in the two regions. Human disturbance is
usually higher in most African forests than their counterparts in Southeast Asia, and
this phenomenon could be responsible for higher liana abundance in Africa than
Southeast Asia, in view of the fact that liana abundance increases with disturbance
in some forests in the tropics (Schnitzer and Bongers 2011). Also, Southeast Asian
forests are wetter than most African forests because they experience relatively
higher amounts of rainfall. The drier nature of forests in Africa may support higher
liana abundance as studies have shown that drier forests usually harbour higher
abundance of lianas in the tropics (Swaine and Grace 2007; DeWalt et al. 2010).
Although liana abundance is usually high in African forests, it can vary from one
part of the continent to another. For example lianas are less common in Afromontane
forests than in lowland Guineo-Congolian forests (cf. Parren 2003). This pattern
6 Liana Assemblages in Tropical Forests of Africa and Southeast Asia… 87

may reflect differences in edaphic, topographic and climatic variations, and human
disturbance levels between the two regions. For example, most Afromontane forest
communities occur above an elevation of 1,200 m above sea level compared to most
Guineo-Congolian forests which are located on lowlands. Therefore, the higher
elevations of Afromontane forests may be partly responsible for its lower liana
abundance compared to the Guineo-Congolian forests in Africa, in view of the fact
that some previous studies from Africa (Addo-Fordjour et al. 2013c; Addo-Fordjour
and Rahmad 2014) and other parts of the tropics (Parthasarathy et al. 2004; Homeier
et al. 2010) revealed this trend.

6.4 Factors Influencing Liana Diversity and Abundance


in Tropical Forests of Africa and Southeast Asia

6.4.1 Forest Disturbance

Generally, lianas respond to disturbance, especially human-induced type, in tropical


forest ecosystems. In Africa and Southeast Asia, human disturbance such as farm-
ing, logging, hunting and fragmentation of forests play a great role in structuring
liana assemblages (Addo-Fordjour et al. 2009b, 2012; cf. Schnitzer and Bongers
2011). Many forests in the two regions undergo continuous human disturbance that
alter liana diversity and abundance in diverse ways (Addo-Fordjour et al. 2008,
2009a, b, 2013a, b). Disturbance is one of the factors that maintain liana diversity
and abundance in tropical forests (Addo-Fordjour et al. 2013b; Anbarashan and
Parthasrathy 2013; Ledo and Schnitzer 2014). In comparison with forests of other
parts of the tropics, African forests harbour considerable liana diversity and abun-
dance (Schnitzer and Bongers 2002). This phenomenon is partly associated with
human disturbance which commonly occurs in African forests. Anthropogenic dis-
turbances often create gaps in tropical forests, providing favourable conditions for
lianas to proliferate (Schnitzer 2005). Forest gap and edge creation is a common
phenomenon in Africa, and this has contributed to higher liana diversity and abun-
dance in African forests. This is particularly so in West Africa where many lianas
are light demanders and so multiply rapidly within forest gaps and edges (cf. Parren
2003). In general, lianas at forest edges undergo a more rapid growth than their
counterparts in forest interior because they are rooted in relatively drier soils and
have better access to high light intensity than those in interior parts of forests
(Schnitzer and Bongers 2011). This explains why forest edges support higher liana
diversity and abundance in tropical forests (Laurance et al. 2001). The aggressive
nature of West African lianas in response to disturbance could have adverse effect
on trees in many ways within some forests. For example, aggressive proliferation of
lianas occurred in the Kakum National Park, Ghana following human disturbance
which has precluded natural regeneration of trees in areas where lianas have formed
monotypic stands (Addo-Fordjour and Rahmad unpublished data; Fig. 6.2).
88 P. Addo-Fordjour and Z.B. Rahmad

Fig. 6.2 Liana proliferation in a disturbed secondary forest in the Kakum National Park, Ghana.
This phenomenon has adversely affected natural regeneration of trees in the forest

Because lianas depend on trees to ascend to the forest canopy, their diversity and
abundance may be also dependent on tree assemblages. For example, studies con-
ducted in Ghana and Malaysia demonstrated that liana abundance was strongly
related to tree abundance and diameter (Addo-Fordjour et al. 2009a, b, 2012). This
suggests that any forest disturbance which affects tree assemblages may likely influ-
ence liana diversity and abundance. Limited availability of trees due to human dis-
turbance was cited as a possible factor responsible for lower liana diversity and
abundance in disturbed secondary forests in Ghana and Malaysia (Addo-Fordjour
et al. 2008, 2013a).

6.4.1.1 Mechanisms That May Explain Liana Assemblage Dynamics


in Disturbed Areas of Africa and Southeast Asia

A couple of mechanisms have been outlined to explain how liana diversity and
abundance increase following forest disturbance. According to Schnitzer and
Bongers (2011), lianas respond to disturbance by recruiting many individuals into
disturbed areas, and then after, use high resources that may prevail in those areas for
their growth. Lianas can employ a variety of reproduction modes in disturbed open
areas which could contribute to higher liana diversity and abundance in those areas
(e.g. Schnitzer et al. 2000; Rutishauser 2011; Ledo and Schnitzer 2014). Schnitzer
and Bongers (2011) outlined various reproduction methods which lianas may use to
recruit individuals into disturbed areas: the use of seed, advance regeneration, lat-
eral growth from the intact understory on the forest floor, and clonal recruitment.
6 Liana Assemblages in Tropical Forests of Africa and Southeast Asia… 89

Out of these methods, clonal recruitment has been reported as the most predominant
method of recruitment in maintaining liana diversity and abundance following
disturbance in tropical forests of North America (Yorke et al. 2013; Ledo and
Schnitzer 2014). However, there is a complete dearth of information about which of
the above mentioned recruitment methods help maintain liana assemblages in Africa
and Southeast Asia. Natural or human disturbance can cause lianas to fall from the
forest canopy onto forest floor (Putz 1984, 2012; Yorke et al. 2013) where some of
them can produce new stems and roots by vegetative means, thereby increasing the
abundance of lianas in the forest (Schnitzer and Bongers 2011; Yorke et al. 2013).
High liana infestations associated with some African and Southeast Asian forests
can trigger the pulling down of trees by lianas. This phenomenon can cause forma-
tion of new forest gaps or increase the size of existing ones thereby providing
favourable conditions for lianas to proliferate and increase in abundance.
Nevertheless, in some tropical forests of Africa and Southeast Asia liana diversity
and abundance decreased in human disturbed areas (Addo-Fordjour et al. 2008,
2009a, b, 2012). This trend has been linked to inadequate number of tree supports
and liana removal due to human disturbance (e.g. Chittibabu and Parthasarathy
2001; Addo-Fordjour et al. 2009a, b; Rahman et al. 2010). In other words, although
human disturbance may result in higher light intensity in disturbed areas, such areas
may not be suitable for lianas as they may harbour less number of trees to support
lianas. This means that the nature and level of human disturbance and how it affects
trees and even lianas themselves will determine the nature of liana dynamics in
disturbed areas (Addo-Fordjour et al. 2009a, b). Considering the fact that human
disturbance is a major phenomenon in tropical forests in Africa and to some extent
Southeast Asia, the above mentioned mechanisms may play a great role in liana
dynamics in these regions.

6.4.2 Environmental Factors

Although lianas have received more attention recently, there is little information
about edaphic factors that affect liana assemblages in tropical forests. The situation
is even worse with regards to Africa and Southeast Asia. In both regions, the studies
which investigated the role of soil properties on liana diversity and abundance
occurred in only one or two countries. In Africa, all the studies that assessed the
relationship between soil properties, and liana diversity and abundance were con-
ducted in only Ghana, and those from Southeast Asia were conducted in Malaysia
and Thailand. The findings of the studies conducted on the above subject matter in
the two regions have been mixed. In a Bornean forest in Malaysia, liana density was
found to be higher on soils with higher fertility and vice versa (DeWalt et al. 2006).
Additionally, some recent studies conducted in Malaysian forests revealed that a lot
of soil properties such as Mg, pH, P and organic matter are related with the abun-
dance and distribution of several individual liana species (Nurfazliza et al. 2012),
and soil P, pH and moisture related positively with total liana abundance, species
90 P. Addo-Fordjour and Z.B. Rahmad

richness and Shannon diversity (Addo-Fordjour et al. 2013c). Similar to the above-
mentioned patterns in Malaysia, a number of soil properties were identified as sig-
nificant correlates of liana diversity and abundance in some tropical forests in Ghana
(Addo-Fordjour et al. 2013b; Addo-Fordjour and Rahmad 2014). Nevertheless,
Lertpanich and Brockelman (2003) observed that soil properties were not signifi-
cant correlates of liana diversity and abundance in Khao Yai National Park, Thailand.
The above mentioned trends in Africa and Southeast Asia mirror those reported
from other parts of the tropics. Thus, on the whole, there is no universal explanation
about how soil properties affect liana assemblages in forests of Africa and Southeast
Asia, and the rest of the tropics.
Similarly, there is no general pattern with respect to the role of rainfall in liana
success in the tropics. Based on several studies conducted in the tropics, Schnitzer
et al. (2005) hypothesised that liana abundance decrease with increasing rainfall,
usually peaking in tropical dry forests, although Africa and Southeast Asia were not
included in the studies he considered. A pantropical study conducted using data sets
from tropical forests of several continents provided evidence to support the above
mentioned hypothesis. Nevertheless, van der Heijden and Phillips (2008) reported
of lack of significant correlation between rainfall and liana abundance in a tropical
forest in Amazonian Peru. Studies that examine the influence of rainfall on liana
diversity and abundance in Africa and Southeast Asia are scanty. In Africa, only one
study reported of a relationship between rainfall and liana abundance, which was
consistent with the hypothesis put forth by Schnitzer et al. (2005). Specifically, the
study indicated that liana abundance was higher in forests that received lower rain-
fall than those which had higher amounts of rainfall in Ghana (Swaine and Grace
2007). Unfortunately, no study has been conducted in Southeast Asia to examine
how rainfall influences liana diversity and abundance. Considering the fact that
Southeast Asia receives high amounts of rainfall, it will be very interesting to con-
duct studies that relate various liana assemblage attributes (species diversity, abun-
dance, basal area etc.) to rainfall.
Topography is an important factor which relates with plant assemblages although
it has no direct influence on them. Associated with topographic variables are a num-
ber of environmental factors such as soil properties, light intensity, temperature etc.
that covary along topographic gradients. For example, a number of soil properties
such as pH, moisture, and exchangeable calcium and magnesium are reported to
decrease with increasing elevational gradients in a number of studies (cf. Moser
et al. 2009). It is actually these environmental factors which directly influence plant
communities but not topography itself (Pausas and Austin 2001; Lookingbill and
Urban 2005). Therefore, topography in a way acts as a surrogate variable for other
environmental variables that covary along it (Pausas and Austin 2001). In this
regard, the presence of relationships between altitude and some attributes of liana
assemblages may be attributed to variations in environmental factors along altitudi-
nal gradients. Liana abundance decreased with increasing altitude in a South African
forest (Balfour and Bond 1993) and a Malaysian forest (Addo-Fordjour et al.
2013c). Additionally, liana diversity related negatively with altitude in the Malaysian
forest (Addo-Fordjour et al. 2013c). Nevertheless, altitude was not a significant cor-
6 Liana Assemblages in Tropical Forests of Africa and Southeast Asia… 91

relate of liana diversity and abundance in forests of Khao Yai National Park, Thailand
(Lertpanich and Brockelman 2003). Similarly, rattan abundance did not vary signifi-
cantly with altitude in a tropical forest in Indonesia although its diversity increased
with altitude in the same sites (Siebert 2005).

6.5 Management of Lianas in Africa and Southeast Asia

Lianas have great impact on tropical forest ecosystems in many different ways
(Tang et al. 2012). They are very useful in tropical forest ecosystems (Bongers et al.
2005; Muthumperumal and Parthasarathy 2013) even though they can also have
adverse effects in them (Parren and Bongers 2001; Bongers et al. 2005). Some for-
esters perceive lianas as nuisance because they are apparently unaware of the posi-
tive roles lianas play in tropical forests. In view of the possessing of dual functions
in forests, lianas are a “necessary evil” plant group which deserves management
attention. Consequently, a number of forest management programmes have inte-
grated liana control and management with the view to reducing their negative effects
in forests. This is achieved through application of silvicultural practices that tend to
reduce liana abundance in treated tropical forest ecosystems. However, there are
other forms of management interventions which have been recommended for imple-
mentation in tropical forest ecosystems with the aim of enhancing liana numbers in
low abundance liana forests (Parren and Bongers 2005). This may be particularly
ideal in areas where liana species of high ecological value are rare or endangered.
Nevertheless, the idea of promoting liana numbers through silvicultural activities
has not received much attention in the tropics. So far, only one study has assessed
the feasibility of artificial regeneration of lianas in a tropical forest (Le Bourlegat
et al. 2013). The study revealed that liana abundance and diversity can be improved
greatly through direct seeding of liana species.
Liana management in Africa and Southeast Asia has, over the years, involved
cutting of lianas either before or after logging. That is, pre-logging and post-logging
liana cutting, respectively. In tropical forests where liana abundance and infestation
are high, the value of trees as timber may be reduced as they suffer damages from
the lianas they carry (cf. Bongers et al. 2005). Lianas can also bind many trees
together making logging operations very difficult (Bongers et al. 2005), and causing
liana-laden trees to pull down other trees during logging. This phenomenon poses
great danger to loggers in such forests. To overcome the abovementioned challenges
and dangers associated with logging, pre-logging liana cutting is practiced in some
tropical forests in Africa and Southeast Asia (see Alvira et al. 2004; Addo-Fordjour
et al. 2009a). Although pre-logging liana cutting has been implemented in a number
of forests in Africa and Southeast Asia, only a limited number of studies have
assessed the effectiveness of this silvicultural operation in reducing or eliminating
the undesirable effects of lianas in forests. Generally, the findings of the studies
have been mixed. For instance, it was demonstrated in a tropical forest in Cameroon
that pre-logging liana cutting did not cause a significant reduction in damages
92 P. Addo-Fordjour and Z.B. Rahmad

associated with logging or gap sizes after logging (Parren and Bongers 2001).
On the other hand, climber cutting prior to logging resulted in considerable
reductions in damages associated with logging in a dipterocarp forest in Malaysia
(Appanah and Putz 1984). Besides the influence of pre-logging liana cutting on
logging damage and forest gap sizes, it also reduces liana abundance and thereby
reduces liana impact on trees (Bongers et al. 2005; Parren and Bongers 2005). This
also forms part of the objectives of pre-logging liana cutting in tropical forests.
Following logging, the forest canopy becomes open and provides favourable
conditions for lianas to proliferate. To avoid vigorous liana proliferation after log-
ging, some forest management systems incorporate cutting of lianas after logging
(i.e. post-logging liana cutting). Unlike pre-logging liana cutting, post-logging liana
cutting has received little attention in Africa and Southeast Asia. There is inadequate
information about the impact of post-logging liana cutting on tree assemblages in
Africa and Southeast Asia, and the whole of the tropics. In Nigeria, trees that were
cleared of lianas in post-logging liana cutting operation increased in diameter, and
tree sapling stocking also improved significantly (cf. Parren and Bongers 2005).
Additionally, there is not much understanding about how post-logging liana cutting
impacts liana assemblages in the two regions. A study demonstrated that liana spe-
cies richness and density were able to increase to pre-treatment levels 40 years after
post-logging liana cutting in a lowland dipterocarp forest in peninsular Malaysia
(cf. Alvira et al. 2004). Nevertheless, liana abundance and infestation were signifi-
cantly reduced in treated forests in Ghana 40 years after post-logging liana cutting
(Foli and Pinard 2009).

6.5.1 Forest Management Systems That Incorporate


Liana Cutting

In Africa and Southeast Asia, several forest management systems have been devel-
oped for managing various tropical forest ecosystems. Most of these systems incor-
porate liana cutting as part of their silvicultural treatments. Although silvicultural
treatment of forests occurs in many parts of Africa, it is more common in West
Africa. In many West African countries such as Ghana and Nigeria, three main types
of forest management systems involving liana cutting have been implemented to
manage forests in the past. These were Tropical Shelterwood System (TSS), Post
Exploitation System (PES) and Selection System (SS). The first silvicultural opera-
tion under the TSS was climber cutting which was done before logging so as to free
crowns of trees, thereby reducing felling damage during logging. Climber cutting
was repeated as the last operation during the 10th year following timber harvesting.
Other silvicultural activities were executed in between the pre-logging and post-
logging liana cutting periods as indicated as follows: opening of the canopy by
removing lower storey non-valuable and larger crowned understorey trees, cleaning
over a number of years, and timber exploitation. Majority of trees were harvested
from the TSS treated stands, leaving only trees of economic benefit (cf.
6 Liana Assemblages in Tropical Forests of Africa and Southeast Asia… 93

Addo-Fordjour et al. 2009a). In the PES the various silvicultural activities outlined
in the TSS were also carried out after logging. Furthermore, understorey and non-
economic trees were removed through cutting or poisoning or both. In the SS, single
matured trees or a small group of matured trees were harvested to liberate immature
trees. All climbers that were within a radius of 4 m of immature trees to be released
were cut before logging (cf. Addo-Fordjour et al. 2009a).
Southeast Asia has a long history of silvicultural methods that integrate liana
cutting. The Malayan Uniform System (MUS), one of the earliest management sys-
tems, was introduced in lowland dipterocarp forests in Malaysia as far back as 1948
(Appanah and Weinland 1993). The silvicultural activities of the MUS involved
cutting of all trees with diameter at breast height (dbh) ≤ 45 cm, poison girdling of
all defective and non-commercial trees, climber cutting, and enrichment planting.
Dipterocarp forests were also managed by another forest management system called
Selective Management System (SMS). This system involved felling of all commer-
cial trees with a dbh ≥ 45 cm for non-dipterocarps and ≥ 50 cm for dipterocarps
(Appanah and Weinland 1993; Appanah 1998; Hassan-Zaki et al. 2004). Pre-felling
and post-felling climber cutting were carried out under the SMS (Appanah 1999).
The MUS was a monocyclic system involving a single harvest whereas the SMS
was a bicyclic system that had two cycles of harvesting with 30 year interval
(Hassan-Zaki et al. 2004). In Indonesia, another type of forest management system
that involved post-logging liana cutting was applied to some tropical forests. The
management system which is referred to as Tebang Pilih Tanam Indonesia (TPTI)
consisted of felling of trees with dbh ≥ 60 cm and subsequent silvicultural treat-
ments. The silvicultural treatments comprised of clearing of all lianas and non-
commercial saplings from the understorey, with the view to improving natural
regeneration of trees and growth of timber species.

6.5.2 Impacts of Liana Cutting on Liana Assemblages


and Forest Biodiversity

Despite the potential of liana cutting in reducing liana abundance, and its conse-
quent impacts on trees and natural regeneration, the practice can have adverse
effects on the overall biodiversity in the forests. First of all, liana cutting reduces
liana species composition and diversity (Parren and Bongers 2005; Addo-Fordjour
et al. 2009a). Though generally many liana species have the capacity to withstand
disturbance thereby maintaining their diversity, a number of them show vulnerabil-
ity towards disturbance (Zagt et al. 2003). Some studies suggested that liana cutting
could reduce liana species composition and diversity in tropical forest ecosystems
(Parren and Bongers 2001; Addo-Fordjour et al. 2009a). This assertion was con-
firmed by recent studies which reported that liana cutting was responsible for a
significant reduction in liana species composition and diversity in tropical forests in
Malaysia and Ghana (Addo-Fordjour et al. 2014a, b). The adverse impact of liana
cutting on liana composition and diversity was apparent even after several years of
94 P. Addo-Fordjour and Z.B. Rahmad

recovery. As lianas form a significant portion of plant diversity in tropical forests


(Addo-Fordjour et al. 2008; Bongers et al. 2005), the overall plant biodiversity of
forests that undergo liana cutting may be affected. As mentioned earlier in this
chapter, lianas provide a source of livelihood for many organisms in tropical forest
ecosystems. For instance, many important dispersers of trees depend on lianas for
food, particularly in the dry season when most trees do not produce flowers and
fruits (cf. Bongers et al. 2005). To this end, lianas are not only key to the survival of
dispersers but also to the survival of trees in tropical forest ecosystems. Lianas pro-
vide a “corridor” through which arboreal animals move from one tree to another
(Putz 2012). These arboreal animals play important roles in dispersing the seeds of
trees on which they live (Clark and Poulsen 2001). From the few but critical impor-
tance of lianas outlined above, it is clear that liana cutting in forests has the potential
of adversely affecting the assemblages of other organisms, and the overall biodiver-
sity in forests.

6.6 Conclusion

The present study revealed that liana species contribute significantly to woody plant
species richness in Africa and Southeast Asia although the contribution of lianas to
woody plant species richness is higher in African tropical forests than Southeast
Asian tropical forests. Similarly, lianas are more abundant in African tropical for-
ests in relation to Southeast Asian tropical forests. However, Southeast Asian tropi-
cal forests exhibit higher rattan assemblages than African tropical forests probably
due to higher humidity and amount of rainfall in Southeast Asia. A number of fac-
tors including human disturbance, soil properties, topographic factors and climatic
factors were identified as the main correlates of liana diversity and abundance in the
two regions. The study also revealed that liana cutting is the main management tool
used in controlling lianas in Africa and Southeast Asia. Even though the main aim
of liana cutting is to reduce liana abundance and its impact on forest ecosystems, it
tends to have adverse impacts liana diversity even after several decades (five
decades) of recovery. Consequently, blanket liana cutting should be reconsidered;
selective liana cutting is recommended.

Acknowledgements We gratefully acknowledge the support of TWAS-USM Postgraduate


Fellowship and Research University Grant (RU) (1001/PBIOLOGI/815086).

References

Addo-Fordjour P, Rahmad ZB (unpublished data) Intercontinental comparison of liana assem-


blages with special reference to tropical forests in Ghana and Malaysia
Addo-Fordjour P, Rahmad ZB (2014) Environmental factors associated with liana community
assemblages in a tropical forest reserve, Ghana. J Trop Ecol. 31:69–79
6 Liana Assemblages in Tropical Forests of Africa and Southeast Asia… 95

Addo-Fordjour P, Anning AK, Atakora EA, Agyei PS (2008) Diversity and distribution of climbing
plants in a semi-deciduous rain forest, KNUST Botanic Garden, Ghana. Int J Bot 4:186–195
Addo-Fordjour P, Anning AK, Larbi JA, Akyeampong S (2009a) Liana species richness, abundance
and relationship with trees in the Bobiri forest reserve, Ghana: impact of management systems.
For Ecol Manage 157:1822–1828
Addo-Fordjour P, Obeng S, Addo MG, Akyeampong S (2009b) Effects of human disturbances and
plant invasion on liana community structure and relationship with trees in the Tinte Bepo forest
reserve, Ghana. For Ecol Manage 258:728–734
Addo-Fordjour P, Rahmad ZB, Shahrul AMS (2012) Effects of human disturbance on liana com-
munity diversity and structure in a tropical rainforest, Malaysia: implication for conservation.
J Plant Ecol 4:391–399
Addo-Fordjour P, Rahmad ZB, Amui J, Pinto C, Dwomoh M (2013a) Patterns of liana community
diversity and structure in a tropical rainforest reserve, Ghana: effects of human disturbance. Afr
J Ecol 51:217–227
Addo-Fordjour P, El Duah P, Agbesi DKK (2013b) Factors influencing liana species richness and
structure following anthropogenic disturbance in a tropical forest, Ghana. ISRN For 2013,
Article ID 920370:11. doi:10.1155/2013/920370
Addo-Fordjour P, Rahmad ZB, Shahrul AMS (2013c) Environmental factors influencing liana
community diversity, structure and habitat associations in a tropical hill forest, Malaysia. Plant
Ecol Divers. doi:10.1080/17550874.2013.782369
Addo-Fordjour P, Rahmad ZB, Shahrul AMS, Asyraf M (2014a) Impacts of forest management on
liana diversity and community structure in a tropical forest in Ghana: implications for conser-
vation. J For Res (Accepted)
Addo-Fordjour P, Rahmad ZB, Shahrul AMS (2014b) Impacts of forest management on commu-
nity assemblage and carbon stock of lianas in a tropical lowland forest, Malaysia. Trop Conserv
Sci 7:244–259
Alvira D, Putz FE, Fredericksen TS (2004) Liana loads and post-logging liana densities after liana
cutting in a lowland forest in Bolivia. For Ecol Manage 190:73–86
Anbarashan M, Parthasrathy N (2013) Diversity and ecology of lianas in tropical dry evergreen
forests on the Coromandel Coast of India under various disturbance regimes. Flora Morphol
Distrib Funct Ecol Plant 208:22–32
Appanah S (1998) Management of natural forests. In: Appanah S, Turnbull JM (eds) A review of
dipterocarps, taxonomy, ecology and silviculture. CIFOR, Bogor, pp 133–149
Appanah S (1999) Trends and issues in tropical forest management: setting the agenda for
Malaysia. FRIM, Kepong
Appanah S, Putz FE (1984) Climber abundance in virgin dipterocarp forest and the effect of
prefelling climber cutting on logging damage. Malays Forester 47:335–342
Appanah S, Weinland G (1993) Will the management system for hill dipterocarp forests, stand up?
J Trop For Sci 3:140–158
Appanah S, Gentry AH, LaFrankie JV (1993) Liana diversity and species richness of Malaysian
rain forests. J Trop For Sci 6:116–123
Baars R, Kelly D, Sparrow AS (1998) Liane distribution within native forest remnants in two
regions of the South Island, New Zealand. N Z J Ecol 22:71–85
Balfour DA, Bond WJ (1993) Factors limiting climber distribution and abundance in a southern
African forest. J Ecol 81:91–100
Bongers F, Schnitzer SA, Traore R (2002) The importance of lianas and consequences for forest
management in West Africa. Bioterre (Special Issue):59–70
Bongers F, Parren MPE, Swaine MD, Traoré D (2005) Forest climbing plants of West Africa:
introduction. In: Bongers F, Parren MPE, Traore´ D (eds) Forest climbing plants of West
Africa: diversity, ecology and management. CAB International, Wallingford, pp 5–18
Boonsermsuk S, Pattanavibool R, Sombun K (2007) Rattan in Thailand, 1st edn. Royal Forest
Department and ITTO Aksornsiam Printing, Bangkok
Caballé G, Martin A (2001) Thirteen years of change in trees and lianas in a Gabonese rainforest.
Plant Ecol 152:167–173
96 P. Addo-Fordjour and Z.B. Rahmad

Chittibabu CV, Parthasarathy N (2001) Liana diversity and host relationships in a tropical ever-
green forest in the Indian Eastern Ghats. Ecol Res 16:519–529
Clark CJ, Poulsen JR (2001) The role of arboreal seed dispersal groups on the seed rain of a
lowland tropical forest. Biotropica 33:606–620
DeWalt SJ, Schnitzer SA, Denslow JS (2000) Density and diversity of lianas along a chronose-
quence in a central Panamanian lowland forest. J Trop Ecol 16:1–19
DeWalt SJ, Ickes K, Nilus R, Harms KE, Burslem DFRP (2006) Liana habitat associations and
community structure in a Bornean lowland tropical forest. Plant Ecol 186:203–216
DeWalt SJ, Schnitzer SA, Chave J, Bongers F, Burnham RJ, Cai Z, Chuyong G, Clark DB, Ewango
CEN, Gerwing JJ, Gortaire E, Hart T, Ibarra-Manríquez G, Ickes K, Kenfacks D, Macía MJ,
Makana J-R, Martínez-Ramos M, Mascaro J, Moses S, Muller-Landau HC, Parren MPE,
Parthasarathy N, Pérez-Salicrup D, Putz FE, Romero-Saltos H, Thomas D (2010) Annual rain-
fall and seasonality predict pantropical patterns of liana density and basal area. Biotropica
42:309–317
Emmons LH, Gentry AH (1983) Tropical forest structure and the distribution of gliding and
prehensile-tailed vertebrates. Am Nat 121:513–523
Ewango CEN (2010) The liana assemblage of a Congolian rainforest: diversity, structure and
dynamics. PhD thesis, Wageningen University, Wageningen
Foli EG, Pinard MA (2009) Liana distribution and abundance in moist tropical forest in Ghana 40
years following silvicultural interventions. Ghana J For 25:1–12
Gentry AH (1991) The distribution and evolution of climbing plants. In: Putz FE, Mooney HA
(eds) The biology of vines. Cambridge University Press, Cambridge, pp 3–49
Ghollasimood S, Hanum IF, Nazre M, Kamziah AK, Awang Noor AG (2011) Vascular plant com-
position and diversity of a coastal hill forest in Perak, Malaysia. J Agric Sci 3:111–126
Hassan-Zaki P, Shinohara T, Nakama Y, Yukutake K (2004) A Selective Management System
(SMS): a case study in the implementation of SMS in managing the dipterocarp forests of
Peninsular Malaysia. Kyushu J For Res 57:39–44
Hawthorne WD, Jongkind C (2006) Woody plants of western African forests: a guide to the forest
trees, shrubs and lianes from Senegal to Ghana. Royal Botanic Gardens, Kew
Homeier J, Englert F, Leuschner C, Weigelt P, Unger M (2010) Factors controlling the abundance
of lianas along an altitudinal transect of tropical forests in Ecuador. For Ecol Manage
259:1399–1405
Ingwell LL, Wright SJ, Becklund KK, Hubbell SP, Schnitzer SA (2010) The impact of lianas on
10 years of tree growth and mortality on Barro Colorado Island, Panama. J Ecol 98:879–887
Jongkind CCH (2005) Checklist of climber species in Upper Guinea. In: Bongers F, Parren MPE,
Traoré D (eds) Forest climbing plants of West Africa: diversity, ecology and management.
CABI Publishing, Wallingford, pp 231–264
Kouame NF (1998) Influence de l’exploitation forestière sur la végétation et la flore de la Forêt
Classée du Haut-Sassandra (Centre-Ouest de la Côte d’Ivoire). Thèse de Doctorat 3ème cycle,
Université de Cocody, Abidjan, 227pp
Kuzee ME (2002) Forest recovery and lianas. PhD thesis, Wageningen University, Wageningen
Langenberger G, Martin K, Sauerborn J (2006) Vascular plant species inventory of a Philippine
lowland rain forest and its conservation value. Biodivers Conserv 15:1271–1301
Laurance WF, Pérez-Salicrup D, Delamônica P, Fearnside PM, D’Angelo S, Jerozolinski A, Pohl
L, Lovejoy TE (2001) Rain forest fragmentation and the structure of Amazonian liana com-
munities. Ecology 82:105–116
Laurance WF, Andrade AS, Magrach A, Camargo JL, Valsko JJ, Campbell M, Fearnside PM,
Edwards W, Lovejoy TE, Laurance SG (2014) Long-term changes in liana abundance and
forest dynamics in undisturbed Amazonian forests. Ecology 96:1604–1611
Le Bourlegat JMG, Gandolif S, Brancalion PHS, Dias TDS (2013) Diversity enhancement of a
forest under restoration through direct seeding of lianas. Hoehnea 40:465–472
Ledo A, Schnitzer SA (2014) Disturbance and clonal reproduction determine liana distribution and
maintain liana diversity in a tropical forest. Ecology 95:2169–2178
6 Liana Assemblages in Tropical Forests of Africa and Southeast Asia… 97

Lertpanich K, Brockelman WY (2003) Lianas and environmental factors in the Mo Singto biodi-
versity research plot, Khao Yai National Park, Thailand. Nat Hist J Chulalongkorn Univ
3:7–17
Lookingbill TR, Urban DL (2005) Gradient analysis, the next generation: towards more
plant-relevant explanatory variables. Can J Forest Res 35:1744–1753
Lü XT, Tang JW, Feng XL, Li MH (2009) Diversity and aboveground biomass of lianas in the
tropical seasonal rain forests of Xishuangbanna, SW China. Rev Biol Trop 57:211–222
Moser KF, Ahn C, Noe GB (2009) The influence of microtopography on soil nutrients in created
mitigation wetlands. Restor Ecol 17:641–651
Muthumperumal C, Parthasarathy N (2013) Diversity, distribution and resource values of woody
climbers in tropical forests of southern Eastern Ghats, India. J For Res 24:365–374
Nabe-Nielsen J (2001) Diversity and distribution of lianas in a neotropical rain forest, Yasuní
National Park, Ecuador. J Trop Ecol 17:1–19
Nurfazliza K, Nizam MS, Supardi MNN (2012) Association of liana communities with their soil
properties in a lowland forest of Negeri Sembilan, Peninsular Malaysia. Sains Malays
41:679–690
Parren MPE (2003) Lianas and logging in West Africa, Tropenbos-Cameroon series 6. Tropenbos
International, Wageningen
Parren M, Bongers F (2001) Does climber cutting reduce felling damage in southern Cameroon?
For Ecol Manage 141:175–188
Parren MPE, Bongers F (2005) Management of climbers in the forest of West Africa. In: Bongers
F, Parren MPE, Traoré D (eds) Forest climbing plants of West Africa: diversity, ecology and
management. CAB International, Wallingford, pp 217–229
Parthasarathy N, Muthuramkumar S, Reddy MS (2004) Patterns of liana diversity in tropical ever-
green forests of peninsular India. For Ecol Manage 190:15–31
Pausas JG, Austin MK (2001) Patterns of plant species richness in relation to different environ-
ments: an appraisal. J Veg Sci 12:153–166
Pinard MA, Putz FE (1994) Vine infestation of large remnant trees in logged forest in Sabah,
Malaysia: biomechemical facilitation in vine succession. J Trop For Sci 6:302–309
Putz FE (1984) The natural history of lianas on Barro Colorado Island, Panama. Ecology
65:1713–1724
Putz FE (2012) Vine ecology. Ecol Inf 24
Putz FE, Chai P (1987) Ecological studies of lianas in Lambir National Park, Sarawak, Malaysia.
J Ecol 75:523–531
Rahman MM, Begum F, Nishat A, Islam KK, Vacik H (2010) Species richness of climbers in natu-
ral and successional stands of Madhupur Sal (Shorea robusta C.F. Gaertn) forest, Bangladesh.
Trop Subtrop Agroecosyst 12:117–122
Rutishauser SE (2011) Increasing liana abundance and biomass in tropical forests: testing mecha-
nistic explanations. MSc. thesis, University of Wisconsin – Milwaukee, Milwaukee
Schnitzer SA (2005) A mechanistic explanation for global patterns of liana abundance and distri-
bution. Am Nat 166:262–276
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
Schnitzer SA, Bongers F (2011) Increasing liana abundance and biomass in tropical forests:
emerging patterns and putative mechanisms. Ecol Lett 14:397–406
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical
forest canopy gaps: evidence for an alternative pathway of gap-phase regeneration. J Ecol
88:655–666
Schnitzer SA, Parren MPE, Bongers F (2004) Recruitment of lianas into logging gaps and the
effects of pre-harvest climber cutting in a lowland forest in Cameroon. For Ecol Manage
190:87–98
Schnitzer SA, Kuzee M, Bongers F (2005) Disentangling above- and below-ground competition
between lianas and trees in a tropical forest. J Ecol 93:1115–1125
98 P. Addo-Fordjour and Z.B. Rahmad

Senbeta F, Schmitt C, Denich M, Demissew S, Vlek PLG, Preisinger H, Woldemariam T, Teketay


D (2005) The diversity and distribution of lianas in the Afromontane rainforests of Ethiopia.
Divers Distrib 11:443–452
Siebert SF (2005) The abundance and distribution of rattan over an elevation gradient in Sulawesi,
Indonesia. For Ecol Manage 210:143–158
Swaine MD, Grace J (2007) Lianas may be favoured by low rainfall: evidence from Ghana. Plant
Ecol 192:271–276
Swaine MD, Hawthorne WD, Bongers F, Toledo-Aceves M (2005) Climbing plants in Ghanaian
forests. In: Bongers F, EParren MPE, Traore´ D (eds) Forest climbing plants of West Africa:
diversity, ecology and management. CAB International, Wallingford, pp 93–108
Tang Y, Kitching RL, Cao M (2012) Lianas as structural parasites: a re-evaluation. Chin Sci Bull
57:307–312
Tobin MF, Wright AJ, Mangan SA, Schnitzer SA (2012) Lianas have a greater competitive effect
than trees of similar biomass on tropical canopy trees. Ecosphere 3:1–11
Toledo-Aceves T, Swaine MD (2008) Above- and below-ground competition between the liana
Acacia kamerunensis and tree seedlings in contrasting light environments. Plant Ecol
196:233–244
Van der Heijden GMF, Phillips OL (2008) What controls liana success in Neotropical forests?
Glob Ecol Biogeogr 17:372–383
Yorke SR, Schnitzer SA, Mascaro J, Letcher SG, Carson WP (2013) Increasing liana abundance
and basal area in a tropical forest: the contribution of long-distance clonal colonization.
Biotropica 45:317–324
Zagt R, Ek R, Raes R (2003) Logging effects on liana diversity and abundance in Central Guyana.
Tropenbos-Guyana reports 2003-1, Tropenbos International, Wageningen
Chapter 7
Diversity of Lianas in Eastern Himalayas
and North-Eastern India

S.K. Barik, D. Adhikari, A. Chettri, and P.P. Singh

Abstract Lianas constitute an important component of plant diversity in a wide


range of ecosystems. The liana diversity has been reported to have critical role in
maintaining ecosystem structure and function, with a predicted greater role to play
under changing climate conditions. However, our understanding on the diversity
pattern of lianas in different forest ecosystems along an elevation gradient is limited.
Studies on lianas in the Eastern Himalayas and North-eastern region of India are
very few. In this paper, we review the status of liana research in the two biodiversity
hotspots viz., the Himalayas and Indo-Myanmar, and conducted primary studies in
selected high diversity forests of the region to understand the patterns of liana diver-
sity along elevation gradient in the Himalayas. The diversity and distribution of
lianas in three major forest types representing the dominant ecological zones of the
region viz. tropical, montane-subtropical and temperate were studied using liana
census protocol of Gerwing et al. (2006). We recorded 196 liana species in three
forest types in the states of Arunachal Pradesh, Sikkim and Meghalaya in north-
eastern India through plot-based field survey. The diversity and abundance of lianas
were greater in sub-tropical and tropical forests than temperate forests. Threat
assessment for liana conservation was suggested for their in situ conservation.

7.1 Introduction

Lianas are woody vines rooted to the ground, while their stems use trees as a vertical
support for climbing and getting access to well-lit space in the forest canopy
(Schnitzer and Bongers 2002; Bongers et al. 2002). They use a variety of climbing
mechanisms to attach themselves to the host trees, which include twining around

S.K. Barik (*) • D. Adhikari • P.P. Singh


Department of Botany, Centre for Advanced Studies in Botany, North-Eastern
Hill University, Shillong 793022, India
e-mail: sarojkbarik@gmail.com
A. Chettri
Department of Botany, Sikkim University, Gangtok 737102, Sikkim, India

© Springer International Publishing Switzerland 2015 99


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_7
100 S.K. Barik et al.

the host stem, using clasping tendrils, and attaching to the host plant through thorns,
spines, adhesive hairs and adventitious roots. Lianas form an important component
of plant diversity in a wide range of ecosystems. They are reported to have critical
role in maintaining ecosystem structure and function. For example, Gentry (1991)
and Schnitzer et al. (2000) have demonstrated that lianas are important contributors
to tropical forest species diversity, structure, and dynamics. They play an important
role in gap-phase regeneration of forests, vitalize the ecosystem-level processes
such as transpiration and carbon sequestration, and are predicted to have a greater
role in forest dynamics in future under changing environmental and climatic condi-
tions (Schnitzer and Bongers 2002).
Identifying the patterns of liana abundance and distribution in different forests
and understanding the underlying causes have been a subject of intensive research
during the past two decades (Schnitzer and Bongers 2002; Schnitzer 2005). The
important generalizations on liana distribution that have emerged from such research
are: (i) the diversity and abundance of lianas peak in the tropical region, and decrease
with increasing latitude, (ii) liana abundance increases with increasing seasonality
and decreasing rainfall, (iii) liana abundance and diversity increase with soil fertility
and disturbance (DeWalt et al. 2006), and (iv) lianas become less diverse as the age
of the forest increases. These generalizations are based on several site-specific stud-
ies followed by meta-analyses (Schnitzer 2005). Most of these studies are from
tropical forests (DeWalt et al. 2000; Muthuramkumar and Parthasarathy 2000;
Chittibabu and Parthasarathy 2001; Nabe-Nielsen 2001; Parthasarathy et al. 2004;
Senbeta et al. 2005; Addo-fordjour et al. 2008; Cai et al. 2009; Yuan et al. 2009;
Muthumperumal and Parthasarathy 2010; Schnitzer et al. 2012; Anbarashan and
Parthasarathy 2013; Naidu et al. 2014) with very few from subtropical (Hegarty
1991; Rice et al. 2004; Malizia and Grau 2006; Campanello et al. 2007; Yuan et al.
2009) and temperate forests (Londré and Schnitzer 2006; Allen et al. 2007; Ichihashi
et al. 2009; Leicht-Young et al. 2010; Chettri et al. 2010). Such generalization therefore
needs validation with more studies from subtropical and temperate forest ecosystems.
Indian Himalayas and north-eastern India, the two important biodiversity
hotspots, with wide elevation range provide an ideal environment to study the diver-
sity and abundance patterns of lianas from tropical to montane subtropical and tem-
perate forest ecosystem. The present study was undertaken to test the hypothesis
that liana diversity and abundance decrease with increasing elevation in a mountain
ecosystem, which is parallel with the global latitudinal gradient. One of our earlier
studies on liana diversity and distribution was confined to Sikkim Himalaya only
(Chettri et al. 2010), without having tropical forest component. The present study
reports for the first time the liana diversity and abundance in tropical, montane sub-
tropical and temperate forests (Champion and Seth 1968) of three north-eastern
states including the Eastern Himalayas.
Incidentally, studies on lianas of Himalayas and Indo-Myanmar biodiversity
hotspots are very few. Chettri et al. (2010) studied 43 liana species from montane,
lower montane and upper montane forests of Sikkim and correlated various environ-
mental factors with liana density and diversity. Other studies on lianas are from the
state of Tripura where the taxonomic accounts of four species viz. Amphineurion
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India 101

marginatum (Roxb.) D.J. Midleton, Bidaria inodora (Loureiro) Decaisne,


Combretum acuminatum Roxb., and Derris feruginea (Roxb.) Bentham was given
by Darlong and Bhattacharyya (2014) and they also highlighted the importance of
liana conservation in Tripura (Darlong and Bhattacharyya 2012).

7.2 Study Area and Methods

The diversity and distribution of lianas in three major forest types representing the
dominant ecological zones of north-eastern India viz. tropical, montane subtropical,
and temperate forests were studied in Sikkim (Sikkim Himalaya), Arunachal Pradesh
(Eastern Himalaya), and Meghalaya (Sub-Himalayan north-eastern state). The stan-
dard protocol for liana census (Gerwing et al. 2006) was followed for the study.
The study was conducted at 20 localities/sites in the three states, 6 sites being in
tropical forest and 7 each in montane subtropical and temperate forests. At each site,
depending on the terrain, accessibility and homogeneity in liana species composi-
tion, an area of 15 ha or 30 ha was demarcated for detailed study. The randomization
was achieved by dividing the demarcated forest area at each site (15 ha or 30 ha)
into 100 × 100 m grids, and drawing 144 computer generated random numbers for
identifying the grid where sampling of lianas was to be made. In each identified
grid, one plot of 25 × 25 m plot was laid. Thus, in total, 144 plots of 25 × 25 m2
dimension were laid randomly in the tropical, subtropical, and temperate forests in
the three states (Table 7.1 and Fig. 7.1).

Table 7.1 Details of the sampling area and sample size in forests of the Himalayas and North-east
India
Elevation Total forest area Number of plots
Forest type range (m) State Locality demarcated (ha) (25 × 25 m) studied
Tropical 0–900 Arunachal Tippi, 30 16
Pradesh Bhalukpong
Sikkim Melli, Kitam 60 16
Meghalaya Nongpoh/ 60 16
Nongkhyllem,
Shella
Montane 900– Arunachal Nechipu, Lumla 60 16
subtropical 1,800 Pradesh
Sikkim Mangan 30 16
Meghalaya Cherrapunjee, 60 16
Umiam, Jowai,
Pynnurshla
Temperate 1,800– Arunachal Bomdila, Jung, 60 16
4,000 Pradesh Thimbu, Mago
Sikkim Yuksom, Tshoka 60 16
Meghalaya Upper Shillong 30 16
102 S.K. Barik et al.

Fig. 7.1 Location of liana sample plots in different climatic zones and forest types in Arunachal
Pradesh, Meghalaya and Sikkim of north-eastern India

The liana diversity and abundance were enumerated in the following ways:
(i) voucher specimens were collected and the individuals were identified to species
level using local and regional floras (Hooker 1872–1897; Kanjilal et al. 1934–1940;
Balakrishnan 1981–1983; Joseph 1982; Haridasan and Rao 1985–1987; Hajra et al.
1996), (ii) the number of individuals under each species was counted, and (iii) the
girth of each individual (≥1 cm) was measured at 130 cm from the rooting point of
the stem. Herbarium was prepared for each species following standard methods.
Further, the liana species were categorized into: (i) true secondary growth (wood or
persistent), (ii) fibrous sub-woody, and (iii) herbaceous climber, based on the
pattern of secondary growth. The lianas were also classified based on climbing
mechanisms i.e. scramblers, twiners, root climbers, tendril climbers, and hook
climbers. Threatened liana species were identified based on the IUCN Red list, Indian
Red Data Book (Nayar and Sastry 1987–1990), and Convention on International
Trade in Endangered Species (CITES).
Plot-based numerical data were pooled to study the diversity and dominance
patterns in different forest types and states. Diversity indices i.e. Shannon’s index,
Simpson’s dominance index, and evenness index for lianas in each forest type
were computed. Shannon’s diversity index was compared between different forest
types and between different states using the bootstrap-pairwise comparison test
using PAST software (Hammer et al. 2001).
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India 103

7.3 Results

7.3.1 Liana Diversity in the Tropical, Subtropical


and Temperate Forests

The liana census resulted in 3,586 individuals belonging to 196 species and 55
families (Table 7.2). These were recorded from the three forest types spread over 20
localities in three states. Fabaceae (22), Apocynaceae (13), and Vitaceae (12) were
the most species-rich families. The liana species richness was highest in the
subtropical forests (104), followed by tropical (74) and temperate (45) forests.
The richness of liana families also followed the same trend.
Overall, 89 % of the total 196 liana species were woody, 3 % belonged to fibrous
category, and the remaining 8 % were herbaceous. The corresponding figures for the
tropical (Woody – 63, Fibrous – 2, Herbaceous – 9), subtropical (Woody – 91,
Fibrous-6, Herbaceous-7) and temperate (Woody – 41, Fibrous – 1, Herbaceous – 3)
forests also showed a similar pattern with dominance of woody species as compared
to the fibrous and herbaceous species (Table 7.2).
Five major climbing strategies for lianas were identified viz. hook climber, root
climber, scrambler, tendril climber, and twiner. Overall, the twiners dominated the
liana community, followed by tendril climbers, scramblers, hook climbers, and
adventitious root climbers. The proportion of twiners decreased with increase in
elevation. However, the relative proportion of scramblers increased with elevation.
The remaining types did not exhibit any definite trend (Fig. 7.2).
The liana species richness ranged between 24 and 63 in the subtropical forests,
while in the tropical and temperate forests it ranged between 19–48 and 12–28,
respectively. The liana density also followed the same trend. The density range of
lianas was highest in the subtropical forests (321–784 ha−1), followed by tropical
(220–460 ha−1) and temperate (241–386 ha−1) forests. The subtropical forests had
highest values for Shannon diversity index, Simpson’s dominance index and
evenness index, which was followed by the tropical and temperate forests. The
Shannon’s diversity index for subtropical forests ranged from 3.08 to 4, and for the
tropical and temperate forests it ranged from 2.8 to 3.71, and 2.3–3.18, respectively.
The Simpson’s dominance index ranged from 0.95 to 0.97 in the subtropical forests,
0.93–0.97 in the tropical forests, and 0.89–0.95 in the temperate forests. Species
evenness index ranged from 0.87 to 0.91 in the subtropical forests, 0.85–0.91 in the
tropical forests, and 0.83–0.89 in the temperate forests (Table 7.3).
Bootstrap-pairwise comparison test revealed that, overall, the Shannon diversity
values significantly differed between tropical, subtropical and temperate forests.
Comparison within similar forest types between different states also yielded a sig-
nificant difference (P < 0.001) in the diversity values. However, comparison of the
Shannon diversity values of the temperate forests in Arunachal Pradesh and Sikkim
did not show a significant difference (Table 7.4).
Exclusive record of a species from a particular climatic zone indicates its
restricted distribution, while occurrence in two or more zones indicates its wide
Table 7.2 List of liana species in Arunachal Pradesh, Meghalaya and Sikkim states of north-eastern India
104

Climbing Forest Elevation


Sl. No. Botanical names Family Category mechanism types (m) States
1 Abrus precatorius L. Fabaceae WV TW Tr 700 Sk
2 Acacia caesia (L.) Willd. Mimosaceae WV HC Tr 900 Meg
3 Acacia concinna DC. Mimosaceae WV SC Str 1,000 Sk
4 Acacia pennata Willd. Mimosaceae WV SC Tr 900 Ar, Sk
5 Actinidia callosa Lindl. Actinidiaceae WV TW Str, Tem 1,700–2,200 Ar,
Meg, Sk
6 Actinidia strigosa Hk.f. & T. Actinidiaceae WV TW Tem 3,100 Sk
7 Adenia trilobata (Roxb.) Engl. Passifloraceae WV TC Tr 900 Meg
8 Ampelocissus barbata (Wall.) Planch. Vitaceae WV TC Tr 900 Meg
9 Ampelocissus latifolia (Roxb.)Planch. Vitaceae WV TC Str 1,200 Meg
10 Ampelocissus sikkimensis Planch. Vitaceae WV TC Str 1,500 Sk
11 Anamirta cocculus (L.) Wight & Arn. Menispermaceae WV TW Str 1,500 Meg
12 Argyreia capitata (Vahl) Choisy Convovulaceae WV TW Str 1,000 Meg
13 Argyreia hookeri C.B.Clarke Convovulaceae WV TW Tr 900 Meg
14 Argyreia nervosa (Burm. f.) Bojer Convovulaceae WV TW Tr 900 Meg
15 Argyreia roxburghii (Wall.) Arn. ex Choisy Convovulaceae WV TW Tr 900 Meg
16 Argyreia wallichii Choisy Convolvulaceae WV TW Str 1,100 Sk
17 Aristolochia cathcartii Hook.f. Aristolochiaceae WV TW Tr 900 Meg
18 Aristolochia griffithii Hook. f. & Thoms. ex Duch. Aristolochiaceae WV TW Str 1,800 Ar, Sk
19 Aristolochia tagala Cham. Aristolochiaceae WV TW Tr 900 Meg
20 Artabotrys caudatus Wall. ex Hook.f. & Thomson Annonaceae WV HC Tr 900 Meg
21 Asparagus racemosus Willd Asparagaceae WV SC Str, Tem 1,200–2,000 Ar,
Meg, Sk
22 Aspidopterys indica (Willd.) W.Theob. Malpighiaceae WV TW Str 1,500 Ar
23 Bauhinia khasiana Baker Caesalpinaceae WV TW Tr 900 Meg
S.K. Barik et al.
Climbing Forest Elevation
Sl. No. Botanical names Family Category mechanism types (m) States
24 Bauhinia nervosa (Wall. ex Benth.) Bake Caesalpinaceae WV TW Tr 900 Meg
25 Bauhinia scandens L. Caesalpinaceae WV TW Tr 900 Meg
26 Bauhinia vahlii W. & A. Caesalpinaceae WV TW Tr, Str 900–1,000 Ar,
Meg, Sk
27 Beaumontia grandiflora Wall. Apocynaceae WV TW Str 1,500 Sk
28 Bridelia stipularis (L.) Blume Phyllanthaceae WV TC Tr 900 Meg
29 Bridelia tomentosa Blume Phyllanthaceae WV TC Tr 900 Meg
30 Butea sp. Fabaceae WV TW Tr 900 Meg
31 Byttneria grandifolia DC. Sterculiaceae WV HC Tr 900 Meg
32 Byttneria pilosa Roxb. Sterculiaceae WV HC Str 1,200 Meg
33 Caesalpinia cucullata Roxb. Fabaceae WV SC Tr 600 Sk
34 Calamus acanthospathus Griff. Arecaceae RA SC Str 1,800 Ar, Sk
35 Calamus flagellum Griff. Arecaceae RA SC Tr, Str 900–1,000 Ar, Sk
36 Calamus guruba Ham. Arecaceae RA SC Tr 800 Sk
37 Calamus inermis Anderson Arecaceae RA SC Str 1,100 Sk
38 Calamus leptospadix Griff. Arecaceae RA SC Str 1,000 Sk
39 Capparis multiflora Hk.f. & T. Capparidaceae WV SC Str 1,900 Sk
40 Cayratia pedata Gagnep. Vitaceae WV TC Str 1,200 Meg
41 Celastrus paniculatus Willd. Celastraceae WV SC Str 1,600–2,300 Ar, Sk
42 Celastrus stylosus Wall. Celastraceae WV SC Tem 2,300 Ar, Sk
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India

43 Chonemorpha macrophylla G.Don Apocynaceae WV TW Str 1,800 Ar, Sk


44 Chonemorpha fragrans (Moon) Alston Apocynaceae WV TW Str 1,500 Meg
45 Cissampelos pareira Linn. Menispermaceae HV TW Tr, Str 800–1,100 Ar,
Meg, Sk
46 Cissus javana DC. Vitaceae HV TC Tr, Str 900–1,000 Meg, Sk
47 Cissus repens Lamk. Vitaceae WV TC Str 1,500–1,600 Ar, Sk
48 Cissus spectabilis Planch. Vitaceae WV TC Tr 600 Ar, Sk
(continued)
105
Table 7.2 (continued)
106

Climbing Forest Elevation


Sl. No. Botanical names Family Category mechanism types (m) States
49 Clematis acuminata DC. Ranunculaceae WV TC Tem 2,700 Ar, Sk
50 Clematis buchananiana DC. Ranunculaceae WV TC Str, Tem 1,500–2,900 Ar,
Meg, Sk
51 Clematis connata DC. Ranunculaceae WV TC Str 1,500 Ar
52 Clematis gouriana Roxb. ex DC. Ranunculaceae WV TC Str 1,500 Ar
53 Clematis montana Ham. Ranunculaceae WV TC Tem 2,900–3,000 Ar, Sk
54 Combretum acuminatum Roxb. Combretaceae WV TW Tr 900 Meg
55 Combretum latifolium Blume Combretaceae WV TW Tr 900 Meg
56 Combretum roxburghii Spreng. Combretaceae WV TW Tr 800 Ar, Sk
57 Connarus paniculatus Roxb. Connaraceae WV TW Str 1,200 Meg
58 Cryptolepis buchananii Roem. & Schult. Apocynaceae WV TW Tr 900 Meg
59 Cryptolepis sinensis (Lour.) Merr. Apocynaceae WV TW Str 1,200 Meg
60 Cyclea bicristata (Griff.) Diels. Menispermaceae WV TW Str 1,500 Meg
61 Cynanchum wallichii Wight Apocynaceae HV TW Tr 900 Meg
62 Dalbergia mimosoides Franch. Fabaceae WV SC Tr 900 Ar, Sk
63 Derris marginata (Roxb.) Benth. Fabaceae WV HC Tr 900 Meg
64 Derris trifoliata Lour. Fabaceae WV HC Tr 900 Meg
65 Desmos dumosus (Roxb.)Safford Annonaceae WV TW Tr 700 Meg
66 Dioscorea alata L. Dioscroeaceae WV TW Str 1,000 Meg
67 Dioscorea bulbifera L. Dioscoreaceae WV TW Str, Tem 1,500–2,000 Ar,
Meg, Sk
68 Dioscorea deltoidea Wall. ex Griseb. Dioscroeaceae WV TW Str, Tem 1,400–1,900 Ar,
Meg, Sk
69 Dioscorea oppositifolia L. Dioscroeaceae WV TW Tr, Str 900–1,000 Ar, Meg
70 Dioscorea pentaphylla L. Dioscroeaceae WV TW Str 1,200 Ar, Meg
S.K. Barik et al.
Climbing Forest Elevation
Sl. No. Botanical names Family Category mechanism types (m) States
71 Dumasia villosa DC. Fabaceae WV TW Str 1,200 Meg
72 Elaeagnus latifolia L. Elaegnaceae WV SC Tem 1,900 Meg
73 Elaeagnus pyriformis Hook.f. Elaegnaceae WV SC Str 1,600 Meg
74 Embelia floribunda Wall. Myrsinaceae WV SC Str 1,500 Ar, Meg
75 Embelia ribes Burm. Myrsinaceae WV SC Str 1,700–1,800 Ar,
Meg, Sk
76 Embelia subcoriacea (C.B.Clarke) Mez Myrsinaceae WV SC Tem 1,900 Meg
77 Entada phaseoloides (L.) Merr. Fabaceae WV TC Str 1,400 Ar, Meg
78 Entada pursaetha DC Fabaceae WV TC Tr 800 Meg
79 Entada rheedei Spreng. Fabaceae WV TC Str 1,000 Meg
80 Entada scandens Benth. Fabaceae WV TC Tr, Str 800–1,100 Ar,
Meg, Sk
81 Erycibe paniculata Roxb. Convolvulaceae WV TW Tr 800 Meg
82 Erythropalum scandens Blume. Erythropalaceae WV TW Tr 800 Meg
83 Fissistigma rubiginosa (A. DC.)Merr Annonaceae WV TW Tr 800 Meg
84 Fissistigma verrucosum (Hook.f. & Thomson) Merr. Annonaceae WV TW Str 1,600 Meg
85 Gnetum montanum Markgr. Gnetaceae WV TW Str 1,500 Sk
86 Gouania leptostachya DC. Rhamnaceae WV TC Tr 700 Sk
87 Hedera helix L. Araliaceae WV RC Str, Tem 1,400–1,900 Ar, Meg
88 Hedera nepalensis K.Koch Araliaceae WV RC Str, Tem 1,600–2,500 Ar,
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India

Meg, Sk
89 Herpetospermum pedunculosum (Ser.) C.B. Clarke. Cucurbitaceae WV TC Tem 2,500 Ar
90 Heterostemma alatum Wight Asclepiadaceae WV TW Str 1,500 Sk
91 Hippocratea arborea Roxb. Celastraceae WV TW Tr 800 Sk
92 Hiptage benghalensis (L.) Kurz Malpighiaceae WV TW Tr, Str 900–1,200 Ar, Sk
93 Hodgsonia heteroclita (Roxb.) Hk.f. & T. Cucurbitaceae WV TC Tr, Str 900–1,100 Ar, Sk
94 Hodgsonia macrocarpa (Blume) Cogn. Cucurbitaceae WV TC Str 1,000 Meg
(continued)
107
Table 7.2 (continued)
108

Climbing Forest Elevation


Sl. No. Botanical names Family Category mechanism types (m) States
95 Holboellia latifolia Wall. Lardizabalaceae WV TW Str, Tem 1,600–3,100 Ar,
Meg, Sk
96 Holmskioldia sanguinea Retz Verbenaceae WV SC Str 1,300–1,400 Ar,
Meg, Sk
97 Hydrangea anomala D.Don Hydrangeaceae WV TW Tem 2,600 Ar, Sk
98 Illigera khasiana C.B. Clarke Hernandiaceae WV TW Str 1,200 Meg
99 Ipomoea indica (Burm.) Merr. Convolvulaceae HV TW Str 1,900 Sk
100 Ipomoea vitifolia Bl. Convolvulaceae WV TW Str 1,000 Meg
101 Jasminum dispermum Wall. Oleaceae WV TW Tem 2,400 Ar, Sk
102 Jasminum lanceolaria Roxb. Oleaceae WV TW Str 1,500 Meg
103 Jasminum nervosum Lour. Oleaceae WV TW Str 1,500 Meg
104 Jasminum scandens Vahl. Oleaceae WV TW Str, Tem 2,000 Ar, Sk
105 Jasminum subglandulosum Kurz. Oleaceae WV TW Str 1,000 Meg
106 Jasminum subtriplinerve Oleaceae WV TW Str 1,500 Meg
107 Kadsura heterociliata (Roxb.) Craib. Schizandraceae WV TW Str 1,400 Meg
108 Kadsura roxburghiana Arn. Schisandraceae WV TW Tem 3,100 Sk
109 Lonicera acuminata Wall. Caprifoliaceae WV TW Tem 2,500 Ar, Sk
110 Lonicera glabrata Wall. Caprifoliaceae WV TW Str 1,900 Sk
111 Lonicera macrantha (D. Don)Spreng. Caprifoliaceae WV TW Str 1,500 Meg
112 Mallotus repandus Muell.-Arg. Euphorbiaceae WV SC Str 1,000 Sk
113 Marsdenia roylei Wight Asclepiadaceae WV TW Str 1,500 Sk
114 Marsdenia tenacissima Moon Asclepiadaceae WV TW Str 1,000 Sk
115 Melodinus cochinchinensis (Lour.) Merr. Apocynaceae WV TW Str 1,500 Meg
116 Melodinus khasianus Hook. f. Apocynaceae WV TW Str 1,500 Meg
117 Melothria heterophylla (Lour.) Cong. Cucurbitaceae WV TC Str 1,500 Meg
S.K. Barik et al.
Climbing Forest Elevation
Sl. No. Botanical names Family Category mechanism types (m) States
118 Merremia umbellata (L.) Hallier f. Convolvulaceae HV TW Str 1,000 Meg
119 Merremia vitifolia (Burm.f.) Hallier f. Convolvulaceae HV TW Tr 900 Meg
120 Mikania micrantha Kunth Asteraceae HV TW Tr 900 Meg
121 Millettia cinerea Benth. Fabaceae WV TW Str 1,500 Meg
122 Millettia pachycarpa Benth. Fabaceae WV TW Tr 700–900 Ar,
Meg, Sk
123 Mucuna macrocarpa Wall. Fabaceae WV TW Tr, Str 900–2,000 Ar,
Meg, Sk
124 Mucuna imbricata Baker Fabaceae WV TW Tr, Str 900–1,600 Ar,
Meg, Sk
125 Mucuna monosperma DC. Fabaceae HV TW Tr 500 Sk
126 Mucuna pruriens (L.) DC. Fabaceae WV TW Tr, Str 900–1,500 Ar,
Meg, Sk
127 Naravelia zeylanica DC. Ranunculaceae WV TC Tr 900 Meg
128 Nepenthes khasiana Hook.f. Nepenthaceae WV TC Str 1,500 Meg
129 Paederia foetida L. Rubiaceae WV TW Str 1,500 Meg
130 Parabaena sagittata Miers Menispermaceae WV TW Tr, Str 900–1,000 Meg, Sk
131 Parabarium micranthum DC. Apocynaceae WV TW Tr 900 Meg
132 Parthenocissus himalayana Planch. Vitaceae WV TC Tem 1,900–2,600 Ar,
Meg, Sk
133 Passiflora napalensis Wall. Passifloraceae WV TC Tem 1,900 Meg
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India

134 Pericampylus glaucus Merr. Menispermaceae WV TW Str 1,500 Sk


135 Periploca calophylla (Wight) Falc. Apocynaceae WV TW Tem 1,900–2,000 Ar, Meg
136 Persicaria chinensis (L.) H. Gross Polygonaceae HV TW Str 1,500 Meg
137 Philadelphus tomentosus Wall. ex G. Don Hydrangeaceae WV TW Tem 2,300 Ar
138 Piper betle L. Piperaceae WV RC Tr 900 Meg
(continued)
109
Table 7.2 (continued)
110

Climbing Forest Elevation


Sl. No. Botanical names Family Category mechanism types (m) States
139 Piper boehmeriaefolium DC. Piperaceae WV RC Str 1,500 Sk
140 Piper longum L. Piperaceae WV RC Str 1,000 Sk
141 Piper mullesua Buch.-Ham. ex D. Don Piperaceae WV RC Str 1,600 Meg
142 Piper peepuloides Miq. Piperaceae WV RC Tr 900 Sk
143 Piper sylvaticum Roxb. Piperaceae WV RC Tr 800 Sk
144 Plectocomia himalayana Griff. Arecaceae RA SC Str 1,900 Sk
145 Plectocomia assamica Griff. Arecaceae RA SC Str 1,000 Ar
146 Poikilospermum suaveolens (Blume) Merr. Urticaceae WV TW Tr 900 Meg
147 Porana paniculata Roxb. Convolvulaceae WV TW Str 1,200 Meg
148 Pottsia laxiflora (Blume) Kuntze Apocynaceae WV TW Str 1,200 Meg
149 Premna scandens Roxb. Verbenaceae WV TW Str 1,000 Sk
150 Pueraria sikkimensis Prain. Fabaceae WV TW Tr, Str 900–1,100 Ar, Sk
151 Pueraria tuberosa (Willd.) DC Fabaceae WV TW Tem 1,900 Meg
152 Pueraria wallichii DC. Fabaceae WV TW Tr 800 Ar, Sk
153 Rhaphidophora decursiva (Roxb.) Schott Araceae HV RC Str 1,500 Meg
154 Rhaphidophora glauca Schott. Araceae HV RC Str 1,800 Ar, Sk
155 Rosa longicuspis Bertol. Rosaceae WV SC Str 1,500 Meg
156 Rubus acuminatus Sm. Rosaceae WV SC Tem 1,900 Meg
157 Rubus ellipticus Sm. Rosaceae WV HC Str 1,800 Ar, Sk
158 Rubus insignis Wirtg. Rosaceae WV SC Tr 900 Meg
159 Rubus niveus Thunb. Rosaceae WV HC Tem 2,700 Ar
160 Rubus paniculatus Sm. Rosaceae WV HC Tem 2,100 Ar, Sk
161 Rubus rugosus Sm. Rosaceae WV SC Tem 1,900 Meg
S.K. Barik et al.
Climbing Forest Elevation
Sl. No. Botanical names Family Category mechanism types (m) States
162 Sabia campanulata Wall. Sabiaceae WV TW Tem 3,000 Sk
163 Sabia paniculata Edgew Sabiaceae WV TW Tr 900 Ar, Sk
164 Schefflera venulosa (Wight & Arn.) Harms Araliaceae WV TW Str 1,500 Meg
165 Schisandra grandiflora Hk. f. & T Schisandraceae WV TW Tem 3,000 Ar, Sk
166 Schisandra neglecta A.C. Sm. Schizandraceae WV TW Tem 1,900 Meg
167 Scindapsus officinalis Schott. Araceae HV RC Tr 500 Sk
168 Smilax aspera L. Smilacaceae WV TC Str 1,800 Ar
169 Smilax lanceifolia Roxb. Smilacaceae WV TC Tr 900 Meg
170 Smilax ovalifolia Roxb. ex D. Don Smilacaceae WV TC Tr 900 Ar, Sk
171 Spatholobus parviflorus (DC.) Kuntze Fabaceae WV TW Tr 700–900 Meg, Sk
172 Stemona tuberosa Lour. Stemonaceae WV TW Tr 600 Meg
173 Stephania glabra (Roxb.) Miers Menispermaceae WV TW Str 1,700 Ar, Sk
174 Stephania glandulifera Miers Menispermaceae WV TW Str 1,800 Ar
175 Strychnos wallichiana Steud. ex A. DC. Loganiaceae WV TC Str 1,000 Meg
176 Tetrastigma lanceolarium (Roxb.)Planch. Vitaceae WV TC Tr 900 Meg
177 Tetrastigma leucostaphylum (Dennst.) Alston Vitaceae WV TC Str 1,600 Meg
178 Tetrastigma rumicispermum (M.A. Lawson) Planch. Vitaceae WV TC Str 1,600–1,800 Ar,
Meg, Sk
179 Tetrastigma serrulatum (Roxb.)Planch Vitaceae WV TC Str, Tem 1,600–1,900 Ar,
Meg, Sk
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India

180 Thladiantha cordifolia (Blume) Cogn. Cucurbitaceae WV TC Tem 2,400 Ar


181 Thunbergia alata Bojer ex Sims Acanthaceae WV TW Tr 800 Ar, Sk
182 Thunbergia coccinea Wall. Acanthaceae WV TW Str 1,400–1,500 Ar,
Meg, Sk
183 Thunbergia fragrans Roxb. Acanthaceae HV TW Str, Tem 1,800–1,900 Ar, Sk
184 Thunbergia grandiflora Roxb. Acanthaceae WV TW Str 1,000 Meg
(continued)
111
Table 7.2 (continued)
112

Climbing Forest Elevation


Sl. No. Botanical names Family Category mechanism types (m) States
185 Tinospora cordifolia (Willd.) Miers Menispermaceae HV TW Tr 700 Meg, Sk
186 Tinospora crispa (L.) Hook. f. & Thomson Menispermaceae WV TW Tr 700 Meg
187 Toddalia asiatica (L.) Lam. Rutaceae WV HC Str, Tem 1,800–1,900 Ar,
Meg, Sk
188 Trachelospermum axillare Hook.f. Apocynaceae WV TW Tem 1,900 Meg
189 Trachelospermum lucidum (D.Don) K.Schum. Apocynaceae WV TW Str, Tem 1,200–1,900 Ar,
Meg, Sk
190 Trichosanthes wallichiana (Ser.) Wight Cucurbitaceae HV TC Str 1,600 Meg
191 Uncaria macrophylla Wall. Rubiaceae WV HC Tr 900 Meg
192 Uncaria sessilifructus Roxb. Rubiaceae WV HC Tr, Str 900–1,300 Ar,
Meg, Sk
193 Uvaria hamiltonii Hook. f. & Thomson Annonaceae WV TW Tr 900 Meg
194 Zanthoxylum oxyphyllum Edgew. Rutaceae WV HC Tem 2,300 Ar, Sk
195 Zanthoxylum sp. Rutaceae WV HC Str 1,700 Meg
196 Zizyphus funiculosus Ham. Rhamnaceae WV SC Str 1,000 Sk
Category: WV Woody vine, HV Herbaceous, RA Rattan. Climbing strategy: HC Hook climber, RC Adventitious root climber, SC Scrambler, TC Tendril
climber, TW Twiner. Forest types: Tr tropical, Str subtropical, Tem temperate. States: Ar Arunachal Pradesh, Meg Meghalaya, Sk Sikkim
S.K. Barik et al.
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India 113

Fig. 7.2 Proportion of liana species following different climbing strategies in the tropical, montane
subtropical and temperate forests of northeastern India. The values inside the bars represent the
relative proportion (in percentage) of liana species under each category of climbing strategy in
different forest types of northeast India. The figures in parentheses represent the absolute number
of liana species under each category

Table 7.3 Species richness, density and diversity of lianas in different forest types of Arunachal
Pradesh, Meghalaya and Sikkim in northeastern India
Number
of species Density Shannon’s Simpson’s Evenness
Forest types States (ha−1) (ha−1) index index index
Tropical Arunachal 19 220 2.8 0.93 0.87
Pradesh
Meghalaya 48 460 3.71 0.97 0.85
Sikkim 22 356 2.99 0.95 0.91
Total tropical 30 345 4.32 0.98 0.85
Subtropical Arunachal 24 321 3.08 0.95 0.91
Pradesh
Meghalaya 63 784 4 0.98 0.87
Sikkim 40 479 3.59 0.97 0.91
Total 42 528 4.72 0.99 0.88
subtropical
Temperate Arunachal 27 386 3.18 0.95 0.89
Pradesh
Meghalaya 12 241 2.3 0.89 0.83
Sikkim 28 339 3.18 0.95 0.86
Total 22 322 4.03 0.97 0.84
temperate
114 S.K. Barik et al.

Table 7.4 Comparison Bootstrap


of Shannon’s diversity Index Comparison (P value)
using bootstrap-pairwise
Between forest types – Overall
comparison test
Tropical vs. subtropical <0.001
Tropical vs. temperate <0.001
Subtropical vs. temperate <0.001
Between states in similar forest types
Tropical
Arunachal Pradesh vs. Meghalaya <0.001
Arunachal Pradesh vs. Sikkim <0.001
Meghalaya vs. Sikkim <0.001
Sub-tropical
Arunachal Pradesh vs. Meghalaya <0.001
Arunachal Pradesh vs. Sikkim <0.001
Meghalaya vs. Sikkim <0.001
Temperate
Arunachal Pradesh vs. Meghalaya <0.001
Arunachal Pradesh vs. Sikkim <0.839
Meghalaya vs. Sikkim <0.001

distribution. The number of species with exclusive recordings was highest in the
subtropical forests (83), followed by tropical (60) and temperate (26) forests.
Thirteen species were common between the tropical and subtropical forests viz.
Bauhinia vahlii, Calamus flagellum, Cissampelos pareira, Cissus javana, Entada
scandens, Hiptage benghalensis, Hodgsonia heteroclita, Hodgsonia macrocarpa,
Mucuna imbricata, Mucuna pruriens, Parabaena sagittata, Pueraria sikkimensis,
and Uncaria sessilifructus. The common species between subtropical and temperate
forests were, Actinidia callosa, Asparagus racemosus, Clematis buchananiana,
Dioscorea bulbifera, Dioscorea deltoidea, Hedera helix, Hedera nepalensis,
Holboellia latifolia, Jasminum scandens, Tetrastigma serrulatum, Thunbergia fra-
grans, Toddalia asiatica, and Trachelospermum lucidum (13 species). No species
was common between tropical and temperate forests (Fig. 7.3 and Table 7.2).
Overall, liana species richness was highest in Meghalaya (123), followed by
Sikkim (90) and Arunachal Pradesh (70) (Table 7.1). Ninety liana species were
recorded exclusively from Meghalaya, while 10 species were restricted to Arunachal
Pradesh, and 32 species to Sikkim. Twenty-three species were common to all the three
states viz. Actinidia callosa, Asparagus racemosus, Bauhinia vahlii, Cissampelos
pareira, Clematis buchananiana, Dioscorea bulbifera, Dioscorea deltoidea, Embelia
ribes, Entada scandens, Hedera nepalensis, Holboellia latifolia, Holmskioldia san-
guinea, Millettia pachycarpa, Mucuna macrocarpa, Mucuna imbricata, Mucuna pru-
riens, Parthenocissus himalayana, Tetrastigma rumicispermum, Tetrastigma
serrulatum, Thunbergia coccinea, Toddalia asiatica, Trachelospermum lucidum and
Uncaria sessilifructus (Table 7.2). Arunachal Pradesh and Sikkim shared the highest
number of species (31), while the numbers of common species between Meghalaya
and the two other states were very low (4 and 6 species).
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India 115

Fig. 7.3 Number of species recorded exclusively in the tropical, subtropical and temperate forests,
and the number of species shared by different forest types

7.3.2 Species Dominance and Dominance Pattern

The dominant species varied considerably with geographical locations i.e. states,
and forest types, with the exception of a few species that were common to the states
and forest types (Table 7.5). The dominant species common to the tropical forests
include Bauhinia vahlii, while in subtropical forests they were Embellia ribes and
Entada phaseoloides, and in temperate forests Holboellia latifolia and Periploca
callophylla were common. The dominance-diversity curves for lianas in tropical,
montane subtropical and temperate forests in three Eastern Himalayas and north-
eastern states of India revealed that the dominance pattern was similar in tropical
and subtropical forests, while it was significantly different in temperate forests
(Fig. 7.4). In tropical and subtropical forests, the lianas exhibited high equitability
or low dominance resembling MacArthur’s broken stick model. In contrast, in
temperate forests the most of the dominance was concentrated in 5 species,
Toddalia asiatica alone sharing 9 % of the total IVI, exhibiting high dominance
or low equitability.
116 S.K. Barik et al.

Table 7.5 Ten dominant liana species in different forest types of Arunachal Pradesh, Meghalaya
and Sikkim in India
Forest types States Name of species
Tropical Arunachal Pradesh Sabia paniculata, Bauhinia vahlii, Entada
scandens, Smilax ovalifolia, Thunbergia alata,
Cissampelos pareira, Millettia pachycarpa,
Mucuna pruriens, Acacia pennata, Calamus
flagellum
Meghalaya Bauhinia khasiana, Bauhinia vahlii, Entada
pursathea, Bauhinia scandens, Stemona tuberosa,
Argyreia nervosa, Bauhinia nervosa, Mikania
micrantha, Tetrastigma lanceolarium,
Poikilospermum suaveolens
Sikkim Thunbergia alata, Acacia pennata, Calamus
guruba, Smilax ovalifolia, Combretum roxburghii,
Sabia paniculata, Tinospora cordifolia, Cissus
spectabilis, Pueraria wallichii, Piper peepuloides
Subtropical Arunachal Pradesh Cissus repens, Mucuna imbricata, Thunbergia
coccinea, Rhaphidophora glauca, Tetrastigma
rumicispermum, Rubus ellipticus, Entada
phaseoloides, Embelia ribes, Embelia floribunda
Meghalaya Zanthoxylum sp., Elaeagnus pyriformes, Entada
phaseoloides, Embellia ribes, Nepenthes khasiana,
Asparagus racemosus, Cayratia pedata,
Raphidophora decursiva, Holboellia latifolia,
Embelia floribunda
Sikkim Calamus flagellum, Pueraria sikkimensis,
Toddalia asiatica, Bauhinia vahlii, Mucuna
pruriens, Calamus acanthospathus, Mallotus
repandus, Cissus javana, Argyreia wallichii,
Ampelocissus sikkimensis
Temperate Arunachal Pradesh Hedera nepalensis, Holboellia latifolia, Periploca
calophylla, Zanthoxyllum oxyphylum,
Tetrastigmata serrulatum, Mucuna macrocarpa,
Hedera helix, Trachelospermum lucidum,
Philadelphus tomentosus, Parthenocissus
himalayana
Meghalaya Toddalia asiatica, Parthenocissus himalayana,
Elaeagnus latifolia, Rubus rugosus, Periploca
callophylla, Pueraria tuberosa, Schisandra
neglecta, Rubus acuminatus, Passiflora
napalensis, Trachelospermum axillare
Sikkim Holboellia latifolia, Hedera nepalensis, Mucuna
macrocarpa, Trachelospermum lucidum, Lonicera
glabrata, Clematis buchananiana, Thunbergia
fragrans, Ipomea indica, Dioscorea bulbifera,
Rubus paniculatus
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India 117

Fig. 7.4 Dominance-diversity curves for lianas in tropical, montane subtropical and temperate
forests in the Eastern Himalayas and northeastern states of India

Table 7.6 Conservation status of liana species of north-eastern India


Conservation status
Sl. No. Species IUCN 3.1 Indian Red Data Book CITES
1 Acacia caesia LC – –
2 Acacia pennata LC – –
3 Bauhinia khasiana LC – –
4 Calamus inermis – Endangered –
5 Cayratia pedata – Rare –
6 Cissus spectabilis – Endangered –
7 Dioscorea deltoidea – – CITES
8 Dumasia villosa LC – –
9 Gnetum montanum LC – CITES
10 Mucuna imbricata LC – –
11 Nepenthes khasiana – – CITES
12 Plectocomia himalayana LC – –
13 Spatholobus parviflorus LC – –

7.3.3 Conservation Status of Recorded Liana Species

Out of the 196 liana species recorded, only 13 species have been assigned conserva-
tion status by the IUCN, Indian Red Data Book (Nayar and Sastry 1987–1990), and
CITES taken together (Table 7.6). The IUCN has placed 8 species under the least
118 S.K. Barik et al.

concern (LC) category, while the Indian Red Data Book has classified 2 species
under endangered and 1 species under rare category. Three species have been
included in the CITES list.

7.4 Discussion

The present study being the first of its kind in the Himalayas, and within a restricted
geographical region i.e. north-eastern India, provides primary data on the distribu-
tion and diversity of liana species along an elevation gradient. The findings refute
the general agreement that the abundance and diversity of lianas are greater in tropi-
cal forests than the subtropical and temperate region. In our sites, the highest num-
ber of species and density were found in the subtropical forests followed by tropical
and temperate forests. The high diversity and abundance of lianas in the subtropical
forests may be attributed to the unique biogeographic location of the study sites and
the prevailing diverse ecological conditions offered by wide variation in the edapho-
climatic and physiographic conditions even within a climatic zone. The study areas
fall in the trijunction of the Indo-Chinese, Indo-Malayan, and the Palaearctic
biogeographic realms, contributing to the presence of floristic elements from each
of these regions. In addition, the interplay of the physio-edapho-climatic factors
might have also led to configuration of innumerable niches resulting in high diversity.
Higher liana density was also recorded by Homeier et al. (2010) at the elevations
of 1,500 and 2,000 m in comparison to lower elevations in the tropical region of
Ecuador. The trend of liana diversity and abundance as observed in the montane
subtropical and temperate forests was similar with that of Chettri et al. (2010) in
Sikkim Himalaya. However, further studies are required to confirm the observed
trends by including more areas under each forest type experiencing varied edapho-
climatic conditions.
The liana species richness of the tropical forests in the present study (30 ha−1) is
comparable with the species richness of tropical semi-evergreen forest (28 ha−1) of
Kalrayan hills in the Eastern Ghats reported by Kadavul and Parthasarathy (1999).
Liana species richness of the subtropical forests (42 ha−1) is the same as that reported
by Hegarty (1991) in an Australian subtropical evergreen forest. Studies on liana
diversity in temperate forests are rather very few and none of the study has quanti-
fied species richness per hectare of forest area. Although difficult to compare, the
liana species richness of the temperate forests in the present study (22 ha−1) seemed
to be higher than the earlier reports. Gianoli et al. (2010) reported 14 liana species
from 0.11 ha of sampled area in the temperate evergreen rainforests of Southern
Chile. In another study in the temperate rain forests, Jimenez-Castillo et al. (2007)
reported a maximum of 5 species per 0.25 ha plot in Chile, and 6 liana species per
0.04 ha plot in New Zealand.
The liana density of the tropical forests in the present study (345 ha−1) is comparable
with the liana density (373 ha−1) in the tropical evergreen forests of Vargalaiar in
Anamalais of Western Ghats (Muthuramkumar and Parthasarathy 2000). The liana
density (528 ha−1) of the subtropical forests in the present study is also comparable
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India 119

with other parts of the world viz., Puerto Rican – El Verde and Bisley subtropical
forests (419 ha−1) (Rice et al. 2004), and Subtropical Ailao mountain area of China
(503 ha−1) (Yuan et al. 2009). The liana density of the temperate forests in the
present study (322 ha−1) is comparable to the density reported by Allen et al. (2007)
in temperate floodplain forests of southeastern United States (293 ha−1), and Londré
and Schnitzer (2006) in temperate deciduous forests of Wisconsin, USA (≈275 ha−1),
but far less than the density reported by Homeier et al. (2010) in temperate forests
of Ecuador (2,810 ha−1).
Dominance-diversity models e.g. lognormal, geometric, and broken stick, have
been globally used by ecologists to describe and compare species abundance pat-
terns in different communities, and to understand how available niche space is
shared amongst the constituent species in these communities (Magurran 2004). In
the present study, the liana abundance pattern in the tropical and subtropical forests
exhibited a close-fit with MacArthur’s broken stick model (MacArthur 1960). This
model assumes that a given number of species that exist in a community have equal
competitive ability, and the inter-specific competition therein determine the niche
boundaries. Hence, populations of these species in the community reach an equilib-
rium without exerting exclusive dominance by any one or a few species. In contrast,
the dominance-diversity curve as obtained for temperate forests exhibited a closer
fit to the geometric model, also known as the niche-preemption model (Tokeshi
1990). This model assumes that the species having highest competitive ability in a
community dominates and utilizes a substantial portion of the resources, and the
remaining resources are used in sequence by the other subordinate species (Whittaker
1972). As a result, an uneven community is generated where the status of the most
abundant species is retained, while the less abundant species lose the competition
for resources, and become rarer over time (Magurran 2004) which has profound
conservation implications.
The climbing mechanism of lianas may be grouped into two major types. Active
mechanism that displays active growth and phototropism, often associated with
the development of specialized morphological modifications/organs i.e. twining,
attachment by tendrils, and attachment using adventitious roots. Passive mecha-
nism, where lianas adhere to the existing structure such as stumps, trunks, poles, and
branches, which come in contact with the liana i.e. scrambling and spines. Each
mechanism provides unique ecological and biomechanical advantages to liana
species depending on the environment in which it grows. The dominance of species
with active mechanisms in tropical forests is a response to the low light environment
created by the dense canopy. This inference is corroborated by earlier studies which
showed that the density and abundance of twiners are more in the late successional
forests than the early successional forests (Muthuramkumar and Parthasarathy 2000;
Dewalt et al. 2000). The proportion of twiners, however, decreased in the subtropi-
cal and temperate forests, where scramblers dominated. This may be attributed to
the open environment with predominance of small-sized trees as compared to those
in tropical forests.
The field survey revealed that most of the recorded liana species are extremely
rare to locate, and have regeneration problems as very few juveniles were encountered
on the forest floor. However, only 13 liana species out of total 196 have so far been
120 S.K. Barik et al.

accorded some conservation status by the earlier workers (IUCN 2014; Nayar
and Sastry 1987–1990, and CITES). From this, it is evident that data for threat
classification for most lianas are non-existent. Hence, it is required to classify all the
species under IUCN categories considering their detailed population status, occu-
pancy and regeneration data. This necessitates further in-depth studies for correct
threat classification of liana life-form, which have been neglected hitherto. Such
classification would help in identifying the liana species which are on the verge of
extinction needing immediate conservation interventions.

References

Addo-Fordjour P, Anning AK, Atakora EA, Agyei PS (2008) Diversity and distribution of climbing
plants in a semi-deciduous rain forest, KNUST Botanic Garden, Ghana. Int J Bot 4:186–195
Allen BP, Sharitz RR, Goebel PC (2007) Are lianas increasing in importance in temperate flood-
plain forests in the southeastern United States? For Ecol Manage 242(1):17–23
Anbarashan M, Parthasarathy N (2013) Diversity and ecology of lianas in tropical dry evergreen
forests on the Coromandel Coast of India under various disturbance regimes. Flora-Morphol
Distrib Funct Ecol Plant 208(1):22–32
Balakrishnan NP (1981–1983) Flora of Jowai, 2 vols, Botanical Survey of India, Howrah
Bongers F, Schnitzer SA, Traore D (2002) The importance of lianas and consequences for forest
management in West Africa. Bioterre 1:59–70
Cai ZQ, Schnitzer SA, Bongers F (2009) Liana communities in three tropical forest types in
Xishuangbanna, South-West China. J Trop For Sci 21:252–264
Campanello PI, Garibaldi JF, Gatti MG, Goldstein G (2007) Lianas in a subtropical Atlantic forest:
host preference and tree growth. For Ecol Manage 242(2):250–259
Champion HG, Seth SK (1968) A revised survey of forest types of India. Govt. of India Press, Delhi
Chettri A, Barik SK, Pandey HN, Lyngdoh MK (2010) Liana diversity and abundance as related to
microenvironment in three forest types located in different elevational ranges of the Eastern
Himalayas. Plant Ecol Diver 3:175–185
Chittibabu CV, Parthasarathy N (2001) Liana diversity and host relationships in a tropical
evergreen forest in the Indian Eastern Ghats. Ecol Res 16(3):519–529
Darlong L, Bhattacharyya D (2012) Some lianas in Tripura, India, demand urgent conservation
efforts. Curr Sci 102(9):1246
Darlong L, Bhattacharyya D (2014) Four new additions of angiospermic liana for the state flora of
Tripura, India. Pleione 8(1):188–192
Dewalt SJ, Schnitzer SA, Denslow JS (2000) Density and diversity of lianas along a chronose-
quence in a central Panamanian lowland forest. J Trop Ecol 16(1):1–19
DeWalt SJ, Ickes K, Nilus R, Harms KE, Burslem DF (2006) Liana habitat associations and com-
munity structure in a Bornean lowland tropical forest. Plant Ecol 186(2):203–216
Gentry AH (1991) The distribution and evolution of climbing plants. In: Putz FE, Mooney HA
(eds) The biology of vines. Cambridge University Press, Cambridge, UK, pp 3–49
Gerwing JJ, Schnitzer SA, Burnham RJ, Bongers F, Chave J, Dewalt SJ, Ewango CEN, Foster R,
Kenfack D, Mart M et al (2006) A standard protocol for liana censuses. Biotropica 38:56–261
Gianoli E, Saldana A, Jiménez‐Castillo M, Valladares F (2010) Distribution and abundance of
vines along the light gradient in a southern temperate rain forest. J Veg Sci 21(1):66–73
Hajra PK, Verma DM, Giri GS (eds) (1996) Materials for the flora of Arunachal Pradesh,
vol I. Botanical Survey of India, Calcutta
Hammer Ø, Harper DAT, Ryan PD (2001) PAST: paleontological statistics software package for
education and data analysis. Palaeontol Electron 4(1):9
Haridasan K, Rao RR (1985–1987) Forest Flora of Meghalaya, 2 vols, Bishen Singh Mehandrapal
Singh, 23-A, Connaught Place, Dehradun
7 Diversity of Lianas in Eastern Himalayas and North-Eastern India 121

Hegarty EE (1991) Vine-host interactions. In: Putz FE, Mooney HA (eds) The biology of vines.
Cambridge University Press, Cambridge, UK, pp 357–375
Homeier J, Englert F, Leuschner C, Weigelt P, Unger M (2010) Factors controlling the abundance
of lianas along an altitudinal transect of tropical forests in Ecuador. For Ecol Manage
259(8):1399–1405
Hooker JD (1872–1897) The flora of British India. Reeve and Co, London
Ichihashi R, Nagashima H, Tateno M (2009) Morphological differentiation of current-year shoots
of deciduous and evergreen lianas in temperate forests in Japan. Ecol Res 24(2):393–403
IUCN (2014) IUCN red list of threatened species. Available at www.iucnredlist.org
Jiménez‐Castillo M, Wiser SK, Lusk CH (2007) Elevational parallels of latitudinal variation in the
proportion of lianas in woody floras. J Biogeogr 34(1):163–168
Joseph J (1982) Flora of Nongpoh and its vicinity. Forest Department, Government of Meghalaya
Kadavul K, Parthasarathy N (1999) Lianas in two tropical semi-evergreen forest sites on the
Kalrayan hills, Eastern Ghats, south India. Trop Biodivers 6:197–208
Kanjilal VN, Kanjilal PC, Das A, De RN, Bor NL (1934–1940) Flora of Assam, 5 vols. Govt.
Press, Shillong
Leicht-Young SA, Pavlovic NB, Frohnapple KJ, Grundel R (2010) Liana habitat and host
preferences in northern temperate forests. For Ecol Manage 260(9):1467–1477
Londré RA, Schnitzer SA (2006) The distribution of lianas and their change in abundance in tem-
perate forests over the past 45 years. Ecology 87(12):2973–2978
MacArthur RH (1960) On the relative abundance of species. Am Nat 94:25–36
Magurran AE (2004) Measuring biological diversity. Blackwell Science, Oxford
Malizia A, Grau HR (2006) Liana–host tree associations in a subtropical montane forest of north-
western Argentina. J Trop Ecol 22(3):331–339
Muthumperumal C, Parthasarathy N (2010) A large-scale inventory of liana diversity in tropical
forests of South Eastern Ghats, India. Syst Biodivers 8(2):289–300
Muthuramkumar S, Parthasarathy N (2000) Alpha diversity of lianas in a tropical evergreen forest
in the Anamalais, Western Ghats, India. Divers Distrib 6(1):1–14
Nabe-Nielsen J (2001) Diversity and distribution of lianas in a Neotropical rain forest, Yasuni
National Park, Ecuador. J Trop Ecol 17:1–19
Naidu MT, Kumar OA, Venkaiah M (2014) Taxonomic diversity of lianas in tropical forests of
Northern Eastern Ghats of Andhra Pradesh, India. Notulae Scientia Biologicae 6(1):59–65
Nayar MP, Sastry ARK (1987–1990) Red Data Book on Indian plants, 3 vols. Botanical Survey of
India, Calcutta
Parthasarathy N, Muthuramkumar S, Reddy MS (2004) Patterns of liana diversity in tropical
evergreen forests of peninsular India. For Ecol Manage 190(1):15–31
Rice K, Brokaw N, Thompson J (2004) Liana abundance in a Puerto Rican forest. For Ecol Manage
190(1):33–41
Schnitzer SA (2005) A mechanistic explanation for global patterns of liana abundance and distribution.
Am Nat 166:262–276
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17(5):223–230
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical
forest canopy gaps: evidence for an alternative pathway of gap‐phase regeneration. J Ecol
88(4):655–666
Schnitzer SA, Mangan SA, Dalling JW et al (2012) Liana abundance, diversity, and distribution on
Barro Colorado Island, Panama. PLoS One 7:e52114
Senbeta F, Schmitt C, Denich M et al (2005) The diversity and distribution of lianas in Afromontane
rain forests of Ethiopia. Divers Distrib 11:443–452
Tokeshi M (1990) Niche apportionment or random assortment: species abundance patterns
revisited. J Anim Ecol 59:1129–1146
Whittaker RH (1972) Evolution and measurement of species diversity. Taxon 21:213–251
Yuan CM, Liu WY, Tang CQ, Li XS (2009) Species composition, diversity, and abundance of lianas
in different secondary and primary forests in a subtropical mountainous area, SW China. Ecol
Res 24(6):1361–1370
Chapter 8
Biodiversity of Lianas and Their Functional
Traits in Tropical Forests of Peninsular India

N. Parthasarathy, P. Vivek, C. Muthumperumal, S. Muthuramkumar,


and N. Ayyappan

Abstract The present study was aimed to investigate the diversity and functional
traits of lianas in four tropical forest types of peninsular India. The work also
intended to study the variation in the proportion of lianas with trees along an altitu-
dinal gradient and to compare the functional traits of lianas and trees sharing similar
environmental conditions. All lianas ≥ 1.5 cm diameter and trees ≥ 10 cm diameter
at breast height were measured and included the inventory. A total of 237 liana spe-
cies were enumerated across the four forest types. The species richness of lianas per
hectare was maximum at semi-evergreen forest (31 ± 5.3) and minimum at dry ever-
green forest sites (21 ± 3.9). Semi-evergreen forest sites of Eastern Ghats also had
the highest density (648 ± 152.7 ha−1) of liana stems and wet-evergreen forest sites
of Western Ghats registered the lowest (261 ± 86.7 ha−1). Dry evergreen forest sites
on the Coromandel Coast recorded the highest basal area (0.75 ± 0.44 m2 ha−1) and
the above ground biomass (22.77 ± 19.1 Mg ha−1) of lianas across the study sites
followed by the semi-evergreen forest sites (0.69 ± 0.3 m2 ha−1 and
10.01 ± 6.2 Mg ha−1). The proportion of liana species richness to total woody species
decreased along the altitudinal gradient in the present study. Majority of lianas in
the study sites were brevi-deciduous by plant type except in wet-evergreen forest
sites. Stem twining was the chief climbing mechanism by species richness and
scramblers formed the most abundant liana group by abundance. Majority of lianas
had microphyllous leaves, whereas trees had mesophyllous leaves as their predomi-
nant leaf type. Flowers of SEF and SDF sites were mostly conspicuous, while the

N. Parthasarathy (*) • P. Vivek


Department of Ecology and Environmental Sciences, Pondicherry University,
Puducherry 605014, India
e-mail: parthapu@yahoo.com
C. Muthumperumal
School of Biological Sciences, Department of Plant Sciences, Madurai Kamaraj University,
Madurai 625021, Tamil Nadu, India
S. Muthuramkumar
V.H.N.S.N. College, Virudhunagar 626001, Tamil Nadu, India
N. Ayyappan
French Institute of Pondicherry, Puducherry 605001, India

© Springer International Publishing Switzerland 2015 123


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_8
124 N. Parthasarathy et al.

flowers of DEF sites were largely inconspicuous. Dispersal by animals (biotic)


formed the major diaspore dispersal strategy of lianas in four tropical forest types of
peninsular India.

Keywords Biodiversity • Climbing mechanism • Eastern Ghats • Functional traits


• Lianas • Plant physiognomic type • Western Ghats

8.1 Introduction

Lianas have generated considerable interest in recent times in recognition of their


importance to forest ecosystems (Addo-Fordjour et al. 2013). Lianas are key
structural component of tropical forests and constitute 10–45 % of all woody
plants and species (Schnitzer and Bongers 2002, 2011). They produce up to 40 %
of the leaves in forests (Kato et al. 1978; Putz 1983) and contribute more than
10 % of fresh aboveground biomass (Putz 1984). Occasionally lianas attain more
than 50 cm in diameter and several hundred meters in length (Schnitzer et al.
2012). However, their richness and abundance differ greatly from one forest to
another and between forest locations (Hegarty and Caballe 1991). These differ-
ences have been demonstrated at the continental scale (Emmons and Gentry
1983). Lianas are common in old-growth or primary forests, and many studies
have documented the increase in liana abundance following forest disturbance
(Schnitzer et al. 2000). Consequently, lianas are more abundant in seasonal dry
forests, where light intensity is high underneath the seasonally deciduous canopy
(Gentry 1991a). In treefall gaps, lianas can dominate regeneration for years
(Schnitzer et al. 2000), and such gaps may play an important role in liana diver-
sity (Schnitzer and Carson 2001). Tree fall dynamics together with host tree iden-
tity and abundance may thus be important factors determining the composition of
liana communities (Ibarra-Manriquez and Martinez-Ramos 2002). Moreover, the
edges of forest fragments exhibit higher liana densities and diversities compared
with undisturbed areas in the interior of a continuous primary forest (Laurance
et al. 2001).
Recent investigations have shown that the abundance of lianas in tropical forests
may increase with global climate change, probably promoted by a higher atmo-
spheric CO2 concentration (Phillips et al. 2002; Wright et al. 2004). Several factors
have been suggested to influence the abundance, species richness and distribution of
lianas in tropical forests. Liana distribution decreases with increasing rainfall (Vidal
et al. 1997; DeWalt et al. 2000), but they tend to increase with soil nutrients (Balfour
and Bond 1993; Putz and Chai 1987), seasonality (Gentry 1991a; Schnitzer 2005)
and natural disturbance such as tree fall gaps, hurricanes etc. (Quigley and Platt
2003; Rice et al. 2004; Garrido-Perez et al. 2008). Lianas exploit trees for physical
support in order to reach the forest canopy (Schnitzer et al. 2012). The increasing of
liana abundance which often pulls down trees has been associated with an increase
8 Biodiversity of Lianas and Their Functional Traits in Tropical Forests… 125

in gap formation (Vidal et al. 1997; Pereira et al. 2002), showing the influence they
have on forest structure (Perez-Salicrup 2001; Schnitzer and Bongers 2002). Lianas
have direct relationship with trees since they use them as direct supports
(Perez-Salicrup and de Meijere 2005; Reddy and Parthasarathy 2006), and are a key
determinant of their diversity and abundance (Laurance et al. 2001;
Addo-Fordjour et al. 2008).
The increasing consensus that ecosystem processes are governed by the
functional traits of species, their abundance and distribution in a community (Dıaz
et al. 2001; McGill et al. 2006) has attracted growing attention on the use of func-
tional trait composition, rather than species richness, in the exploration of
biodiversity–ecosystem functioning relationships. Experimental biodiversity–
Ecosystem functioning research has demonstrated the importance of biodiversity
for a number of ecosystem processes such as plant productivity, but it remains a
central challenge to identify the underlying mechanisms (Hooper et al. 2005).
Functional traits are morphological, physiological and phenological features mea-
surable at the individual level which modulate plant performance and individual
fitness via their effects on growth, survival and reproductive output (Violle et al.
2007). Plant traits largely determine how individual plant species contribute to pro-
cesses at the community-level (Roscher et al. 2012). The concept of functional trait
diversity is based on the assumption that with increasing trait dissimilarity among
species the diversity in resource use strategies increases as well and species overlap
along resource axes decreases (Tilman et al. 1997). Liana life-form has five poten-
tial climbing modes categorized by Putz and Chai (1987) and DeWalt et al. (2000).
There are (1) scramblers, (2) twiners, (3) root climbers, (4) tendril climbers, and (5)
hook climbers. Stem twining was reported to be the chief climbing mechanism and
composed the largest proportion of species richness, abundance and basal area in
many tropical regions (Putz and Chai 1987; Nabe-Nielsen 2001; Parthasarathy et al.
2004; Senbeta et al. 2005; Reddy and Parthasarathy 2006; Edwards et al. 2007). A
general tendency for twining plants to ascend their hosts in right-handed helices,
corresponding to anti-clockwise motion in the direction of growth (Hashimoto
2002). Leaf traits are thought to play important roles in carbon assimilation, water
relations and energy balance (Castellanos 1991). Our current knowledge on vine
leaf traits is relatively poor (see Castellanos et al. 1989; Molina-Freaner and
Tinoco-Ojanguren 1997).
Lianas are important for the diet of animals at times when flowers and fruits from
trees are scarce. Some lianas have no flowering and fruiting peaks, whereas others
that have peaks reproduce at different times than co-occurring trees (Putz and
Windsor 1987). Lianas reproduce by seeds or by sprouting, which enables them to
colonize and become established in a broad range of habitats (Nabe-Nielsen and
Hall 2002). Lianas display a wide range of diaspore types such as anemochory,
zoochory, barochory and hydrochory. Seed dispersal syndromes are often correlated
with seasonality and precipitation, with wind dispersed seeds common in highly
seasonal forests and far less common in aseasonal forests (Gentry 1982, 1991a). In
general, drier tropical forests tend to be more wind dispersed, while wetter forests
126 N. Parthasarathy et al.

typically have more bird and mammal dispersed seeds. The wind dispersed seeds
also tend be associated with the high-light, drier conditions of an open, disturbed
space in forest gaps and along permanent edges. It has been demonstrated that there
is a lower percentage of wind dispersal of vines in Old World forests than in
Neotropical forests (Gentry 1991a). Under the changing climatic conditions, lianas
are increasing in tropical forests (Reference) and it signifies their impacts on
ecosystems experiencing changes to a greater extent in all aspects, along the
decades. Hence, we need to understand the trait biology of climbing plants which
majorly contribute to ecosystem functions. Hence, in this paper, we compare the
liana diversity and biomass across wet to dry tropical forests and we analyze the
traits of woody climbers as compared to tree communities in tropical forests of
peninsular India.

8.2 Materials and Methods

8.2.1 Study Area and Forest Types

Liana and tree diversity inventories were carried out in 40 different sites in four
major tropical forest types viz., wet-evergreen forest (hereafter referred as WEF),
semi-evergreen forest (SEF), seasonally dry forest (SDF), and dry evergreen forest
(DEF), distributed in Western Ghats, Eastern Ghats, and the Coromandel Coast of
peninsular India (Fig. 8.1). Tropical wet evergreen forest (WEF) sites include
Varagalaiar, Valparai, Agumbe (in the Western Ghats), and the hills of Kalrayans,
Pachamalais and Shervarayans (in the southernmost Eastern Ghats). The Varagalaiar
and Valparai plateaus are located within the Indira Gandhi National Park and
Wildlife Sanctuary, Anamalais, Western Ghats, south India. Of the five sites in
Valparai, two sites, Tata Finlay and Injipara, are surrounded by coffee and tea plan-
tations and are subject to greater human interference than the other sites. These sites
benefit from both southwest (June–September) and northeast monsoon (October–
December) rains.
Among the Eastern Ghats sites, liana inventories were made in the hill ranges
of Bodamalai, Chitteri, Kalrayan, the Kolli hills, Pachamalai, and Shervarayan,
which are situated in the Salem district of Tamil Nadu, south India. The princi-
pal vegetation of all sites is the SDF (seasonal dry forest, in which we include
tropical mixed and dry deciduous and thorn forests) at lower altitudes and the
SEF to WEF (particularly in the high ranges of Kalrayan, Kolli hills, Pachamalai
and Shervarayan). These sites receive rainfall between August and December,
with a mean annual value of 1,057.9 mm. Most sites are subject to various forest
disturbances, such as land-use change for cultivation (e.g., cereals, pineapple,
tapioca, etc.), illegal timber extraction, and collection of non-timber forest pro-
duce, soil removal (for ore mining and road construction), cattle grazing, tour-
ism, and plant invasions. The DEF sites on the Coromandel Coast are distributed
in Cuddalore, Pudukottai and Villupuram districts of Tamil Nadu. All the 18
8 Biodiversity of Lianas and Their Functional Traits in Tropical Forests… 127

Fig. 8.1 Map showing the location of study sites in four tropical forest types of peninsular India

DEF sites are protected as sacred groves or temple forests protected by the local
people on the religious grounds.
Of the 40 study sites, DEF sites on the Coromandel Coast and SDF sites of the
Eastern Ghats experience longer dry seasons (6–8 months), followed by the SEF
sites of Eastern Ghats; the dry season is shortest (3–4 months) in the WEF sites.
Among the four forest types, the forest stature is short (7 m) in DEF sites, moder-
ately tall (12–16 m) in SDF and SEF sites, and tallest in WEF sites of the Western
Ghats (35 and 25 m respectively at Varagalaiar and Agumbe).
128 N. Parthasarathy et al.

8.2.2 Sampling and Data Analysis

All lianas ≥1.5 cm diameter, rooted inside the plot were measured at 1.3 m from the
rooting point and all trees ≥10 cm diameter at breast height were measured in 40
sites of four tropical forest types. This standard method allowed us to validly com-
pare liana diversity across four forest types. Data on liana density were converted to
individuals per hectare by interpolation. We converted liana density for Injipara,
Lower Manamboli, and Tata Finlay from 0.8 to 1 ha. Liana aboveground biomass
was calculated using the allometric equation of Schnitzer et al. (2006):

AGB = exp ⎡⎣ −1.484 + 2.657 ln ( D )⎤⎦ , where D is the diameter.

Liana and tree functional traits were analyzed in the field and also by referring to the
pertinent manuals (Gamble and Fischer 1915–1935; Matthew 1991). We assigned
five different climbing mechanisms (the stem and branch twiners, scramblers, ten-
dril climbers, hook climbers and root climbers) for lianas. The diaspore dispersal
modes were classified as biotic (zoochory) and abiotic (anemochory and autochory)
types. Leaf types were classified (Webb 1959 with some modifications) as lepto-
phyll (<0.25 cm2), nanophyll (0.25–2.25 cm2), microphyll (2.25–20.25 cm2), noto-
phyll (20.25–45 cm2), mesophyll (45–182.25 cm2), macrophyll (182.25–1640.2 cm2)
and megaphyll (>1,640.2 cm2). Flower types were classified (Gentry 1991a) as con-
spicuous (>1 cm) and inconspicuous flowers (<1 cm).

8.3 Results

8.3.1 Taxonomic Diversity

Liana diversity inventory yielded a total of 237 species enumerated from four tropi-
cal forest types of peninsular India (Table 8.1) and among them, the semi-evergreen
forest (SEF) harbored a maximum of 126 liana species (31 ± 5.3 species ha−1) and
the dry evergreen forest (DEF) the minimum (56 species, 21 ± 3.9 species ha−1). The
mean species richness respectively for wet evergreen forest (WEF) and seasonally
dry forest was 30 ± 6.2 and 25 ± 4.8 species ha−1 (Table 8.1). Fifteen liana species,
including Carissa spinarum, Combretum albidum and Ziziphus oenoplia occurred
in all four forest types, while 29 species and 57 species were represented in any of
the three and two forest types respectively (Appendix). A total of 136 species was
exclusive to any one of the forest type. Vitaceae was the most speciose liana family,
represented by 18 species in four tropical forest types of peninsular India followed
by Apocynaceae and Papilionaceae with 17 species each. The Papilionaceae domi-
nated the WEF (10 species) and the DEF (5) sites, while Asclepiadaceae and
Apocynaceae were the most species-rich families in SDF (8 species) and SEF (12
species) sites (Appendix). The SEF sites of the Eastern Ghats recorded the highest
density of liana stems (648 ± 152.7) ha−1, while the WEF sites of the Western Ghats,
8 Biodiversity of Lianas and Their Functional Traits in Tropical Forests… 129

Table 8.1 Liana species richness, density and aboveground biomass in four tropical forest types
of peninsular India
Wet-evergreen Semi-evergreen Seasonally dry Dry evergreen
Variable forest (WEF) forest (SEF) forest (SDF) forest (DEF)
Total area sampled (ha) 38.4 21.5 31.5 18
Diameter threshold (cm) ≥1.5 ≥1.5 ≥1.5 ≥1.5
Species richness (on plot) 116 126 70 56
Species richness ha−1 30 ± 6.2 31 ± 5.3 25 ± 4.8 21 ± 3.9
Density (individuals ha−1) 261 ± 86.7 648 ± 152.7 595 ± 138.8 522 ± 242
Basal area (m2 ha−1) 0.42 ± 0.19 0.69 ± 0.3 0.21 ± 0.06 0.75 ± 0.44
Aboveground biomass 186.86 200.25 49.98 409.9
(on plot Mg)
Aboveground biomass 5.19 ± 2.9 10.01 ± 6.2 1.78 ± 1.4 22.77 ± 19.1
(Mg ha−1)

registered the lowest density (261 ± 86.7 ha−1). The mean liana density in SDF and
DEF sites was 595 ± 138.8 ha−1 and 522 ± 242 ha−1respectively (Table 8.1). Although,
slightly lower in the number of liana stems than the SEF and the SDF sites, the DEF
sites on the Coromandel Coast, registered a maximum mean basal area value of
0.75 ± 0.44 m2 ha−1 followed by SEF (0.69 ± 0.3 ha−1) and WEF (0.42 ± 0.1 m2 ha−1)
sites (Table 8.1). Furthermore, the DEF sites accounted for 48.3 % (186.8 Mg) of
the total liana aboveground biomass on the plot (AGB) and the SDF sites contrib-
uted the minimum of 49.9 Mg (Table 8.1). The DEF sites also recorded the highest
mean liana AGB value on a per hectare basis (22.77 ± 19.1 Mg ha−1) and it decreased
in the order of DEF > SEF > WEF > SDF (Table 8.1).

8.3.2 Relationship Between Altitude and Mean Proportion


of Liana and Tree Species

The regression analysis performed using altitude and the mean proportion of liana
and tree species of the four forest types revealed that the mean proportion of lianas
decreased over trees with increasing altitude (R2 = 0.558). Whereas, the proportion
of tree species increased with increasing altitude (Fig. 8.2). The proportion of lianas
against tree species was particularly higher in the DEF sites with lower altitude and
it tends to decrease in WEF sites, where the proportion of lianas was relatively
lesser than the trees, with increasing altitude.

8.3.3 Liana Functional Traits

Among the total of 62,054 liana individuals and 237 species inventoried, 64.6 % of
liana individuals represented by 112 species (47 % of the total liana species rich-
ness) were brevi-deciduous by plant type. The evergreen species comprised 31.1 %
130 N. Parthasarathy et al.

0.9
R2 = 0.557
Proportion of trees and lianas 0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1 Liana species richness
Tree species richness
0
0 500 1000 1500
Altitude (m)

Fig. 8.2 Relationship between altitude (m) and mean proportion of tree and liana species richness
at the unified one-hectare scale for thirty sites of tropical forests in peninsular India

(89 species; 37.3 %) and the deciduous species contributed just 4.2 % of the total
liana abundance across the four forest types (Figs. 8.3 and 8.4). However, an analy-
sis by forest type revealed that the proportion of species abundance under the three
plant physiognomic types varied (Fig. 8.4). For instance, the WEF sites had a higher
proportion (60.4 %) of evergreen species, whereas the other forest types predomi-
nantly harbored greater abundance of brevi-deciduous plant species than the ever-
green and deciduous species.
We distinguished five different climbing mechanisms employed by lianas in four
tropical forest types of peninsular India and among them stem twining was the chief
climbing mechanism by species richness, adopted by 103 species (29.5 % of the
total liana abundance) (Fig. 8.3). The scramblers have shown greater dominance by
constituting 53.7 % of the total liana abundance (Fig. 8.4). Tendril climbers, hook
climbers and root climbers contributed 9 %, 4 % and 2 % respectively of the total
liana abundance. The root climbers, represented by just 3 species are confined to the
WEF and SEF sites (Fig. 8.3).
We recognized four different leaf types among 237 liana species, in which, lianas
with microphyllous leaves dominated the peninsular Indian forests forming 60.1 %
of the total abundance and 48.5 % of the total liana species richness (Figs. 8.3 and
8.4). Lianas with mesophyllous (83 species; 35 % of total species richness) and
notophyllous (29 species; 12.2 %) leaf types formed 27.3 % and 7 % of the total
liana abundance. Lianas with macrophyllous leaves were encountered only from the
WEF sites where they composed 18.3 % of the total abundance. Across the four
forest types, lianas with conspicuous flowers formed 59.6 % of the total liana abun-
dance and 57.9 % of the total species richness. Inconspicuous flowers were pre-
8 Biodiversity of Lianas and Their Functional Traits in Tropical Forests… 131

Fig. 8.3 Percentage of liana species richness under various plant functional traits (plant type,
climbing mechanism, leaf type, flower type, fruit type and dispersal mode) in four tropical forest
types of peninsular India

dominant in DEF sites (63.1 %) and comprised 40.4 % of total liana abundance
(Fig. 8.4). The majority of liana species (60.7 %) and individuals (54.9 %) in the
four studied forest types were fleshy-fruited and meant for biotic mode of diaspore
dispersal (Figs. 8.3 and 8.4). Dry-fruited lianas including pod, samara and follicle
comprised 43.1 % of the total liana abundance.

8.3.4 Comparison of Liana and Tree Functional Traits

A comparison of liana and tree functional traits in four forest types revealed that
unlike lianas, majority of tree species were evergreen by plant physiognomic type
(Fig. 8.5), especially in WEF (79 % of total tree species richness) and SEF sites
(57 %) (Fig. 8.5a, b). However, in SDF and DEF sites, the brevi-deciduous tree
132 N. Parthasarathy et al.

Fig. 8.4 Percentage of liana abundance under various plant functional traits (plant type, climbing
mechanism, leaf type, flower type, fruit type and dispersal mode) in four tropical forest types of
peninsular India

species dominated, which is on par with that of the lianas in the respective forest
types (Fig. 8.5c, d). There was not much variation between tree and liana species
richness with regard to the leaf types, except that the species with microphyllous
leaves dominated the liana life-form in all the forest types, whereas leaves of trees
are largely mesophyllous (Fig. 8.5). Lianas in all the forest types had conspicuous
flowers as their predominant flower type, whereas trees had more of inconspicuous
flowers in WEF (54 %) and DEF (53 %) sites. The liana species bearing fleshy-fruits
dominated all the four tropical forest types, while for trees, the proportion of fleshy-
fruited species was greater in WEF (78 %), SDF (67 %) and DEF (89 %) sites
(Fig. 8.5). In SEF sites, the species richness of dry-fruited (52 %) and fleshy-fruited
tree species (48 %) varied marginally (Fig. 8.5b).
8 Biodiversity of Lianas and Their Functional Traits in Tropical Forests… 133

Fig. 8.5 Comparison of liana and tree species richness under various plant functional (plant type,
leaf type, flower type, fruit type and dispersal mode) traits in four tropical forest types of peninsu-
lar India
134 N. Parthasarathy et al.

8.4 Discussion

Among the four tropical forest types the mean liana species richness was least for
DEF sites when compared to the other three forest types of peninsular India. Gentry
(1982, 1988) also suggested that forests in dry areas are less diverse than wet areas.
The greater species richness in SEF and SDF sites than that of the DEF with similar
rainfall pattern and length of dry season indicates that the observed species richness
may be related to other factors besides precipitation, such as forest disturbance
(Anbarashan and Parthasarathy 2013). The liana species composition in the present
study sites showed high degree of heterogeneity, as just 15 among 237 species enu-
merated were common to all four forest types. The distribution of these 15 liana
species in all four forest types indicates their wide range of ecological amplitude
and adaptation. Furthermore, a total of 136 species were exclusively represented in
any one of the forest types which further signifies the differences in ecological
needs and demands among the inventoried liana species (Yuan et al. 2009). These
varying species composition of lianas in different forest types may further be attrib-
uted to the large ecological and functional differences across species.
The highest density of lianas in DEF, SEF and SDF sites could be attributed to the
extent of anthropogenic pressure coupled with relatively lesser rainfall and seasonality
(Gentry 1991a). Several other studies (Swaine and Grace 2007; Schnitzer 2005;
Schnitzer and Bongers 2011; DeWalt et al. 2009) have also explained the greater abun-
dance of lianas in forests with lowest rainfall and highest seasonality or length of dry
months. Besides precipitation and seasonality, another possible reason could be the
light which is supposed to be the important factor influencing the liana proliferation
(Putz 1984; Schnitzer et al. 2000) and thus, the DEF, SEF and SDF sites with pro-
nounced dry seasons may have higher light penetrations and thereby induce the ger-
mination, growth and survival of lianas in the understory (Kurzel et al. 2006). The
reason for lower density of lianas in WEF sites can be explained by the fact that lianas
do not regenerate successfully under dense and shady canopy. Moreover, Hartshorn
and Hammel (1994) suggested the possible exclusion of lianas by epiphytes and stran-
glers in wet forests. Hence, lianas may compete better in seasonal than in the asea-
sonal forests (Restom and Nepstad 2001; Schnitzer and Bongers 2002; Restom and
Nepstad 2003). Families such as Vitaceae, Apocynaceae and Papilionaceae dominated
the peninsular Indian forest sites as in most of the Asian tropics (Putz and Chai 1987;
Cai et al. 2009; Chettri et al. 2010) and the distribution of the families Fabaceae and
Apocynaceae are reported pantropically (Anbarashan and Parthasarathy 2013).
Lianas in general may constitute up to 25 % of the woody species richness in tropics
(Schnitzer and Bongers 2002) and the contribution of lianas to the woody species rich-
ness in present study sites declined along the increasing altitudinal gradient, supporting
also the findings of Bhattrai and Vetaas (2003), Jiminez-Castillo et al. (2007) and Addo-
Fordjour et al. (2013). Jiminez-Castillo et al. (2007) attributed the freezing temperature
at the higher altitudes to play a role besides other environmental factors in restricting
liana species richness, but in the present study it may be explained as the influence of
environmental variables at medium elevation that possibly co-vary along the altitudinal
gradient (Addo-Fordjour et al. 2013). In the present study, majority of lianas in DEF,
SDF and SEF sites are brevi-deciduous which corroborates with the observation of
8 Biodiversity of Lianas and Their Functional Traits in Tropical Forests… 135

other studies that reported the co-existence of evergreen and deciduous species in dry
forests (Sarmiento 1972 and Mateucci 1987). Stem twining was the chief climbing
mechanism in the four tropical forest types of peninsular India, which is in accordance
with the results of earlier studies conducted across the tropics (DeWalt et al. 2000;
Chalmers and Turner 1994; Senbeta et al. 2005 and Cai et al. 2009). The reason for the
greater abundance of scramblers in SDF and SEF sites could be assigned to the low
canopy height (Ding 2006) and increased edge proximity coupled with greater exten-
sion of Lantana camara (armed scrambler) invasion. Different climbing mechanisms
have different trellis requirements (Putz 1984; Hegarty and Caballe 1991; Nabe-Nielsen
2001). For example, the particular abundance of tendril climbers in DEF sites may be
assigned to the availability of many suitable small-sized host stems. Similarly, the
restricted presence of root climbers only in the WEF and SEF sites can be reasoned for
the facts that root climbers can climb without any regard to the diameter of the host
(Putz 1984; Putz and Holbrook 1991) and they are generally shade-tolerant under the
dense canopy of these tall-statured forests. According to Hegarty and Caballe (1991) the
tendril climbers are early successional and root climbers as late successional species.
The dominance of lianas with microphyllous leaves, especially in the SDF, DEF
and SEF sites could serve as adaptation strategy by lianas in these seasonal forests
with pronounced dry period. In contrast, WEF sites had majority of lianas with
mesophyllous leaves which corroborates with the results of Ewango in Congolian
rain forests (2010). In general, the leaves with smaller size are associated with the
maintenance of favorable leaf temperature and greater water-use efficiency under
high solar radiation and low water availability (Parkhurst and Loucks 1972; Givinish
and Vermeij 1976). Majority of lianas in the study sites had conspicuous flowers
visited by birds and large Lepidopterans and those of inconspicuous flowers by bees
and Dipterans. The greater abundance of conspicuous flowers in SEF and SDF sites
supports the hypothesis of Gentry (1982, 1991a) who proposed the dominance of
conspicuous flowers in dry forests and inconspicuous flowers in wet forests. But, in
contrast, the DEF sites had majority of inconspicuous flowers. Cai et al. (2009) also
reported the dominance of conspicuous flowers in three tropical forest types of
South-west China. According to Gentry (1982), lianas in dry forests are more prone
to be wind-dispersed, but in the present study sites, zoochorous mode of dispersal
was the predominant across all the forest types and this signifies the importance of
lianas for the dependent faunal assemblage. In contrast, Cai et al.(2009) and Senbeta
et al. (2005) reported the dominance of abiotic mode of dispersal in tropical forests
of China and Ethiopia respectively. Several other studies (Ewango 2010; Addo-
Fordjour et al. 2008) have also reported the predominance of biotic mode of seed
dispersal. The contrasting leafing phenology strategies of trees and lianas in tropical
forest types of peninsular India reflects the different vegetative traits and leafing
phenology between trees and liana life-forms in the same ecosystem (Roldan 1999).
Most of the liana-life forms had the microphyllous leaves and trees had mesophyl-
lous leaves, which could offer a competitive advantage for lianas over trees, particu-
larly in the DEF and SDF sites which experiences 6–8 months dry period. Therefore,
the patterns of liana diversity and functional traits in the tropical forest types of
peninsular India are mainly predicted by the rainfall and the extent of seasonality
besides the effect of elevational gradient and anthropogenic impacts.
136

Appendix

List of 237 liana species enumerated in peninsular Indian tropical forests with their distribution in different forest types (1 Wet evergreen forest (WEF), 2 Semi-
evergreen forest (SEF), 3 Seasonally dry forest (SDF), 4 Dry evergreen forest (DEF) and their functional traits (Plant type: Bd brevi-deciduous, D deciduous,
E evergreen. Climbing mechanism: St stem twiner, Tc tendril climber, Hc hook climber, SA scrambler-armed, SU scrambler-unarmed, Rc root climber. Leaf
type: Mi microphyll (2.25–20.25 cm2), No notophyll (20.25–45 cm2), Me mesophyll (45–182.25 cm2), Ma macrophyll (182.25–1,640.2 cm2). Flower type: Cn
conspicuous, Ic inconspicuous. Dispersal mode: Bi biotic, Ab abiotic).
Liana functional traits
Occurrence in WEF, Plant Climbing Leaf Flower Dispersal
Liana species Family SEF, SDF, DEF type mechanism type type mode
Abrus precatorius L. Papilionaceae 2,3,4 D St Mi Cn Ab
Acacia caesia (L) Willd. Mimosaceae 2,3,4 Bd SA Me Cn Ab
Acacia canescens Grah. Mimosaceae 2 Bd SA Me Cn Ab
Acacia columnaris Craib Mimosaceae 2 Bd SA Me Cn Ab
Acacia hohenackeri Craib Mimosaceae 1 Bd SA Me Cn Ab
Acacia pennata (L.) Willd Mimosaceae 1,2,3 Bd SA Me Cn Ab
Acacia sinuata (Lour.) Merr. Mimosaceae 1,2 Bd SA Me Cn Ab
Acacia torta Roxb. Mimosaceae 1,2,3 Bd SA Me Cn Ab
Adenia hondala (Gaertn.) Wilde Passifloraceae 1 Bd Tc Mi Cn Bi
Adenia wightiana (Wall.ex Wight & Arn.) Engler Passifloraceae 4 Bd Tc Mi Cn Bi
Aganosma cymosa (Roxb.) G. Don var. lanceolata Hook. f. Apocynaceae 1,2 Bd St Mi Cn Ab
Aganosma cymosa (Roxb.) G.Don var. cymosa Hook. f. Apocynaceae 1,2,3,4 Bd St Mi Cn Ab
Alangium salvifolium (L.f.) Wang. ssp. hexapetalum (Lam.) Wang. Alangiaceae 1 Bd SA Me Cn Bi
Allophylus concanicus Radlk. Sapindaceae 1 Bd SU Me Ic Bi
Ampelocissus araneosa (Dalz. & Gibs.) Planch. Vitaceae 2 D Tc Me Ic Bi
Ampelocissus arnottiana Planch. Vitaceae 3 D Tc Me Ic Bi
Ampelocissus eriocladus (Wight & Arn.) Planch. Vitaceae 1 D Tc Me Ic Bi
N. Parthasarathy et al.
8

Ampelocissus tomentosa (Heyne ex Roth) Planch. Vitaceae 2,3,4 D Tc Me Ic Bi


Anamirta cocculus (L.) Wight Menispermaceae 1,2 E St Me Ic Bi
Ancistrocladus heyneanus Wall. ex Graham Ancistrocladaceae 1 E HC MA CN AB
Anodendron paniculatum A.DC. Apocynaceae 1,2 E St Mi Cn Ab
Anodendron rhinosporum Thw. Apocynaceae 1 E St Mi Cn Ab
Argyreia cuneata (Willd.) Ker-Gawl. Convolvulaceae 2 E St Mi Cn Bi
Argyreia daltoni Clarke Convolvulaceae 2 E St Me Cn Ab
Argyreia elliptica (Roth) Choisy Convolvulaceae 1 E St Mi Cn Bi
Argyreia kleiniana (Roem. & Schultes) Raizada Convolvulaceae 2 E St Me Cn Bi
Argyreia pilosa Arn. Convolvulaceae 2 E St Mi Cn Bi
Argyreia populifolia Choisy Convolvulaceae 1 E St Me Cn Ab
Argyreia sericea Dalz. Convolvulaceae 3 E St No Cn Bi
Aristolochia indica L. Aristolochiaceae 3,4 E St Mi Cn Ab
Aristolochia tagala Cham. Aristolochiaceae 2 E St No Cn Ab
Artabotrys zeylanicus J.D.Hook. & Thoms. Annonaceae 1 E Hc Me Cn Bi
Asparagus racemosus Willd. Liliaceae 4 D SA Mi Ic Bi
Azima tetracantha Lam. Salvadoraceae 4 E SA Mi Ic Bi
Basella alba L. Basellaceae 3 E St Mi Ic Bi
Bauhinia phoenicea Wight & Arn. Caesalpiniaceae 1 Bd St Me Cn Ab
Bragantia wallichii Br. Aristolochiaceae 1 E St Me Cn Bi
Bridelia scandens (Roxb.) Willd. Euphorbiaceae 1 D SA Mi Ic Bi
Butea parviflora Roxb. Papilionaceae 1,2 D SU Me Ic Ab
Biodiversity of Lianas and Their Functional Traits in Tropical Forests…

Caesalpinia crista L. Caesalpiniaceae 2,3 D SA Me Cn Ab


Caesalpinia cucullata Roxb. Caesalpiniaceae 1,2 Bd SA Me Cn Ab
Calamus gamblei Becc.ex Becc. & J.D.Hook. Arecaceae 1 E Gc Ma Ic Bi
Calamus pseudo-tenuis Beccari ex Beccari & Hook. f. Arecaceae 1 E SA Ma Ic Bi
(continued)
137
Liana functional traits
138

Occurrence in WEF, Plant Climbing Leaf Flower Dispersal


Liana species Family SEF, SDF, DEF type mechanism type type mode
Calamus thwaitesii Becc. & J.D.Hook. Arecaceae 1 E Gc Ma Ic Bi
Calycopteris floribunda Lam. Combretaceae 1,4 D St Mi Cn Ab
Canavalia virosa (Roxb.) Wight & Arn. Papilionaceae 2,3,4 D St Me Cn Ab
Cansjera rheedii Gmel. Opiliaceae 1,2,3,4 Bd St Mi Ic Bi
Canthium angustifolium Roxb. Rubiaceae 1 D SA Mi Ic Bi
Canthium rheedii DC. Rubiaceae 1 D SA Mi Ic Bi
Capparis brevispina DC. Capparaceae 2,4 E SA Mi Cn Bi
Capparis divaricata Lam. Capparaceae 3,4 E SA Mi Cn Bi
Capparis fusifera Dunn Capparaceae 1 Bd SA Mi Cn Bi
Capparis moonii Wight Capparaceae 1 Bd SA Mi Cn Bi
Capparis rotundifolia Rottb. Capparaceae 4 E SA Mi Cn Bi
Capparis sepiaria L. var. retusella Thwaites Capparaceae 2 Bd SA Mi Cn Bi
Capparis sepiaria L. var. sepiaria Capparaceae 2,3,4 Bd SA Mi Cn Bi
Capparis shevaroyensis Sun.-Ragh. Capparaceae 1 E SA Mi Cn Bi
Capparis zeylanica L. Capparaceae 2,3,4 Bd SA Mi Cn Bi
Carissa carandas L. Apocynaceae 2,3 Bd SA Mi Cn Bi
Carissa gangetica Stapf Apocynaceae 3 Bd SA Mi Cn Bi
Carissa inermis Vahl Apocynaceae 1 Bd SA Mi Cn Bi
Carissa paucinervia A. DC. Apocynaceae 2,3 Bd SA Mi Cn Bi
Carissa salicina Lam. Apocynaceae 3 Bd SA Mi Cn Bi
Carissa spinarum L. Apocynaceae 1,2,3,4 Bd SA Mi Cn Bi
Cayratia carnosa (Wall. ex Wight & Arn.) Gagnep Vitaceae 1 Bd Tc Me Ic Bi
Cayratia japonica (Thunb.) Gagnep. Vitaceae 2 Bd Tc Me Ic Bi
Cayratia pedata (Lam.) Juss.ex Gagnep. Vitaceae 1,2,4 Bd Tc Me Ic Bi
Cayratia roxburghii (Wight & Arn.) Gagnep. Vitaceae 2 Bd Tc Me Ic Bi
N. Parthasarathy et al.
8

Cayratia sp. Vitaceae 1 Bd Tc Me Ic Bi


Cayratia tenuifolia (Wight &Arn.) Gagnep. Vitaceae 1 Bd Tc Mi Ic Bi
Celastrus paniculatus Willd. Celastraceae 1,2,3 D SU Mi Ic Bi
Ceropegia thwaitesii J.D.Hook. Asclepiadaceae 1 D St No Cn Ab
Chilocarpus atrovirens (G.Don) Bl. Apocynaceae 1 Bd St Me Cn Bi
Chonemorpha fragrans (Moon) Alston Apocynaceae 1 Bd St Me Cn Ab
Cissus gigantea (Bedd.) Planch. Vitaceae 2 Bd Tc Me Ic Bi
Cissus glauca Roxb. Vitaceae 1 Bd Tc Me Ic Bi
Cissus glyptocarpa (Thw.) Planch. Vitaceae 1 Bd Tc Me Ic Bi
Cissus heyneana (Wall. ex Lawson) Planch. Vitaceae 2,3 D Tc Me Ic Bi
Cissus quadrangularis L. Vitaceae 2,3,4 E Tc Mi Ic Bi
Cissus vitiginea L. Vitaceae 2,3,4 Bd Tc Me Ic Bi
Clematis gouriana Roxb. ex DC. Ranunculaceae 1,2 D SU Me Cn Ab
Clerodendrum inerme (L.) Gaertn. Verbenaceae 4 Bd SU Mi Cn Bi
Coccinia indica (L.) Voigt Cucurbitaceae 3,4 D Tc No Cn Bi
Cocculus hirsutus (L.) Diels Menispermaceae 2,3 Bd St Mi Ic Bi
Combretum acuminatum Roxb. Combretaceae 2,3 Bd St No Ic Ab
Combretum albidum G. Don Combretaceae 1,2,3,4 Bd St No Ic Ab
Combretum latifolium Bl. Combretaceae 1 Bd St No Ic Ab
Connarus sclerocarpus (Wight & Arn.) Schellenb. Connaraceae 1 Bd St Ma Ic Bi
Coscinium fenestratum (Gaertn.) Coleb. Menispermaceae 1 E St Me Ic Bi
Cosmostigma racemosum (Roxb.) Wight Asclepiadaceae 1,2 E St No Cn Ab
Biodiversity of Lianas and Their Functional Traits in Tropical Forests…

Croton caudatus Geiseler Euphorbiaceae 1 D SU Mi Ic Ab


Cryptolepis buchanani Roemer & Schultes Asclepiadaceae 2 Bd St No Cn Ab
Dalbergia conjesta Graham ex Wight & Arn. Papilionaceae 2 D St Me Ic Ab
Dalbergia rubiginosa Roxb. Papilionaceae 1,2 D St Me Ic Ab
Decalepis hamiltonii Wight & Arn. Asclepiadaceae 2,3 D St Mi Ic Ab
(continued)
139
Liana functional traits
140

Occurrence in WEF, Plant Climbing Leaf Flower Dispersal


Liana species Family SEF, SDF, DEF type mechanism type type mode
Derris benthamii var. benthamii (Thw.) Thw. Papilionaceae 1 Bd St Me Cn Ab
Derris brevipes (Benth.) Baker Papilionaceae 1 Bd St Me Cn Ab
Derris heyneana (Wight & Arn.) Benth. Papilionaceae 1 Bd St Me Cn Ab
Derris ovalifolia (Wight & Arn.) Benth. Papilionaceae 4 Bd St Me Cn Ab
Derris scandens (Roxb.) Benth. Papilionaceae 2,3,4 Bd St Mi Cn Ab
Derris thyrsiflora Benth. Papilionaceae 1 Bd St Me Cn Ab
Derris trifoliata Lour. Papilionaceae 1 Bd St Me Cn Ab
Desmos viridiflorus Saff. Annonaceae 2 Bd SU No Cn Bi
Dioscorea pentaphylla L. Dioscoreaceae 2 D St Me Ic Ab
Dioscorea oppositifolifa L. Dioscoreaceae 2,3,4 D St No Ic Ab
Diploclisia glaucescens (Blume) Diels Menispermaceae 1,2 E St No Ic Bi
Dolichos trilobus L. Papilionaceae 2 D St Mi Cn Ab
Ehretia canarensis (Clarke) Gamble Boraginaceae 1 E SU Mi Ic Bi
Elaeagnus conferta Roxb. Elaeagnaceae 1 E SA Mi Ic Bi
Elaeagnus indica Servattaz Elaeagnaceae 1,2 Bd SU Mi Cn Bi
Ellertonia rheedii Wight Apocynaceae 2 E St Mi Cn Ab
Embelia basaal A.DC. Myrsinaceae 1,2 E SU Mi Cn Bi
Embelia ribes Burm.f. Myrsinaceae 2 Bd SU No Ic Bi
Entada pursaetha DC. Mimosaceae 1,2 E SU No Cn Hc
Erycibe paniculata Roxb. Convolvulaceae 1 E St Mi Cn Bi
Erythropalum populifolium (Arn.) Mast. Erythropalaceae 1 E St Me Ic Bi
Gloriossa superba L. Liliaceae 4 E Tc Me Cn Bi
Gnetum ula Brongn. Gnetaceae 1,2 E St No Ic Bi
Grewia flavescens Juss. Tiliaceae 2,3 Bd SU Mi Cn Bi
Grewia heterotricha Mast. Tiliaceae 2 Bd SU Mi Cn Bi
N. Parthasarathy et al.
8

Grewia opposistifolia Buch. -Ham. Tiliaceae 2,3 Bd SU Mi Cn Bi


Grewia rhamnifolia Heyne ex Roth Tiliaceae 1,2,3,4 Bd SU Mi Cn Bi
Gymnema elegans Wight & Arn. Asclepiadaceae 2 Bd St Mi Ic Ab
Gymnema hirsutum Wight & Arn. Asclepiadaceae 2 Bd St Mi Ic Ab
Gymnema sylvestre (Retz.) R.Br.ex Schultes Apocynaceae 2,3,4 Bd St Mi Ic Ab
Gymnema tingens (Roxb.) Wight & Arn. Asclepiadaceae 2 E St Mi Ic Ab
Gymnopetalum cochinchinense Kurz Cucurbitaceae 1 D Tc Me Cn Bi
Hemidesmus indicus (L.) R.Br. Asclepiadaceae 1 E St Mi Ic Ab
Hippocratea bourdillonii Gamble Hippocrateaceae 1 E St Mi Ic Ab
Hiptage benghalensis (L.) Kurz Malpighiaceae 1,2,3 Bd SU Me Cn Ab
Hugonia ferruginea Wight & Arn. Linaceae 1 Bd Hc Mi Cn Bi
Hugonia mystax L. Linaceae 2,3,4 Bd Hc Mi Cn Bi
Ichnocarpus frutescens (L.) R. Br. Apocynaceae 2,3,4 Bd St Mi Ic Ab
Icnocarpus publiflorus Hook. f. Apocynaceae 2 Bd St Mi Ic Ab
Ipomoea asarifolia (Desr.) Roem. & Schultes Convolvulaceae 2 Bd St No Cn Ab
Ipomoea staphylina Roem & Schultes Convolvulaceae 1,2,3,4 Bd St Me Cn Ab
Jasminum angustifolium (L.) Willd. Oleaceae 1,2,3,4 E St Mi Cn Bi
Jasminum auriculatum Vahl Oleaceae 2,3 Bd St Mi Cn Bi
Jasminum azoricum L. var. azoricum Oleaceae 1,2 E St No Cn Bi
Jasminum cuspidatum Rottl. Oleaceae 1,2 E St Mi Cn Bi
Jasminum malabaricum Wight Oleaceae 2,3 Bd St No Cn Bi
Jasminum multiflorum (Burm. f.) Andr. Oleaceae 1,2,3 E St Mi Cn Bi
Biodiversity of Lianas and Their Functional Traits in Tropical Forests…

Jasminum ritchiei Clarke Oleaceae 1 E St Mi Cn Bi


Jasminum rottlerianum Wall. ex A.DC. Oleaceae 1 E St Mi Cn Bi
Jasminum sambac (L.) Ait. Oleaceae 1 E St Mi Cn Bi
Jasminum scandens Vahl Oleaceae 1 E St Mi Cn Bi
Jasminum sessiliflorum Vahl Oleaceae 2,3,4 E St Mi Cn Bi
(continued)
141
Liana functional traits
142

Occurrence in WEF, Plant Climbing Leaf Flower Dispersal


Liana species Family SEF, SDF, DEF type mechanism type type mode
Jasminum trichotomum Heyne ex Roth Oleaceae 3 Bd St Mi Cn Bi
Kedrostis courtallensis (Arn.) Jeffrey Cucurbitaceae 2 D Tc Me Cn Bi
Kunstleria keralense Mohanan & Nair Papilionaceae 1 E St Ma Cn Ab
Lantana camara L. Verbenaceae 1,2,3 Bd SA Mi Cn Bi
Loeseneriella obtusifolia (Roxb.) A.C. Smith Hippocrateaceae 2,3 Bd SU No Ic Ab
Luvunga sarmentosa (Blume) Kurz Rutaceae 1 E SA Me Cn Bi
Maerua oblongifolia (Forsk.) A. Rich. Capparaceae 2,3 Bd SU Mi Cn Bi
Marsdenia brunoniana Wt. & Arn. Asclepiadaceae 1,2 Bd St Me Cn Ab
Marsdenia tenacissima (Roxb.) Moon Asclepiadaceae 2 E St Me Cn Ab
Maytenus heyneana (Roth) Raju & Babu Celastraceae 1,2 D SA MI IC BI
Maytenus royleanus (Wallich ex M. Lawson) M.A. Rau Celastraceae 2 D SA MI IC BI
Mimosa intsia L. Mimosaceae 1,2,3 Bd SA Me Cn Ab
Morinda reticulata Gamble Rubiaceae 1 E St Mi Cn Bi
Morinda umbellata L. Rubiaceae 1,2 E St Mi Cn Bi
Moullava spicata (Dalz.) Nicolson Caesalpiniaceae 1 Bd SA Me Cn Ab
Mucuna atropurpurea DC. Papilionaceae 1,2 D St Me Cn Ab
Mucuna pruriens (L.) DC. Papilionaceae 1,2 D St Me Cn Ab
Mussaenda belilla Buch.-Ham. Rubiaceae 1 E SU Mi Cn Bi
Mussaenda hirsutissima (Hook.f.) Hutch. ex Gamble Rubiaceae 2 E St Mi Cn Bi
Myxopyrum serratulum A.W. Hill Oleaceae 1 E SU Me Ic Bi
Naravelia zeylanica (L.) DC. Ranunculaceae 1,2 D Tc No Cn Ab
Neonotonia wightii (Wight & Arn.) Lackey Papilionaceae 2,3 D St Me Cn Ab
Olax scandens Roxb. Olacaceae 1,4 E SU Mi Cn Bi
Opilia amentacea Roxb. Opiliaceae 1 Bd SU Mi Ic Bi
Pachygone ovata (Poir.) Miers ex Hook. Menispermaceae 2,3,4 Bd St Mi Ic Bi
N. Parthasarathy et al.
8

Paramignya armata (Thw.) Oliver Rutaceae 1 Bd SA Me Cn Bi


Paramignya beddomei Tanaka Rutaceae 2 Bd SA Mi Cn Bi
Paramignya monophylla Wight Rutaceae 1 Bd SA Mi Cn Bi
Parsonsia alboflavescens (Dennst.) Mabberley Apocynaceae 1 E St Mi Cn Ab
Passiflora calcarata Mast. Passifloraceae 1 Bd Tc Mi Cn Bi
Passiflora subpeltata Ortega Passifloraceae 1 Bd Tc Mi Cn Bi
Phyllanthus reticulatus Poir. Euphorbiaceae 1,3 Bd SU Mi Ic Bi
Piper mullesua Buch.-Ham.ex D. Don Piperaceae 1 E St Mi Ic Bi
Piper nigrum L. Piperaceae 1,2 E Rc Me Ic Bi
Piper sp. Piperaceae 1 E Rc Mi Ic Bi
Pisonia aculeata L. Nyctaginaceae 1,2,3 Bd SA Mi Ic Ab
Plecospermum spinosum Trecul Moraceae 1,2,3,4 Bd SA Mi Cn Bi
Polygonum chinense L. Polygonaceae 1 E SU Mi Ic Bi
Pothos scandens L. Araceae 1 E Rc Mi Cn Bi
Premna corymbosa (Burm.f.) Rottl. & Willd. Verbenaceae 2,3,4 Bd SU Mi Ic Bi
Premna sp. Verbenaceae 2 Bd SU Mi Ic Bi
Premna villosa Clarke Verbenaceae 1,2,3 Bd SU Me Ic Bi
Premna wightiana Schauer Verbenaceae 1,2 Bd SU Me Ic Bi
Pseudaidia speciosa (Bedd.) Tirveng. Rubiaceae 1,2 E St Me Cn Bi
Pterolobium hexapetalum (Roth) Sant. & Wagh. Caesalpiniaceae 1,2,3,4 Bd SA Me Cn Ab
Pyrenacantha volubilis Wight Icacinaceae 4 E St Mi Ic Bi
Reissantia indica (Willd.) Halle Celastraceae 3,4 Bd SU Mi Ic Ab
Biodiversity of Lianas and Their Functional Traits in Tropical Forests…

Rivea hypocrateriformis (Desr.) Choisy Convolvulaceae 3,4 Bd St No Cn Bi


Rourea minor (Gaertn.) Alston Connaraceae 1 E St Ma Cn Bi
Rubus ellipticus Smith Rosaceae 1,2 E SA Me Cn Bi
Rubus micropetalus Gard. Rosaceae 1 Bd SA Me Cn Bi
Rubus niveus Thunb. Rosaceae 2 E SA Me Cn Bi
(continued)
143
Liana functional traits
144

Occurrence in WEF, Plant Climbing Leaf Flower Dispersal


Liana species Family SEF, SDF, DEF type mechanism type type mode
Sageretia filiformis (Schult.) Don Rhamnaceae 1,2 Bd SA Mi Ic Bi
Salacia beddomei Gamble Celastraceae 1 E St Me Ic Bi
Salacia chinensis L. Celastraceae 1,2,4 Bd SA Mi Ic Bi
Salacia malabarica Gamble Hippocrateaceae 1 E SU Me Ic Bi
Salacia oblonga Wal. ex Wight & Arn. Hippocrateaceae 1 Bd SA No Ic Bi
Salacia sp. Hippocrateaceae 1 E St Me Ic Bi
Sarcostemma acidum (Roxb.) Voigt Asclepiadaceae 2,3,4 E SU - Cn Ab
Sarcostigma kleinii Wight & Arn. Icacinaceae 1 E St Mi Ic Bi
Schefflera venulosa (Wight & Arn.) Harms. Araliaceae 1 E St Ma Ic Bi
Scutia myrtina (Brum.f.) Kurz Rhamnaceae 1,2,3,4 E SA Mi Ic Bi
Secamone emetica (Retz.) R.Br. ex Schultes Asclepiadaceae 1,2,3,4 E St Mi Ic Ab
Smilax perfoliata Lour. Smilacaceae 1 Bd Tc No Cn Bi
Smilax sp. Liliaceae 1 Bd Tc No Cn Bi
Smilax zeylanica L. Liliaceae 1,2 Bd Tc No Cn Bi
Solanum seaforthianum Andr. Solanaceae 2 E St Me Cn Bi
Strychnos dalzellii C.B. Clarke Loganiaceae 1 E St Mi Ic Bi
Strychnos minor Dennst. Loganiaceae 4 E St Mi Ic Bi
Strychnos vanprukii Craib Loganiaceae 1 E St Mi Ic Bi
Symphorema involucratum Roxb. Verbenaceae 4 D SU No Ic Bi
Tetrastigma leucostaphylum (Dennst.) Alston Vitaceae 1 E Tc Ma Ic Bi
Tetrastigma sulcatum (Lawson) Gamble Vitaceae 1 E Tc Ma Ic Bi
Tiliacora acuminata (Lam.) Hook. f. & Thoms. Menispermaceae 4 D St Me Ic Bi
Tinospora cordifolia (Willd.) J.D.Hook. & Thoms. Menispermaceae 1,2,3,4 D St Me Ic Bi
Toddalia asiatica (L.) Lam. Rutaceae 1,2,3,4 Bd SA Mi Ic Bi
Trichosanthes anaimalaiensis Bedd. Cucurbitaceae 1,2 D Tc Me Cn Bi
N. Parthasarathy et al.
8

Trichosanthes tricuspidata Lour. Cucurbitaceae 4 E Tc Me Cn Bi


Tylophora indica (Burm. f.) Merr. Asclepiadaceae 1,4 E St No Ic Ab
Tylophora subramanii Henry Asclepiadaceae 2 Bd St No Cn Ab
Unona viridiflora Bedd. Annonaceae 1 E SU Me Cn Bi
Uvaria narum (Dunal) Wall. ex Wight Annonaceae 1 E SU No Cn Bi
Ventilago bombaiensis Dalz. Rhamnaceae 1 Bd St Mi Ic Ab
Ventilago goughii Gamble Rhamnaceae 2 Bd St Mi Ic Ab
Ventilago madraspatana Gaertn. Rhamnaceae 1,2,3,4 Bd St Mi Ic Ab
Wattakaka volubilis (L.f.) T. Cooke Asclepiadaceae 2,3,4 Bd St Me Cn Ab
Zanthoxylum ovalifolium Wight Rutaceae 1,2 E SA Me Cn Ab
Zanthoxylum tetraspermum Wight & Arn. Rutaceae 1,2 E SA Me Cn Ab
Ziziphus horrida Roth. Rhamnaceae 2,3 Bd SA Mi Ic Bi
Ziziphus oenoplia (L.) Mill. Rhamnaceae 1,2,3,4 Bd SA Mi Ic Bi
Ziziphus rugosa Lam. Rhamnaceae 1,2 Bd SA Mi Ic Bi
Biodiversity of Lianas and Their Functional Traits in Tropical Forests…
145
146 N. Parthasarathy et al.

References

Addo-Fordjour P, Anning AK, Atakora EA, Agyei PS (2008) Diversity and distribution of climbing
plants in a semi-deciduous rain forest, KNUST Botanic Garden, Ghana. Int J Bot 4:186–195
Addo-Fordjour P, Rahmad ZB, Shahrul AMS (2013) Environmental factors influencing liana com-
munity diversity, structure and habitat associations in a tropical hill forest, Malaysia. Plant Ecol
Divers 7:1–12
Anbarashan M, Parthasarathy N (2013) Diversity and ecology of lianas in tropical dry evergreen
forests on the Coromandel Coast of India under various disturbance regimes. Flora 208:22–32
Balfour DA, Bond WJ (1993) Factors limiting climber distribution and abundance in a southern
African forest. J Ecol 81:93–99
Bhattarai KR, Vetaas OR (2003) Variation in plant species richness of different life-forms along a
subtropical elevation gradient in the Himalayas, east Nepal. Global Ecol Biogeogr
12:327–340
Cai ZQ, Schnitzer SA, Bongers F (2009) Liana communities in three tropical forest types in
Xishuangbanna, South-West China. J Trop For Sci 21:252–264
Castellanos AE (1991) Photosynthesis and gas exchange of vines. In: Putz FE, Mooney HA (eds)
The biology of vines. Cambridge University Press, Cambridge, pp 181–202
Castellanos AE, Mooney HA, Bullock SH, Jones C, Robichaux R (1989) Leaf, stem and metamer
characteristics of vines in a tropical deciduous forest in Jalisco, Mexico. Biotropica 21:41–49
Chalmers AC, Turner JC (1994) Climbing plants in relation to their supports in a stand of dry rain
forest in the Hunter Valley, New South Wales. Proc Linn Soc NSW 114:73–89
Chettri A, Barik SK, Pandey HN et al (2010) Liana diversity and abundance as related to microen-
vironment in three forest types located in different elevational ranges of the eastern Himalayas.
Plant Ecol Divers 3:175–185
DeWalt SJ, Schnitzer SA, Denslow JS (2000) Density and diversity of lianas along a chronose-
quence in central Panamanian lowland forest. J Trop Ecol 16:1–19
DeWalt S, Schnitzer SA, Chave J et al (2009) Annual rainfall and seasonality predict pan-tropical
patterns of liana density and basal area. Biotropica 42:309–317
Dıaz S, Noy-Meir I, Cabido M (2001) Can grazing response of herbaceous plants be predicted
from simple vegetative traits? J Appl Ecol 38:497–508
Ding Y (2006) Study on recovery ecology of the degraded tropical forest vegetation in Hainan
Island, South China. In: Chinese Academy of Forestry, Beijing (in Chinese)
Edwards W, Moles AT, Franks P (2007) The global trend in plant twining direction. Glob Ecol
Biogeogr 16:795–800
Emmons LH, Gentry AH (1983) Tropical forest structure and the distribution of gliding and
prehensile-tailed vertebrates. Am Nat 121:513–524
Ewango CEN (2010) The liana assemblage of a Congolian rainforest: diversity, structure and
dynamics. Ph.D. thesis, Wageningen University
Gamble JS, Fischer CEC (1915–1935) Flora of the Presidency of Madras, vols I–III. Adlard and
Son, London
Garrido-Perez EI, Dupuy JM, Duran-Garcia R et al (2008) Effects of lianas and Hurricane Wilma
on tree damage in the Yucatan Peninsula, Mexico. J Trop Ecol 24:559–562
Gentry AH (1982) Patterns of Neotropical plant species diversity. In: Hecht MK, Wallace B,
Prance GT (eds) Evolutionary biology. Plenum Press, New York, pp 1–84
Gentry AH (1988) Tree species richness of upper Amazonian forests. Proc Natl Acad Sci U S A
85(1):156–159
Gentry AH (1991a) Breeding and dispersal systems of lianas. In: Putz FE, Mooney HA (eds) The
biology of vines. Cambridge University Press, Cambridge, pp 393–426
Gentry AH (1991b) The distribution and evolution of climbing plants. In: Putz FE, Mooney HA
(eds) The biology of vines. Cambridge University Press, Cambridge, pp 3–49
Givinish TJ, Vermeij GJ (1976) Sizes and shapes of liane leaves. Am Nat 110:743–778
8 Biodiversity of Lianas and Their Functional Traits in Tropical Forests… 147

Hartshorn GS, Hammel BE (1994) Vegetation types of floristic patterns. In: Mcdade LA, Bawa
KS, Hespenheide HA, Harthsorn GS (eds) La Selva: ecology and natural history of Neotropical
rainforest. The University of Chicago Press, Chicago, pp 73–89
Hashimoto T (2002) Molecular genetic analysis of left-right handedness in plants. Philos Trans R
Soc B Biol Sci 357:799–808
Hegarty EE, Caballe G (1991) Distribution and abundance of vines in forest communities. In: Putz
FE, Mooney HA (eds) The biology of vines. Cambridge University Press, Cambridge,
pp 313–334
Hooper DU, Chapin FS, Ewel JJ, Hector A, Inchausti P, Lavorel S, Lawton JH, Lodge DM, Loreau
M, Naeem S, Schmid B, Setala H, Symstad AJ, Vandermeer J, Wardle DA (2005) Effects of
biodiversity on ecosystem functioning: a consensus of current knowledge. Ecol Monogr
75:3–35
Ibarra-Manriquez G, Martinez-Ramos M (2002) Landscape variation of liana communities in a
Neotropical rain forest. Plant Ecol 160:91–112
Jiminez-Castillo M, Wiser SK, Lusk CH (2007) Elevational parallels of latitudinal variation in the
proportion of lianas in woody floras. J Biogeogr 34:163–168
Kato R, Tadakai Y, Ogalya H (1978) Plant biomass and growth increment studies in Pasoh Forest.
Malay Nat J 30:211–224
Kurzel BP, Schnitzer SA, Carson WP (2006) Predicting liana crown location from stem diameter
in three Panamanian lowland forests. Biotropica 38:262–266
Laurance WF, Perez-Salicrup P, Delamonica P et al (2001) Rain forest fragmentation and the struc-
ture of Amazonian liana communities. Ecology 82:105–116
Mateucci S (1987) The vegetation of Falcon state, Venezuela. Vegetation 70:67–91
Matthew KM (1991) An excursion flora of Central Tamil Nadu, India. Oxford and IBH, New Delhi
McGill BJ, Enquist BJ, Weiher E, Westoby M (2006) Rebuilding community ecology from func-
tional traits. Trends Ecol Evol 21:178–185
Molina-Freaner F, Tinoco-Ojanguren C (1997) Vines of a desert plant community in Central
Sonora, Mexico. Biotropica 29:46–56
Nabe-Nielsen J (2001) Diversity and distribution of lianas in a Neotropical rain forest, Yasuni
National Park, Ecuador. J Trop Ecol 17:1–19
Nabe-Nielsen J, Hall P (2002) Environmentally induced clonal reproduction and life history traits
of the liana Machaerium cuspidatum in an Amazonian rain forest, Ecuador. Plant Ecol
162:215–226
Parkhurst D, Loucks O (1972) Optimal leaf size in relation to environment. J Ecol 60:505–537
Parthasarathy N, Muthuramkumar S, Reddy MS (2004) Patterns of liana diversity in tropical ever-
green forests of peninsular India. For Ecol Manage 190:15–31
Pereira R Jr, Zweede J, Asner GP et al (2002) Forest canopy damage and recovery in reduced-
impact and conventional selective logging in eastern Para, Brazil. For Ecol Manage
168:77–89
Perez-Salicrup DR (2001) Effect of liana cutting on tree regeneration in a liana forest in Amazonian
Bolivia. Ecology 82:389–396
Perez-Salicrup DR, de Meijere W (2005) Number of lianas per tree and number of trees climbed
by lianas at Los Tuxtlas, Mexico. Biotropica 37:153–156
Phillips O, Martinez VA et al (2002) Increasing dominance of large lianas in Amazonian forests.
Nature 418:770–774
Putz FE (1983) Liana biomass and leaf area of a “Tierra Firme” Forest in the Rio Negro Basin,
Venezuela. Biotropica 15:185–189
Putz FE (1984) The natural history of Lianas on Barro Colorado Island, Panama. Ecology
65:1713–1724
Putz FE, Chai P (1987) Ecological studies on lianas in Lambir National Park, Sarawak, Malaysia.
J Ecol 75:523–531
Putz FE, Holbrook NM (1991) Biomechanical studies of vines. In: Putz FE, Mooney HA (eds) The
biology of vines. Cambridge University Press, Cambridge, UK, pp 73–97
148 N. Parthasarathy et al.

Putz FE, Windsor DM (1987) Liana phenology on Barro Colorado Island, Panama. Biotropica
19:334–341
Quigley MF, Platt WJ (2003) Composition and structure of seasonally deciduous forests in the
Americas. Ecol Monogr 73:87–106
Reddy MS, Parthasarathy N (2006) Liana diversity and distribution on host trees in four inland
tropical dry evergreen forests of peninsular India. Trop Ecol 47:109–123
Restom TG, Nepstad DC (2001) Contribution of vines to the evapotranspiration of a secondary
forest in eastern Amazonia. Plant Soil 236:155–163
Restom TG, Nepstad DC (2003) Seedling growth dynamics of a deeply-rooting liana in a second-
ary forest in eastern Amazonia. For Ecol Manage 190:109–118
Rice K, Brokaw N, Thompson J (2004) Liana abundance in a Puerto Rican forest. For Ecol Manage
190:33–41
Roldan AI (1999) Seasonal changes in liana cover in the upper canopy of a Neotropical dry forest.
Biotropica 31:186–192
Roscher C, Schumacher J, Gubsch M, Lipowsky A, Weigelt A, Buchmann N, Schmid B, Schulze
E-D (2012) Using plant functional traits to explain diversity–productivity relationships. PLoS
One 7(5):e36760. doi:10.1371/journal.pone.0036760
Sarmiento G (1972) Ecological and floristic convergences between seasonal plant formations of
tropical and subtropical South America. J Ecol 60:367–410
Schnitzer SA (2005) A mechanistic explanation for global patterns of liana abundance and distri-
bution. Am Nat 166:262–276
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
Schnitzer SA, Bongers F (2011) Increasing liana abundance and biomass in tropical forests:
emerging patterns and putative mechanisms. Ecol Lett 14:397–406
Schnitzer SA, Carson WP (2001) Treefall gaps and the maintenance of species diversity in a tropi-
cal forest. Ecology 82:913–919
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical
forest canopy gaps: evidence for an alternate pathway of gap-phase regeneration. J Ecol
88:655–666
Schnitzer SA, DeWalt SJ, Chave J (2006) Censusing and measuring lianas: a quantitative compari-
son of the common methods. Biotropica 38:581–591
Schnitzer SA, Mangan SA, Dalling JW (2012) Liana abundance, diversity, and distribution on
Barro Colorado Island, Panama. PLoS One 7:e52114
Senbeta F, Schmitt C, Denich M et al (2005) The diversity and distribution of lianas in Afromontane
rain forests of Ethiopia. Divers Distrib 11:443–452
Swaine MD, Grace J (2007) Lianas may be favored by low rainfall: evidence from Ghana. Plant
Ecol 192:271–276
Tilman D, Knops J, Wedin D, Reich P, Ritchie M, Siemann E (1997) The influence of functional
diversity and composition on ecosystem processes. Science 277:1300–1302
Vidal E, Johns J, Gerwing JJ et al (1997) Vine management for reduced-impact logging in eastern
Amazonia. For Ecol Manage 98:105–114
Violle C, Navas M-L, Vile D, Kazakou E, Fortunel C, Hummel I, Gamier E (2007) Let the concept
of trait be functional! Oikos 116:882–892
Wright SJ, Calderon O, Hernandez A et al (2004) Are lianas increasing in importance in tropical
forests? A 17-year record from Panama. Ecology 85:484–489
Yuan CM, Liu WY, Tang CQ et al (2009) Species composition, diversity and abundance of lianas
in different secondary and primary forests in a subtropical mountainous area, SW China. Ecol
Res 24:1361–1370
Chapter 9
The Contribution of Lianas to Forest Ecology,
Diversity, and Dynamics

Stefan A. Schnitzer

Abstract Lianas are a common component of forests worldwide and they contribute
to forest ecology, diversity, and dynamics. Lianas can have both positive and
negative effects in forests. Lianas can be an important resource for animals, as food
(in the form of nectar, pollen, fruits, leaves, or sap), providing nesting sites, shelter
and, by climbing among many tree crowns, lianas can provide aerial highways for
many animal species. By contrast, lianas also compete intensively with trees, reduc-
ing tree recruitment, growth, reproduction, and survival, as well as tree diversity and
forest-level carbon sequestration. While the inclusion of lianas in ecological studies
have lagged behind that of trees, over the past three decades, the study of liana ecol-
ogy has grown significantly, revealing many important contributions of lianas to
forest ecology. In this chapter, I review the state of knowledge about the ecology of
lianas and their contribution to forest ecology, diversity, and dynamics.

9.1 Introduction

Lianas are a common component of forests world-wide where they contribute to


many aspects of forest ecology, diversity, and dynamics. Lianas can add consider-
ably to forest structure and woody plant diversity, especially in seasonally dry low-
land tropics, where lianas can represent in excess of 25 % of the woody stems and
more than one-third of the woody species (e.g., Schnitzer et al. 2012, 2015a). Acting
as an important resource for animals by providing nesting sites and shelter, and in
addition, by climbing and intertwining themselves within many tree crowns, lianas
can serve as aerial pathways for many animal species (Yanoviak 2015). Lianas are
also a source of food in the form of nectar, pollen, fruits, leaves, or sap, thus benefit-
ting many animal species (Arroyo-Rodriguez et al. 2015).

S.A. Schnitzer (*)


Department of Biological Sciences, Marquette University,
PO Box 1881, Milwaukee, WI 53201-1881, USA
Smithsonian Tropical Research Institute, Apartado,
Balboa 2072, Republic of Panama
e-mail: S1@uwm.edu

© Springer International Publishing Switzerland 2015 149


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_9
150 S.A. Schnitzer

By contrast, a considerable amount of evidence now indicates that lianas also


have large negative impacts on some aspects of tropical forest functioning. Lianas
compete intensely with trees, reducing tree growth, survival, regeneration, and fecun-
dity. By limiting tree growth and increasing tree mortality, lianas can reduce the
amount of carbon that trees sequester and store, thereby lowering the carbon storage
capacity of forests (Durán and Gianoli 2013; van der Heijden et al. 2013; Schnitzer
et al. 2015a). Lianas may also deplete soil nutrients and soil moisture, further affect-
ing tree communities and ecosystem functioning (Powers 2015). Moreover, lianas
are now increasing in abundance and biomass in neotropical forests (Schnitzer and
Bongers 2011; Schnitzer 2015), an outcome that may have substantial consequences
for forest diversity, composition, and ecosystem-level functioning.
While the study of liana ecology has lagged behind that of trees, over the past
three decades our knowledge of the role of lianas in tropical forests has grown sig-
nificantly (Schnitzer et al. 2015b). In this chapter, we review the state of knowledge
on the ecology of lianas and their contribution to tropical forest ecology, diversity,
and dynamics. We divide the chapter into three main sections. (1) Positive interac-
tions between lianas and animals in tropical forests - an aspect of liana ecology that
is often overlooked, yet is important and worthy of discussion. (2) The negative inter-
actions between lianas and trees in tropical forests. Lianas have long been suspected
of having a negative effect on tree growth and reproduction, but it is only recently
that we have gained a solid and quantitative understanding of the magnitude of that
effect. The role of lianas on tree and whole-forest biomass uptake is a relatively new
area of study and, although lianas themselves store a relatively small amount of for-
est biomass, recent evidence demonstrates that lianas may have a surprisingly large
effect on the biomass dynamics of trees. (3) The distribution of liana stem density
and species diversity within and among tropical forests. Determining the factors that
control the distribution of organisms is a fundamental goal of ecology and recent
studies are revealing that the seemingly different factors responsible for local and
regional liana distributions may share surprising mechanistic similarities. Although
lianas may also play a key role in the ecology of both temperate forests (Ladwig and
Meiners 2015) and higher-altitude tropical forests (Jimenez-Castillo et al. 2007), here
we restrict the discussion to lianas of lowland tropical forests because of the clear
importance of lianas in these forests.

9.2 Positive Contributions of Lianas to Forest Ecosystems

Lianas have a number of positive effects in forest ecosystems, and these effects can
be viewed in the context of the contribution of lianas to: (1) species diversity;
(2) stem density; (3) forest structure; and (4) resources used by wildlife. Lianas
contribute substantially to plant species diversity, and the evolution of climbing may
be a key innovation that allows for the diversification (and thus high diversity) of
9 The Contribution of Lianas to Forest Ecology, Diversity, and Dynamics 151

climbing plant lineages (Gianoli 2004). On Barro Colorado Island, Panama (BCI),
lianas comprised 35.5 % of the woody species even though they were only 25 % of
the woody stems. Thus, lianas significantly expand the vascular plant diversity in
lowland tropical forests.
The contribution of lianas to plant stem density and the architectural complexity
of tropical forests can be considerable. Lianas commonly contribute 25 % of the
rooted woody stems in tropical forests (Parthasarathy et al. 2015; Schnitzer et al. 2015a)
and, in most tropical forests, lianas are found in the crowns of ~75 % of the canopy
trees (Ingwell et al. 2010; van der Heijden et al. 2013). Lianas use the structure of
trees to ascend into the forest canopy (Putz 1984; Schnitzer et al. 2012, 2015a). By
the time that a liana stem in a seasonal tropical forest reaches 3 cm in diameter, it
has a 90 % probability of having its crown in the forest canopy; lianas reach the for-
est canopy of wetter forests at even smaller diameters (Kurzel et al. 2006).
Lianas, however, typically do not climb straight up to the forest canopy; rather,
they loop through the understory where they add considerably to the non-vertical
structure and architectural complexity to the forest. Once in the canopy, larger lia-
nas often cross from tree crown to tree crown, adding connections to the forest
canopy that are used by many arboreal animal species to move throughout the forest
(Arroyo-Rodriguez et al. 2015; Yanoviak 2015). By draping from the side of tree
crowns, lianas can fill spaces between trees, adding even more structure to tropical
forests. Lianas also serve to tie the forest canopy together, which can reduce the
number of trees that fall during wind-storms (Garrido-Perez et al. 2008). In the
understory, liana stems form dense tangles following disturbance (e.g., Schnitzer
et al. 2000; Ledo and Schnitzer 2014), providing structure, forage, and an escape
from predators for many bird and small mammal species (e.g., Lambert and Halsey
2015; Michel et al. 2015).
Lianas also contribute substantially to the suite of resources that are important
to forest animals. Both vertebrates and invertebrates consume liana leaves, flowers,
and fruits throughout the year. However, the phenology of many liana species is
asynchronous with respect to trees, thus providing a rich resource base for many
animal species when trees are producing few leaves, fruits or seeds. Specifically,
many liana species produce leaves and flowers during the dry season, which may
provide a particularly critical resource for animals that depend on a steady supply
of fruits, seeds, and young leaves. Indeed, many primate species switch their diet to
include far more lianas during the dry season (Arroyo-Rodriguez et al. 2015).
Lianas are the source of many critical ingredients for products that are consumed
by humans throughout the world. These products include melons, passion fruit,
gourds, and a multitude of legumes; stimulants and medicines (e.g., guarana from
Paullinia cupana and curare from Strychnos toxifera and Chondrodendron tomento-
sum); and the fundamental ingredients for the production of wine and cognac from
grapes (Vitis) and beer from hops (Humulus lupulus) (Schnitzer 2015). Therefore,
lianas make a wide variety of contributions to forest plant diversity, structure, and
wildlife, and they have also established themselves as an important component of
human society.
152 S.A. Schnitzer

9.3 Negative Competitive Effects of Lianas


in Forest Ecosystems

More than two-dozen studies over the past 20 years have demonstrated unequivo-
cally that lianas reduce the performance of tropical forests. Lianas compete intensely
with trees, reducing tree recruitment, growth, fecundity, and survival (e.g., Schnitzer
et al. 2005; Ingwell et al. 2010; Schnitzer and Carson 2010; van der Heijden et al.
2013; Kainer et al. 2014). Grauel and Putz (2004) removed lianas in portions of a
seasonally inundated monodominant forest in the Darien region of Panama and
reported that after 5 years, canopy tree diameter growth had doubled compared to
canopy trees in control plots. In a forest in Côte d’Ivoire in West Africa, tree sap-
lings grew more than five-times faster in plots when lianas were removed compared
to plots in which saplings were competing with lianas (Schnitzer et al. 2005). In a
liana removal experiment in treefall gaps in central Panama, Schnitzer and Carson
(2010) reported that removing lianas for 8 years increased tree recruitment by 45 %
and tree growth by 55 %. This finding translates to mean annual effect of lianas on
tree recruitment and growth of 5.6 % and 6.9 %, respectively. Two additional studies
in central Panama demonstrated that lianas actively reduce tree growth performance
and that immediately following liana cutting, sap velocity and growth of canopy
trees increased substantially and significantly (Tobin et al. 2012; Alvarez-Cansino
et al. 2015). Lianas also appear to increase tree mortality, and two studies provide
compelling evidence that when lianas are able to heavily infest trees, the probability
of mortality increases significantly (Phillips et al. 2005; Ingwell et al. 2010).
Competition from lianas may also reduce the diversity of tree species regenerat-
ing in treefall gaps. Schnitzer and Carson (2010) reported that shade-tolerant tree
species diversity increased over 65 % faster in treefall gaps without lianas than in
control gaps with lianas. Gaps provide a heterogeneous and resource-rich environ-
ment that could allow tree species to partition resource (reviewed by Brokaw and
Busing 2000). Gaps also promote the density and diversity of lianas (e.g., Putz
1984; Schnitzer et al. 2000; Schnitzer and Carson 2001; Ledo and Schnitzer 2014).
Lianas in gaps compete intensely with trees, suppressing tree colonization, survival,
density, and ultimately their diversity (Schnitzer and Carson 2010). Biotic interfer-
ence in the form of liana competition reduced tree species richness in gaps to the
point that it may have precluded the possibility of tree community level niche
partitioning, as had been the theoretical prediction (Schnitzer and Carson 2000,
2010).

9.3.1 Lianas and Forest Carbon Dynamics

By competing with trees, which typically store more than 90 % of aboveground forest
carbon, lianas can dramatically reduce rates of forest carbon uptake (Durán and
Gianoli 2013; van der Heijden et al. 2013). Forest areas with high liana densities tend
9 The Contribution of Lianas to Forest Ecology, Diversity, and Dynamics 153

to have far lower biomass than forest areas with low liana densities, suggesting that
lianas reduce whole-forest carbon accumulation and storage (e.g., Chave et al.
2001; Durán and Gianoli 2013). This relationship was confirmed in a liana removal
experiment in central Panama, where Schnitzer et al. (2014) demonstrated that lia-
nas reduced tree biomass accumulation in gaps by 180 % over an 8-year period (a
22 % annual reduction in biomass accumulation). By contrast, lianas themselves
accumulated and stored very little carbon, and were able to compensate for only
24 % of the biomass accumulation that they displaced in trees (Schnitzer et al.
2014). Scaling up to the whole-forest level, the annual loss in biomass due to lianas
in gaps can be nearly 20 % of total forest biomass (Schnitzer et al. 2014). Consistent
with these findings, van der Heijden and Phillips (2009) observed that lianas in the
Peruvian Amazon reduced annual tree biomass increase by approximately 10 %,
and that annual liana biomass increase compensated for only around 29 %of the
biomass displaced in trees. Consequently, lianas can dramatically reduce the capac-
ity of tropical forests to uptake and store carbon.
An implicit assumption of most theories is that competition among plants is a
zero-sum-game in terms of carbon uptake. That is, one plant’s carbon loss is another
plant’s carbon gain. This assumption does not appear to hold when there is competi-
tion among contrasting growth forms that inherently store different amounts of
carbon. The example with lianas illustrates that competition between lianas and
trees results in a net carbon loss for tropical forests. Tropical forests store over 30 %
of aboveground terrestrial forest carbon worldwide, and nearly all of this carbon is
stored in trees. Therefore, the negative effects of lianas on tree carbon accumulation
and storage have ramifications for the global carbon cycle. Furthermore, the ongoing
increase in liana abundance in neotropical forests (Schnitzer and Bongers 2011;
Schnitzer 2015) will likely amplify the effects of lianas on global carbon dynamics.

9.3.2 The Effects of Lianas on Tree Species Diversity


and Community Composition

Lianas may alter tree species diversity and community composition by competing
more intensely with some tree species than others (Schnitzer and Bongers 2002; van
der Heijden and Phillips 2009; Schnitzer and Carson 2010). Indeed, a small subset
of gap specialist tree species, particularly those that grow extremely rapidly and
have monopodial growth with few branches and large leaves are able to escape colo-
nization by lianas (e.g., Putz 1984; Clark and Clark 1990; Schnitzer and Carson
2010; van der Heijden et al. 2015). However, lianas appear to compete with the vast
majority of tree species. For example, Alvarez-Cansino et al. (2015) examined the
effect of large-scale liana cutting on sap-flow of canopy trees in a forest in central
Panama and reported that lianas had a negative effect on all seven species examined.
In this same experimental infrastructure, Martinez-Izquierdo et al. (in review)
examined the effects of liana cutting on seedlings of 14 well-replicated tree species
154 S.A. Schnitzer

and found that all species responded similarly. However, the real test of whether
lianas influence tree species diversity and community composition is whether the
effect of lianas on tree population growth rates varies with tree species identity.

9.4 The Global Distribution of Lianas

Determining the mechanisms responsible for the abundance and distribution of


organisms is one of the fundamental goals of ecology and lianas are an interesting
functional group of species to investigate this goal because of their clear distribu-
tional pattern. Liana density and diversity vary predictably within and among for-
ests, and the predominant determinant of liana distribution may depend on spatial
scale (Schnitzer 2005). Within forests, liana diversity and distribution appear to be
maintained by disturbance, mostly in form of treefall gaps (Schnitzer et al. 2000;
Schnitzer and Carson 2001; Dalling et al. 2012; Ledo and Schnitzer 2014). Among
forests, however, the relative distribution and diversity of lianas (compared to trees)
is maintained by the amount and seasonality of rainfall. Below I provide evidence
for both of these patterns, along with the most likely ecological explanations.

9.4.1 Forest-Level (Local) Distribution of Liana Density


and Diversity

Disturbance appears to be the main determinant for the distribution of liana species
and the maintenance of liana diversity at the local (within forest-level) scale. The
most common form of repeatable natural disturbance in many tropical forests is the
death of a canopy tree and the resulting treefall gap (e.g., Schnitzer et al. 2000).
Lianas accumulate in large numbers in treefall gaps, where they grow rapidly in the
high light environment. For example, treefall gaps explained the distribution of more
than 50 % of the liana species in the BCI 50-ha forest dynamics plot in Panama,
(Dalling et al. 2012; Schnitzer et al. 2012, 2015a; Ledo and Schnitzer 2014). The
combination of disturbance and positive density dependence (a measure of liana spa-
tial clumping) explained the within-forest distribution of nearly 80 % of the liana
species on the BCI 50 ha plot (Ledo and Schnitzer 2014). Furthermore, liana diversity
was always higher in treefall gaps than in non-gap sites, even when controlling for
area and stem density (Schnitzer and Carson 2001; Ledo and Schnitzer 2014).
Lianas may respond rapidly to disturbance because of their ability colonize by
creating many clonal stems in the high light environment following disturbance.
The majority of liana species on the BCI 50 ha plot produced significantly more
clonal stems in disturbed areas than in intact, undisturbed forest (Schnitzer et al.
2012, 2015a, Ledo and Schnitzer 2014). Ledo and Schnitzer (2014) tested the rela-
tive role of disturbance against two other putative diversity maintenance mechanisms
9 The Contribution of Lianas to Forest Ecology, Diversity, and Dynamics 155

(niche partitioning and negative density dependence) and found that disturbance
explained the vast majority of local liana diversity and distribution, whereas the
other two mechanisms explained very little variation in liana distribution and diversity.
In summary, disturbance appears to be the main determinant of local liana distribution
and the maintenance of liana diversity in tropical forests.

9.4.2 Pan-tropical Distribution of Relative Liana Density


and Diversity

Expanding the scale of liana distribution from within lowland tropical forests to
among lowland tropical forests, liana density increases as mean annual precipitation
decreases. Thus, lianas peak in density in seasonally dry forests, where mean annual
rainfall is relatively low and the number of dry months is high (Schnitzer 2005;
DeWalt et al. 2010, 2015). By contrast, the density of trees (and most other plant
functional groups) increase with increasing precipitation in tropical forests (Fig. 9.1;
from Schnitzer 2005). This pattern was first demonstrated by Schnitzer (2005) using
data collected by A.H. Gentry, who surveyed lianas (2.5 cm diameter) in small
transects (0.1 ha) in forests around the world (Africa, n = 8; Asia, n = 4; Central
America, n = 9; and South America, n = 45). Even considering the very low level of
within forest sampling (and thus high variation among forests), the Gentry data

500

TREES
400
Liana and tree density
(individuals 0.1 ha-1)

300

200

LIANAS
100

0
0 1000 2000 3000 4000 5000 6000 7000 8000
Mean annual precipitation (mm)

Fig. 9.1 The change in lianas and trees (2.5 cm diameter) with mean annual precipitation in
lowland tropical forests from Africa (n = 8), Asia (n = 4), Central America (n = 9), and South
America (n = 45). Triangles represent trees and circles represent lianas. Closed symbols are neo-
tropical sties (Central and South America) and open symbols are old-world sites (Africa and Asia).
Data were collected by A.H. Gentry in 10 small (2 × 50 m) transects (0.1 ha total) in each forest
(Figure from Schnitzer (2005))
156 S.A. Schnitzer

revealed a clear decrease in liana density with increasing mean annual rainfall
(Fig. 9.1). This pattern was confirmed by DeWalt et al. (2010, 2015) using data from
29 relatively well-sampled forest plots around the world: Africa (n = 3), Asia (n = 9),
Mexico (n = 2), Central America (n = 4), and South America (n = 11). Thus, liana density
increases with increasing seasonality and decreasing precipitation among forests.
The distribution of lianas occurs within the context of their tree competitors,
and thus examining how liana density changes relative to trees is more useful to
determine the abundance and distribution of lianas. The density of lianas relative to
trees increases steeply with increasing seasonality and decreasing precipitation
(Schnitzer et al. in prep). The unique distribution of lianas may be explained by the
ability of lianas to grow more than competing trees during the dry season, which
may confer a competitive advantage for lianas (Schnitzer 2005). This hypothesis is
now supported empirically by a number of studies that show that lianas are more
physiologically active than trees during the dry season relative to the wet season
(Cai et al. 2009; Zhu and Cao 2009; Wyka et al. 2013; Alvarez-Cansino et al. 2015;
Chen et al. 2015), and thus the pattern of liana distribution appears to be controlled
predominantly by the ability of lianas to grow when their tree competitors are
largely dormant. By contrast, lianas may be in much lower abundance in aseasonal,
wet forests, where water is rarely limiting and radiation is relatively low year-round,
and thus lianas gain no such advantage (Schnitzer 2005). Integrated over many
decades, the dry season growth advantage may allow lianas to increase in abundance
relative to trees in seasonal forests compared to non-seasonal wet forests, thus
explaining their pan-tropical distribution.
Liana diversity (species richness) relative to that of trees follows the same pattern
among forests as relative liana density. For example, using the Gentry dataset,
Schnitzer et al. (in prep) found that liana species richness (relative to tree species
richness) increases sharply with increasing seasonality and decreasing precipitation
among forests. The same pattern was also reported from Ghana, where lianas varied
from 30 % of the vascular plant species in the wetter forests (2,000 mm annual
rainfall) to 43 % of the vascular plant species in the drier forests (1,000 mm annual
rainfall; Swaine and Grace 2007). Consequently, the distribution of liana density
and diversity may be explained by their competitive advantage in seasonal tropical
forests.

9.4.3 Reconciling Within- and Among-Forest


Liana Distributions

The peak in liana density and diversity in treefall gaps (within-forests) and in
seasonal forests (among-forests) may both be explained by the unique ecological,
anatomical, and physiological traits of lianas (Schnitzer 2005). While there is a lot
of functional diversity among liana species, the liana growth form tends to converge
upon certain shared characteristics. For example, nearly all lianas have thin stems
with a relatively large mass of leaves that are typically deployed at the very top of
the forest canopy (Wyka et al. 2013). Lianas also have extraordinarily large vessel
9 The Contribution of Lianas to Forest Ecology, Diversity, and Dynamics 157

elements modified for highly efficient water movement, which enables them to
effectively supply water to their leaves (e.g., Wyka et al. 2013). Lianas may also
have deep and efficient root and vascular systems, which allow them to acquire
more soil water and nutrients and suffer less water-stress during the dry season
compared to many tree species (Restom and Nepstad 2004; Andrade et al. 2005;
Chen et al. 2015). Thus, the unique anatomical and physiological traits of lianas
may allow them to exploit seasonal forests more effectively than can trees, thus
explaining their large-scale, among-forest distribution.
The same factors that explain high liana density and diversity among forests may
also explain lianas at the local (within-forest) scale. The hot, dry environmental
conditions in treefall gaps may resemble those of tropical forest experiencing a
seasonal drought. Hot, dry conditions in treefall gaps and during seasonal droughts
results in high vapor pressure deficit (VPD), which causes water stress in plants.
The ability of lianas to tolerate and even to grow under high VPD conditions may
provide the mechanism that allows lianas to peak in density and diversity in treefall
gaps and during seasonal droughts, thus reconciling the causes for the distribution
of lianas across local, regional, and global scales.
In summary, the study of lianas has historically lagged behind that of trees;
however, lianas are now widely recognized as a key component of forests worldwide,
particularly in seasonal tropical areas, where liana density and diversity peak rela-
tive to trees. Lianas compete intensely with trees, decreasing tree recruitment,
growth, fecundity, and diversity. Lianas also significantly reduce forest-level bio-
mass accumulation, which may have profound effects on the global carbon cycle.
The effects of lianas are likely to increase in neotropical forests, where liana abun-
dance is increasing relative to that of trees. While, at first pass, it is tempting to
remedy the negative effects of lianas with large-scale liana cutting treatments, that
recommendation would not be prudent. Lianas contribute significantly to tropical
plant diversity and they provide an important source of food and shelter for animals,
as well as serving as connections among trees that many animals use to traverse the
forest canopy. The combination of both positive and negative effects of lianas illus-
trates their importance to the dynamics and ecology of tropical forest ecosystems.

Acknowledgments This work was made possible by financial support from US National Science
Foundation grants DEB-0613666, NSF-DEB 0845071, and NSF-DEB 1019436. I thank
N. Parthasarathy for inviting me to write this chapter and N. Parthasarathy and A Ercoli, for helpful
comments on the manuscript.

References

Alvarez-Cansino L, Schnitzer SA, Reid J, Powers JS (2015) Liana competition with tropical trees
varies with seasonally, but not with tree species identity. Ecology (in press)
Andrade JL, Meinzer FC, Goldstein G, Schnitzer SA (2005) Water uptake and transport in lianas
and co-occurring trees of a seasonally dry tropical forest. Trees Struct Funct 19:282–289
Arroyo-Rodriguez V, Asensio N, Dunn JC, Cristobal-Azkarate J, Gonzalez-Zamora A (2015) Use
of lianas by primates: more than a food resource. In: Schnitzer SA, Bongers F, Burnham RJ,
Putz FE (eds) The ecology of lianas. Wiley-Blackwell, Oxford, pp 407–426
158 S.A. Schnitzer

Brokaw NVL, Busing RT (2000) Niche versus chance and tree diversity in forest gaps. Trends Ecol
Evol 15:183–188
Cai Z-Q, Schnitzer SA, Bongers F (2009) Seasonal differences in leaf-level physiology give lianas
a competitive advantage over trees in a tropical seasonal forest. Oecologia 161:25–33
Chave J, Riéra B, Dubois M (2001) Estimation of biomass in a neotropical forest in French Guiana:
spatial and temporal variability. J Trop Ecol 17:79–96
Chen Y-J, Cao K-F, Schnitzer SA, Fan Z-X, Zhang J-L, Bongers F (2015) Water-use advantage of
lianas over trees in seasonal tropical forests. New Phytol 205:128–136
Clark DB, Clark DA (1990) Distribution and effects on tree growth of lianas and woody hemiepi-
phytes in a Costa Rican tropical wet forest. J Trop Ecol 6:321–331
Dalling JW, Schnitzer SA, Baldeck C, Harms KE, John R, Mangan SA, Lobo E, Yavitt JB,
Hubbell SP (2012) Resource-based habitat associations in a neotropical liana community.
J Ecol 100:1174–1182
DeWalt SJ, Schnitzer SA, Chave J, Bongers F, Burnham RJ, Cai ZQ, Chuyong G, Clark DB,
Ewango CEN, Gerwing JJ, Gortaire E, Hart T, Ibarra-Manríquez G, Ickes K, Kenfack D, Macía MJ,
Makana JR, Mascaro J, Martínez-Ramos M, Moses S, Muller-Landau HC, Parren MPE,
Parthasarathy N, Pérez-Salicrup DR, Putz FE, Romero-Saltos H, Thomas D (2010) Annual
rainfall and seasonality predict pan-tropical patterns of liana density and basal area. Biotropica
42:309–317
DeWalt SJ, Schnitzer SA, Alves LE, Bongers F, Burnham RJ, Cai ZQ, Carson WP, Chave J, Chuyong
G, Costa F, Ewango CEN, Gallagher RV, Gerwing JJ, Gortaire E, Hart T, Ibarra-Manríquez G,
Ickes K, Kenfack D, Letcher SG, Macía MJ, Makana JR, Martínez-Ramos M, Mascaro J,
Muthumperumal C, Muthuramkumar S, Nogueira A, Parren MPE, Parthasarathy N, Pérez-
Salicrup DR, Putz FE, Romero-Saltos H, Reddy MS, Sainge MN, Thomas D, van Melis J (2015)
Biogeographical patterns of liana abundance and diversity. In: Schnitzer SA, Bongers F, Burnham
RJ, Putz FE (eds) The ecology of lianas. Wiley-Blackwell Publishing, Oxford, pp 131–148
Durán SM, Gianoli E (2013) Carbon stocks in tropical forests decrease with liana density. Biol Lett
9:20130301
Garrido-Pérez EI, Dupuy JM, Durán-García R, Gerold G, Schnitzer SA, Ucan-May M (2008)
Structural effects of lianas and hurricane Wilma on trees in Yucatan peninsula, Mexico. J Trop
Ecol 24:559–562
Gianoli E (2004) Evolution of a climbing habit promotes diversification in climbing plants. Proc R
Soc Lond B 271:2011–2015
Grauel WT, Putz FE (2004) Effects of lianas on growth and regeneration of Prioria copaifera in
Darien, Panama. For Ecol Manage 190:99–108
Ingwell LL, Wright SJ, Becklund KK, Hubbell SP, Schnitzer SA (2010) The impact of lianas on
10 years of tree growth and mortality on Barro Colorado Island, Panama. J Ecol 98:879–887
Jimenez-Castillo M, Wiser S, Lusk C (2007) Elevational parallels of latitudinal variation in the
proportion of lianas in woody floras. J Biogeogr 34:163–168
Kainer KA, Wadt LHO, Staudhammer CL (2014) Testing a silvicultural recommendation: Brazil
nut responses 10 years after liana cutting. J Appl Ecol. doi:10.1111/1365-2664.12231
Kurzel BP, Schnitzer SA, Carson WP (2006) Predicting liana crown location from stem diameter
in three Panamanian forests. Biotropica 38:262–266
Ladwig LM, Meiners SJ (2015) The role of lianas in temperate tree communities. In: Schnitzer SA,
Bongers F, Burnham RJ, Putz FE (eds) The ecology of lianas. Wiley-Blackwell, Oxford,
pp 188–202
Lambert TD, Halsey M (2015) Relationship between lianas and arboreal mammals: examining the
Emmons-Gentry hypothesis. In: Schnitzer SA, Bongers F, Burnham RJ, Putz FE (eds) The
ecology of lianas. Wiley-Blackwell Publishing, Oxford, pp 398–406
Ledo A, Schnitzer SA (2014) Disturbance and clonal reproduction determine liana distribution and
maintain liana diversity in a tropical forest. Ecology 95:2169–2178
Michel NL, Robinson D, Sherry TW (2015) Liana-bird relationships: a review. In: Schnitzer SA,
Bongers F, Burnham RJ, Putz FE (eds) The ecology of lianas. Wiley-Blackwell Publishing,
Oxford, pp 362–397
9 The Contribution of Lianas to Forest Ecology, Diversity, and Dynamics 159

Parthasarathy N, Muthuramkumar S, Muthumperumal C, Vivek P, Ayyappan N, Reddy MS (2015)


Liana composition and diversity among tropical forest types on peninsular India. In: Schnitzer
SA, Bongers F, Burnham RJ, Putz FE (eds) The ecology of lianas. Wiley-Blackwell Publishing,
Oxford, pp 36–49
Phillips OL, Vasquez Martinez R, Mendoza AM, Baker TR, Vargas PN (2005) Large lianas as
hyperdynamic elements of the tropical forest canopy. Ecology 86:1250–1258
Powers JS (2015) Reciprocal interactions between lianas and forest soil. In: Schnitzer SA,
Bongers F, Burnham RJ, Putz FE (eds) The ecology of lianas. Wiley-Blackwell, Oxford,
pp 175–187
Putz FE (1984) The natural history of lianas on Barro Colorado Island, Panama. Ecology
65:1713–1724
Restom TG, Nepstad DC (2004) Seedling growth dynamics of a deeply-rooting liana in a second-
ary forest in eastern Amazonia. For Ecol Manage 190:109–118
Schnitzer SA (2005) A mechanistic explanation for global patterns of liana abundance and distri-
bution. Am Nat 166:262–276
Schnitzer SA (2015) Increasing liana abundance and biomass in neotropical forests: causes and
consequences. In: Schnitzer SA, Bongers F, Burnham RJ, Putz FE (eds) The ecology of lianas.
Wiley-Blackwell Publishing, Oxford, pp 451–464
Schnitzer SA, Carson WP (2000) Have we missed the forest because of the trees? Trends Ecol Evol
15:376–377
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
Schnitzer SA, Bongers F (2011) Increasing liana abundance and biomass in tropical forests:
emerging patterns and putative mechanisms. Ecol Lett 14:397–406
Schnitzer SA, Carson WP (2001) Treefall gaps and the maintenance of species diversity in a
tropical forest. Ecology 82:913–919
Schnitzer SA, Carson WP (2010) Lianas suppress tree regeneration and diversity in treefall gaps.
Ecol Lett 13:849–857
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical
forest canopy gaps: evidence for an alternative pathway of gap-phase regeneration. J Ecol
88:655–666
Schnitzer SA, Kuzee M, Bongers F (2005) Disentangling above- and below-ground competition
between lianas and trees in a tropical forest. J Ecol 93:1115–1125
Schnitzer SA, Mangan SA, Dalling JW, Baldeck CA, Hubbell SP, Ledo A, Muller-Landau H,
Tobin MF, Aguilar S, Brassfield D, Hernandez A, Lao S, Perez R, Valdes O, Yorke SR
(2012) Liana abundance, diversity, and distribution on Barro Colorado Island, Panama.
PLoS One 7:e52114
Schnitzer SA, van der Heijden GMF, Mascaro J, Carson WP (2014) Lianas reduce biomass accu-
mulation in a tropical forest. Ecology 95:3008–3017
Schnitzer SA, Bongers F, Burnham RJ, Putz FE (eds) (2015a) The ecology of lianas. Wiley-
Blackwell, Oxford, 504 pp. ISBN 978-1-118-39249-2
Schnitzer SA, Putz FE, Bongers F, Kroening K (2015b) The past, present, and potential future of
liana ecology. In: Schnitzer SA, Bongers F, Burnham RJ, Putz FE (eds) The ecology of lianas.
Wiley-Blackwell Publishing, Oxford, pp 3–12
Swaine MD, Grace J (2007) Lianas may be favoured by low rainfall: evidence from Ghana.
Plant Ecol 192:271–276
Thomas D, Burnham RJ, Chuyong G, Kenfack D, Nsangy Sainge M (2015) Liana abundance and
diversity in Cameroon’s Korup National Park. In: Schnitzer SA, Bongers F, Burnham RJ, Putz
FE (eds) Ecology of lianas. Wiley-Blackwell, Oxford, pp 13–22
Tobin MF, Wright AJ, Mangan SA, Schnitzer SA (2012) Lianas have a greater competitive effect
than trees of similar biomass on tropical canopy trees. Ecosphere 3, Article 20: 1–11.
http://dx.doi.org/10.1890/ES11-00322.1
van der Heijden GMF, Phillips OL (2009) Liana infestation impacts tree growth in a lowland
tropical moist forest. Biogeosciences 6:2217–2226
160 S.A. Schnitzer

van der Heijden GMF, Schnitzer SA, Powers JS, Phillips OL (2013) Liana impacts on carbon
cycling, storage and sequestration in tropical forests. Biotropica 45:682–692
van der Heijden G, Phillips O, Schnitzer SA (2015) Effects of lianas on forest level biomass. In:
Schnitzer SA, Bongers F, Burnham RJ, Putz FE (eds) The ecology of lianas. Wiley-Blackwell
Publishing, Oxford, pp 164–174
Wyka TP, Oleksyn J, Karolewski P, Schnitzer SA (2013) Phenotypic correlates of the lianescent
growth form – a review. Ann Bot 112:1667–1681
Yanoviak S (2015) Effects of lianas on canopy arthropod community structure. In: Schnitzer SA,
Bongers F, Burnham RJ, Putz FE (eds) The ecology of lianas. Wiley-Blackwell, Oxford,
pp 345–361
Zhu S-D, Cao K-F (2009) Hydraulic properties and photosynthetic rates in co-occurring lianas and
trees in a seasonal tropical rainforest in southwestern China. Plant Ecol 204:295–304
Chapter 10
Liana Diversity and Their Ecosystem
Services in Tropical Dry Evergreen Forest
on the Coromandel Coast of India

N. Parthasarathy, P. Vivek, and K. Anil

Abstract Lianas constitute one of the most important components and play a vital
role in structural and functional aspects of tropical forests. Few studies have reported
the ecosystem services offered by lianas. This chapter presents the research conducted
in eighteen tropical dry evergreen forest (TDEF) sites on the Coromandel Coast of
India to investigate various ecosystem services rendered by the liana life-form. Lianas
in the studied sites contributed a total of 56 species to forest plant biodiversity and
added 15,224 individuals to forest stand. At individual sites, liana species richness
ranged from 11 species ha−1 to 31 species ha−1 and the density contribution ranged
from 408 individuals ha−1 to 1,658 individuals ha−1. The proportion of lianas to total
woody species (lianas + trees) richness ranged from 0.38 to 0.53 and the proportion
of stem density varied from 0.29 to 0.62. To the total forest aboveground biomass,
lianas contributed 411.26 Mg. Further, lianas in our sites provide valuable resources
to various faunal groups chiefly the forest foliar herbivores, florivores and frugivores.
Among the leaf-eaters, beetles and lepidopteran larvae formed the major foliar
herbivores. Bees and butterflies are the major floral resource users, while birds
consumed major fruit resources of TDEF ecosystem. Lianas also offer important
livelihood to local people by providing various edible products and feed for cattle.
Further lianas offer valuable medicinal resources, 49 species being used for the
treatment of various ailments. Sustainable use of resources offered by lianas would
help long-term species survival and hence, the need for conservation of liana
species is emphasized useful to, forest wealth and human health.

Keywords Aboveground biomass • Biodiversity • Ecosystem services • Foliar


herbivores • Lianas • Medicinal values • NTFP

N. Parthasarathy (*) • P. Vivek • K. Anil


Department of Ecology and Environmental Sciences, Pondicherry University,
Puducherry 605 014, India
e-mail: parthapu@yahoo.com

© Springer International Publishing Switzerland 2015 161


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_10
162 N. Parthasarathy et al.

10.1 Introduction

Lianas are important but under-researched plant life-form common to most tropical
forests. Although several studies have reported the role of lianas, especially their
deleterious effects on host trees and their implications on tropical forest structure,
regeneration and dynamics (Putz 1984; Schnitzer et al. 2005; Ingwell et al. 2010),
the positive aspects of lianas remain under-studied in tropical forest ecosystems. It
is estimated that lianas could commonly compose up to 25 % of the total woody
plant density and species richness in tropical forests and may comprise as much as
44 % in some forests (Gentry 1991; Perez-Salicrup 2001), thereby, attesting their
substantial contribution to the whole forest biodiversity. Lianas also play a major
role at the ecosystem level by contributing to the whole forest carbon budget of
tropical forests, representing as much as 10 % of the fresh aboveground biomass
(Putz 1984). Most economic appraisals of tropical forests focus on timber as the
major forest produce, however, a large variety of non-timber forest products
(NTFPs) are also used by millions of people around the world. Liana species con-
stitute a very important group of NTFPs, as is becoming clear over the last few
decades (Abbiw 1990; Malaisse 1997; Van Andel 2000; Tra Bi 1997; Van Valkenburg
1997). Lianas are used by local people in many different ways, for example, in Cote
d’Ivoire, Tra Bi et al. (2002) reported that 114 species of lianas are used by the local
populations, especially for medicinal purposes. Vine extracts have traditionally
been prescribed for treating various medical ailments including cancer, dental
problems, eye inflammations, malaria, snakebite and urinary tract infection (Tra Bi
et al. 2005).
Lianas provide valuable habitat and connections among tree canopies that
provide a pathway for arboreal animals to transverse the tree tops (Emmons and
Gentry 1983; Schnitzer and Carson 2001). Lianas greatly contribute to canopy
closure after tree fall and help stabilize the microclimate underneath (Schnitzer and
Bongers 2002), giving shade-tolerant species a chance to establish. Unlike trees,
climbers show no peaks in flowering and fruiting (Putz and Windsor 1987; Heideman
1989), which implies that lianas might form an essential part of animal diet in times
of scarcity of fruits and flowers (Bongers et al. 2005). Their leaves are used as
oviposit hosts by lepidopterans and as foliar resources by many forest fauna, chiefly
the insects (Odegarrd 2000). Martins (2009) found that lianas constituted 27 %
of the diet for the brown howling monkey (Aouatta guariba) and 33 % of the diet
for the Southern muriqui (Brachyteles arachnoides). In this chapter, we discuss
the ecosystem services rendered by liana communities and the need for liana
conservation in tropical dry evergreen forest (TDEF) on the Coromandel Coast of
India. We recognized three major ecosystem services provided by liana communities
in TDEF ecosystem:
1. Biodiversity as an ecosystem service
2. Ecological service by providing resource for faunal groups and
3. Economic goods service – NTFPs.
10 Liana Diversity and Their Ecosystem Services in Tropical Dry Evergreen Forest… 163

10.2 Materials and Methods

10.2.1 Study Area

The study was conducted in eighteen tropical dry evergreen forest (TDEF) sites
located on the Coromandel Coast of peninsular India (refer Fig. 8.1 in Chap. 8, this
volume). The TDEF is a forest type with restricted geographical distribution globally
(Parthasarathy et al. 2008) and in India, classified under type 7/C1 of Champion
and Seth (1968), that occur as patches along the Coromandel Coast. These are
short-statured, two- to three-layered, tree-dominated and liana-dense forests without
distinct shrub layer and with sparse understory vegetation. The canopy is about
10–12 m in height, dominated by large trees such as Pterospermum canescens Roxb.
and Lannea coromandelica (Houtt.) Merr., while the sub-canopy is composed of
medium-statured trees such as Canthium dicoccum (Gaertn.) Teijsm. & Binn., Drypetes
sepiaria (Wight and Arn.) Pax & Hoffm., Garcinia spicata and lower-canopy is com-
posed of small trees such as Memecylon umbellatum and Glycosmis mauritiana
Yuich. Tanaka etc. Major lianas include Strychnos lenticellata Dennst., Combretum
albidum G. Don., Reissantia indica (Willd.) Halle, Carissa spinarum L., Hugonia
mystax L., and Grewia rhamnifolia Heyne ex Roth. Ecbolium viride (Forsskal)
Alston and Sansevieria roxburghiana Schultes & Schultes f. are the major native
perennial herbs present in this forest type (Parthasarathy et al. 2008).
The rainfall is tropical dissymmetric type with most of the showers received during
the north-east monsoon and little, inconsistent rainfall during the south-west
monsoon. The TDEF experiences 6–8 months of dry period. The mean annual rain-
fall and temperature in the nearest towns of Cuddalore (11.749 N and 79.750 E),
Nagapattinam (10.766 N and 79.840 E) and Pudukottai (10.379 N and 78.823 E) are
1,184 mm (36.9 °C and 20.8 °C), 1,346 mm (36.1 °C and 21.6 °C) and 919 mm
(34.9 °C and 22.3 °C) respectively. All the study sites are protected as sacred groves
(sacred forests) owing to the traditional and religious beliefs of the local people,
except site Point Calimere Wildlife Sanctuary, which is the largest TDEF site that
exists in India.

10.2.2 Methodology

Biodiversity inventory of lianas was carried out in 18 one-hectare study plots


distributed one in each of the 18 tropical dry evergreen forest sites on the Coromandel
Coast of India. Each one-hectare plot was further divided into one hundred 10 × 10 m
quadrats to facilitate the inventory. All lianas, rooted inside the quadrat with
diameter ≥1 cm from their rooting point were measured and enumerated. All the
enumerated liana species were recognized to species level using regional floras
(Gamble and Fischer 1915–1935; Matthew 1991) and confirmed with the specimens
lodged in the herbarium of Department of Ecology and Environmental Sciences,
164 N. Parthasarathy et al.

Pondicherry University. The above ground biomass (AGB) of lianas was estimated
using the regression equation (Schnitzer et al. 2006) AGB = exp [−1.484 + 2.657)
ln (D)], where D is stem diameter. To assess the ecological services rendered by the
enumerated 56 liana species of tropical dry evergreen forest, direct observation on leaf,
flower and fruit resource users were made in the field using binoculars and cameras,
from August 2012 to July 2014. A minimum of 2–4 individuals for each species
(for rare & sub-dominant species) to a maximum of 10 individuals (for common and
dominant species) were studied for this. We collected the leaf eaters and reared them
in laboratory until the adults emerged, for identification. Foliar herbivores were also
captured in photos and videos when fed upon leaves. Economic good services of
lianas were assessed in the field and also by conducting participatory rural appraisals
(PRAs) with the local people residing in the forest fringes and nearby villages.
Medicinal values of lianas were assessed by interacting with the local traditional
healers and also by referring to the available literature (Khare 2007).

10.3 Results

10.3.1 Biodiversity Service

Lianas contributed 56 species to the plant diversity of tropical dry evergreen forest
(TDEF) enumerated from the 18 sites on the Coromandel Coast of peninsular India.
At individual site, lianas contributed 11–31 species with a mean value of 24.4 ± 4.7
species ha−1 (Table 10.1a). Lianas in the eighteen study sites contributed 37 % of the
total woody species richness (trees + lianas) and this proportion was maximum
(53 %) in two TDEF sites SL and PC 2 (Table 10.2). The mean liana density in the
study sites was 846 individuals ha−1 that totaled to 15,224 individuals for 18 sites
(408–1,658 individuals ha−1). The contribution of lianas to the total woody species
density in the study sites ranged from 62 % at site SV to 29 % at site AK and the
mean liana contribution across the study sites was 43 %. Strychnos lenticellata
(2,757 individuals), Combretum albidum (2138) and Reissantia indica (1138) were
the top three most abundant liana species that together formed 40 % of the total liana
abundance (Table 10.3). Whereas, two species were represented by a single individual
and eight species were represented by less than ten individuals. Lianas scored a total
basal area of 16.02 m2 for the 18 sites with a mean value of 0.89 m2 ha−1. At individual
sites, liana basal area varied from 0.20 m2 ha−1– 1.85 m2 ha−1 (Table 10.1a). Lianas
comprised 4.4 % of the total woody species basal area across the study sites with the
maximum contribution (8.3 %) made by site VV. The mean aboveground biomass
(AGB) contribution of lianas in the study sites was 22.85 Mg ha−1 (411.21 Mg for the
18 sites) and it ranged from 2.25 to 42.14 Mg ha−1 across the 18 study sites
(Table 10.1a). Derris scandens (69.57 Mg), Combretum albidum (45.16 Mg) and
Acacia caesia (39.77 Mg) were the top three liana species that contributed maximum
AGB and accounted for 60.2 % of the total liana above ground biomass (Table 10.3).
10 Liana Diversity and Their Ecosystem Services in Tropical Dry Evergreen Forest… 165

Table 10.1 Consolidated details of ecosystem services rendered by lianas species enumerated from
18 tropical dry evergreen forest sites distributed 1 ha in each site on the Coromandel Coast of India
a. Biodiversity service
Variable On plot (18 ha) Mean ha−1 Range
Liana species richness 56 24.4 11–31
Liana abundance & density 15,224 846 408–1,658
Basal area (m2 ha−1) 16.02 0.89 0.20–1.85
Aboveground biomass (Mg) 411.21 22.85 2.25–42.14
b. Ecological service
Reward Number of liana species (%)
Leaf resource for a total of 21 foliar herbivores 40 (71.4)
Floral resource for 30 florivores 38 (67.9)
Fruit resource for 14 frugivores 24 (42.9)
c. Economic goods service
Importance Number of liana species (%) Abundance (%)
Edible 10 (17.86) 1,890 (12.42)
Medicinal 49 (87.50) 11,584 (76.14)
Cattle feed 17 (30.36) 7,880 (51.79)

10.3.2 Ecological Services

Of the total 56 liana species, 40 species (71.4 %) offered leaf resources to 21 species
of foliar herbivores that include lepidopteran larvae, beetles and grasshoppers
(Table 10.1b). Beetles formed the predominant foliar herbivores of TDEF and
utilized leaves of 23 species (41 %) of lianas followed by lepidopteran larvae
(15 species; 26.7 %) and grasshoppers (5 species; 8.7 %) (Table 10.3, Plate 10.1).
Species such as Combretum albidum, Reissantia indica, Cissus vitiginea, Grewia
rhamnifolia and Cayratia pedata were heavily folivored. Thirty-eight liana species
(67.9 %) offered floral resources to 30 flower visitors (Tables 10.1b and 10.3).
The most common floral resource users were the hymenopterans (55.3 %) and
lepidopterans (42.8 %) (Plate 10.1). Among them, the bees and butterflies are the
predominant liana flower users (utilizing 22 liana species each; 38.5 %), followed
by flies, ants, moths and bats. The common dwarf honey bee, Apis florea was the
major floral resource user in TDEF (Tables 10.1b and 10.3). Among the total 56
liana species, six species (Abrus precatorius, Dioscorea oppositifolia, Plecospermum
spinosum, Strychnos lenticellata, Tiliacora acuminata and Wattakaka volubilis) were
exclusively visited by bee species and two species (Ampelocissus tomentosa and
Jasminum angustifolium) exclusively by butterflies and one species (Olax scandens)
by blister beetle (Tables 10.1b and 10.3). Frugivore observation in TDEF revealed
that 17 species of lianas (42.9 %) are utilized by 12 species of birds and two species
166

Table 10.2 List of 18 tropical dry evergreen forest sites on the Coromandel Coast, their site code and proportional contribution of lianas to total forest woody
species richness, density and basal area
Species richness (Number of species
ha−1) Density (Number of stems ha−1) Basal area (m2 ha−1)
Lianas Trees Lianas Trees Lianas Trees
Site ≥1 cm ≥10 cm Proportion ≥1 cm ≥10 cm Proportion ≥1 cm ≥10 cm Proportion
Site name code diameter gbh of lianas diameter gbh of lianas diameter gbh of lianas
Araiyapatti AP 26 35 0.43 792 807 0.50 1.09 19.1 0.054
Arasadikuppam AK 29 31 0.48 1,163 2,815 0.29 0.58 17.6 0.032
Karisakkadu KR 23 30 0.43 515 596 0.46 0.67 21.6 0.030
Karukkai KA 22 20 0.52 941 845 0.53 0.61 13.3 0.044
Kothattai KT 20 25 0.44 552 661 0.46 0.23 15.9 0.014
Kuzhandhaikuppam KK 28 26 0.52 497 1,251 0.28 1.37 14.6 0.086
Maanadikuppam MK 11 18 0.38 408 786 0.34 0.2 8.2 0.024
Maramadakki MM 21 28 0.43 596 724 0.45 1.68 15.5 0.098
Oorani OR 24 30 0.44 812 1,286 0.39 1.85 27.3 0.063
Point Calimere 1 PC 1 28 37 0.43 672 790 0.46 0.68 29.5 0.023
Point Calimere 2 PC 2 31 27 0.53 720 803 0.47 0.44 20.4 0.021
Purangani PR 22 20 0.52 919 948 0.49 0.65 10.6 0.058
Puthupet PP 28 29 0.49 835 1,338 0.38 0.99 37.6 0.026
Shanmuganathapuram SP 27 26 0.51 775 1,663 0.32 0.44 22.1 0.020
Silattur SL 25 22 0.53 1,213 1,211 0.50 0.77 14.3 0.051
Sunayakkadu SN 28 27 0.51 701 841 0.45 0.95 18 0.050
Suran Viduthi SV 27 37 0.42 1,455 888 0.62 1.06 19.1 0.053
Vanniyan Viduthi VV 20 30 0.40 1,658 1,693 0.49 1.76 19.5 0.083
Total 56 92 0.37 15,224 19,946 0.43 16.02 344.2 0.044
N. Parthasarathy et al.
Table 10.3 List of a total of 56 liana species and the ecosystem services rendered in tropical dry evergreen forest (TDEF) on the Coromandel Coast of India
(for expansion of faunal groups see Table 10.4)
Biodiversity service – Ecological service (resource
contribution to user – faunal group) Economic good service
Medicinal use
Total Aboveground Foliar Plant Cattle
Sl. no. Liana species (family) abundance1 biomass (kg) herbivores Florivore Frugivore Edible part Ailment/treatment feed
2
1 Abrus precatorius L. 26 11.0 GH2 BE1,2 Seed Uterine stimulant
(Papilionaceae)
2
2 Acacia caesia (L) Willd. 314 39,770.9 BT6 BE1,4, BT3 Bark Anti-inflammatory
(Mimosaceae)
3 Adenia wightiana (Wall.ex 4 5.9 Leaf Tuber Peptic ulcer
Wight & Arn.) Engler
(Passifloraceae)
4 Aganosma cymosa (Roxb.) 18 0.0 L36
G. Don var. cymosa Hook. f.
(Apocynaceae)
5 Ampelocissus tomentosa 19 11.5 BT2 BF4 B12 Root Bone fracture
(Heyne ex Roth) Planch.
(Vitaceae)
6 Aristolochia indica L. 5 0.0 Entire Oxytocic &
(Aristolochiaceae) abortifacient
7 Asparagus racemosus Willd. 131 37.8 BE1,2 Root Galactogogue &
(Liliaceae) nerve tonic
8 Azima tetracantha Lam. 25 0.4 Entire Asthma &
(Salvadoraceae) bronchitis
10 Liana Diversity and Their Ecosystem Services in Tropical Dry Evergreen Forest…

9 Calycopteris floribunda 21 5,363.7 Leaf Anti-dysenteric &


Lam. (Combretaceae) jaundice
(continued)
167
Table 10.3 (continued)
168

Biodiversity service – Ecological service (resource


contribution to user – faunal group) Economic good service
Medicinal use
Total Aboveground Foliar Plant Cattle
Sl. no. Liana species (family) abundance1 biomass (kg) herbivores Florivore Frugivore Edible part Ailment/treatment feed
10 Canavalia virosa (Roxb.) 11 5.7 BT4 AN2, BE6, Seeds
Wight & Arn. BF12, TH1
(Papilionaceae)
11 Cansjera rheedii Gmel. 53 3,477.4 BT5 AN2, BF2,3 Leaf Leaf Poisonous bites
(Opiliaceae)
2
12 Capparis brevispina DC. 237 1,626.3 BT6 AN2, MO5, B8 Root Tooth ache
(Capparaceae) TH1, BA1
13 Capparis divaricata Lam. 1 0.0 Fruits
(Capparaceae)
2
14 Capparis rotundifolia Rottb. 53 695.2 BT3 AN2, MO5, B11 Root Head ache
(Capparaceae) TH1, BA1
15 Capparis sepiaria L. 23 1,506.3 BT1 AN2, MO5, B11 Leaf Anti-septic &
var.sepiaria (Capparaceae) TH1 anti-pyretic
2
16 Capparis zeylanica L. 140 3,101.0 BT3 AN2, MO5, B8 Leaf Applied on
(Capparaceae) TH1 swelling and boils
2
17 Carissa spinarum L. 800 2,394.0 BT6 BF5,6, BT2 B7 Fruits Root Pugative & cardio
(Apocynaceae) tonic
18 Cayratia pedata (Lam.) 7 2.0 BT4 BE1, BF9 B11 Leaf Astringent
Juss.ex Gagnep. (Vitaceae)
19 Cissus quadrangularis L. 491 938.5 L2 F2, BE1, B10 Entire Scurvy & fracture
(Vitaceae) BF5, W2 healing
20 Cissus vitiginea L. 309 410.2 L5 F2, BE1,3, B11 Fruit Relief of flatulence
(Vitaceae) W2, BF10
N. Parthasarathy et al.
21 Clerodendrum inerme (L.) 13 0.1 Tender Stomach ache
Gaertn. (Verbenaceae) leaf
2
22 Coccinia grandis (L.) Voigt 334 820.0 GH3, L6 BF2, BE1,2 B5 Fruits Root Head ache
(Cucurbitaceae)
2
23 Combretum albidum G. Don 2,138 45,164.1 GH1, L7 F1, BE1,2,5, Bark Treating skin
(Combretaceae) TH1, BF2,3 diseases
24 Derris ovalifolia (Wight & 500 261.3 BT4, GH2 BE1,2, BF6
Arn.) Benth. (Papilionaceae)
25 Derris scandens (Roxb.) 603 69,578.6 BT1 BE3 Bark Treating skin
Benth. (Papilionaceae) wounds
26 Dioscorea oppositifolifa L. 1 0.0 GH4 BE5 Tubers Entire Applied on
(Dioscoreaceae) swellings
27 Gloriossa superba L. 4 0.0 L30 Root Antidote for snake
(Lilliaceae) bite
2
28 Grewia rhamnifolia Heyne 632 14,223.4 L4 F3, BF2,4, B11, 4, 6, Fruit Stomach disorders
ex Roth (Tiliaceae) BE1,2 BU1
29 Gymnema sylvestre (Retz.) 352 795.1 L31 BE5,3, BF2 Leaf Anti-diabetic
R.Br.ex Schultes
(Asclepiadaceae)
2
30 Hugonia mystax L. 637 19,732.2 L4 BF11, BE1, B2 Root Anti-inflammatory
(Linaceae)
2
31 Ichnocarpus frutescens (L.) 162 0.2 L12 BF1,8,11
R. Br. (Apocynaceae)
32 Ipomoea staphylina Roem. 39 668.3 Leaf Stomach disorders
& Schultes
(Convolvulaceae)
2
33 Jasminum angustifolium (L.) 640 1,258.5 L30 BF2 B1 Leaf Emetic & treating
10 Liana Diversity and Their Ecosystem Services in Tropical Dry Evergreen Forest…

Willd. (Oleaceae) ringworm


34 Jasminum sessiliflorum Vahl 109 46.6
(Oleaceae)
(continued)
169
Table 10.3 (continued)
170

Biodiversity service – Ecological service (resource


contribution to user – faunal group) Economic good service
Medicinal use
Total Aboveground Foliar Plant Cattle
Sl. no. Liana species (family) abundance1 biomass (kg) herbivores Florivore Frugivore Edible part Ailment/treatment feed
35 Olax scandens Roxb. 23 27.5 BT3 BT4 B9 Bark Anaemia &
(Olacaceae) diabetes
36 Pachygone ovata (Poir.) 143 1,287.5 BT3 F1, BF4,8, B1 Fruits Leaf Analgesic
Miers ex Hook. BE1,3 properties
(Menispermaceae)
37 Plecospermum spinosum 66 4,214.8 BT5 BE1 B12 Stem- Cholera
Trecul (Moraceae) thorn
38 Premna corymbosa 254 1,127.6 BT6 F3, W2, B13 Leaf Galactogogue
(Burm.f.) Rottl. & Willd. BF11,13
(Verbenaceae)
39 Pterolobium hexapetalum 91 4,744.2 Leaf Anti-fungal
(Roth) Sant. & Wagh.
(Caesalpiniaceae)
40 Pyrenacantha volubilis 375 190.7 BT5 BE1,3, BF13 B9 seed Breast cancer
Wight (Icacinaceae)
2
41 Reissantia indica (Willd.) 1,138 15,276.4 BT5 BF1,3,5 Leaf Wound healing
Halle (Celastraceae)
2
42 Rivea hypocrateriformis 84 628.0 BT6 BF4, MO6, Leaf Astringent
(Desr.) Choisy F3
(Convolvulaceae)
43 Salacia chinensis L. 16 637.9 BT1 F3, AN2, B11 Root Diabetes
(Celastraceae) BE1
44 Sarcostemma acidum 73 19.6 Stem Emetic
(Roxb.) Voigt
N. Parthasarathy et al.

(Asclepiadaceae)
45 Scutia myrtina (Brum.f.) 110 2.9 Leaf Ointment for
Kurz (Rhamnaceae) parturition
46 Secamone emetica (Retz.) 382 497.5 Fruits Leaf Head-ache
R.Br. ex Schultes
(Asclepiadaceae)
47 Strychnos lenticellata 2,757 7,362.4 BT4 BE1,4 B2, BU3
(Dennst.) Hill (Loganiaceae)
48 Symphorema involucratum 62 6.3 Flower Anti-inflammatory
Roxb. (Verbenaceae)
49 Tiliacora acuminata (Lam.) 16 0.0 BT2 BE1,3 B11 Entire Spasmolytic
Hook. f. & Thoms.
(Menispermaceae)
2
50 Tinospora cordifolia (Willd.) 208 86.8 BT3 F2, BE1,2,3, B3 Stem Anti-pyretic
J.D. Hook. & Thoms. BF4,2,7, TH1
(Menispermaceae)
51 Toddalia asiatica (L.) Lam. 40 0.3 L12 BE1,2,3,4,5, B11 Bark Anti-spasmodic
(Rutaceae) BF1,2,3,5,8
52 Trichosanthes tricuspidata 2 0.0 Fruit Asthma
Lour. (Cucurbitaceae)
53 Tylophora indica (Burm. f.) 7 6.7 L31 BE1, BF9 Leaf Bronchial asthma
Merr. (Asclepiadaceae)
2
54 Ventilago madraspatana 176 5,849.0 BT5 BF1,10, BE5 Bark Skin diseases
Gaertn. (Rhamnaceae)
55 Wattakaka volubilis (L.f.) 106 163.1 L3 BE5 Leaf Cough & asthma
T. Cooke (Asclepiadaceae)
2
56 Ziziphus oenoplia (L.) Mill. 161 2,245.8 L28 F1,2, TH1 B11 Fruits Root Hyper acidity
(Rhamnaceae)
10 Liana Diversity and Their Ecosystem Services in Tropical Dry Evergreen Forest…

1
Abundance contribution of liana species in the total eighteen 1-ha plots
2
Liana species used as cattle feed
171
172 N. Parthasarathy et al.

Table 10.4 List of faunal groups that utilize various liana resources (floral, fruit and foliar) in
tropical dry evergreen forest on the Coromandel Coast of India, provided with their common name,
scientific name and species code
Faunal group Common name Scientific name Code
Floral resource users
Mammals
Bats Short-nosed fruit bat Cynopterus sphinx BA1
Coleopterans
Beetles Jewel beetle Sternocsera chrysis BT3
Blister beetle Mylabris pustulata BT4
Dipterans
Flies Hover fly Volucella sp. F1
Green bottle fly Calliphora sp. F2
Blow fly Chrysomya sp. F3
Hymenopterans
Ants Weaver ant Oecophylla samaragdina AN2
Bees Dwarf honey bee Apis florea BE1
Indian rock bee Apis dorsata BE2
Oriental honey bee Apis cerana BE3
Blue banded bees Amegilla zonata BE4
Stingless bees Trigone iridipennis BE5
Carpenter bee Xylocopa sp. BE6
Wasps Mason wasp Eumenes conica W2
Lepidopterans
Butterflies Common crow Euploea core BF1
Plain tiger Danus chrysippus BF2
Blue tiger Limniace leopardus BF3
Common evening brown Melanitis leda BF4
Common leopard Phalantha phalantha BF5
Lemon pansy Junonia lemonias BF6
Common sailor Neptis hylas BF7
Tawny coster Acraea violae BF8
Indian cupid Chilades parrhasius BF9
Yellow orange tip Ixias pyrene BF10
Common jezebel Delias eucharis BF11
Common wanderer Pareronia valeria BF12
Crimson rose Tros hector BF13
Moths Tiger moth Amata passalis MO5
Unidentified Unidentified MO6
Thysanopterans
Thrips Thrips 1 Thrips hawaiensis TH1
(continued)
10 Liana Diversity and Their Ecosystem Services in Tropical Dry Evergreen Forest… 173

Table 10.4 (continued)


Faunal group Common name Scientific name Code
Fruit resource users
Birds White-headed babbler Turdoides affinis B11
White-browed bulbul Pyconotus luteolus B10
Coppersmith barbet Megalaima haemacephala B12
Asian koel Eudynamys scolopacea B2
Rose-ringed parakeet Psittacula krameri B8
Purple sunbird Cinnyris asiaticus B6
House sparrow Passer domesticus B3
Purple-rumped sunbird Nectarinia zeylonica B4
Eurasian golden oriole Oriolus oriolus B13
Orange-headed thrush Zoothera citrina B5
Red-vented bulbul Pyconotus cafer B7
Tickells flowerpecker Dicaeum erythrohynchos B9
Bugs Jewel bug Chrysocoris stolli BU1
Man-faced bug Catacanthus incarnatus BU3
Leaf resource users
Larva Moth larva 2 Unidentified L2
Moth larva 3 Unidentified L3
Moth larva 4 Unidentified L4
Moth larva 5 Unidentified L5
Moth larva 6 Unidentified L6
Moth larva 7 Unidentified L7
Moth larva 12 Unidentified L12
Moth larva 28 Unidentified L28
Yellow orange tip Ixias pyrene L30
Common grass yellow Eurema hecabe L31
Painted hand maiden moth Euchromia polymena L36
Beetle Beetle 1 Unidentified BT1
Leaf beetle Chrysochus auratus BT2
Beetle 2 Unidentified BT3
Beetle 3 Unidentified BT4
Jewel beetle Sternocera chrysis BT5
June beetle Phyllophaga sp. BT6
Grasshoppers Grasshopper 1 Unidentified GH1
Grasshopper 2 Unidentified GH2
Grasshopper 3 Unidentified GH3
Grasshopper 4 Melanoplus sp. GH4
174 N. Parthasarathy et al.

Plate 10.1 (a) Common crow throng for nectar from tubular flowers of Acacia caesia; (b)
Common jezebel accesses nectar from Lantana camara; (c) Bee approaching Canavalia virosa flower;
(d) Jewel beetle relishes leaves of Acacia caesia; (e) Moth larvae heavily punctures Cayy atia pedata
leaves; (f) Butterfly larva in Ziziphus oenoplia; (g) Drupes of Tiliacora acuminata attract vertebrates
for dispersal; (h) Man-faced bug sucks fruit-sap from Benkara malabarica; (i) Zoochorous fruits
of Ziziphus oenoplia; (j) Medicinally values Tinospora cordifolia also used as cattle feed; (k) Rind
of hesperidium berry of Tricosanthes tricuspidata used for treating migraine; (l) Multi-useful
(fruits– edible, leaves–cattle feed, flower and fruit- medicinal) armed scrambler Carissa spinarum

of bugs (Table 10.3, Plate 10.1). Common frugivorous birds in TDEFs that consumed
fruits of various liana species include the white-headed babbler (19.6 %), coppersmith
barbet and flowerpecker (Table 10.3). Fruit resource use by bugs is less known and
in TDEF, it includes jewel bug that consumes fruit sap of Grewia rhamnifolia
and man-faced bug that takes juice from the predominant liana species Strychnos
lenticellata (Table 10.3).
10 Liana Diversity and Their Ecosystem Services in Tropical Dry Evergreen Forest… 175

10.3.3 Economic Services

Liana species enumerated in the present study sites offer three major economic
good services such as edible materials, medicinally valued products and cattle
feed (Table 10.1c). Of the total 56 liana species, the fruits and /leaves of ten
species are edible for human consumption; especially, the fruits of Zizyphus oeno-
plia, Coccinia grandis and Carissa spinarum (Plate 10.1) are widely collected
and fruits of Pachygone ovata are collected to a lesser extent. Annually, during
January – February, tubers of the yam, Dioscorea oppositifolia are collected for
local use. The edible lianas in the study sites comprised 17.9 % of the total species
richness and 12.42 % (1,890 individuals) of the total abundance, thus rewarding a
valuable economic service to the local people. Among the 56 liana species, a total
of 49 species are medicinally useful for treating various major and minor ailments
(Table 10.3). Lianas such as Gymnema sylvestre, Tinospora cordifolia and Salacia
chinensis are used for treating diabetes. Several other liana species are commonly
used to treat hepatitis (Abrus precatorius), anemia (Olax scandens), rheumatism
(Premna corymbosa), respiratory disorders (Reissantia indica) and skin diseases
(Tiliacora acuminata and Ventilago madraspatana). Lianas are also used for their
anti-septic, astringent and anti-inflammatory properties: Capparis sepiaria,
Cayratia pedata and Hugonia mystax to mention a few. Leaves are the most
widely used plant part followed by root and bark for their medicinal value.
Seventeen liana species are useful as cattle feed and their abundance was 51.7 %
in the liana community. Combretum albidum, Carissa spinarum and Capparis
brevispina are the most preferred cattle feed collected in large quantity. Among
the 56 liana species enumerated, 52 lianas are economically beneficial by providing
at least one of the three economic goods services (Table 10.3). Lianas such as
Carissa spinarum, Coccinia grandis and Ziziphus oenoplia are multi-beneficial
species and they provide valuable edible fruits for human consumption, their plant
parts are also medicinally valuable and they further comprise one of the major diet
of cattle feed from the forest.

10.4 Discussion

Few studies are available on resource values and ecosystem services of lianas
(Bongers et al. 2002; Tra Bi et al. 2005), despite their importance in providing
livelihood for local people and valuable resources for various faunal groups. Lianas
in the present study contributed substantially to the whole forest stem density and
woody species richness. It is estimated that lianas could comprise about 25 % of
species diversity (Gentry and Dodson 1987; Schnitzer and Bongers 2002) and
10–45 % of the woody species density (Gentry 1991; Perez-Salicrup 2001; Senbeta
et al. 2005). In contrast, lianas in Indian TDEFs comprised 37 % of the woody species
richness and 43 % of the woody species density there by making a substantial con-
tribution to the whole forest diversity. Lianas also added considerably to the carbon
176 N. Parthasarathy et al.

budget of the TDEF ecosystem, despite their greater abundance in smaller stem size
class. Thus, lianas play a major role in maintaining the biodiversity and carbon stock
of the whole forest. Lianas, however in larger proportions may have a devastating
effect on trees and reduce the ability of the forest to sequester carbon.
Lianas in the tropical dry evergreen forest provide foliar resources for many
Lepidopteran larvae by serving as oviposit hosts and further the leaves of are fed
upon by diverse foliar herbivores such as Coleoptera, Lepidopteran larva and
Orthoptera. These foliar herbivores are reported to be the most important con-
sumers in various other tropical forests too (Angulo-Sandoval and Aide 2000;
Williams-Linera and Herrera 2003; Angulo-Sandoval et al. 2004; Mazia et al. 2012).
Coleopterans are the major leaf-eaters (of 41 % of liana species) in tropical dry
evergreen forest as also reported in many other tropical forests (Murali and Sukumar
1993; Varanda and Pais 2006). Lianas in the studied sites also provided valuable
floral rewards to various faunal groups, especially to the bees and butterflies. The
proportion of species with multiple floral visitors is 83 % and that of exclusive
species is 17 % in the TDEF. Many plant species in tropical community seem to be
generalist in their visitation (Bawa et al. 1985; Momose et al. 1998). A high ratio of
multiple floral visitors can enhance community stability and organization (Ramirez
2004). Emmons and Gentry (1983) estimated that on average 21 % of plant species
eaten by wide variety of tropical primates were climbers. In the present study, 16
species provided fruit resource for many vertebrate fauna, which reveals the
dependency of various faunal groups on lianas. Thus, the lianas in TDEF provide a
wide array of resources to the faunal assemblage and fauna in turn play a major role
in pollination and seed dispersal ecology of liana species.
Local people use a large number and kind of lianas, all over the world for a large
variety of uses (Bongers et al. 2005). In this study, we recognized 52 liana species
as economically useful for the local people, which are at least two-fold lesser than
the reports of Tra Bi (1997), who found 114 liana species beneficial to the local
population in Ghana. However, in the present study sites, 92 % of lianas species
produced economic benefits to the dependent local population, especially the
medicinal and health benefits. Lianas in TDEFs are also one of the major diet for
cattle in nearby villages. Thus, lianas provide livelihood support to the local people
inhabiting around the forest areas. The edible fruits and other edible plant parts are
not used for commercial purpose in TDEF, but are used largely by the local people
for their consumption. Such low level of resource extraction from local forests will in
long run be useful for sustainable development of resources and in maintenance of
forest wealth and human health as well.

10.5 Conclusion

Lianas in the tropical dry evergreen forest contribute substantially to the whole
forest diversity, which reflects the role that lianas could play in forest structure and
functioning. Further, lianas support a rich insect and avifauna, many of which help
10 Liana Diversity and Their Ecosystem Services in Tropical Dry Evergreen Forest… 177

in functional ecology (pollination and seed dispersal etc.) of the forest community.
Data on different plant resource (leaf, flower and fruit) users and their food-plants
provide a better understanding of the forest biotic interactions and such information
will be helpful for biological conservation and management of tropical forests.
Lianas in TDEFs also offer valuable resources to the dependent local people thereby
they meet various local livelihoods. An optimal utilization of these economically
and culturally important species will help sustainable resource development and
species survival in long run in order to benefit from various ecosystem services
rendered by liana communities of Indian TDEF, realizing the need for conservation
of plant resources in this and similar tropical forests.

Acknowledgements We thank the Department of Science and Technology, Indian National


Science Academy and Ministry of Environment and Forests, New Delhi, India for funding various
phases of this study through project.

References

Abbiw DK (1990) Useful plants of Ghana. West African uses of wild and cultivated plants. Royal
Botanical Gardens, Kew
Angulo-Sandoval P, Aide TM (2000) Leaf phenology and leaf damage of saplings in the Luquillo
Experimental Forest, PuertoRico. Biotropica 32(41):5–422
Angulo‐Sandoval P, Fernández‐Marín H, Zimmerman JK (2004) Changes in patterns of understory
leaf phenology and herbivory following hurricane damage. Biotropica 36:60–67
Bawa KS, Bullock SH, Perry DR et al (1985) Reproductive biology of tropical lowland rain forest
trees. II. Pollination systems. Am J Bot 72:346–356
Bongers F, Schnitzer SA, Traore D (2002) The importance of lianas and consequences for forest
management in West Africa. Bioterre, Rev. Int. Sci. de la Vie et de la Terre, No. Especial: 59–70
Bongers FJJM, Parren MPE, Swaine MD, Traoré D (2005) Forest climbing plants of West Africa:
introduction. In: Forest climbing plants of West Africa: diversity, ecology and management.
CABI Publishing, Wallingford, pp 5–18
Champion SH, Seth SK (1968) A revised survey of the forest types of India. Manager of
Publications, Delhi
Emmons LH, Gentry AH (1983) Tropical forest structure and the distribution of gliding and
prehensile-tailed vertebrates. Am Nat 121:513–524
Gamble JS, Fischer CEC (1915–1935) Flora of Presidency of Madras. Adlard and Son, London
Gentry AH (1991) The distribution and evolution of climbing plants. In: Mooney HA, Putz FE
(eds) The biology of vines. Cambridge University Press, Cambridge, pp 349–369
Gentry AH, Dodson CH (1987) Contribution of non-trees to species richness of tropical rain forest.
Biotropica 19:149–156
Heideman PD (1989) Temporal and spatial variation in the phenology of flowering and fruiting in
a tropical rainforest. J Ecol 77:1059–1079
Ingwell LL, Joseph Wright S, Becklund KK et al (2010) The impact of lianas on 10 years of tree
growth and mortality on Barro Colorado Island, Panama. J Ecol 98:879–887
Khare CP (ed) (2007) Indian medicinal plants: an illustrated dictionary. Springer, New York
Malaisse F (1997) Se nourrir en foret claire africaine. Approche ecologique et nutritionelle. Les
Presses Agronomique de Gembloux, gembloux, and Centre Technique de Coopération Agricole
et Rurale, CTA, Wageningen
Martins MM (2009) Lianas as a food resource for brown howlers (Alouatta guariba) and southern
muriquis (Brachyteles arachnoides) in a forest fragment. Anim Biodivers Conserv 32:51–58
178 N. Parthasarathy et al.

Matthew KM (1991) An excursion flora of central Tamil Nadu. CRC Press, India
Mazia N, Chaneton EJ, Dellacanonica C et al (2012) Seasonal patterns of herbivory, leaf traits and
productivity consumption in dry and wet Patagonian forests. Ecol Entomol 37:193–203
Momose K, Yumoto T, Nagamitsu T et al (1998) Pollination biology in a lowland dipterocarp
forest in 3232 Sarawak, Malaysia. I. Characteristics of the plant-pollinator community in a
lowland dipterocarp forest. Am J Bot 85:1477–1501
Murali KS, Sukumar R (1993) Leaf flushing phenology and herbivory in a tropical dry deciduous
forest, southern India. Oecologia 94(114):119
Odegarrd F (2000) The relative importance of trees versus lianas as host for phytophagous beetles
(Coleopteran) in tropical forest. J Biogeogr 27:283–296
Parthasarathy N, Selwyn MA, Udayakumar M (2008) Tropical dry evergreen forests of peninsular
India: ecology and conservation significance. Trop Conserv Sci 1(2):89–110
Perez-Salicrup DR (2001) Effect of liana cutting on tree regeneration in a liana forest in Amazonian
Bolivia. Ecology 82:389–396
Putz FE (1984) The natural history of Lianas on Barro Colorado Island, Panama. Ecology
65:1713–1724
Putz FE, Windsor DM (1987) Liana phenology on Barro Colorado Island, Panama. Biotropica
19:334–341
Ramirez N (2004) Pollination specialization and time of pollination on a tropical Venezuelan plain:
variations in time and space. Bot J Linn Soc 145:1–16
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
Schnitzer SA, Carson WP (2001) Treefall gaps and maintenance of species diversity in a tropical
forest. Ecology 82:913–919
Schnitzer SA, Kuzee ME, Bongers F (2005) Disentangling above- and below-ground competition
between lianas and trees in a tropical forest. J Ecol 93:1115–1125
Schnitzer SA, DeWalt SJ, Chave J (2006) Censusing and measuring lianas: a quantitative compari-
son of the common methods. Biotropica 38:581–591
Senbeta F, Schmitt C, Denich M et al (2005) The diversity and distribution of lianas in Afromontane
rain forests of Ethiopia. Divers Distrib 11:443–452
Tra Bi FH (1997) Utilisations des plantes, par l’homme, dans les Forets Classees du Haut Sassandra
et du Scio, en Côte d’Ivoire. These, 3emeCycle, Universite de Cocody, Abidjan
Tra Bi FH, Kouame FN, Traore D (2002) Utilisation des lianes dans deux Forêts Classées de
l’Ouest de la Côte d’Ivoire. Chapitre dans le livre Bongers & Traoré
Tra Bi FH, Kouame FN, Traore D (2005) Utilisation of climbers in two forest reserves in west Cote
d’Ivoire. In: Bongers F, Parren MPE, Traore D (eds) Forest climbing plants of West Africa:
diversity, ecology, and management. CABI Publishing, Wallingford, pp 167–182
Van Andel T (2000) Non-timber forest products of the North-West District of Guyana.
Tropenbos-Guyana Series 8a, Tropenbos Foundation, Wageningen
Van Valkenburg JLCH (1997) Non-timber forest products of East Kalimantan: potential for sus-
tainable forest use. Tropenbos Series 16. Ph.D. thesis, Wageningen University, The Netherlands
Varanda EM, Pais MP (2006) Insect folivory in Didymopanax vinosum (Apiaceae) in a vegetation
mosaic of Brazilian Cerrado. Braz J Biol 66:671–680
Williams‐Linera G, Herrera F (2003) Folivory, herbivores, and environment in the understory of a
tropical montane cloud forest. Biotropica 35:67–73
Chapter 11
A Review of Biotechnological Approaches
to Conservation and Sustainable Utilization
of Medicinal Lianas in India

Shaily Goyal, Varsha Sharma, and Kishan Gopal Ramawat

Abstract Medicinal lianas of India are source of several important secondary


metabolites which have great potential for future therapeutics. Because of their
overexploitation using destructive harvesting methods and/or with some
reproductive barriers, most of them have become threatened species. Recent bio-
technological developments depicting their biological activity, standardizing the
micropropagation methods, and development of callus and cell cultures for the pro-
duction of useful metabolites have opened up novel ways for their conservation by
reducing pressure on their natural populations. Though considerable success has
been achieved in developing the micropropagation protocols as well as bioreactor
level production of secondary metabolites for some lianas, yet for many lianas
development of these technologies remained a challenge. This article reviews the
status of biotechnology of medicinal lianas of India and discusses the novel combi-
nations to achieve the desired results. The review is expected to help in deciding the
priorities for further technology development for production of useful metabolites
and multiplication of lianas in mass scale.

Keywords Medicinal lianas • Bioactive molecules • Biological activity •


Micropropagation • Secondary metabolites

S. Goyal
Laboratory of Bio-Molecular Technology, Department of Botany,
M. L. Sukhadia University, Udaipur 313001, Rajasthan, India
5360, Rome Drive, Erie, PA, USA
V. Sharma
Laboratory of Bio-Molecular Technology, Department of Botany,
M. L. Sukhadia University, Udaipur 313001, Rajasthan, India
Department of Biochemical Engineering & Biotechnology,
Indian Institute of Technology (IIT), Hauz Khas, New Delhi 110016, India
K.G. Ramawat (*)
M.L. Sukhadia University, Botany Department, Udaipur 313001, Rajasthan, India
e-mail: kg_ramawat@yahoo.com

© Springer International Publishing Switzerland 2015 179


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_11
180 S. Goyal et al.

11.1 Introduction

Lianas are woody climbers rooted in the soil that are incapable of autonomous verti-
cal growth above a certain height and must rely on external support (Richards 1996).
Most of them have low wood densities and invest their resources in growth of length
(Putz 1984), and remain thin-stemmed. This provides higher hydraulic conductivity
in their stems. It is generally observed that they have a very high canopy and stem
biomass ratio (Schnitzer and Bongers 2002). For instance, lianas present in tropical
rain forest in Panama accounts for less than 10 % of total forest biomass but up to
40 % of the leaf biomass of the forests (Putz 1984). Liana density and diversity
increase dramatically with decreasing latitude. Besides contributing to density and
diversity, they play a significant role in forest structure and functional aspects of
tropical forests (Putz and Mooney 1991). The peculiar characteristics of lianas
always attracted researchers. They show typical climbing mechanisms (Darwin
1865; Isnard and Silk 2009), biomechanical characteristics (Rowe et al. 2004),
extreme stem hydraulic capacities (Gartner et al. 1990; Ewers et al. 1991), anatomi-
cal modifications (Bowling and Vaughn 2009) and extraordinary developmental
plasticity (Lee and Richards 1991). Lianas too play a vital role in structuring the
forest. They have great impact on the diversity and overall density of tropical forests
(Gentry and Dodson 1987). Many studies suggest mechanisms by which lianas alter
the forest diversity and regeneration (Schnitzer and Carson 2001), promoting some
species while harming others (Schnitzer et al. 2000). Apart from the aesthetic and
ecological values lianas have immense medicinal properties.
The plant-based traditional medicine systems continue to play an essential role
in health care, with about 80 % of the world’s inhabitants relying mainly on tradi-
tional medicines for their primary health care. Now-a-days there is an increasing
interest in drugs of natural origin and are considered as green medicines and these
are always considered being safe. The advantage of natural drugs is their easy avail-
ability, economic and less or no side effects. Thus, the demand for medicinal plants
is continuously increasing not only in developing countries but also in developed
countries as drug, food supplement (nutraceuticals) and cosmetics (Ramawat and
Goyal 2008; Ramawat et al. 2009). The more effective the natural drug more is its
demand and the chances of non-availability increases. Only a small percentage of
medicinal plants are solely cultivated while others are collected from natural habi-
tats. Population growth, urbanization and the unrestricted collection of medicinal
plants from the wild is resulting in an over-exploitation of natural resources
(Table 11.1). Over-exploitation and/or destructive harvesting to meet such demand
in fact threaten the very survival of many species. There are a number of constraints
for the propagation and conservation through conventional methods like vegetative
and seed propagation. The propagation in its natural habitat is becoming a hard
phenomenon. Therefore, biotechnology helps in conservation of such plants and
meets the growing requirements (Arora et al. 2010; Sharma et al. 2011a). The lianas
selected for detailed study in this article are traditionally used in Indian system of
11 A Review of Biotechnological Approaches to Conservation and Sustainable… 181

Table 11.1 Utilization of medicinal lianas for traditional medicines (After Ramawat and Goyal
(2008), Ved and Goraya (2007))
Plant species Plant part used Demand (tonnes/year)
Abrus precatorius Seeds <100a
Celastrus paniculatus Seeds <100a
Embelia ribes Roots and fruits <100
Glycirrhiza glabra Roots 1,328
Hemidesmus indicus Roots 1,614a
Pueraria tuberosa Tuber 100a
Tinospora cordifolia Stem 2,932a
Tylophora indica Stem 300a
a
= Mostly collected from wild

medicine and are well- established medicinal plants. A detailed account of the
biotechnological interventions made to protect and propagate the medicinally
important lianas of India such as Abrus precatorius, Cayratia trifolia, Celastrus
paniculatus, Clitoria ternatea, Cocculus hirsutus, Glycirrhiza glabra, Hemidesmus
indicus, Pueraria tuberosa, Tylophora indica, Tinospora cordifolia and Vitis vinif-
era has been given.

11.2 Scientific Validation of Biological Activities of Lianas

The medicinal plants form the backbone of traditional system of medicine. Plant-
based treatments are being recommended for curing various diseases by traditional
medical practitioners around the world (Arora et al. 2010; Ramawat and Goyal
2008). The phytochemicals present in the medicinal plants are getting attention day-
by-day for their active role in the prevention of several human diseases. In the last
few decades, very intense pharmacological studies have resulted in establishing
medicinal plants as potential sources of new compounds of therapeutic value and as
a source of lead compounds in drug development. Some selected medicinal lianas
of India are presented in Fig. 11.1.
Medicinal liana species such as A. precatorius, C. trifolia, C. paniculatus, C.
hirsutus, G. glabra, H. indicus, T. indica and V. vinifera show effective anticancer-
ous or chemopreventive properties. Today with the change in lifestyle like inappro-
priate food habits, lack of exercises and work pressure, incidence of cardiovascular
diseases and diabetes has increased. Plants like C. trifolia, C. paniculatus, E. ribes,
T. cordifolia and V. vinifera are quite effective against cardiovascular diseases while
most of the lianas undertaken for this study exhibit anti-diabetic properties. Species
like C. trifolia, P. tuberosa, G. glabra, H. indicus, T. cordifolia, T. indica and V.
vinifera are hepatoprotective. Pharmacological activities of selected lianas are
182 S. Goyal et al.

Fig. 11.1 Some selected medicinal lianas of India

presented in Table 11.2. It is evident that biological activity varies from antimicro-
bial and antioxidant to hepatoprotective, neuroprotective and anticancerous.
Therefore, in view of the important activities performed by these plants, investiga-
tions must be continued to explore and utilize the full potential of medicinal plants.
Plants may offer alternatives to expensive medicines in future, hence more evalua-
tion on other biological activities using state of art techniques is required.

11.3 Why Biotechnology of Lianas?

The science of biotechnology has provided us the required tools and techniques for
mass cloning, improvement in medicinal plants and increased production of useful
metabolites (Arora et al. 2010; Sharma et al. 2011a). Conventional methods of prop-
agation have their own shortcomings with respect to commercial level production
and quality control. Seeds are highly heterozygous and resulting seedlings show
great variations in growth and yield. Majority of plants are not amenable to vegeta-
tive propagation through cutting and grafting, thus limiting the multiplication of
11

Table 11.2 Some selected medicinal lianas of India, their bioactive molecules and biological activity
Common name
Plant name (plant parts used) Bioactive molecules Biological activities References
Abrus precatorius Chirmi/Crab eye Abrine, abricin, arbidin, Antioxidant Gul et al. (2013), Pal et al.
(Leguminosae) (Leaves, root) precatorine, choline (2009)
Anticancerous (hepatocellular Mukhopadhyay et al. (2014)
carcinoma)
Antidiabetic Monago and Alumanah (2005)
Antibacterial activity Mistry et al. (2010), John et al.
(2012)
Antiprotozoal activity Hata et al. (2013)
Potential anti- cancerous agent Lee et al. (2014)
Cayratia trifolia Khatta nimba (fruits) Piceid, resveratrol, viniferin Cardioprotective Waffo-Teguo et al. (2008)
(Vitaceae) Hepatoprotective Guru Kumar et al. (2011)
Neuroprotective Wu et al. (2009)
Antiulcerogenic Gupta et al. (2012)
Anticancerous Delmas et al. (2006)
Celastrus paniculatus Malkangni/black oil plant Celapanin, celapanigin, Seed oil- antidepressant-like Valecha and Dhingra (2014)
(Celastraceae) (seed) celapagin, celastrine activity
paniculatine Anti- inflammatory activity Younus and Agarwal (2013)
Antiproliferative activity Weng et al. (2013)
A Review of Biotechnological Approaches to Conservation and Sustainable…

Hypolipidemic Patil et al. (2010)


Cortico-hippocampal Chakrabarty et al. (2012)
neurodegeneration and
oxidative stress
183

(continued)
Table 11.2 (continued)
184

Common name
Plant name (plant parts used) Bioactive molecules Biological activities References
Clitoria ternatea Aparajit/butterfly-pea Rutin, delphidin, Antidiabetic Verma et al. (2013)
(Fabaceae) (root, seed and leaves) kaempferol, quercetin and Anti-inflammatory, analgesic, Devi et al. (2003)
malvidin antipyretic properties
Antioxidant activity Kamkaen and Wilkinson (2009)
Hepatoprotective Nithianantham et al. (2011)
Nootropic, anxiolytic, Jain et al. (2003)
antidepressant, anticonvulsant,
antistress activity
Cocculus hirsutus Vasanvel or tana Shaheenine, cohirsinine, Antihyperglycemic activity Badole et al. (2006)
(Menispermaceae) (stem, root, leaves) hirsutine, jamtinine, Antimicrobial activity Kuchana et al. (2014)
cohirsine, dehydrocohirsine, Aphrodisiac Patil et al. (2014)
cohirsitine, syringasesinol,
Anticancerous Thavamani et al. (2014)
isotriboline, triboline and
haiderine Acute and chronic diuretic effect Badole et al. (2009)
Nephroprotective Gadapuram et al. (2013)
Combretum albidum Manjakody, buffalo calf Caryophyllene, linalool, Antimicrobial- Escherichia Sahu et al. (2014a), Sreedhar
(Combretaceae) (stem bark, leaves, fruit) β-phellandrene, phytol, coli, Klebsiella pneumonia, et al. (2012, 2013a, b)
b-sitosterol, ursolic acid Proteus mirabilis,
and gallic acid Pseudomonas aeruginosa,
Bacillus subtilis and
Salmonella typhimurium
Leaves have anti-ulcerous Ramasubramaniaraja and
properties Niranjan Babu (2011)
Combretum Yada teega (leaves) Significant antibacterial Shrisha et al. (2011)
latifolium activity against Escherichia
(Combretaceae) coli, Salmonella typhi,
Staphylococcus aureus and
Bacillus cereus
S. Goyal et al.

Stem shows antioxidant activity Dao et al. (2012)


11

Embelia ribes Vidanga/false black Embelin, embelinol, Neuroprotective Ansari et al. (2008)
(Myrsinaceae) pepper (fruit) embeliaribyl ester, embeliol Antihyperglycemic Bhandari et al. (2013)
and vilangin Embelin effective Joy et al. (2014)
photodynamic therapeutic
against tumor in vivo
Cardioprotective Sahu et al. (2014b)
Antidepressant-like activity Gupta et al. (2013),
Kazmi et al. (2013)
Protective effect of embelin Thippeswamy et al. (2011)
against ulcerative colitis
Anthelmintic activity Mohandas et al. (2013)
Cosmetic agent Radhakrishnan et al. (2011)
Glycyrrhiza glabra Licorice/Mulaithi (roots) Glycyrrhizin, 18beta- Antiallergic effects Shin et al. (2007)
L., (Leguminosae) glycyrrhetinic acid, Antioxidant, anti-genotoxic Siracusa et al. (2011)
isoliquiritin, liquiritigenin, and anti-inflammatory activity
licoagrodin, Isoliquiritigenin inhibits Zheng et al. (2014)
licoagrochalcones, human breast cancer
licoagroaurone, metastasis
licochalcone C, kanzonol Y,
Estrogenic activities Simons et al. (2011)
glyinflanin B and
glycyrdione A Hepatoprotective Yin et al. (2011), Lin et al.
(2008)
Cadium induced genotoxicity Dirican and Turkez (2014)
Glycyrrhizin effective against Yoshida et al. (2014)
pseudomonal burn wound
infection
A Review of Biotechnological Approaches to Conservation and Sustainable…

(continued)
185
186

Table 11.2 (continued)


Common name
Plant name (plant parts used) Bioactive molecules Biological activities References
Hemidesmus indicus Indian sarsaparilla/ 2-hydroxy 4-methoxy Properties of an antiresorptive Pompo et al. (2014), Jain
(Apocynaceae) anantamool (root) benzoic acid (HMBA), drug (Jain 2003)
hemidesmol, hemidesterol, Anti-ulcerogenic and Anoop and Jegadeesan (2003),
hemidescine, emidine antienterobacterial activity Das and Devaraj (2006)
Antioxidant Jayaram and Dharmesh (2011)
Hepatoprotective Saravanan and Nalini (2007)
Cytodifferentiating, cytotoxic Ferruzzi et al. (2013)
and cytostatic activities, potent
antileukemic activity
Tyrosinase (monophenolase) Kundu and Mitra (2014)
inhibitory activity
Pueraria tuberosa Vidarikand/Indian Kudzu Puerarin, genistin, Antioxidant Pandey and Tripathi (2010)
(Leguminosae) (tuber) genistein, daidzein, diadzin, Hepatoprotective Chen et al. (2013)
tuberosin Antidiabetic effect Wu et al. (2013)
Puerarin prevents the Zhang et al. (2013), Zhu et al.
dysfunction of the neuronal (2014)
cholinergic system;
neuroprotective
Nephroprotective Tripathi et al. (2012)
Anti-osteoporotic Michihara et al. (2012)
Reduces alcohol intake Penetar et al. (2012)
S. Goyal et al.
11

Tinospora cordifolia Giloya/Guduchi (all Berberine, palmatine, Antidiabetic Sangeetha et al. (2013)
(Menispermaceae) parts) tinosporin, tinocordiside, Antimicrobial activity Bonvicinia et al. (2014)
cordioside, cordifolioside Antioxidant activity Naik et al. (2014)
A, cordifolioside B,
Cardioprotective Rao et al. (2005)
tinocordifolin
Hepatoprotective Dhanush et al. (2013)
Protect against cytotoxicity Masuma et al. (2014)
and DNA damage in PC12
cells
Tylophora indica Indian ipecac (leaves, Tylophorine, tylophorinine, Antiasthmatic Gupta et al. (1979)
(asclepiadaceae) stem) tylophorinidine, Antifeedant and antimicrobial Reddy et al. (2009)
(+)-septicine, activity
isotylocrebrine, Antiangiogenic and antitumor Saraswati et al. (2013)
tylophorinicine activity
Hepatoprotective Mujeeb et al. (2009)
Vitis vinifera Grape vine (fruit) cis-resveratrol, trans- Antihyperglycaemic effects Moura et al. (2012)
(Vitaceae) resveratrol, cis-piceid, Antioxidant and Fortes Gris et al. (2011)
trans-piceid, and tyrosol hypolipidemic Activity
Hepatoprotective Liu et al. (2012)
Neuroprotective Lakshmi et al. (2014)
Chemopreventive Filip et al. (2011)
Antimicrobial activity Oliveiraa et al. (2013)
A Review of Biotechnological Approaches to Conservation and Sustainable…
187
188 S. Goyal et al.

desired cultivars. Commercial production demands disease-free uniform medicinal


plants with high yield. To meet these demands biotechnology has proved to be of
great help. Biotechnology in plant science is generally regarded as the use of in vitro
techniques in conjunction with molecular tools. In vitro culture and micropropaga-
tion have many advantages like year-round disease-free plantlets production, rapid
mass cloning of genotypes, provide ready protocol for the application of genetic
engineering tools for plant improvement, plant material generated are “synchro-
nized”, miniaturized and relatively homogenous in terms of size, cellular composi-
tion and physiological state. These methods also help micropropagation of those
species which are difficult to multiply by conventional propagation methods.
Besides these benefits, the callus-derived plants exhibit huge genetic variation that
could be exploited for developing superior clones/varieties particularly in vegeta-
tively propagated plant species (Dass and Ramawat 2009).
Modern biotechnological techniques including, tissue and cell cultures, and
bioreactor-mediated bioproduction of phytochemicals provide a potential alterna-
tive to the mass harvesting of plants for the purpose of obtaining crude drug
extracts. In vitro culture protocols can be adapted to produce the compounds of
interest which are simpler to isolate than from in vivo plants. Greater product yield
in a short period of time is possible with in vitro cultures, relative to cultivated
plants (Reppert et al. 2008).
Various tissue culture protocols have been developed for a wide range of liana spe-
cies, which include several threatened species. In species such as Abrus precatorius,
Celastrus paniculatus and Clitoria ternatea, conventional propagation through seeds
is rather slow and not very efficient because of poor seed viability coupled with low
germination or high mortality of seedlings (Ananth et al. 2011). Similarly, Glycyrrhiza
glabra is rarely propagated by seeds due to its hard testa (Gupta et al. 1997), Tylophora
indica due to its limited seed availability (Chandrasekhar et al. 2006) and Embelia
ribes due to development of abortive embryos, and the slow germination (Anon 1990).
In case of Pueraria tuberosa, wide-spread harvesting for its tubers restricts its repro-
duction and regeneration. Due to such propagation difficulties and overexploitation,
species like C. paniculatus, P. Tuberose and Tinospora cordifolia are categorized
under ‘Nearly threatened’ (Balaguru et al. 2006), ‘threatened’ (Rathore and Shekhawat
2009) and enlistedas ‘threatened plant species of India’ (Bhat et al. 2013), respec-
tively. C. ternatea is included in the list of rare species by International Union for
Conservation of Nature and Natural Resources (Pandey et al. 1993). United states
Development Agency also selected C. ternatea with other 16 leguminous plants for
immediate conservation (Morris 1999). E. ribes is one of the 32 medicinal plant spe-
cies identified by the National Medicinal Plant Board, Govt. of India, New Delhi, as
being important for large-scale cultivation because of their commercial use (Anon
2008). Tylophora indica has now been depleted from its primary habitat and recog-
nized as an endangered plant species (Islam 2009). Biotechnological approaches to
conservation and utilization of medicinal lianas are presented in Table 11.3. There is
a need to develop alternative technology for the production of their bioactive mole-
cules and standardize the micropropagtion protocol of these plants for their conserva-
tion and sustainable utilization.
Table 11.3 Recent, selected examples of biotechnological approaches made to lianas towards their conservation
11

Plant species Inoculum used Medium In vitro study Reference


A. Production of secondary metabolites
Abrus precatorius Callus, cell cultures, MS medium + NAA Glycyrrhizin production Dixit and Vaidya (2010),
fungal elicitators; hairy (1g/l) + Kn (1g/l) + Sucrose Karwasara et al. (2010)
root culture 30 g/l
Cayratia trifolia Cell cultures MS medium + angiosperm Enhanced stilbenes production, Arora et al. (2009, 2010)
parasite ∼14-fold higher; piceid
Root cultures Cuscuta reflexa and increased ∼200-fold
salicylic acid
Glycyrrhiza uralensis Tween 80 elicitor, MS liquid medium without 9- and 11-fold higher Zhang et al. (2011)
hairy roots plant growth regulators licochalcone A content
Glycrrhiza glabra Plantlets Treatment with methyl Glycyrrhizin, 4-fold increase Shabani et al. (2009)
jasmonate and salicylic acid
Pueraria tuberosa Callus and cell cultures MS + 2,4,5-T + Kn Cell culture, isoflavonoids (Goyal and Ramawat
(0.1 mg/l) + additives production, shake flasks and 2008a, b; Goyal et al.
bioreactor, morphactin, nutrients 2011), Sharma et al. (2009,
and elicitation 2011a, b), Sharma and
Ramawat (2013)
B. Micropropagation
Abrus precatorius Node MS + BAP Callus Biswas et al. (2007)
(5.0 mg/l) + NAA (0.5 mg/l)
Callus MS + BAP (3.0 mg/l) + Kn Shoot regeneration -do-
(0.5 mg/l) + NAA (0.5 mg/l)
In-vitro shoots ½ MS + IBA (1.0 mg/l) Rooting, the rooted plantlets -do-
were transferred to soil after
A Review of Biotechnological Approaches to Conservation and Sustainable…

proper acclimatization
Celastrus paniculatus Nodal explants MS medium + 1.0 mg/l Clonal propagation, Lal and Singh (2010),
BAP micropropagation and Ananth et al. (2011),
assessment of genetic stability Senapati et al. (2013)
189

(continued)
190

Table 11.3 (continued)


Plant species Inoculum used Medium In vitro study Reference
Clitoria ternatea Nodal explants MS + 8.9 mM BA Shoot multiplication Rout (2005), Singh and
Tiwari (2010)
Embelia ribes Hypocotyl segments MS medium + additives Micropropagation system via Dhavala and Rathore
ascorbic acid, citric acid direct shoot organogenesis; (2010), Preetha et al.
cysteine, glutamine survival up to 95 %; regeneration (2012), Dhavala et al.
1.13 µM TDZ through seeds of diverse origin in (2013)
ex-situ and in situ
Glycyrrhiza glabra Leaf explants MS liquid medium Rhizogenesis from leaf explants, Gupta et al. (2013), Duhan
+0.01 mg/l NAA stolon and shoot regeneration et al. (2014)
Glycyrrhiza glabra Axillary micro MS medium + 0.1 mg/l IAA Re growth of encapsulated micro Mehrotra et al. (2012)
shoots-encapsulated in shoots: 95 % survival,
alginate beads genetically similar
Hemidesmus indicus Axillary shoot MS medium + NAA Micropropagation (indirect Siddique and Bari (2010)
(1 g/l) + Kn (2 g/l) organogenesis)
Maerua oblongifolia Nodal shoot explants MS medium 2.0 mg/l 58 shoots/explants; 80 % plants Rathore and Shekhawat
BAP + additives (25.0 mg/l transferred pots in nursery (2011)
adenine
sulphate + 25.0 mg/l citric
acid + 50.0 mg/l ascorbic
acid)
Pueraria tuberosa Node MS + TDZ Shoot culture in flasks and Sharma et al. (2011a)
(0.25 mg/l) + IBA Growtek bioreactor, Puerarin
(0.05 mg/l) biosynthesis in shoots
S. Goyal et al.
Plant species Inoculum used Medium In vitro study Reference
11

Tinospora cordifolia Node WPM + BAP (8.87 µM) Multiple shoot regeneration Raghu et al. (2006)
In-vitro shoots ½ MS + IAA (2.85 µM) Rooting, rooted plantlets were -do-
successfully transferred to sand Bhat et al. (2013)
and established with 80 %
survival rate
Tylophora indica Long term (6 years) MS basal medium Transgenic plants: genetic Roychowdhury et al. (2013)
cultures, field grown transformation by Agrobacterium
normal and rhizogenes; morphological,
transformed plants biochemical and molecular
characterisation: stable after 6
years
Tylophora indica Leaf MS medium + 2.0 mg/l 19 shoot/explants; regenerants Haque and Ghosh (2013)
BAP true-to-type
C. Other biotechnological studies
Glycyrrhiza glabra Cell cultures Artemisinin derivatives: Biotransformation products Gaur et al. (2013)
α-artemether and
dihydroartemisinin
Hemidesmus indicus Shoot cultures 1/2 MS + 20 g/l sucrose 18–22 months storage George et al. (2010),
Somatic embryos Synthetic seeds from somatic Cheruvathur et al. (2013)
embryos
All the medium recipes were formulated using Murashige and Skoog [S,1962] basal medium; 2,4-D 2,4-diphenoxyacetic acid, 2IP 2, isopentenyl adenine, BAP
6-benzylamino purine, CH casein hydrolysate, IAA indole-3-acetic acid, IBA indole-3-butyric acid, KN kinetin, NAA α naphthalene acetic acid, CH coconut
water, AS adenine sulphate, PGR plant growth regulators, TDZ thidiazuron, ppm parts per million
A Review of Biotechnological Approaches to Conservation and Sustainable…
191
192 S. Goyal et al.

11.4 Production of Secondary Metabolites

Plant cell and tissue culture remained a favored methodology for biotechnological
production of natural therapeutics or secondary metabolites (SMs). Production
through cell and tissue culture so far has been commercially successful in species
such as Taxus baccata, Lithospermum erythrorhizon and Panax ginseng in which
cell cultures are used for production of taxol (an anticancerous diterpene), shikonin
(a natural red dye) and ginsenosides (an adaptogen), respectively (Ramawat and
Goyal 2014). Tremendous research and development efforts have advanced a num-
ber of other in vitro-derived secondary products to semi-commercial status, includ-
ing vanillin and rosmarinic acid production in cell cultures (Karuppusamy 2009).
Such type of mass scale production becomes the alternative technology for metabo-
lites production without harming the plants in nature, thus contributes towards their
conservation.
Plant cell, callus and organized cultures such as shoot cultures, root cultures and
transformed root cultures are employed to obtain production of those SMs which
cannot be obtained from in vivo grown plants because of one or the other reason.
Their production is also not possible through microbial system or chemical synthe-
sis. Small plant parts or explants are used to initiate in vitro callus cultures. In gen-
eral, growth conditions are optimized followed by optimization of production
conditions. Both for growth and production of SMs, primary metabolites are uti-
lized making them a limiting factor. Therefore, growth conditions and production
conditions should be defined separately (Fig. 11.2). Cell suspension cultures are
raised by transferring friable callus into liquid medium and agitating the flasks on
rotary shakers. The plant growth regulators influence the biomass and cell
differentiation (Suri and Ramawat 1995) and the reduction in biomass is always
correlated with increased accumulation of secondary metabolites (Ramawat and
Mathur 2007). Cell cultures are fast growing as compared to callus cultures but
become heterogeneous in terms of productive and non-productive cells. This
requires selection of cells through cloning by transferring small volume of cell sus-
pension cultures on plates containing medium. The culture plates may or may not
contain a selection agent. This provides opportunity to select highly productive
clones by analyzing the growing clone, and cultures obtained from such clones
should be more stable (Dass and Ramawat 2009). Such cultures can be used for all
types of studies related to production of SMs and are best suited for the treatment of
elicitors (Fig. 11.2). Once the basic works related to optimization, selection and use
of production medium are standardized, special techniques like transformation or
elicitation can be employed. Optimized production system is used for scale-up tech-
nology towards industrial production of selected SM. Biotic and abiotic elicitors are
frequently used to stimulate the biosynthetic activity to optimize the SMs accumu-
lation in plant cell cultures (Zhao and Verpoorte 2007). Such cultures are explored
for commercial production (Savitha et al. 2006) using elicitors, e.g. the cell cultures
of Glycyrrhiza echinata, Cicer arietinum, Pueraria lobata and P. thomsonii for iso-
flavonoid production (Luczkiewicz 2008; Maojun et al. 2006). The details of basic
11 A Review of Biotechnological Approaches to Conservation and Sustainable… 193

Medicinal Plants

Explant

Callus and cell cultures


Physico-chemical factors
Optimal growth
Precursors
Enhanced Secondary metabolites

Optimal production of secondary metabolites

Production medium
Selection by Cloning

Biosynthetic pathway manipulation, immobilization,,


Elicitation
Transformation: hairy roots, gene transfer

Scale up technology using Bioreactor


Growth, permealization, process engineering

Fig. 11.2 Schematic presentation of various strategies for the production of secondary metabo-
lites and development of scale-up technology

cell cultures and optimization of SMs are beyond the scope of this paper and can be
found elsewhere (Merillon and Ramawat 2007a, b). The commercial production of
a metabolite is a hard to achieve goal requiring several basic studies of nutrient
optimization and production at high level (Goyal and Ramawat 2008a, b; Goyal
et al. 2011). Different plant cultures may have unusual growth and accumulation
requirements and thus may require modified procedures for culture. Here, we sum-
marize the efforts made by various scientific groups for developing the high yield-
ing cultures of different medicinally important lianas.
194 S. Goyal et al.

11.4.1 Production Through Unorganized/Organized Cultures

The presence and production of SMs have been shown in callus and / cell suspen-
sion cultures of T. cordifolia, P. tuberosa, C. trifolia, G. glabra, and A. precatorius.
Accumulation of berberine and jatrorrhizine (protoberberine alkaloids) was
observed in both callus and cell suspension cultures of T. cordifolia. The alkaloids
were also exudates into the surrounding nutrient medium by non-immobilized and
immobilized cultures (Chintalwar et al. 2003). The occurrence of stilbenes was
observed in cell suspension cultures of C. trifolia as influenced by plant growth
regulators (Roat and Ramawat 2009a). A maximum yield of isoflavonoids ~82-folds
(80 mg/l) was obtained in P. tuberosa cultures grown in morphactin and N6-(2-
Isopentenyl) adenine (2iP) supplemented medium (Goyal and Ramawat 2008a).
These cultures were scaled up to 2 L stirred tank bioreactor (Sharma et al. 2009).
Treatment of A. precatorius cell cultures with 100 µg/l oxytocin (a peptide animal
hormone), improved glycyrrhizin production up to 34.27 mg/l on the dry cell weight
basis (Karwasara et al. 2011). Besides these unorganized cultures, shoot and root
cultures were also established in H. indicus and C. trifolia respectively. H. indicus
shoot cultures accumulated lupenol, vanillin and rutin (Misra et al. 2005) while the
C. trifolia roots accumulated stilbenes (Arora et al. 2009).
In another study, the biotransformation of b-artemether by cell suspension cul-
tures of G. glabra was reported. The major biotransformation product was charac-
terized as a tetrahydrofuran (THF) – acetate derivative. This was an effort to
synthesize more efficient artemisinin derivatives/analogs with improved clinical
efficiency (Patel et al. 2011). Other report on Glycyrrhiza showed that extracts of its
cell culture possessed a similar anti-inflammatory effect to those of field-cultivated
equivalent, through the enhancement of the superoxide dismutase activity of plasma
and liver tissues. Thus, it was proved that use of a cell culture of liquorice instead of
field cultivation would be potentially profitable (Man et al. 2013).
To improve the productivity of the plant cell systems significantly, they are treated
with different elicitors. Elicitors are compounds that stimulate plant cell secondary
metabolism and are usually derived from components of fungal or plant cell walls. It
is well known that upon treatment with elicitors, an array of defense reactions,
including the accumulation of a range of plant defensive secondary metabolites in
plants or in cell cultures, occurs. Feeding of elicitors has been proven to be an effec-
tive way to enhance secondary metabolites in plant cell and tissue cultures (Goyal
et al. 2011). Cell cultures of A. precatorius, C. trifolia, P. tuberose and V. vinifera
were treated with elicitors to enhance the production of respective SM. In a study, A.
precatorius cultures were elicited by the elicitors prepared from the fungi (Aspergillus
niger and Rhizopus stolonifer), yeast extract, salicylic acid, ascorbic acid, and euge-
nol to induce and enhance the synthesis of glycyrrhizin. A. niger elicitation resulted
in 53.62 mg/l glycyrrhizin, which was 5.22 times higher in productivity in compari-
son to control cultures (Karwasara et al. 2010). Similarly, elicitation of C. trifolia and
P. tuberosa cell cultures using salicylic acid (SA), methyl jasmonate (MJ), yeast
extract and ethrel, enhanced the stilbenes (4–6 fold) and isoflavonoids (14-fold) pro-
duction, respectively (Roat and Ramawat 2009b; Goyal and Ramawat 2008b).
11 A Review of Biotechnological Approaches to Conservation and Sustainable… 195

Both C. trifolia and P. tuberosa cell cultures were also elicited by an angiospermic
parasite, Cuscuta reflexa. In case of P. tuberosa, 8 days of elicitation was optimal for
the high accumulation of isoflavonoid, giving productivity of ~4 mg/l/day. The cul-
tures were scaled-up in a 2-l bioreactor where maximum yield of ~91 mg/l of isofla-
vonoid was recorded, which was about 19 % higher than the control cultures. Cuscuta
elicitation increased puerarin content in particular up to 11 mg/l which was 580 %
higher than the value recorded in the control cultures (Goyal et al. 2011). Likewise,
in C. trifolia, enhanced yield of stilbenes (~8-fold) was recorded under the influence
of Cuscuta. Co-treatment of Cuscuta and salicylic acid further increased total stil-
bene yield up to 50.1 mg/l, which was ~14-folds higher than in control cultures.
Piceid content increased ~200-folds in such cultures (Arora et al. 2010).
Combination of elicitation with other factors also increases metabolite produc-
tion in callus cultures of T. cordifolia. The callus cultures were immobilized using
sodium alginate and calcium chloride and cultured in MS medium along with BA
and 2,4-D and 3 % sucrose. The immobilized cultures, when subjected to elicitation
and cell permeabilization with chitosan followed by in situ removal of the second-
ary metabolites by addition of resin, showed a 10-fold increase in the production of
arabinogalactan (0.490 % dry weight) as compared to respective controls devoid of
resin and chitosan. This indicates higher accumulation of secondary metabolites
using elicitors in combination with nutrients and effectors. Similarly, in case of V.
vinifera a number of reports are available for increased production of its metabolites
using an array of elicitors, alone or in combination with each other or with different
elements. The effects of two synthetic elicitor indanoyl-isoleucine (In-Ile),
N-linolenoyl-L-glutamine (Lin-Gln) and one biotic elicitor insect saliva (from
Manduca sexta larvae) was studied in plant cell cultures of V. vinifera. The predomi-
nant phenolic acid 3-O- glucosyl-resveratrol was significantly stimulated by saliva
by 7-fold and the production of 4-(3,5-dihydroxy-phenyl)-phenol was significantly
stimulated by In-Ile with 6.4-fold (Cai et al. 2011). Another group of workers
reported bioproduction of 13C-labelled trans-resveratrol in plant cell culture of V.
vinifera. Two elicitors (SA and JA) and adsorbents (HP2MGL) were introduced into
the V. vinifera cell suspension culture, and the results indicated that they could work
synergistically to improve the production of trans-resveratrol (2.6 g/l) (Yue et al.
2011). Use of cyclodextrins (a carbohydrate polymer), methyl jasmonate and UV
irradiation dosage (Belchí-Navarro et al. 2012) and the effect of low-energy ultra-
sound (US), used alone or in combination with methyl jasmonate (MeJA)
(Santamaria et al. 2012) resulted in the accumulation of trans-resveratrol and vinif-
erin, respectively.

11.4.2 Production Through Transformed Cultures

It is observed that cultures generally require exogenously added phytohormones


and have a tendency to lose their biosynthetic capability after a short period, while
Agrobacterium- transformed cultures are found to be fast growing, biochemically
and genetically more stable over long periods. Crown galls and hairy root cultures
196 S. Goyal et al.

are significant steps for the metabolic manipulation of plant secondary compounds
and applicable for large-scale production of phytopharmaceuticals (Tré mouillaux-
Guiller 2013). In many plant species, transgenic cell cultures or roots have shown to
accumulate higher amount of SMs in comparison to untransformed cultures. High
amount of glycyrrhizin producing and fast growing cell lines of A. precatorius
transformed with Agrobacterium tumefaciens MTCC-431 were reported. These cell
lines when elicited with A. niger (7.5 % v/v) enhanced the glycyrrhizin content by
4.9-fold indicating the prospective of the combination of elicitation methodology
with transformed cell cultures (Kundu et al. 2012). Other lianas such as T. indica,
and C. ternatea also proved to be an efficient transformation system. Intact shoots
of T. indica inoculated at the nodes with A. rhizogenes strain A4 gave an optimal
transformation frequency of up to 60 %. The production of tylophorine, the major
alkaloid of the plant, varied substantially among the nine root clones studied.
Interestingly, in liquid culture, the culture medium also accumulated tylophorine up
to concentrations of ~ 9.78 mg/l (Chaudhari et al. 2005). An unusual biflavonoid
named licoagrodin was also isolated from the hairy root cultures of G. glabra along
with three prenylated retrochalcones and four known prenylated flavonoids and
other glycosides (Li et al. 2000). Independent transformed root somaclones (rhizo-
clones) of C. ternatea were established by A. rhizogenes. The major compound
isolated and purified from the transformed root extracts was taraxerol whose pro-
duction was 4-fold higher than that in natural roots (Swain et al. 2012). Production
strategies should employ combination of effectors for higher yields and new bio-
transformation works should be carried out using new cell cultures and new bioac-
tive molecules. Combined use of elicitors, sucrose, precursors, and retardants using
transformed cultures may prove to be efficient strategy to enhance production of
SMs in root/cell cultures.

11.5 Micropropagation of Lianas

Micropropagation refers to in vitro mass production of plant propagules from any


plant part or cell. Such propagules are used to raise whole plants. Micropropagation
exploits the inherent “totipotency” of plant cells. This potentiality has been exploited
through the culture of protoplasts, cells, tissues and organs in vitro. There are sev-
eral advantages of plantlet regeneration through plant biotechnological methods
using organogenesis or embryogenesis, as compared to conventional methods of
propagation. The advantages include the efficiency of process (the formation of
plantlet in fewer steps, simultaneously with reduction in labor, time and cost), the
potential for the production of much higher number of plantlets and the morpho-
logical and cytological uniformity of the plantlets. Micropropagation through cul-
ture of explants having pre-existing meristem is a powerful option which allows
multiplying genetically stable and true-to-type progeny of the species that are
threatened and difficult to propagate as well as crops of horticulture importance.
Regeneration methods

Explant
Axillary bud break

Regeneration from leaf explants

Callus culture

Organogenesis

Cell suspension culture

Multiple shoots

Somatic embryos formation


Somatic embryos germination

Fig. 11.3 Various methods of regeneration towards micropropagation using direct regeneration
through axillary bud break from explants or via callus/cell cultures

Micropropagation/Clonal propagation techniques using shoot tip and nodal


segments are must for mass-scale multiplication within short period and limited
space. Different methods to achieve multiplication are depicted in a schematic
presentation (Fig. 11.3) are defined as follows.
198 S. Goyal et al.

1. Axillary budding- ensures genetic stability and the production of plants from
axillary buds has proved to be the most reliable method of in vitro propagation.
Shoots and subsequently plantlets formation from axillary bud (pre-existing
meristem) ensures genetic fidelity and termed as ‘clonal propagation’ (clone
constitutes vegetatively derived individuals from a plant, e.g. by cuttings,
explants).
2. Adventitious budding- in response to external stimulus, pre-existing meristem as
well as meristematic cells of the explant produce shoot buds. Shoot buds origi-
nating from other than existing meristems are known as adventitious buds such
as those produced by stem cortex and petiole. In highly proliferating explants,
adventitious shoots are produced giving a gregarious appearance to the explants.
3. Organogenesis-This is the process by which cells and tissues are forced to
undergo changes which lead to the production of a unipolar structure, namely a
shoot or root primordium, whose vascular system is often connected to the par-
ent tissue. This system is commonly produced in callus cultures, but can be pro-
duced directly from the explants.
4. Somatic embryogenesis-Somatic embryogenesis is a process whereby a cell or
group of cells from somatic tissue forms an embryo. The development of somatic
embryos nearly replicates the process of zygotic embryo formation.
Different regeneration pathways such as direct or indirect somatic embryos or
callus-mediated shoot regeneration were explored in different lianas species to get
optimum multiplication rate.

11.5.1 Shoots/Organogenesis

There are many reports in which nodal segment mediated in vitro propagation was
developed in the following lianas: A. precatorius, C. paniculatus, C. ternatea, C.
hirsutus, E. ribes, G. glabra, H. indicus, P. tuberosa, T. indica and T. cordifolia. In
most of these plants maximum number of shoot proliferation with higher frequency
was observed in nodal explants grown on MS medium with 6-benzylaminopurine
(BAP) in different concentrations. For example, in T. indica highest number of
shoots, 7 shoots/explant (Rani and Rana 2010); G. glabra, 12 shoots/explant (Gupta
et al. 2014); C. ternatea, 10 shoots/explant (Singh and Tiwari 2010); T. cordifolia,
4–5 shoots (Sivakumar et al. 2014). In another study, on T. cordifolia 6.3 shoots/
explant were obtained with WP medium supplemented with BA (Raghu et al. 2006).
Similarly, (Meena et al. 2012) reported 45–46 shoots/ explants, from the nodal seg-
ments of T. cordifolia on MS medium containing BAP along with additives like
adenine sulphate and glutamine. Furthermore, BAP was also effective with combi-
nation of other plant growth regulators. In a micropropagation work on C. panicu-
latus and P. tuberosa maximum number of shoots 47–50 and 35–40 were achieved
on MS medium supplemented BAP and indole-3-acetic acid (IAA) (Phulwaria et al.
2013; Rathore and Shekhawat 2009).
11 A Review of Biotechnological Approaches to Conservation and Sustainable… 199

Micropropagation of some other lianas with optimal shoot numbers was observed
on MS medium supplemented with thidiazuron (TDZ) alone or in combination with
other growth hormones or additives. For example, in E. ribes highest number of
shoots, 5–6 shoots/explant (Dhavala and Rathore 2010); in C. ternatea, 15–20
shoots/explant (Ismail et al. 2011); and in A. precatorius, 15 shoots/explant (Perveen
et al. 2013) were observed on MS medium containing TDZ with indole-3-butyric
acid (IBA); TDZ alone; and TDZ and α-naphthalene acetic acid (NAA, in presence
of PVP), respectively. In case of H. indicus highest shoot multiplication rate of 8
shoots/explant with a 95 % frequency was achieved on MS medium supplemented
with kinetin and NAA (Patnaik and Debata 1996).
Besides axillary bud proliferation using nodal explants, organogenesis was also
observed in many of the lianas. Callus-mediated organogenesis depends on various
factors. The cells, although undifferentiated, contain all the genetic information
present in parent plant. By suitable manipulation of growth regulators and contents
of the medium, it is possible to initiate the development of roots, shoots and com-
plete plant from callus cultures. An efficient protocol was obtained for in vitro prop-
agation of A. precatorius through induction of organogenesis in nodal segment
derived callus tissue. The callus differentiated into de novo shoots on an average of
6–7 shoots/culture (Biswas et al. 2007). Similarly, in a report on E. ribes, leaf
explants produced regenerative callus on MS medium supplemented with
2,4-dichlorophenoxy acetic acid (2,4-D) and BAP. The frequency of shoot bud for-
mation was highest (24 shoots/explant) on MS medium containing TDZ and IAA
(Raghu et al. 2011). Likewise, a work on C. ternatea showed an efficient system
where organogenic callus derived from leaf explants developed about 74 shoots per
culture when grown on MS medium supplemented with BAP and IAA (Arumugam
and Panneerselvam 2012). Root explants of T. indica cultured on MS medium sup-
plemented with BA or 2-isopentenyladenine (2iP) produced organogenic nodular
meristemoids (NMs), which showed direct shoot bud formation and somatic
embryogenesis, when maintained on induction medium. In 42 % of the explants,
NMs developed shoot buds and about 18–19 shoots per gram of NM tissue were
obtained (Chaudhari et al. 2004). Direct induction of nodular meristemoids from the
cut edge of the ‘young leaves’ reached the maximum of 25–26 shoots per explant of
T. indica (Haque and Ghosh 2013). Same number of shoot frequency was also
noticed by Nayeem et al. (2014) where adventitious roots were used as an explant.
Furthermore, direct shoot formation from hypocotyl segment was also observed in
E. ribes. A high frequency (84 %) adventitious shoot induction was observed from
hypocotyls on MS medium supplemented with additives and TDZ after 4 weeks of
culture (Annapurna and Rathore 2010). Propagation through semi-hardwood cut-
tings derived from 10 to 15 year old plants of C. paniculatus was achieved by treat-
ing with different concentrations of IAA, IBA and NAA. After 90 days of treatment,
the highest rooting (57 %) response with highest number of roots of about 77 was
obtained in IAA. Rooted cuttings exhibited 100 % survival in the experimental field
(Raju and Prasad 2007).
An attempt was made to achieve protoplast culture and plant regeneration in T.
indica using leaf mesophyll cells and cultures. Callus pieces of 0.2–0.4-mm size
200 S. Goyal et al.

were transferred on to MS medium supplemented with 2,4-D to obtain friable callus


and subsequently transferred to MS medium supplemented with TDZ and NAA
where optimum shoot regeneration (44 %) occurred with an average number of 12
shoots per callus. Whole plants were recovered following rooting of shoots (Thomas
2009). V. vinifera is the liana exploited to great extent through in vitro techniques.
In a study, different approaches to produce transgenic grapevines based on regen-
eration via embryogenesis from anther tissue of V. vinifera and three embryogenic
culture types (embryogenic callus, proliferating embryos, and a suspension) were
established. The three culture types were inoculated with A. tumefaciens or were
bombarded with microprojectiles. Transgenic plants were produced only from
Agrobacterium transformation experiments (Franks et al. 1998).

11.5.2 Somatic Embryogenesis

Somatic embryogenesis mostly occurs indirectly via an intervening callus phase/


cell suspension culture or directly from initial explant and can grow into seedlings
on suitable medium. This pathway has offered a great potential for the production
of plantlets and its biotechnological manipulation. Plant regeneration via somatic
embryogenesis has been demonstrated in C. ternatea, E. ribes and T. indica.
Amongst these, T. indica showed direct embryogenesis. An efficient procedure was
developed for inducing somatic embryogenesis from mature leaves of T. indica
(Chandrasekhar et al. 2006). Leaf sections were initially cultured on MS medium
supplemented with TDZ and 2,4-D for induction of somatic embryos and 18 globu-
lar embryos were developed from each explant. After 6 weeks of time, 2 % sodium
alginate encapsulated embryos were placed on MS basal medium and 65 % of the
synthetic seeds formed plantlets. The percent of survival during hardening was
65–70 %.
Unlike T. indica, indirect somatic embryogenesis was achieved in C. ternatea
and E. ribes, using cotyledonary explants and leaf explants, respectively (Kumar
and Thomas 2012; Raghu et al. 2011). In C. ternatea, optimum embryogenic callus
(75 %) was induced on MS medium supplemented with 2, 4-D. On subculturing the
callus frequency of 83 % with 37 embryos per gram callus, was observed when the
MS medium was supplemented with BA, NAA, and ABA. Synthetic seeds were
produced by encapsulating embryos in calcium alginate gel which showed germina-
tion up to 92 %. The synthetic seeds were stored at 4 °C and lab conditions (25 ± 2 °C)
up to 5 months. The synthetic seeds kept at 4 °C showed 86 % viability even after 5
months of storage. Plantlets of both somatic embryos and synthetic seeds origin
were transferred to soil successfully. In the case of E. ribes, embryos were devel-
oped for the increased production of embelin content as it is well established that
the organogenic cultures accumulate more metabolites than the unorganized cul-
tures (Sharma et al. 2011a). The highest differentiation of embryos (56.5 %) was
seen after 6 weeks of culture. HPLC-UV assay demonstrated the highest embelin
content (5.33 % w/w) in the embryogenic callus cultures. In almost all the above
11 A Review of Biotechnological Approaches to Conservation and Sustainable… 201

mentioned micropropagation experiments, plantlets derived via somatic embryo-


genesis or shoot organogenesis, were successfully hardened and transferred to
the field with a success frequency ranging from 60 to 100 %. The data also show
potentiality of artificial seeds delivering somatic embryos to the field. However,
more refinement of the methods is necessary for their efficient exploitation.
Micropropagation protocols have been developed for A. precatorius, C. panicu-
latus, C. ternatea, C. hirsutus, E. ribes, G. glabra, H. indicus, P. tuberosa, T. indica
and T. cordifolia. Among these plants C. paniculatus (~48 shoots/explants), C. hir-
sutus (~45 shoots/explant), E. ribes (~65 shoots/explant), C. ternatea (~74 shoots/
culture), P. tuberosa (~37 shoots/culture) and T. indica (~67 shoots/culture) showed
highly efficient micropropagation systems. Micropropagation systems like that
achieved in E. ribes and T. indica where direct organogenesis has been developed,
are quite promising for mass scale production as they are less time consuming and
cost effective methods. Further, the studies undertaken by (Haque and Ghosh 2013),
on T. indica showed that after 18 months of field transfer, ~69 % of regenerated
plants produce flowers and ~29 % produce fruits. Seeds of these plants show ~78 %
viability, more or less similar with in vivo grown plants with ~72 % viability on in
vitro germination, which ensures the fertility of the regenerants. Moreover, result of
RAPD markers analysis indicated that all the regenerants were true-to-type and
there were no somaclonal variations among these regenerants. Such types of studies
are pre-requisites before elevating the laboratory scale protocols to commercial
level. Comprehensive and complete protocols are need of the hour for all the medic-
inal plants in general and lianas in particular for their efficient conservation and
sustainable utilization.

11.6 Conclusions

Lianas constitute an important plant life-form common to most of the tropical for-
ests. Lianas contribute substantially to the diversity of forests, provide food and are
widely used by the local people for medicine due to presence of bioactive molecules
(Bongers et al. 2002). Moreover, biological activity and clinical studies have to be
encouraged which can establish these plants as a potential source of standard drugs.
Though several works describe in vitro growth and production of SMs, remark-
able achievements have been attained in case of P. tuberosa and A. precatorius. P.
tuberosa cultures were scaled-up in a 2-l bioreactor where maximum yield of
~91 mg/l of isoflavonoid was recorded. Such type of laboratory scale production of
SM achieved in medicinal plants gives lead but necessitates much more experimen-
tation at the bioreactor level using combination of elicitors, growth retardants, pre-
cursors and sugar to enhance yields towards commercial production of
phytochemicals. The biotransformation of b-artemether by cell suspension cultures
of G. glabra is another new avenue. This was an effort to synthesize highly efficient
artemisinin derivatives/analogs with improved clinical efficiency. Comparison of
biological activities of in vitro plants’ to the in vivo plants’ extracts helps in proving
202 S. Goyal et al.

and establishing the in vitro cultures as an efficient alternative to in vivo plants.


Scientific validation of traditional medicinal plants using high through put screen-
ing methods will prove their age old usage.
It has also been found that genetically transformed cultures like Agrobacterium
mediated cultures are fast growing and biochemically and genetically more stable
over long periods. Transformation protocols with high yielding lines were success-
fully established in A. precatorius, T. indica, G. glabra and C. ternatea. Efficient
use of biotechnological techniques like in vitro cryopreservation, DNA fingerprint-
ing, micropropagation and ecorestoration support the in situ conservation of lianas.
Synthetic seeds are a potential method for delivering somatic embryos to the field
which can be used for germplasm conservation. Thus the biotechnological tech-
niques help in replenishments through macro- and micro-propagation methods, pro-
duction of metabolites in high yielding cell lines through cell suspension cultures.
Molecular markers and metabolomic provide information about diversity and
metabolites production. Therefore, biotechnological approaches continue to serve
as key methods towards the understanding, production and conservation of lianas in
particular and medicinal plants in general.

Acknowledgements Research programmes on medicinal plants in the Department of Botany


were supported by the University Grants Commission (UGC), New Delhi under SAP-DSA
Phase-II, Department of Science and Technology (DST) and Department of Biotechnology (DBT),
Govt. of India New Delhi to Prof. K.G. Ramawat.

References

Ananth A, Lakshman K, Rajasekaran PE (2011) In vitro propagation of Celastrus paniculatus


Willd., a threatened medicinal plant. J Genet Evol 4:20
Annapurna D, Rathore TS (2010) Micropropagation of Embelia ribes Burm f. through prolifera-
tion of adult plant axillary shoots. In Vitro Cell Dev Biol Plant 46(2):180–191
Anon (1990) The ayurvedic pharmacopoeia of India, part I, vol II. Ministry of Health and Family
Welfare, Government of India, pp 123–124
Anon (2008) National medicinal plants board. Ministry of Health and Family Welfare press
releases, 27 Feb 2008. Available online at http://pib.nic.in/release/release.asp?relid= 35664
Anoop A, Jegadeesan M (2003) Biochemical studies on the anti-ulcerogenic potential of
Hemidesmus indicus R.Br. var. indicus. J Ethnopharmacol 84(2–3):149–156
Ansari MN, Bhandari U, Islam F, Tripathi CD (2008) Evaluation of antioxidant and neuroprotec-
tive effect of ethanolic extract of Embelia ribes Burm in focal cerebral ischemia/reperfusion-
induced oxidative stress in rats. Fundam Clin Pharmacol 22:305–314
Arora J, Roat C, Goyal S, Ramawat KG (2009) High stilbenes accumulation in root cultures of
Cayratia trifolia (L.) Domin grown in shake flasks. Acta Physiol Plant 31:1307–1312
Arora J, Goyal S, Ramawat KG (2010) Enhanced stilbene production in cell cultures of Cayratia
trifolia through co-treatment with abiotic and biotic elicitors and sucrose. In Vitro Cell Dev
Biol Plant 46(5):430–436
Arumugam M, Panneerselvam R (2012) In vitro propagation and antibacterial activity of Clitoria
ternatea Linn. Asian Pac J Trop Biomed 2:S870–S875
Badole S, Patel N, Bodhankar S, Jain B, Bhardwaj S (2006) Antihyperglycemic activity of aqueous
extract of leaves of Cocculus hirsutus (L.) Diels in alloxan-induced diabetic mice. Indian J
Pharmacol 38:49–53
11 A Review of Biotechnological Approaches to Conservation and Sustainable… 203

Badole SL, Bodhankar SL, Patel NM, Bhardwaj S (2009) Acute and chronic diuretic effect of etha-
nolic extract of leaves of Cocculus hirsutus (L.) Diles in normal rats. J Pharm Pharmacol
61(3):387–393
Balaguru B, John Britto SJS, Nagamurugan N, Natarajan D, Soosairaj S (2006) Identifying
conservation priority zones for effective management of tropical forests in Eastern Ghats of
India. Biodivers Conserv 15:1529–1543
Belchí-Navarro S, Almagro L, Lijavetzky D et al (2012) Enhanced extracellular production of
trans-resveratrol in Vitis vinifera suspension cultured cells by using cyclodextrins and methyl-
jasmonate. Plant Cell Rep 31:81–89
Bhandari U, Chaudhari HS, Khanna G, Najmi AK (2013) Antidiabetic effects of Embelia ribes
extract in high fat diet and low dose streptozotocin-induced type 2 diabetic rats. Front Life Sci
7:186–196
Bhat TM, Singh M, Tafazul M (2013) Need and importance of conservation of Tinospora cordifo-
lia-A threatened Medicinal plant. Indo Am J Pharm Res. ISSN NO. 3(5): 3515–3518
Biswas A, Roy M, Miah MAB et al (2007) In vitro propagation of Abrus precatorius L. – a rare
medicinal plant of Chittagong Hill tracts. Plant Tissue Cult Biotechnol 17(1): 59–64
Bongers F, Schnitzer SA, Traore D (2002) The importance of lianas and consequence for forest
management in West Africa. Bioterre, Rev. Inter. Sci.de la vie Terre, No Special: 59–70
Bonvicinia F, Mandroneb M, Antognonic F et al (2014) Ethanolic extract of Tinospora cordifolia
and Alstonia scholaris show antimicrobial activity towards clinical isolates of methicillin resis-
tant and carbapenemase producing bacteria. Nat Prod Lett 28:1438–1445
Bowling AJ, Vaughn KC (2009) Gelatinous fibers are widespread in coiling tendrils and twining
vines. Am J Bot 96:719–727
Cai Z, Riedel H, Thaw Saw NM et al (2011) Effects of pulsed electric field on secondary metabo-
lism of Vitis vinifera L. cv. Gamay Fréaux suspension culture and exudates. Appl Biochem
Biotechnol 164(4):443–453
Chakrabarty M, Bhat P, Kumari S et al (2012) Cortico-hippocampal salvage in chronic aluminium
induced neurodegeneration by Celastrus paniculatus seed oil: neurobehavioural, biochemical,
histological study. J Pharmacol Pharmacother 3:161–171
Chandrasekhar T, Hussain TM, Gopal GR, Rao JVS (2006) Somatic embryogenesis of Tylophora
indica (Burm.f.) Merril., an important medicinal plant. Int J Appl Sci Eng 4(1):33–40
Chaudhari KN, Ghosh B, Jha S (2004) The root: a potential new source of competent cells for
high-frequency regeneration in Tylophora indica. Plant Cell Rep 22:731–740
Chaudhari KN, Ghosh B, Tepfer D, Jha S (2005) Genetic transformation of Tylophora indica with
Agrobacterium rhizogenes A4: growth and tylophorine productivity in different transformed
root clones. Plant Cell Rep 24:25–35
Chen X, Rong L, Liang T et al (2013) Puerarin improves metabolic function leading to hepatopro-
tective effects in chronic alcohol-induced liver injury in rats. Phytomedicine 20:849–852
Cheruvathur MK, Najeeb N, Thomas TD (2013) In vitro propagation and conservation of Indian
sarsaparilla, Hemidesmus indicus L. R. Br through somatic embryogenesis and synthetic seed
production. Acta Physiol Plant 35(3):771–779
Chintalwar GJ, Gupta S, Roja G, Bapat VA (2003) Protoberberine alkaloids from callus and cell
suspension cultures of Tinospora cordifolia. Pharm Biol 41(2):81–86
Dao PTA, Hai NX, Nhan NT, Quan TL, Mai NTT (2012) Study on DPPH free radical scavenging
and lipid peroxidation inhibitory activities of Vietnamese medicinal plants. Nat Prod Sci
18:1–7
Darwin C (1865) On the movements and habits of climbing plants. Longman, Green, Longman,
Roberts & Green/Williams & Norgate, London
Das S, Devaraj SN (2006) Antienterobacterial activity of Hemidesmus indicus R. Br root extract.
Phytother Res 20(5):416–421
Dass S, Ramawat KG (2009) Studies on somatic cell variability in Commiphora wightii (Arnott.)
Bhandari for guggulsterone production. Nat Prod Radiance 8(5):532–536
Delmas D, Lancon A, Colin D et al (2006) Resveratrol as a chemopreventive agent: a promising
molecule for fighting cancer. Curr Drug Targets 7:423–442
204 S. Goyal et al.

Devi BP, Boominathan R, Mandal SC (2003) Anti-inflammatory, analgesic and antipyretic proper-
ties of Clitoria ternatea root. Fitoterapia 74:345–349
Dhanush KB, Suguna R, Satyanarayana ML et al (2013) Hepatoprotective effects of Tinospora
cordifolia extract in streptozotocin induced damage in diabetic rats. Indian J Vet Pathol
37(2):153–158
Dhavala A, Rathore TS (2010) Micropropagation of Embelia ribes Burm f. through proliferation
of adult plant axillary shoots. In Vitro Cell Dev Biol Plant 46:180–191
Dhavala A, Srivastva A, Rathore TS (2013) Impact of population structure, growth habit and seed-
ling ecology on regeneration of Embeliaribes Burm.f. – approaches toward a quasi in-situ con-
servation strategy. Am J Plant Sci 4:28–35
Dirican E, Turkez H (2014) In vitro studies on protective effect of Glycyrrhiza glabra root extracts
against cadmium-induced genetic and oxidative damage in human lymphocytes. Cytotechnology
66(1):9–16
Dixit AK, Vaidya S (2010) Agrobactirium rhizogenes induced hairy root development and its effect
on production of glycyrrhizin in Abrus precatorius (L). Int J Curr Res 6:33–38
Duhan MP, Kajla S, Chaudhury A (2014) Optimization of media for establishment of Glycyrrhiza
glabra micropropagation. Ann Agri Bio Res 19(2):197–201
Ewers FW, Fisher JB, Fichtner K (1991) Water flux and xylem structure in vines. In: Putz FE,
Mooney HA (eds) The biology of vines. Cambridge University Press, Cambridge,
pp 127–179
Ferruzzi L, Turrini E, Burattini S et al (2013) Hemidesmus indicus induces apoptosis as well as
differentiation in a human promyelocytic leukemic cell line. J Ethnopharmacol 147:84–91
Filip A, Clichici S, Daicoviciu D et al (2011) Chemopreventive effects of Calluna vulgaris and
Vitis vinifera extracts on uvb-induced skin damage in skh-1 hairless mice. J Physiol Pharmacol
62:385–392
Franks T, He DG, Thomas M (1998) Regeneration of transgenic shape Vitis vinifera L. Sultana
plants: genotypic and phenotypic analysis. Mol Breed 4(4):321–333
Gadapuram TK, Murthy JS, Rajannagari RR, Kandati V, Choda PK, Shukla RN (2013)
Ephroprotective activity of Cocculus hirsutus leaf extract in 5/6 nephrectomized rat model. J
Basic Clin Physiol Pharmacol 24(4):299–306
Gartner BL, Bullock SH, Mooney HA, Brown VB, Whitbeck JL (1990) Water transport properties
of vine and tree stems in a tropical deciduous forest. Am J Bot 77:742–749
Gaur R, Patel S, Verma RK et al (2013) Biotransformation of artemisinin derivatives by Glycyrrhiza
glabra, Lavandula officinalis, and Panax quinquefolium. Med Chem Res 23(3):1202–1206
Gentry AH, Dodson C (1987) Contribution of non-trees to species richness of tropical rain forest.
Biotropica 19:149–156
George S, Geetha SP, Anu A, Indra B (2010) In vitro conservation studies in Hemidesmus indicus,
Decalpis hamiltoni and Ulteria salicifolialin. In: Proceedings of the 22nd Kerala Science
Congress, Kerala state council for science technology and environment, Thiruvanthapuram,
pp 258–259
Goyal S, Ramawat KG (2008a) Synergistic effect of morphactin on cytokinin-induced production
of isoflavonoid in cell cultures of Pueraria tuberosa (Roxb. ex. Willd.) DC. Plant Growth Regul
55:175–181
Goyal S, Ramawat KG (2008b) Ethrel treatment enhanced isoflavonoids accumulation in cell sus-
pension cultures of Pueraria tuberosa, a woody legume. Acta Physiol Plant 30:849–853
Goyal S, Sharma V, Ramawat KG (2011) Marked effect of Cuscuta on puerarin accumulation in
cell cultures of Pueraria tuberosa grown in shake flasks and a bioreactor. Plant Biotechnol Rep
5(2):121–126
Gris EF, Mattivi F, Ferreira EA et al (2011) Stilbenes and tyrosol as target compounds in the
assessment of antioxidant and hypolipidemic activity of Vitis vinifera red wines from Southern
Brazil. J Agric Food Chem 59(14):7954–7961
Gul MZ, Ahmad F, Kondapi AK et al (2013) Antioxidant and antiproliferative activities of Abrus
precatorius leaf extracts – an in vitro study. BMC Complement Altern Med 13:53
Gupta S, George P, Gupta V et al (1979) Tylophora indica in bronchial asthma – a double blind
study. Indian J Med Res 69:981–989
11 A Review of Biotechnological Approaches to Conservation and Sustainable… 205

Gupta V, Kak A, Singh BB (1997) Studies on seed germination and seedling vigour in liquorice
(Glycyrrhiza glabra). JMAPS 2:412–413
Gupta J, Kumar D, Gupta A et al (2012) Evaluation of gastric anti-ulcer activity of methanolic
extract of Cayratia trifolia in experimental animals. Asian Pac J Trop Dis 2(2):99–102
Gupta A, Maheshwari DK, Khandelwal G (2013) Antibacterial activity of Glycyrrhiza glabra roots
against certain gram-positive and gram-negative bacterial strains. J Appl Nat Sci
5(2):459–464
Gupta M, Duhan P, Kajla S (2014) Optimization of media for establishment of Glycyrrhiza glabra
micropropagation. Ann Agri Bio Res 19:197–201
Guru Kumar D, Sonumol VM, Rathi MA (2011) Hepatoprotective activity of Cayratia trifolia (L.)
domin against nitrobenzene induced hepatotoxicity. Lat Am J Pharm 30(3):546–549
Haque SM, Ghosh B (2013) Field evaluation and genetic stability assessment of regenerated plants
produced via direct shoot organogenesis from leaf explant of an endangered ‘Asthma Plant’
(Tylophora indica) along with their in vitro conservation. Natl Acad Sci Lett 36(5):551–562
Hata Y, Raith M, Ebrahimi SN et al (2013) Antiprotozoal isoflavan quinones from Abrus precato-
rius ssp. Africanus. Planta Med 79:492–498
Islam AS (2009) Utilization of indigenous medicinal plants and their conservation in Bangladesh.
In: Akerele O, Heywood V, Synge H (eds) The conservation of medicinal plants. Cambridge
University Press, New York, p 323
Ismail N, Rani U, Batra A (2011) Crucial role of TDZ in the quick regeneration of multiple shoots
of Clitoria ternatea. Int J Pharm Sci Rev Res 6(2):23–26
Isnard S, Silk WK (2009) Moving with climbing plants from Charles Darwin’s time into the 21st
century. Am J Bot 96:1205–1221
Jain SK (2003) Indian Sarsaparilla (Hemidesmus, Anantmul), In: Medicinal plants. National Book
Trust, New Delhi, pp 95–96
Jain NN, Ohal CC, Shroff SK, Bhutada RH et al (2003) Clitoria ternatea and the CNS. Pharmacol
Biochem Behav 75:529–536
Jayaram S, Dharmesh SM (2011) Assessment of antioxidant potentials of free and bound phenolics
of Hemidesmus indicus (L) R.Br against oxidative damage. Pharmacogn Res 3(4):225–231
John DB, Benjamin PJ, Rathna K, Herin DS (2012) Abrus precatorius L. a medicinal plant with
potential as antibacterial agent. J Pharm Res 5:1207–1209
Joy B, Kumar N, Soumya MS, Radhika AR, Vibin M, Abraham A (2014) Embelin
(2,5-dihydroxy-3-undecyl-p-benzoquinone): a bioactive molecule isolated from Embelia
ribes as an effective photodynamic therapeutic candidate against tumor in vivo.
Phytomedicine 21(11):1292–1297
Kamkaen N, Wilkinson JM (2009) The antioxidant activity of Clitoria ternatea flower petal
extracts and eye gel. Phytother Res 23:1624–1625
Karuppusamy S (2009) A review on trends in production of secondary metabolites from higher
plants by in vitro tissue, organ and cell cultures. J Med Plant Res 3:1222–1239
Karwasara VS, Jain R, Tomar P, Dixit VK (2010) Elicitation as yield enhancement strategy for
glycyrrhizin production by cell cultures of Abrus precatorius Linn. In Vitro Cell Dev Biol Plant
46:354–362
Karwasara VS, Tomar P, Dixit VK (2011) Oxytocin influences the production of glycyrrhizin from
cell cultures of Abrus precatorius Linn. Plant Growth Regul 65:401–405
Kazmi GM, Afzal M, Upadhyay G, Singh R, Habtemariam S (2013) Antidepressant-like activity
of Embelin isolated from Embelia ribes. Phytopharmacology 4(1):87–95
Kuchana V, Sampathi S, Begum R et al (2014) Antimicrobial activity of roots of Cocculus hirsutus
by agar plate diffusion method. Int J Phytopharmacol 5(4):320–322
Kumar GK, Thomas TD (2012) High frequency somatic embryogenesis and synthetic seed pro-
duction in Clitoria ternatea Linn. Plant Cell Tiss Org Cult 110:141–151
Kundu A, Mitra A (2014) Evaluating tyrosinase (monophenolase) inhibitory activity from fragrant
roots of Hemidesmus indicus for potent use in herbal products. Ind Crops Prod 52:394–399
Kundu A, Jawali N, Mitra A (2012) Shikimate pathway modulates the elicitor-stimulated accumu-
lation of fragrant 2-hydroxy-4-methoxybenzaldehyde in Hemidesmus indicus roots. Plant
Physiol Biochem 56:104–108
206 S. Goyal et al.

Lakshmi BV, Sudhakar M, Anisha M (2014) Neuroprotective role of hydroalcoholic extract of Vitis
vinifera against aluminium-induced oxidative stress in rat brain. Neurotoxicology 41:73–79
Lal D, Singh N (2010) Mass multiplication of Celastrus paniculatus wild: an important medicinal
plant under in vitro conditions via nodal segments. Int J Biodivers Conserv 2:140–145
Lee DW, Richards JH (1991) Heteroblastic development in vines. In: Putz FE, Mooney HA (eds)
The biology of vines. Cambridge University Press, Cambridge, pp 205–243
Lee WY, Chen KC, Chen HY et al (2014) Potential mitochondrial isocitrate dehydrogenase R140Q
mutant inhibitor from traditional Chinese medicine against cancers. BioMed Res Int 1–10.
10.1155/2014/364625
Li W, Asada Y, Yoshikawa K (2000) Flavonoid constituents from Glycyrrhiza glabra hairy root
cultures. Phytochemistry 55(5):447–456
Lin A, Liu Y, Huang Y et al (2008) Glycyrrhizin surface-modified chitosan nanoparticles for
hepatocyte-targeted delivery. Int J Pharm 359:247–253
Liu T, Zhao J, Ma L et al (2012) Hepatoprotective effects of total triterpenoids and total flavonoids
from Vitis vinifera L against immunological liver injury in mice. Evid Based Complement
Altern Med. doi:10.1155/2012/96938
Luczkiewicz M (2008) Research into isoflavonoid: phytoestrogens in plant cell cultures. In:
Ramawat KG, Merillon JM (eds) Bioactive molecules and medicinal plants. Springer,
Heidelberg, pp 54–84
Man S, Wang J, Gao W et al (2013) Chemical analysis and anti-inflammatory comparison of the
cell culture of Glycyrrhiza with its field cultivated variety. Food Chem 136:513–517
Maojun XU, Jufang D, Muyuan Z (2006) Nitric oxide mediates the fungal elicitor-induced puera-
rin biosynthesis in Pueraria thomsonii Benth suspension cells through a salicylic acid (SA)-
dependent and a jasmonic acid (JA) – dependent signal pathway. Sci China C Life Sci
49:1–11
Masuma R, Okuno T, Shahabuddin M et al (2014) Effect of Tinospora cordifolia on the reduction
of ultraviolet radiation‐induced cytotoxicity and DNA damage in PC12 cells. J Environ Sci
Health 49(6):416–421
Meena MK, Singh N, Patni V (2012) In vitro multiple shoot induction through axillary bud of
Cocculus hirsutus (L.) Diels: a threatened medicinal plant. Afr J Biotechnol 11(12):2952–2956
Mehrotra S, Khwaja O, Kukreja AK, Rahman L (2012) ISSR and RAPD based evaluation of
genetic stability of encapsulated micro shoots of Glycyrrhiza glabra following six months of
storage. Mol Biotechnol 52:262–268
Merillon JM, Ramawat KG (2007a) Biotechnology for medicinal plants: research need. In:
Ramawat KG, Merillon JM (eds) Biotechnology secondary metabolites. Sci Pub, Enfield,
pp 1–20. Enfield, New Hampshire, USA
Merillon JM, Ramawat KG (2007b) Understanding the regulatory mechanism of secondary metab-
olites production. In: Ramawat KG, Merillon JM (eds) Biotechnology secondary metabolites.
Sci Pub, Enfield, pp 233–254. Enfield, New Hampshire, USA
Michihara S, Tanaka T, Uzawa Y, Moriyama T (2012) Puerarin exerted anti-osteoporotic action
independent of estrogen receptor-mediated pathway. J Nutr Sci Vitaminol 58:202–209
Misra N, Misra P, Datta SK et al (2005) In vitro biosynthesis of antioxidants from Hemidesmus
indicus R. Br. cultures. In Vitro Cell Dev Biol 41:285–290
Mistry K, Mehta M, Mendpara N, Gamit S, Shah G (2010) Determination of antibacterial activity
and MIC of crude extract of Abrus precatorius L. Adv Biotech 10:25–27
Mohandas S, Sreekumar TR, Prakash V (2013) Anthelmintic activity of Vidangadi Churna. Asian
J Pharm Clin Res 6(3):94–95
Monago CC, Alumanah EO (2005) Antidiabetic effect of chloroform -methanol extract of Abrus
Precatorius Linn seed in alloxan diabetic rabbit. J Appl Sci Environ Manag 9:85–88
Morris JB (1999) Legume genetic resources with novel value added industrial and pharmaceutical
use. In: Janick J (ed) Prospective on new crops and new uses. ASHS Press, Alexandria,
pp 196–200
Moura RS, Costa GFD, Moreira ASB et al (2012) Vitis vinifera L. grape skin extract activates the
insulin-signalling cascade and reduces hyperglycaemia in alloxan-induced diabetic mice. J
Pharm Pharmacol 64:268–276
11 A Review of Biotechnological Approaches to Conservation and Sustainable… 207

Mujeeb M, Aeri V, Bagri P, Khan SA (2009) Hepatoprotective activity of methanolic extract of


Tylophora indica leaves. Int J Green Pharm 3(2):125–127
Mukhopadhyay S, Panda PK, Das DN et al (2014) Abrus agglutinin suppresses human hepatocel-
lular carcinoma in vitro and in vivo by inducing caspase-mediated cell death. Acta Pharmacol
Sin 35:814–824
Naik D, Dandge C, Rupanar S (2014) Determination of chemical composition and evaluation of
antioxidant activity of essential oil from Tinospora cordifolia (Willd.) leaf. J Essent Oil Bear
Plants 17(2):228–236
Nayeem A, Panchakshararadhya RM, Basappa VA (2014) In vitro plant regeneration using adven-
titious roots as explants in Tylophora indica. Asian J Plant Sci Res 4(1):15–18
Nithianantham K, Shyamala M, Chen Y et al (2011) Hepatoprotective potential of Clitoria ternatea
leaf extract against paracetamol induced damage in mice. Molecules 16(12):10134–10145
Oliveiraa DA, Salvador AA, Smânia A et al (2013) Antimicrobial activity and composition profile
of grape (Vitis vinifera) pomace extracts obtained by supercritical fluids. J Biotechnol
164:423–432
Pal RS, Ariharasivakumar G, Girhepunje K, Upadhyay A (2009) In vitro antioxidant activity of
phenolic and flavonoids compounds extracted from seeds of Abrus precatorius. Int J Pharm
Pharm Sci 2(1):136–140
Pandey N, Tripathi YB (2010) Antioxidant activity of tuberosin isolated from Pueraria tuberosa
Linn. J Inflamm 7:47
Pandey NK, Tewari KC, Tewari RN, Joshi GC, Pande VN, Pandey G (1993) Medicinal plants of
Kumaon Himalaya, strategies for conservation. In: Dhar U (ed) Himalaya biodiversity conser-
vation strategies. Himavikas Publication, Nainital, pp 293–302
Patel S, Gaur R, Upadhyaya M et al (2011) Glycyrrhiza glabra (Linn.) and (L.) cell suspension
cultures-based biotransformation of b-artemether. J Nat Med 65:646–650
Patil RH, Prakash K, Maheshwari VL (2010) Hypolipidemic effect of Celastrus paniculatus in
experimentally induced hypercholesterolemic Wistar rats. Indian J Clin Biochem 25:405–410
Patil SA, Sujaya M, Patil SB (2014) Aphrodisiac and phytochemical studies of Cocculus hirsutus
extracts in albino rats. Asian Pac J Reprod 3:23–29
Patnaik J, Debata BK (1996) Micropropagation of Hemidesmus indicus (L.) R. Br. through axillary
bud culture. Plant Cell Rep 15(6):427–430
Penetar DM, Toto LH, Farmer SL et al (2012) The isoflavone puerarin reduces alcohol intake in
heavy drinkers: a pilot study. Drug Alcohol Depend 126:251–256
Perveen S, Anis M, Aref IM (2013) Lipid peroxidation, H2O2 content, and antioxidants during
acclimatization of Abrus precatorius to ex vitro conditions. Biol Plant 57(3):417–424
Phulwaria M, Rai MK, Patel AK, Kataria V, Shekhawat NS (2013) A genetically stable rooting
protocol for propagating a threatened medicinal plant – Celastrus paniculatus. AoB Plants
5:pls054. doi:10.1093/aobpla/pls054
Pompo DG, Poli F, Mandrone M et al (2014) Comparative “in vitro” evaluation of the antiresorp-
tive activity residing in four Ayurvedic medicinal plants. Hemidesmus indicus emerges for its
potential in the treatment of bone loss diseases. J Ethnopharmacol 154(2):462–470
Preetha TS, Hemanthakumar AS, Krishnan PN (2012) Effect of plant growth regulators on high
frequency in vitro multiplication of a vulnerable woody medicinal climber Embelia ribes
Burm. f. J Med Plant Res 6(23):4011–4018
Putz FE (1984) The natural history of lianas on Barro Colorado Island, Panama. Ecology
65:1713–1724
Putz FE, Mooney HE (1991) The biology of vines. Cambridge University Press, Cambridge, 526pp
Radhakrishnan N, Kavitha V, Raja STK et al (2011) Embelin-A natural potential cosmetic agent. J
Appl Cosmetol 29:99–107
Raghu AV, Geetha SP, Martin G et al (2006) In vitro clonal propagation through mature nodes of
Tinospora cordifolia (Willd.) hook. f. & thoms: an important Ayurvedic medicinal plant. In
Vitro Cell Dev Biol Plant 42:584–588
Raghu AV, Unnikrishnan K, Geetha SP et al (2011) Plant regeneration and production of embelin
from organogenic and embryogenic callus cultures of Embelia ribes Burm. f.a vulnerable
medicinal plant. In Vitro Cell Dev Biol Plant 47:506–515
208 S. Goyal et al.

Raju NL, Prasad MNV (2007) Cytokinin-induced high frequency shoot multiplication in Celastrus
paniculatus Willd., a red listed medicinal plant. Med Aromat Plant Sci Biotechnol 1:133–137
Ramasubramaniaraja R, Niranjan Babu M (2011) Peptic ulcer and phytochemistry: an overview. J
Pharm Res 4(1):156–160
Ramawat KG, Goyal S (2008) Indian herbal drugs scenario in global perspectives. In: Ramawat
KG, Merillon JM (eds) Biotechnology: bioactive molecules. Springer, Heidelberg,
pp 323–345
Ramawat KG, Goyal S (2014) Comprehensive biotechnology. S. Chand, New Delhi, pp 261–272
Ramawat KG, Mathur M (2007) Factors affecting production of secondary metabolites. In:
Ramawat KG, Merillon JM (eds) Biotechnology secondary metabolites. Sci Pub, Enfield,
pp 59–101. Enfield, New Hampshire, USA
Ramawat KG, Dass S, Mathur M (2009) The chemical diversity of bioactive molecules and thera-
peutic potential of medicinal plants. In: Ramawat KG (ed) Herbal drugs: ethnomedicine to
modern medicine. Springer, Heidelberg, pp 7–32
Rani J, Rana S (2010) In Vitro propagation of Tylophora indica-influence of explanting season,
growth regulator synergy, culture passage and planting substrate. J Am Sci 6(12):385–392
Rao PR, Kumar VK, Viswanath RK et al (2005) Cardioprotective activity of alcoholic extract of
Tinospora cordifolia in ischemia-reperfusion induced myocardial infarction in rats. Biol Pharm
Bull 28:2319–2322
Rathore MS, Shekhawat NS (2009) Micropropagation of Pueraria tuberosa (Roxb. Ex Willd.) and
determination of puerarin content in different tissues. Plant Cell Tiss Org Cult 99:327–334
Rathore MS, Shekhawat NS (2011) Micropropagation of Maerua oblongifolia: a rare ornamental
from semi arid regions of Rajasthan, India. J Dev Biol Tissue Eng 3(8):92–98
Reddy K, Balaji M, Reddy PU et al (2009) Antifeedant and antimicrobial activity of Tylophora
indica. Afr J Biochem Res 3:393–397
Reppert A, Yousef GG, Rogers RB, Lila MA (2008) Isolation of radiolabeled isoflavones from
Kudzu (Pueraria lobata) root cultures. J Agric Food Chem 56:7860–7865
Richards PW (1996) The tropical rain forest: an ecological study. Cambridge University Press,
Cambridge, p 575
Roat C, Ramawat KG (2009a) Elicitor-induced accumulation of stilbenes in cell suspension cul-
tures of Cayratia trifolia (L.) Domin. Plant Biotechnol Rep 3:135–138
Roat C, Ramawat KG (2009b) Morphactin and 2iP markedly enhance accumulation of stilbenes in
cell cultures of Cayratia trifolia (L.) Domin. Acta Physiol Plant 31:411–414
Rout GR (2005) Micropropagation of Clitoria ternatea Linn. (Fabaceae) – an important medicinal
plant. In Vitro Cell Dev Biol Plant 41:516–519
Rowe NP, Isnard S, Speck T (2004) Diversity of mechanical architecture in climbing plants: an
evolutionary perspective. J Plant Growth Regul 23:108–128
Roychowdhury D, Ghosh B, Chaubey B et al (2013) Genetic and morphological stability of six-
year-old transgenic Tylophora indica plants. Nucleus 56:81–89
Sahu BD, Anubolu H, Koneru M, Kumar JM, Kuncha M, Rachamalla SS, Sistla R (2014a)
Cardioprotective effect of embelin on isoproterenol-induced myocardial injury in rats: possible
involvement of mitochondrial dysfunction and apoptosis. Life Sci 107:59–67
Sahu MC, Patnaik R, Padhy RN (2014b) In vitro combinational efficacy of ceftriaxone and leaf
extract of Combretum albidum G. Don against multidrug-resistant Pseudomonas aeruginosa
and host-toxicity testing with lymphocytes from human cord blood. J Acute Med 4(1):26–37
Sangeetha MK, Mohana Priya CD, Vasanthi HR (2013) Anti-diabetic property of Tinospora cordi-
folia and its active compound is mediated through the expression of Glut-4 in L6 myotubes.
Phytomedicine 20:246–248
Santamaria AR, Innocenti M, Mulinacci N et al (2012) Enhancement of viniferin production in
Vitis vinifera L. cv. Alphonse Lavallée Cell Suspensions by low-energy ultrasound alone and in
combination with methyl jasmonate. J Agric Food Chem 60(44):11135–11142
Saraswati S, Kanaujia PK, Kumar S et al (2013) Tylophorine, a phenanthraindolizidine alkaloid
isolated from Tylophora indica exerts antiangiogenic and antitumor activity by targeting vascu-
lar endothelial growth factor receptor 2–mediated angiogenesis. Mol Cancer 12:82
11 A Review of Biotechnological Approaches to Conservation and Sustainable… 209

Saravanan N, Nalini N (2007) Antioxidant effect of Hemidesmus indicus on ethanol-induced hepa-


totoxicity in rats. J Med Food 10(4):675–682
Savitha BC, Timmaraju R, Bhagyalaksami N, Ravishankar GA (2006) Different biotic and abiotic
elicitors influence betalain production in hairy root cultures of Beta vulgaris in shake flask and
bioreactor. Process Biochem 41:50–60
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17:223–230
Schnitzer SA, Carson WP (2001) Treefall gaps and the maintenance of species diversity in a tropi-
cal forest. Ecology 82:913–919
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical
forest canopy gaps: evidence for an alternative pathway of gap-phase regeneration. J Ecol
88:655–666
Senapati SK, Aparajita S, Rout GR (2013) Micropropagation and assessment of genetic stability in
Celastrus paniculatus: an endangered medicinal plant. Biology 68:627–632
Shabani L, Ehsanpour AA, Asghari G et al (2009) Glycyrrhizin production by in vitro cultured
Glycyrrhiza glabra elicited by methyl Jasmonate and salicylic acid. Russ J Plant Physiol
56:621–626
Sharma V, Ramawat KG (2013) Isoflavonoids. In: Ramawat KG, Merillon JM (eds) Natural prod-
ucts, phytochemistry, botany, metabolism, Springer-Verlag, Berlin, Heidelberg, Germany, vol
2. pp 1849–1865.
Sharma V, Goyal S, Ramawat KG (2009) Scale up production of isoflavonoids in cell suspension
cultures of Pueraria tuberosa grown in shake flasks and bioreactor. Eng Life Sci 9:267–271
Sharma V, Goyal S, Ramawat KG (2011a) Increased puerarin biosynthesis during in vitro shoot
formation in Pueraria tuberosa grown in Growtek bioreactor with aeration. Physiol Mol Biol
Plants 17:87–92
Sharma V, Goyal S, Ramawat KG (2011b) Biotechnology and agroforestry in Indian arid regions.
In: Lichtfouse E (ed) Genetics, biofuels and local farming system, vol 7, Sustainable Agriculture
Reviews. Springer, Dordrecht/Heidelberg, pp 309–345
Shin YW, Bae EA, Lee B et al (2007) In vitro and in vivo antiallergic effects of Glycyrrhiza glabra
and its components. Planta Med 73(3):257–261
Shrisha DL, Raveesha KA, Nagabhushan (2011) Bioprospecting of selected medicinal plants for
antibacterial activity against some pathogenic bacteria. J Med Plant Res 5(17):4087–4093
Siddique NA, Bari MA (2010) Plant regeneration from axillary shoot segments derived callus in
Hemidesmus indicus,(L.) R. Br.(Anantamul) an endangered medicinial in Bangladesh. Plant
Sci 5(1):61–67
Simons R, Vincken JP, Mol LAM et al (2011) Agonistic and antagonistic estrogens in licorice root
(Glycyrrhiza glabra). Anal Bioanal Chem 401:305–313
Singh J, Tiwari KV (2010) High-frequency in vitro multiplication system for commercial propaga-
tion of pharmaceutically important Clitoria ternatea L. – a valuable medicinal plant. Ind Crop
Prod 32(3):534–538
Siracusa L, Saija A, Cristani M et al (2011) Phytocomplexes from liquorice (Glycyrrhiza glabra
L.) leaves – chemical characterization and evaluation of their antioxidant, anti-genotoxic and
anti-inflammatory activity. Fitoterapia 82:546–556
Sivakumar V, Rajan MSD, Sadiq AM et al (2014) In vitro micropropagation of Tinospora cordifo-
lia (Willd.) Miers ex Hook. F. & Thoms – an important medicinal plant. J Pharmacogn
Phytochem 3(2):5–10
Sreedhar S, Prakash Kumar U, Reema Shree AB (2012) Pharmacognostic analysis of stem bark of
Combretum Albidum G. Donan unexplored medicinal plant. Pharmacogn J 4(28):13–16
Sreedhar S, Nitha B, Rema Shree AB (2013a) Antimicrobial activity of stem bark of Combretum
albidum G. Don; a traditional medicinal liana. Int J Pharm Sci Res 4(8):3184–3188
Sreedhar S, Kumarb UP, Girijaa TP, Remashree AB (2013b) Pharmacognostic standardisation of
Combretum albidum G. Don leaf; medicinally important liana. Pharmacogn J 5(6):247–255
Suri SS, Ramawat KG (1995) In vitro hormonal regulation of laticifers differentiation in Calotropis
procera. Ann Bot 75:477–480
210 S. Goyal et al.

Swain SS, Rout KK, Chand PK (2012) Production of triterpenoid anti-cancer compound taraxerol
in agrobacterium-transformed root cultures of butterfly pea (Clitoria ternatea L.). Appl
Biochem Biotechnol 168:487–503
Thavamani BS, Mathew M, Palaniswamy DS (2014) Anticancer activity of Cocculus hirsutus
against Dalton’s lymphoma ascites (DLA) cells in mice. Pharm Biol 52(7):867–872
Thippeswamy BS, Mahendran S, Biradar MI et al (2011) Protective effect of embelin against ace-
tic acid induced ulcerative colitis in rats. Eur J Pharmacol 654(1):100–105
Thomas TD (2009) Isolation, callus formation and plantlet regeneration from mesophyll proto-
plasts of Tylophora indica (Burm. f.) Merrill: an important medicinal plant. In Vitro Cell Dev
Biol Plant 45:591–598
Tré mouillaux-Guiller J (2013) Hairy root culture: an alternative toterpenoid expression platform. In:
Ramawat KG, Merillon JM (eds) Natural products. Springer, Berlin/Heidelberg, pp 2941–2970
Tripathi YB, Nagwani S, Mishra P et al (2012) Protective effect of Pueraria tuberosa DC. embed-
ded biscuit on cisplatin-induced nephrotoxicity in mice. J Nat Med 66:109–118
Valecha R, Dhingra D (2014) Antidepressant-like activity of Celastrus paniculatus seed oil in mice
subjected to chronic unpredictable mild stress. Br J Pharm Res 4:576–593
Ved DK, Goraya GS (2007) Demand and supply of medicinal plants in India. NMPB/FRLHT, New
Delhi/Bangalore
Verma PR, Itankar PR, Arora SK (2013) Evaluation of antidiabetic antihyperlipidemic and pancre-
atic regeneration, potential of aerial parts of Clitoria ternatea. Rev Bras Farm 23:819–829
Waffo-Teguo P, Krisa S, Richard T, Merillon JM (2008) Grapevine stilbenes and their biological
effects. In: Ramawat KG, Merillon JM (eds) Bioactive molecules and medicinal plants.
Springer, Heidelberg, pp 24–55
Weng JR, Yen MH, Lin WU (2013) Cytotoxic constituents from Celastrus paniculatus induce
apoptosis and autophagy in breast cancer cells. Phytochemistry 94:211–219
Wu Z, Xu Q, Zhang L et al (2009) Protective effect of resveratrol against kainate-induced temporal
lobe epilepsy in rats. Neurochem Res 34:1393–1400
Wu K, Liang T, Duan X et al (2013) Anti-diabetic effects of puerarin, isolated from Pueraria
lobata (Willd.), on streptozotocin-diabetogenic mice through promoting insulin expression and
ameliorating metabolic function. Food Chem Toxicol 60:341–347
Yin G, Cao L, Xu P et al (2011) Hepatoprotective and antioxidant effects of Glycyrrhiza glabra
extract against carbon tetrachloride (CCl4) induced hepatocyte damage in common carp
(Cyprinus carpio). Fish Physiol Biochem 37:209–216
Yoshida S, Lee JO, Nakamura K et al (2014) Effect of glycyrrhizin on Pseudomonal skin infections
in human-mouse chimeras. PLoS One 9(1):e83747
Younus M, Agarwal M (2013) Anti-inflammatory activity of Celastrus Paniculatus seeds. Int J
Chem Appl 5:39–44
Yue X, Zhang W, Deng M (2011) Hyper-production of 13C-labeled trans-resveratrol in Vitis vinif-
era suspension cell culture by elicitation and in situ adsorption. Biochem Eng J 53(3):292–296
Zhang HC, Liu JM, Chen HM et al (2011) Up-regulation of licochalcone a biosynthesis and secre-
tion by tween 80 in hairy root cultures of Glycyrrhiza uralensis Fisch. Mol Biol 47:50–56
Zhang Y, Chen Y, Shan Y, Wang D, Zhu C, Xu Y (2013) Effects of puerarin on cholinergic enzymes
in science the brain of ovariectomized guinea pigs. Int J Neurosci 123:783–791
Zhao J, Verpoorte R (2007) Manipulating indole alkaloid production by Catharanthus roseus cell
cultures in bioreactors: from biochemical processing to metabolic engineering. Phytochem Rev
6:435–457
Zheng H, Li Y, Wang Y et al (2014) Downregulation of COX-2 and CYP 4A signaling by
isoliquiritigenin inhibits human breast cancer metastasis through preventing anoikis resistance,
migration and invasion. Toxicol Appl Pharmacol 280(1):10–20
Zhu G, Wang X, Wu S et al (2014) Neuroprotective effects of puerarin on 1-methyl-4-phenyl-
1,2,3,6-tetrahydropyridine induced Parkinson’s disease model in mice. Photother Res
28:179–186
Chapter 12
Biological Invasion of Vines, Their Impacts
and Management

SM. Sundarapandian, C. Muthumperumal, and K. Subashree

Abstract Invasive species, from their natives enter into new areas where they
establish, proliferate, spread and affect the natural communities. They alter the local
biodiversity, cause changes in hydrology and ecosystem functions. In this chapter,
we review the global distribution of invasive vines, their impacts, widely-used con-
trol measures and future prospects. According to Global Invasive Species Database
(2013) and other sources, a total of 55 vines are considered as world’s worst invasive
species. Approximately 29 % of the vines have drifted from Asia to North America,
South America, Europe, Oceania and Africa. However, only 6 % of invasive vines
have invaded from North America to Asia and Oceania. Approximately, 21 %, 19 %
and 17 % of the invasive vines have spread to Europe, Asia and South America
respectively. A concise account on world’s top ten invasive climbers is provided in
this chapter. A detailed review on the ecology of two prominent invasive vines in
India viz. Mikania micrantha and Lantana camara are also included. Different con-
trol measures viz. physical, chemical, biological and cultural methods are in prac-
tice to contain the vigorous growth of several invasive vines. However, an integrated
approach has been proven to be most successful. Though invasive vines are noxious
and notorious to the environment, they also have some ecological and economic
benefits as ornamentals, edibles, medicinal plants etc. Considering their impacts on
environment on one side and the economic values on the other side, wiser manage-
ment of these species is emphasized.

SM. Sundarapandian (*) • K. Subashree


Department of Ecology and Environmental Sciences, Pondicherry University,
Puducherry 605014, India
e-mail: smspandi@yahoo.co.in
C. Muthumperumal
School of Biological Sciences, Department of Plant Sciences, Madurai Kamaraj University,
Madurai 625021, Tamil Nadu, India

© Springer International Publishing Switzerland 2015 211


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_12
212 SM. Sundarapandian et al.

12.1 Introduction

Invasion by exotic species has been an issue of serious concern all over the world as
they impact the structural integrity of any natural ecosystem. The term “invasive”
has many definitions, but in this context, it applies to species that enter into new
areas where they proliferate, spread and persist to cause negative impact to the envi-
ronment, economy and human health (Zheng et al. 2006). Alien species invasion
decreases species richness and abundance of native flora and fauna, which in turn,
alters biological communities spatially (Olden and Poff 2003; Sax and Gaines
2003; Winter et al. 2009) and ecosystem functions and services (Richardson et al.
2000; Hulme 2007; Vila et al. 2009, 2011; Pejchar and Mooney 2009; Pysek and
Richardson 2010; Pysek et al. 2012). Invasive or non-native plants comprise 8–47 %
of the total flora of United States of America (Reardon 2006). Once invasive plants
become established, they are not easily suppressed or eliminated, as they often
possess characteristics that favor their population increase.
In the past few centuries, humans have moved thousands of plants out of their
natural ranges for many purposes and in recent decades, many of these have become
naturalized or invasive (Binggeli 1996; Binggeli et al. 1998; Richardson 1998,
2011a, b; Richardson and Rejmanek 2004; Williams and Cameron 2006).“Tens
Rule” was proposed by Williamson and Fitter (1996) based on statistical analysis
of British plants and animals. According to this rule, on an average, one in ten of
introduced species escapes cultivation. Only one in ten of these escapes become
naturalized and in turn, only one in ten of these naturalized species may become
invasive. Due to numerous introductions over years, even this relatively smaller
percentage of the total plants introduced has constituted a great number.
Climate change may also lead to biotic invasions, and the effects need to be better
understood (Kueffer and Daehler 2009). Atmospheric CO2 increase may also accelerate
the spread of native and alien invasive vines in dense forest ecosystems that are currently
somewhat resistant to biotic invasions (Granados and Koerner 2002; Dietz et al. 2006;
Pauchard et al. 2009; Kueffer and Daehler 2009). Bioclimatic modeling predicts that
tropical vines may invade even temperate zones due to climate change (e.g. Cryptostegia
grandiflora invasion in northern New South Wales (Kriticos et al. 2003)).
Many woody plants have naturalized or become invasive only in the recent past
and their invasion ecology is largely unknown. Invasion ecology requires a com-
parative and an accurate assessment of the survival status of the invasive species,
thriving under various climatic conditions in different regions of the world. This
would help in development and design of general models pertaining to appropriate,
sustainable management policies. One best way to predict the invasiveness of an
introduced species is to assess whether it has invaded elsewhere in the globe, assum-
ing it had time to invade where it had been introduced (Rejmanek et al. 2005;
Gordon et al. 2010).This chapter deals with the global distribution of invasive vines.
We review the origin and current distribution of invasive vines, merits and demerits
of their invasion in ecosystems, their control measures and future prospects. Further,
we provide an updated crisp account on the worlds’ worst top ten invasive vines and
dealt in detail two major invasive vines.
12 Biological Invasion of Vines, Their Impacts and Management 213

12.2 Invasive Vines

Alien vines may act as transformer species (Richardson et al. 2000), changing the
microenvironment where they become dominant. Natural and man-made canopy
disturbances in rainforests pave the way to invasion of exotic vines (Floyd 1989),
which may smother both existing canopy vegetation as well as ground layer (Groves
and Willis 1999; Greenberg et al. 2001; Kriticos et al. 2003; Timmins and Reid 2000).
In addition, change in microenvironment, particularly intensity of light penetra-
tion might suppress native species regeneration. Canopy branches may break due to
the sheer weight of overtopping exotic vines (Harden and Fox 1988; Harden et al.
2004). Some species like Mikania micrantha tend to form dense ground cover
carpets, suppressing native flora.
Alien vine invasions change the diversity and abundance of plant-dwelling inver-
tebrates (Ernst and Cappuccino 2005). Dense smothering blankets of exotic vines
entangle and restrict the movement of some native fauna, thereby affecting their
ability to access water and other resources. Asparagus spp. forms masses of root
tubers which modify the biota of soil and litter, affecting litter decomposition rates
and also nutrient cycling, leading to competition for resources with native flora
(Raymond 1996; Groves and Willis 1999; Timmins and Reid 2000; Willis et al.
2003). Infestation by alien vines creates a humid microclimate at ground level, which
favours attack by pathogens, especially in riparian zones. Many exotic vines exhibit
a ruderal strategy by their ability to reproduce prolifically, both by asexual, vegeta-
tive propagules (for example Anredera cordifolia) and sexually via seeds. Seeds of
vines with fleshy fruits are readily dispersed by birds, for example Asparagus spp.
(Stansbury 2001) and Solanum spp.

12.3 Diversity and Distribution of Invasive Vines

12.3.1 Global Distribution

From the global species invasive database (http://www.issg.org/database/) and other


literary sources, a total of 55 vines are considered as most invasive species across
the globe (Table 12.1). It can be discerned that a sizeable proportion (~29 %) of the
vines have migrated from Asia to mainly North America, and then to South America,
Europe, Oceania and Africa. In contrast, only around 6 % of the vines have migrated
from North America to Asia and Oceania. About 13 % of the vines have spread from
South America to North America and to eastern continents like Asia, Africa and
Oceania. Around 11 % of the vines share the nativity of both North and South
Americas, from where they have expanded their boundaries of invasion to Oceania,
Asia, Europe, Africa and Caribbean Islands. One such important scrambler is
Lantana camara which has now become naturalized across the globe. Approximately,
only 10 % of the vines from Africa have invaded Oceania, Asia, North America and
South America. Vines whose native range overlaps with Asia and Oceania (6 %)
Table 12.1 List of invasive vine species with details on their nativity, continents invaded, habitat ecology, impacts on ecosystems, control measures and
214

resource values reported, in various studies


Invasive vine Native Continent(s)/area Control Resource
species Family continent(s) invaded Habitat Impacts on ecosystem measures value
Abrus precatorius Fabaceae Africa, North America, Agricultural areas, Alters soil nutrients Physical Medicinal
L. Asia, South America natural forests, and exerts allelopathy and
Oceania grasslands, ruderal/ chemical
disturbed sites, scrub/
shrub lands
Akebia quinata Lardizabalaceae Asia Europe, North Riparian zones, urban Outcompetes native Physical Edible and
(Houtt.) Dcne America areas, wetlands flora for light, water, and medicinal
nutrients, and space chemical
Ampelopsis Vitaceae Asia North America Riparian zones, Forms thick mats and Physical Ornamental
brevipedunculata ruderal/disturbed sites, outcompetes native and
(Maxim) Trautv. urban areas flora for light, water, chemical
nutrients and space
Anredera cordifolia Basellaceae South Africa, Oceania, Natural forests, Smothers native flora Physical, Ornamental
(Ten.) Steenis America Europe, North plantations, riparian and capable of break chemical
America zones branches as well and
causing gap formation cultural
Antigonon leptopus Polygonaceae North Tropics Ruderal/disturbed, Smothers native flora Physical Ornamental
Hook. & Arn. America urban areas
Buddleja Scrophulariaceae Africa Tropics/ Natural forests, Causes throat allergies Physical Ornamental
madagascariensis naturalized plantations, ruderal/ and toxic to cattle and
Lam. disturbed, urban areas chemical
Cardiospermum Sapindaceae North Oceania Natural forests, Forms thick mats and Physical Medicinal
grandiflorum Sw. America, riparian zones, urban outcompetes native and
South areas flora for light, water, chemical
America, nutrients, and space
Africa
SM. Sundarapandian et al.
Invasive vine Native Continent(s)/area Control Resource
species Family continent(s) invaded Habitat Impacts on ecosystem measures value
12

Celastrus Celastraceae Asia North America, Agricultural areas, Overtops native flora Physical, Ornamental
orbiculatus Thunb. Oceania coastland, natural and seeds have a chemical
forests, planted forests, unique ability to and
range/grasslands, germinate under a biological
riparian zones, ruderal/ variety of light
disturbed sites, scrub/ conditions
shrub lands, urban
areas
Clematis terniflora Ranunculaceae Asia North America, Coastland, natural Potential to climb Not Medicinal
DC Europe forests, ruderal/ 7.6–9.1 m and smother specified
disturbed sites fully grown trees
Clematis vitalba L. Ranunculaceae Asia, Oceania, North Agricultural areas, Smothers the tallest Physical, Medicinal
Europe America coastland, natural forest trees, forming a chemical
forests, planted forests, dense, light-absorbing and
range/grasslands, canopy that suppresses biological
riparian zones, ruderal/ all vegetation beneath it
disturbed sites, scrub/ and causes breakage of
shrub lands, urban branches
areas
Coccinia grandis Cucurbitaceae Africa, Oceania Agricultural areas, Smothering vine with Physical, Edible
(L.) Voigt Asia, natural forests, planted extensive tuberous chemical
Oceania forests, ruderal/ root system and acts as and
Biological Invasion of Vines, Their Impacts and Management

disturbed sites a host for melon fly biological


and ring spot virus
Cryptostegia Apocynaceae Africa Oceania, Asia, Agricultural areas, Poisonous and forms Biological Ornamental
grandiflora (Roxb. North America natural forests, planted impenetrable thickets
ex R. Br.) R. Br. forests, range/
grasslands, riparian
zones, ruderal/
disturbed sites, scrub/
shrub lands, water
courses, wetlands
215

(continued)
Table 12.1 (continued)
216

Invasive vine Native Continent(s)/area Control Resource


species Family continent(s) invaded Habitat Impacts on ecosystem measures value
Cryptostegia Apocynaceae Africa Asia, Oceania, Agricultural areas, Prevents the successful Physical, Economic
madagascariensis North America coastland, estuarine establishment of native chemical (in making
Bojer ex Dcne. habitats, natural flora and ropes and
forests, range/ biological fishing lines)
grasslands, urban areas,
water courses
Cynanchum Apocynaceae Asia, North America Shaded woodlands, Forms dense stands, Physical, Ornamental
rossicum (Kleopov) Europe roadsides, dry displaces native chemical
Barbarich grasslands, pastures, vegetation, exerts and
coastal shores, allelopathy and serves biological
agricultural areas, as a host for various
natural forests, planted pests
forests, range/
grasslands, riparian
zones, ruderal/
disturbed, scrub/shrub
lands, urban areas,
wetlands
Delairea odorata Asteraceae Africa Europe, Oceania Coastland, lakes, Forms dense mats that Physical, Not assessed
Lem. natural forests, range/ prevents seedling chemical,
grasslands, riparian establishment of native cultural
zones, ruderal/ plants and competes and
disturbed sites, scrub/ for soil nutrients and biological
shrublands, water water. Causes soil
courses, wetlands erosion along
watercourses due to its
shallow root system
SM. Sundarapandian et al.
Invasive vine Native Continent(s)/area Control Resource
species Family continent(s) invaded Habitat Impacts on ecosystem measures value
12

Dioscorea Dioscoreaceae Africa, North America, Ruderal/disturbed sites, Forms impenetrable Physical, Edible and
bulbifera L. Asia, South America urban areas mats and smothers chemical medicinal
Oceania seedlings and trees. and
Spreads fast by biological
anthropogenic control
dispersal of the bulbils
Dioscorea Dioscoreaceae Asia North America Natural forests, Smothers native flora Physical, Edible and
oppositifolia L. riparian zones, ruderal/ and is capable of chemical medicinal
disturbed sites, urban breaking branches as and
areas, wetlands well as causing gap biological
formation
Dipogon lignosus Fabaceae Africa Asia, Oceania, Natural forests, Smothers native flora Physical Economic
(L.) Verdc. North America, ruderal/disturbed sites, and is capable of and (nitrogen
South America urban areas breaking branches as chemical fixer)
well as causing gap
formation
Dolichandra Bignoniaceae North Asia, Oceania, Natural forests, planted Smothers and disrupts Physical, Ornamental
unguis-cati (L.) America, Caribbean Islands forests, range/ native plant chemical, and
A.H. Gentry South grasslands, ruderal/ communities biological medicinal
America disturbed sites, urban and
areas cultural
Epipremnum Araceae Asia, North America Not assessed Smothers native flora Not Not assessed
Biological Invasion of Vines, Their Impacts and Management

pinnatum (L.) Oceania and toxic specified


Engl.
Euonymus fortunei Celastraceae Asia Africa, Europe, Natural forests, Outcompetes native Physical Ornamental
(Turcs.) North America, riparian zones, ruderal/ flora for light, water, and
Hand.-Mazz. South America disturbed sites, urban nutrients, and space chemical
areas and along
sidewalks
(continued)
217
Table 12.1 (continued)
218

Invasive vine Native Continent(s)/area Control Resource


species Family continent(s) invaded Habitat Impacts on ecosystem measures value
Fuchsia Onagraceae South Oceania, Europe, Natural forests, planted Outcompetes native Physical Not assessed
magellanica Lam. America North America forests, ruderal/ flora for light, water, and
disturbed nutrients, and space chemical
Hedera helix L. Araliaceae Africa, North America, Coastland, estuarine Prevents the successful Physical, Ornamental
Asia, Oceania habitats, natural establishment of native chemical
Europe forests, planted forests, flora and
riparian zones, urban biological
areas, wetlands
Hiptage Malpighiaceae Asia North America, Natural forests Smothers and disrupts Not Ornamental
benghalensis (L.) Oceania native plant specified
Kurz. communities
Ipomoea aquatica Convolvulaceae Asia Africa, Australia, Lakes, riparian zones, Outcompetes native Chemical Edible and
Forsk. South America water courses, wetlands flora for light, water, and medicinal
nutrients, and space biological
Ipomoea cairica Convolvulaceae Africa, Oceania Agricultural areas, Smothers and disrupts Not Not assessed
(L.) Sweet Asia, North natural forests, planted native plant specified
America forests, ruderal/ communities. Also
disturbed sites, open exerts allelopathy
woodlands and coastal
sand dunes
Ipomoea purpurea Convolvulaceae South East Africa/ Waste lands, cultivated Outcompete native Physical, Ornamental
(L.) Roth America naturalised ground, riparian area, species for nutrients, chemical
wetlands and coastal water and sunlight. and
habitats Parts of this plant, cultural
including the seed are
poisonous if ingested
SM. Sundarapandian et al.
Invasive vine Native Continent(s)/area Control Resource
species Family continent(s) invaded Habitat Impacts on ecosystem measures value
12

Lantana camara L. Verbenaceae North Naturalized Agricultural areas, Smothers and disrupts Physical Ornamental,
America, coastland, natural native plant and medicinal
South forests, planted forests, communities biological and
America range/grasslands, economic
riparian zones, ruderal/ (making
disturbed sites, scrub/ furniture,
shrub lands, urban firewood,
areas, wetlands mulching,
etc.)
Lonicera japonica Caprifoliaceae Asia North America, Natural forests, Outcompetes native Physical Forage and
Thunb. South America, ruderal/disturbed, flora for light, water, and medicinal
Europe urban areas, thickets, nutrients, and space chemical
fence rows and
well-drained forest
sites
Lygodium Lygodiaceae Africa, North America Estuarine habitats, lakes, Smothers and disrupts Chemical Medicinal
japonicum Thunb. Asia, natural forests, planted native plant and
ex Murr.) Sw. Europe forests, riparian zones, communities biological
ruderal/disturbed sites,
water courses, wetlands
Merremia peltata Convolvulaceae Asia, North Africa Agricultural areas, Forms impenetrable Physical, Not assessed
(L.) Merr. America, natural forests, planted mats and smothers chemical,
Biological Invasion of Vines, Their Impacts and Management

Oceania forests, range/ seedlings and trees biological


grasslands, riparian and
zones, ruderal/ cultural
disturbed sites, scrub/
shrub lands, urban
areas
(continued)
219
Table 12.1 (continued)
220

Invasive vine Native Continent(s)/area Control Resource


species Family continent(s) invaded Habitat Impacts on ecosystem measures value
Merremia tuberose Convolvulaceae North Asia, Oceania Agricultural areas, Outcompetes native Physical, Ornamental
(L.) Rendle America coastland, natural flora for light, water, chemical
forests, planted forests, nutrients, and space and
ruderal/disturbed sites, and is toxic biological
scrub/shrublands,
urban areas
Mikania micrantha Asteraceae North Asia, Oceania Agricultural areas, Outcompetes native Chemical Not assessed
(L.) Kunth. America, coastland, natural and cash crops and
South forests, planted forests, biological
America ruderal/disturbed sites,
scrub/shrub lands,
urban areas
Mimosa diplotricha Fabaceae South Oceania, Asia, Not specified Forms a thorny mat Physical, Not assessed
C. Wright ex America Africa over the natural chemical
Sauvalle vegetation, preventing and
animals from biological
accessing and utilising
natural vegetation
Paederia foetida L. Rubiaceae Asia North America, Natural forests, Invades pasture land Physical Not assessed
South America ruderal/disturbed sites, and causes problems and
wetlands along highways and on chemical
power lines
Passiflora edulis Passifloraceae South Africa Agricultural areas, Not assessed Not Edible
Sims America natural forests, ruderal/ specified
disturbed sites, scrub/
shrub lands
SM. Sundarapandian et al.
Invasive vine Native Continent(s)/area Control Resource
species Family continent(s) invaded Habitat Impacts on ecosystem measures value
12

Passiflora foetida Passifloraceae South Asia, North Agricultural areas, Forms a dense ground Not Edible and
L. America America ruderal/disturbed sites cover which prevents specified medicinal
or delays the
establishment of other
species
Passiflora Passifloraceae North Oceania, Europe Ruderal/disturbed sites Not assessed Not Not assessed
maliformis L. America, specified
South
America
Passiflora suberosa Passifloraceae South Asia Natural forests, range/ Not assessed Not Not assessed
L. America grasslands, ruderal/ specified
disturbed sites, scrub/
shrub lands
Passiflora Passifloraceae South Africa, Oceania, Agricultural areas, Suppresses tree Physical, Edible and
tarminiana America Asia, North natural forests, planted regeneration, topples chemical ornamental
Coppens & America forests, scrub/ shallow-rooted trees and
V.E. Barney shrublands and kills standing trees biological
blocking sunlight
Persicaria Polygonaceae Asia, North Asia, North Natural forests, planted Causes mechanical Physical, Not assessed
perfoliata (L.) America America forests, ruderal/ damage to native trees chemical
H. Gross disturbed sites, urban due to its enormous and
areas, wetlands weight biological
Biological Invasion of Vines, Their Impacts and Management

Pueraria montana Fabaceae Asia, Africa, Europe, Agricultural areas, Smothers native plants Physical, Edible and
(Willd.) Maesen & Oceania North America natural forests, planted by forming dense chemical ornamental
S. Almeida forests, range/grasslands, blankets and causes and
riparian zones, ruderal/ mechanical damage to biological
disturbed sites, scrub/ native trees due to its
shrub lands, urban areas, enormous weight
thickets and deciduous
forests
(continued)
221
Table 12.1 (continued)
222

Invasive vine Native Continent(s)/area Control Resource


species Family continent(s) invaded Habitat Impacts on ecosystem measures value
Rubus discolor Rosaceae Europe North America, Agricultural areas, Forms large Chemical Forage
Weihe & Nees Africa, Oceania planted forests, riparian impenetrable thickets and
zones, ruderal/ of prickly canes biological
disturbed sites, water blocking access to
courses humans and livestock
Rubus ellipticus Rosaceae Asia North America, Agricultural areas, Forms large Physical, Edible,
Sm. Europe natural forests, range/ impenetrable thickets chemical medicinal
grasslands, ruderal/ of prickly canes and and
disturbed sites blocking access to biological economic
humans and livestock (dyeing)
Rubus moluccanus Rosaceae Asia, Africa, North Natural forests, planted Not specified Physical, Edible,
L. Oceania America forests, ruderal/ chemical medicinal
disturbed sites, and and
wetlands biological economic
(dyeing)
Rubus niveus Rosaceae Asia Oceania, North Agricultural areas, Forms large Physical, Edible,
Thunb. America, South natural forests, riparian impenetrable thickets biological medicinal
America, Africa zones, ruderal/ of prickly canes and and
disturbed sites, scrub/ blocking access to cultural economic
shrublands humans and livestock (dyeing)
Rubus ulmifolius Rosaceae Africa, North America, Agricultural areas, Outcompetes native Not Forage
Schott Europe South America, riparian zones, ruderal/ flora for light, water, specified
Oceania disturbed sites nutrients, and space
Sechium edule Cucurbitaceae North North America, Agricultural areas, Not assessed Not Edible
(Jacq.) Sw. America Oceania ruderal/disturbed sites Specified
Solanum Solanaceae North Oceania, Europe Natural forests, Smothers and disrupts Not Not assessed
seaforthianum America, ruderal/disturbed sites native plant specified
Andr. South communities
SM. Sundarapandian et al.

America
Invasive vine Native Continent(s)/area Control Resource
species Family continent(s) invaded Habitat Impacts on ecosystem measures value
12

Syngonium Araceae North Oceania, Asia, Natural forests, planted Forms a dense mat on Physical, Ornamental
podophyllum America, Africa forests, ruderal/ the forest floor and on chemical
Schott South disturbed sites, scrub/ trees and
America shrublands, urban integrated
areas, wetlands control
Thunbergia Acanthaceae Indian Naturalised Closed forests, forest Threatens remnant Physical Ornamental
fragrans Roxb. Sub- overseas in margins, watercourses, vegetation in the wet and
continent south-eastern urban bushland, tropic, degrades creek chemical
and USA, Caribbean, disturbed sites, and river banks
south- Mascarenes and roadsides
eastern several Pacific
Asia islands
Thunbergia Acanthaceae Asia Africa, Oceania, Natural forests, planted Forms large Not Ornamental
grandiflora Roxb. North America forests, ruderal/ impenetrable thickets specified
disturbed sites of prickly canes
blocking access to
humans and livestock
Thunbergia Acanthaceae Native to Australia Watercourses, Alert list for Physical Ornamental
laurifolia Lindl. southern disturbed forests, forest environmental weeds and
China, margins, open in Australia. Climbs chemical
Taiwan and woodlands, roadsides, native vegetation,
south- fence-lines, gardens smothering, shading
Biological Invasion of Vines, Their Impacts and Management

eastern out and killing the


Asia understory. Often pulls
down mature trees
with the weight of the
vine
(continued)
223
Table 12.1 (continued)
224

Invasive vine Native Continent(s)/area Control Resource


species Family continent(s) invaded Habitat Impacts on ecosystem measures value
Wisteria floribunda Fabaceae Asia North America Natural forests, Causes mechanical Physical Ornamental
(Willd.) DC. riparian zones, ruderal/ damage to native trees and
disturbed sites, urban due to its enormous chemical
areas weight
Wisteria sinensis Fabaceae Asia North America Natural forests, Smothers and disrupts Physical Ornamental
(Sims) DC. riparian zones, ruderal/ native plant and
disturbed sites, urban communities chemical
areas
SM. Sundarapandian et al.
12 Biological Invasion of Vines, Their Impacts and Management 225

have migrated across to North America, Africa and Europe. Another 6 % of the
vines which share a common nativity from Africa, Asia and Oceania have invaded
North and South Americas. Just 2 % of the vines have migrated from Europe to
other continents. It is difficult to narrow down the exact origin of many vines
(Hedera helix, Ipomoea cairica, Abrus precatorius, Persicaria perfoliata, Merremia
peltata, Cardiospermum grandiflorum, etc.) as their native range overlaps across
many continents.
According to Zhang and Chen (2011), land area of a country is not responsible
for the number of invasive species. The authors have proposed that higher similarity
of invasive species in different countries in same or different continents could be
due to geographical adjacency, climate and environmental similarity. Moreover,
island countries have been noted to hold more invasive species than non-island
countries. It is evident from Table 12.1 that most of the notorious vines across the
world with different nativities have invaded North America (about 61 %). Around
51 % of the invasive vines have spread to Oceania that include Australia and
other islands of tropical Pacific Ocean. Approximately, 21 %, 19 % and 17 % of the
invasive vines have migrated to Europe, Asia and South America respectively.
Antigonon leptopus, Buddleja madagascariensis, Abrus precatorius, Passiflora
foetida and Merremia tuberosa have naturalized in tropics. More infestation by
invasive vines in a continent could be regarded as an indicator of more habitat
destruction in that continent.

12.3.2 Regional Distribution of Invasive Vines

Vines have invaded more in the United States of America (32 species) followed by
Australia (23 species) (Table 12.1). A biodiversity-rich country like Brazil has nine
invasive vines. India, South Africa and United Kingdom have been infested with
five invasive vines. Anning and Yeboah-Gyan (2007) reported four vines from 43
invasive species enumerated in Ashanti Region of Ghana. Various aspects have been
studied on invasive vines in different parts of the world and region (Table 12.2).
Ecology, distribution of 10, 6.6 and 3 invasive vines have been studied respectively
in USA, Australia, Africa and worldwide respectively (Table 12.2). The medicinal
potential of invasive vines has been unraveled through anti-microbial and anti-
diabetic studies by several workers, especially in USA and Australia. Impacts and
control measures of invasive vines have been extensively studied in USA, Australia,
Africa and worldwide.

12.3.3 Invasive Vine Distribution by Plant Families

It is apparent from Table 12.1 that most of the invasive vines belong to Fabaceae
(~11.5 %). Passifloraceae and Rosaceae are the next dominant families, each con-
tributing to 9.6 %. These are followed by Convolvulaceae and Apocynaceae which
Table 12.2 Aspects studied on invasive vines in various countries
226

Name of the liana Area studied Aspect studied Source


Abrus precatorius , Acetosa sagittata, Araujia sericifera, New South Wales Establishment of exotic vines and Scientific Committee and New
Aristolochia elegans, Asparagus aethiopicus, Asparagus scramblers South Wales Government 2011.
africanus, Asparagus asparagoides, Asparagus (http://www.environment.nsw.
plumosus, Asparagus scandens, Asystasia gangetica, gov.au/determinations/
Caesalpinia decapetala, Cardiospermum grandiflorum, ExoticVinesKtp.htm)
Delairea odorata, Dipogon lignosus, Dioscorea
bulbifera, Hedera helix, Ipomoea alba, Ipomoea cairica,
Ipomoea indica, Ipomoea purpurea, Lathyrus tingitanus,
Lonicera japonica, Macfadyena unguis-cati, Passiflora
suberosa, Passiflora subpeltata, Passiflora toriminiana,
Senecio angulatus, Senecio macroglossus, Solanum
jasminoides, Solanum seaforthianum, Sollya
heterophylla, Thunbergia alata, Thunbergia grandiflora,
Tradescantia fluminensis, Vinca major
Amphicarpa bracteata, Bignonia capreolata, Decumaria Georgia, USA Documentation of non-native Loewenstein and Loewenstein
barbara, Gelsemium sempevirens, Vitis rotundifolia, plants in riparian forests of Georgia 2005
Toxicodendron radicans, Lonicera japonica,
Parthenocissus quinquefolia, Puearia lobata
Anredera cordifolia Australia Ecology, distribution and control Johnson 2011
measures
Buddleja madagascariensis Worldwide Distribution patterns, ecology and Starr et al. 2003a
control measures
Calopogonum mucunoides, Cardiospermum Ashanti region, Ghana Diversity and distribution of Anning and Yeboah-Gyan 2007
grandiflorum, Cardiospermum halicacabum, invasive weeds
Centrosema pubescens, Clematis vitalba
Cardiospermum grandiflorum Queensland, Australia; Invasion history and ecology and Carroll et al. 2005;
Argentina and South bio-control agents Kay et al. 2010
Africa
SM. Sundarapandian et al.
12

Celastrus orbiculatus Connecticut; New Comparison of responses to Leicht-Young et al. 2007; Pisula
Jersey; Michigan; different environmental gradients and Meiners 2010; Michigan
Piedmont; Virginia by invasive vs. native species; Department of Natural
assessment of allelopathic Resources 2012 (http://www.
potential; distribution, impacts and michigan.gov/documents/dnr/
control measures; colonization Oriental_Bittersweet_389123_7.
dynamics; differential responses to pdf); Robertson et al. 1994;
light quality Leicht and Silander 2006;
Harrington et al. 2003
Clematis vitalba New Zealand Influence of seed germination on Bungard et al. 1997; Ogle et al.
distribution; impact of Clematis 2000
vitalba and control measures
Coccinea grandis Tamil Nadu, Kerala Drug for diabetes; antimicrobial Ramachandran and
activity from the leaves Subramaniam 1983; Satheesh
and Murugan 2011
Convolvulusarvensis, Ipomoea hederacea, Ipomoea Korea Ecology and management of Kim 2012
lacunosa, Ipomoea purpurea, Ipomoea triloba, Wisteria floribunda
Jacquemontia taminifolia, Polygonum convolvulus,
Polygonum dumetorum, Quamoclit coccinea, Sicyos
angulatus, Vicia dasycarpa, Wisteria floribunda
Cryptostegia grandiflora Queensland, Australia; Biological control using a rust, Tomley and Evans 2004;
worldwide Maravalia cryptostegiae; impact of Kriticos et al. 2003
Biological Invasion of Vines, Their Impacts and Management

climate change on the potential


distribution and relative abundance
Cryptostegia madagascariensis Koolan Island, Western Biological inventory Keighery et al. 1995
Australia
Delairea odorata Sydney, Australia; Distribution patterns, ecology and Sydney Weeds Committees 2012
Mauna Kea, Hawaii; control measures (http://sydneyweeds.org.au/
worldwide wp-cms/wp-content/uploads/
Weed-Fact-Sheet-Cape-Ivy.pdf);
Banko et al. 2002; Starr et al. 2003b
227

(continued)
Table 12.2 (continued)
228

Name of the liana Area studied Aspect studied Source


Dioscorea bulbifera Florida, USA Chemical control of bulbils Duxbury et al. 2003
Dioscorea oppositifolia Illinois Spatial distribution, growth Thomas et al. 2006
patterns and habitat constraints
Dipogon lignosus Temperate Australia Evaluation of the risk of introduced Emms et al. 2004
legumes to natural ecosystems of
Australia
Epipremnum aureum, Monstera deliciosa, Philodendron Texas, USA Habitat and landscape utilization Arnold 2013
hederaceum
Euonymus fortunei USA Analysis of distribution, ecological Remaley 2005
threat and management options
Hedera helix Michigan Distribution, taxonomic status, Palazzola and Burnham 2013
ethnomedicinal uses (http://climbers.lsa.umich.
edu/?p=172)
Hiptage benghalensis Mascarene Islands; Distribution status and potential of Baret et al. 2006; Tassin et al.
Re’Union Island, invasive, alien plants; presented a 2006; Starr et al. 2003c; Vitelli
Indian Ocean; ranking of invasive woody plant et al. 2009
worldwide; species for management; an
Queensland, Australia overview on distribution patterns,
ecology and control measures
Ipomoea aquatica Worldwide; Oregon Medicinal vegetable; risk Austin 2007; Harwood and
assessment of Ipomoea aquatica Sytsma 2003
Ipomoea cairica Baiyun Mountain, Acceleration of expansion of invasive Wang et al. 2011
Guangzhou, South China species by elevated temperature
Lantana camara Australia, India and Invasion and management strategies; Bhagwat et al. 2012; Baars and
South Africa biological control measures Neser 1999
Lathyrus tingitanus Temperate Australia To evaluate the risk of introduced Emms et al. 2004
legumes to natural ecosystems of
Australia
SM. Sundarapandian et al.
12

Lonicera japonica South Carolina; Comparison of the effects of Schierenbeck et al. 1994;
Charlotte, USA; USA; herbivory on growth and biomass Barden and Matthews 1980;
New Jersey, USA; allocation of native and introduced Leatherman 1955; Pisula and
worldwide except Asia species of Lonicera; changes in Meiners 2010; Schierenbeck
abundance after prescribed burning 2004
in a Piedmont pine forest;
ecological life history of Lonicera
japonica; assessment of allelopathic
potential; ecology, distribution, life
history and control measures
Lygodium japonicum Florida, USA Control measures Bohn et al. 2011
Lygodium microphyllum and Lygodium japonicum Florida, USA Reproductive biology Lott et al. 2003
Macfadyena unguis-cati Central Queensland, Invasive liana to simulated Raghu et al. 2006
Australia herbivory: implications for its
biological control; biological
control using simulated herbivory
Merremia peltata Pacific region Biological control Paynter et al. 2006
Mimosa diplotricha Southwestern Ethiopia Status of invasion and control Wakjira 2011
measures
Paederia foetida Florida, USA Control status, distribution and Pemberton and Pratt 2002
biological control measures
Biological Invasion of Vines, Their Impacts and Management

Paederia scandens China Influence of disturbance on Ye et al. 2013


spreading of Paederia scandens
Passiflora edulis Uganda Control measures Ismail 2006
Persicaria perfoliata Pennsylvania Integration of management Lake et al. (2011; 2014)
techniques for control
Pueraria montana Philadelphia, USA Biological control using a Ding et al. 2006
Chrysomelid beetle, Gonioctena
tredecimmaculata
(continued)
229
Table 12.2 (continued)
230

Name of the liana Area studied Aspect studied Source


Rubus ellipticus Hawaii Impact and spreading of Rubus Stratton 1996
ellipticus
Thunbergia grandiflora and Thunbergia laurifolia Queensland, Australia Declaration as pest plants, habitat, Department of Agriculture,
distribution and control measures Fisheries and Forestry 2013
Vincetoxicum nigrum Europe Identification of biological control Weed et al. 2010
agents
Vincetoxicum rossicum Toronto; Ontario Invasion, distribution and taxonomic Kricsfalusy and Miller 2010;
status; management practices; decline Anderson 2012; Ernst and
in pollinators and arthropods due to Cappuccino 2005; Averill et al.
the presence of Vincetoxicum 2010; Weed et al. 2010
rossicum; establishment of
Vincetoxicum; identification of
biological control agents
SM. Sundarapandian et al.
12 Biological Invasion of Vines, Their Impacts and Management 231

constitute 7.6 % and 5.7 % respectively. The other families constitute only a
smaller percentage. The legume family has always been regarded as a major con-
tributor to the notorious, naturalized flora of the world (Binggeli 1996; Pysek
1998; Wu et al. 2003). The successful establishment of Fabaceae may be due to
root nodulation and floral characteristics such as the quite common buzz pollina-
tion, in addition to explosive fruit dehiscence. Most of the vines of Passifloraceae
are tendrillar, which facilitate climbing onto the host plant. Moreover, floral mim-
icry is also quite prevalent in this family, paving way for wide dispersal. The
Rosaceae (e.g. Rubus discolor, Rubus ellipticus, Rubus moluccanus, Rubusniveus
and Rubus ulmifolius) are characterized by large, showy flowers that attract polli-
nators which facilitate pollination even in newly introduced areas. The floral mor-
phology of Convolvulaceae (e.g. Ipomoea aquatica, Ipomoea cairica, Merremia
peltata and Merremia tuberosa) attracts pollinators as well as humans and hence
are spread naturally as well as introduced ornamentally. Many members of
Apocynaceae (Cryptostegia grandiflora, Cryptostegia madagascariensis and
Cynanchum rossicum) are introduced artificially for ornamental and edible pur-
poses (de Araujo et al. 2014).

12.4 World’s Worst Top Ten Invasive Vines

A total of 55 invasive vines were distributed across the globe. Their impacts are
varied with the physiographic features of the places, where invaded. The level of
native diversity also alters invasiveness. However, several species are well-
established in the introduced habitat and affect the environment, of which ten spe-
cies are dealt here.

12.4.1 Celastrus orbiculatus Thunb. (Celastraceae)

The oriental bittersweet (Celastrus orbiculatus) introduced as an ornamental from


eastern Asia in 1860 is a deciduous woody vine (vine) that had become naturalized
in the northeastern USA (Sorrie and Somers 1999). Bittersweet produces attractive,
colourful fruits that are readily dispersed by birds. This exotic vine is seen as a pest
by many land managers as it readily overtops native flora along roadsides and forest
gaps (Dreyer et al. 1987; McNab and Meeker 1987). This plant is reported to have
grown with an average length of 4.5 m in 12 weeks under partial shade, whereas,
certain individuals grew up to 9 m (Ellsworth 2003). Germination of bittersweet
seeds occurs in various light intensities (Patterson 1974). Patterson (1975) found
that shade-grown plants adapted their photosynthetic ability as that of sun-grown
plants within 8 days. This ability helps the vine to rapidly establish in perturbed
areas. Luxuriant growth over stumps and saplings leads to problems in silviculture
by arresting natural regeneration and planted stocks (McNab and Meeker 1987). In
232 SM. Sundarapandian et al.

addition, as the vine climbs on from one tree to another, there is a high risk of mul-
tiple tree fall due to logging or wind throw (Putz 1991).

12.4.2 Cynanchum rossicum (Kleopov) Barbarich


(Apocynaceae)

The Dog-strangler viz., Cynanchum rossicum is a perennial twiner that was intro-
duced 100 years ago from Europe, which has then become naturalized in northeast-
ern United States and southeastern Canada. This species usually colonizes natural
uplands, woodlands, pastures, fallow land and other wastelands and vine is cur-
rently expanding in natural and semi-natural forests at an alarming rate (DiTommaso
et al. 2005). They colonize densely, amassing existing vegetation, besides reducing
faunal diversity. Reproduction via polyembryony and wind-dispersed seeds enhance
further invasion. Use of herbicides is in practice to control this weed, because
mechanical removal is cumbersome and there are no known natural enemies
(DiTommaso et al. 2005).

12.4.3 Euonymus fortunei (Turcs.) Hand.-Mazz. (Celastraceae)

Euonymus fortunei, also known as winter creeper, a native of Asia is an evergreen


vine that has spread to Africa, Europe, North America and South America. It has a
great potential to dominate the herbaceous layer and forms large patches that can
cover up to several acres. Once the ground cover is dominated, they ascend on trees,
where they blossom and produce seeds. Their fruits and seeds are often dispersed by
floods and birds. Although they prefer moist habitats, they also grow in shady as
well as sunny areas. When the area of infestation is small, they are removed mechan-
ically. Chemical control by foliar application of 2 % Tryclopyr + surfactant have
also proven to be successful (Anonymous 2006).

12.4.4 Lonicera japonica Thunb. (Caprifoliaceae)

Lonicera japonica, also called as Japanese honeysuckle, a native of East Asia is a


fast-growing woody vine that can form a thick ground cover and also climb trees.
It has invaded most of North America, South America and Europe. They thrive pro-
fusely in sunny areas, although they can grow in shade too and their plentiful seeds
produced are readily dispersed by birds. In small or sensitive areas, this vine is
removed mechanically by digging and for large populations, foliar application of
glyphosate is the most sought way of control (Anonymous 2006).
12 Biological Invasion of Vines, Their Impacts and Management 233

12.4.5 Pueraria montana (Willd.) Maesen & S. Almeida


(Fabaceae)

Pueraria montana, commonly called Kudzu and a native of China and Japan, is a
woody vine that exhibits vigorous growth in a wide variety of conditions in most
soil types. Their most preferred habitats are ruderal sites with bright sunlight such
as roadsides, abandoned fields and forest edges. This vine suppresses native trees by
covering them under a thick blanket of their leaves. Once this vine gets firmly estab-
lished, it grows rapidly at an average rate of one foot per day as much as about 60 ft
per season. It has been suggested that for successful and long-term control of this
vine, their extensive root system must be destroyed. Mowing and cutting followed
by herbicide application is the most recommended control method (www.wvnps.
org/FightingInvasives.pdf).

12.4.6 Persicaria perfoliata (L.) H. Gross (Polygonaceae)

Persicaria perfoliata, commonly referred to as devil’s tail, a native of Asia is an


annual trailer that generally colonizes disturbed sites, open areas, wetlands, forest
edges and roadsides. Although bright sunlight and moist soils are most conducive
for the establishment of this vine, it can also survive in partial shade and dry soils.
In open areas, they form dense mats smothering herbs and young tree seedlings,
thereby preventing forest regeneration. Plant diversity has been found to be greatly
reduced in those areas where this vine is prominent and wildlife has also been
affected due to diminished food and habitat sources. As the roots of this vine are
weak, hand pulling is done for smaller areas. Repeated mowing is also found to be
effective. For larger areas, triclopyr or glyphosate are often used as chemical control
agents. Biologically, a weevil, Rhinoncomimus latipes has been found to be host-
specific to this vine. While adult weevils feed on the foliage, young larvae feed
within the nodes and suppress growth. (www.wvnps.org/FightingInvasives.pdf)

12.4.7 Dioscorea bulbifera L. (Dioscoreaceae)

Dioscorea bulbifera, commonly known as air potato vine, a native of Asia and
Africa can reach even up to 65 ft infesting shrubs and trees. They spread vigorously
by dangling potato-like tubers at leaf axils as well as by underground tubers. They
spread rapidly in open to semi-shady sites. Although the vines die during winter,
they persist and spread by underground tubers which give rise to numerous aerial
shoots with the onset of spring (Miller 2003). They are often controlled either physi-
cally by digging or biologically by the release of the beetle Lilioceris cheni.
234 SM. Sundarapandian et al.

12.4.8 Hedera helix L. (Araliaceae)

Hedera helix, also known as English ivy, a native of Eurasia is an evergreen woody
vine that can reach up to 90 ft by their clinging aerial roots and can also form a
dense ground cover by trailing. They grow profusely in a wide range moisture and
soil conditions, but thrive best in moist open forests. They build up on trees, decreas-
ing their vigour and increasing the chance of wind throw. Also, they act as favour-
able sites of reproduction for bacterial leaf scorch that infects oaks (Quercus spp.),
elms (Ulmus spp.), and maples (Acer spp.) (Miller 2003). Often an integrated
approach involving hand pulling and herbicide treatment are used to control the
spread of this vine (Waggy 2010).

12.4.9 Wisteria sinensis (Sims) DC. and Wisteria floribunda


(Willd.) DC. (Fabaceae)

Wisteria sinensis and W. floribunda, also called as Chinese and Japanese wisterias
respectively are deciduous, high-climbing woody vines. They colonize in both
wet and dry sites and they root at nodes. They form dense infestations wherever
they are planted (Miller 2003) and they tend to establish and spread along forest
edges, disturbed areas and riparian zones. When they occur in riparian zones, the
seeds are water-dispersed. Also, canopy gap formation which occurs when the
wisteria topples a large tree, favours the growth of wisteria seedlings. Cutting,
digging and herbicide application are the most widely followed control measures
(Stone 2009).

12.4.10 Dolichandra unguis-cati (L.) L. G.


Lohmann(Bignoniaceae)

Dolichandra unguis-cati, commonly known as cat’s claw vine, and a native of North
America and South America, has spread to Asia, Oceania and Caribbean Islands.
Although, they grow relatively small, they are more persistent with tuberous roots.
They cling firmly onto the host plants with clawed tendrils. This liana thrives well
in bright sunlight as well as in partial shade. It produces winged seeds, which are
easily dispersed by wind. This obnoxious vine can form a dense ground cover as
well as smother native trees. This vine is controlled by physical, chemical and cul-
tural control measures (Johnson 2011).
12 Biological Invasion of Vines, Their Impacts and Management 235

12.5 Detailed Review Analysis on Species Biology


and Invasion Ecology of Two Selected Invasive Vines

12.5.1 Mikania micrantha (L.) Kunth.(Asteraceae)

Mikania micrantha, commonly called as mile-a-minute weed for its rapid growth
during spring and summer is a perennial vine that has been listed as one of the
world’s top ten worst weeds (Cronk and Fuller 1995) and one of the ten most inva-
sive plants in the world (Lowe et al. 2001). M. micrantha has also been listed as one
of the most noxious invasive species across the globe (Holm et al. 1977; Waterhouse
2003). A native of Central and South America, this vine was first introduced in
Taiwan for soil conservation in the 1970s, but it is now being complained of as an
aggressive invader between the latitudes 22–24 °N and longitudes 111–117 °W, in
the altitudinal range of 50–200 m (Zhang et al. 2004; Tripathi et al. 2012). In China,
commonly called as ‘the plant killer’ for its capacity to smother other species, this
neotropical vine was reported to have first invaded Hong Kong in 1910, Yunnan
province in 1950s and the coast of Guangdong province in the 1980s via sea route
(Xie et al. 2000; Wang et al. 2001; Ye and Zhou 2001; Zhang et al. 2004; Du et al.
2006; Li et al. 2006; Tripathi et al. 2012; Shen et al. 2013) and has since then spread
to several tropical countries. In southern China, this vine has invaded a wide range
of habitats, leading to huge economic and environmental losses (Li et al. 2000,
2002, 2006; Wang et al. 2004; Zhang et al. 2004). Zhang et al. (2004) has stated that
M. micrantha invasion in China has accelerated in the last decade due to policy
reforms and economic development. This vine, originally introduced as a ground
cover in Indonesia has then expanded its boundaries to Pacific, Southeast Asia and
New Guinea (Waterhouse 1994). Now, this weedy vine has been reported from
Mauritius, India, Sri Lanka, Bangladesh, Southeast Asia and the Pacific (Deng et al.
2004), infesting agricultural lands and plantations (Choudhury 1972; Parker 1972;
Holm et al. 1977). M. micrantha is considered as a serious pest in plantation
crops and commercial forests, from Mauritius to West Africa and also across Asia
(Hills 1999; Zhang et al. 2004). Tiwari et al. (2005) have nominated the vine as one
of the six of “high risk posed” weed in Nepal, where Sapkota (2007) has also
regarded it as a threat to coconut, oil palm, banana and cacao plantations. Moreover,
as stated by Yu and Yang (2011), the seeds of this weed have a tendency to germi-
nate in the bare soils of coastal areas causing extensive damage to mangroves. This
invasive vine was also reported from Queensland, Australia as a major pest with a
great potential of invading the wet tropics and other humid regions of Northern
Australia. However, it is seldom considered as a weed in its native range (Central
and South America) as its natural enemies exert a significant pressure and restrict
the occurrence and abundance of this vine (www.cabi.org; Ram 2008). This aggres-
sive vine is now cited as an alien invasive species in 31 countries or states in the
world, viz., American Samoa, Australia, Bangladesh, China, Cook Islands, Fiji, French
Polynesia, Guam, India, Indonesia, Malaysia, Mauritius, Micronesia, Federated
States of Micronesia (FSM), Nepal, New Caledonia, Niue, Northern Mariana
236 SM. Sundarapandian et al.

Islands, Palau, Papua New Guinea, Philippines, Samoa, Solomon Islands, Sri Lanka,
Thailand, Tokelau, Tonga, Tuvalu, Vanuatu, Wallis, Futuna (Global Invasive Species
Database 2013; Ram 2008).
M. micrantha was introduced in India during second world war to as a way of
concealing air fields and as a ground cover for tea plantations (Tripathi et al. 2012).
It has then spread to moist tropics, subtropics and north-eastern parts of India. It is
now reported to be widely prevalent in the states of Assam, West Bengal, Orissa,
Union territory of Andaman and Nicobar Islands and in Tamil Nadu (Swamy and
Ramakrishnan 1987a). However of late, the infestation by this vine has been
reported from even southernmost regions of India like Cauvery delta and beyond.
This noxious vine possesses a vigorous vegetative and sexual reproductive capacity
(Swamy and Ramakrishnan 1987a) as well as phenotypic plasticity (Prabu et al.
2014), but cannot tolerate dense shade (Holm et al. 1977). M. micrantha spreads
profusely by rooting at nodes, ramets, perennating rosettes and by efficient mode of
seed dispersal by wind, water and animals especially from October to April (Ismail
and Mah 1993; Swamy and Ramakrishnan 1987b). This notorious vine is known to
colonize and grow vigorously in forests, riverbanks, disturbed sites and roadsides
and once they firmly establish, they smother the native flora, reducing the availabil-
ity of light beneath its canopy (Huang et al. 2000; Tripathi et al. 2012).
M. micrantha also secretes allelochemicals such as dihydromikanolide and
deoxymikanolide which inhibit seed germination and seedling growth (Shao et al.
2005). They also have profound effects on soil microbial community and litter
decomposition of native plants (Ni et al. 2006; Chen et al. 2007). Mechanical
removal, application of herbicides and biological control by Cuscuta campestris
and Puccinia spegazzinii greatly suppress the dense growth of this vine (Tripathi
et al. 2012). Despite the negative impacts, the amassing of M. micrantha need not
be much despaired as according to Chen et al. (2009), as their presence has been
known to enrich the soil by increasing the availability of carbon, nitrogen, phos-
phate and ammonia. Utilization of this vine as a medicinal plant would help in
drastic reduction of the existing populations of this notorious vine as well as limit
the expanse of further invasion.
In India, Swamy and Ramakrishnan (1987c, d) studied the influential role of
Mikania micrantha in the forests of north-east India during secondary succession
following slash and burn agriculture, which is a predominant land use system.
Despite posing problems due to invasion, this weed is known to have a conservation
role in restoring a disturbed ecosystem (Ramakrishnan 1984). M. micrantha is
known to have significantly contributed to litter production that increases with fal-
low age up to four years because the entire aboveground biomass of M. micrantha
produced during the growing season (March–November) is converted to litter due
to the death of all the aboveground parts from December to January. The drastic
decline after 4 years is due to its suppression by fast-growing shrubs and trees
(Swamy 1986). This vine also plays a crucial role in nutrient cycling during these
early phases of succession as it conserves substantial level of potassium in its bio-
mass. Therefore, M. micrantha which dominates the fallows for only 4 years could
restore a disturbed ecosystem only under long slash and burn cycles.
12 Biological Invasion of Vines, Their Impacts and Management 237

Swamy and Ramakrishnan (1987a, 1988a) have analysed the reproductive


potential, growth and allocation patterns of M. micrantha in successional communi-
ties. The weed exhibited a peak vigour up to July in younger fallows up to 3-years,
followed by a decline in older fallows and the reproduction via ramets outnumbered
that from seeds. The allocation of biomass and nutrients in M. micrantha follows a
ruderal strategy as in young fallows, the resources are mainly allocated to ramet root
system and vegetatively reproducing stolons and seeds, while in older fallows, more
allocation is towards the rosette root system (a perennating organ). Swamy and
Ramakrishnan (1987b, 1988b) have also extensively studied the effects of fire on
population dynamics, growth and allocation strategies of M. micrantha. Fire is
found to enhance the population size of M. micrantha in 2- to 8-year old fallows and
the sexually reproductive ramets were higher in burnt than in unburnt fallows.
However, reduced regeneration via perennating rosettes in the post-burn phase
implies damage to rosette caudices by fire. The plant vigour was found to be more
stable in burnt plots, but declined in unburnt plots. The nutrient uptake was also
found to increase with age in burnt than unburnt fallows. This could be due to
differences in nutrient availability in the soils of burnt and unburnt plots. The ruderal
nature of M. micrantha is also reflected in its better blooming at a low nutrient cost
in burnt than in unburnt fallows. Further it is concluded that M. micrantha is well-
adapted to survive after fire and the fire also boosts its chances of successful estab-
lishment during early succession.

12.5.2 Lantana camara L. (Verbenaceae)

Lantana camara is a noxious invasive vine and has spread most parts of the world.
It has a great genetic diversity (650 varieties) and wide geographic expansion (60
countries) and great ecological tolerance (http://www.issg.org/database/species/
ecology.asp?si=56). The Invasive Species Specialist Group (ISSG) have regarded
this pantropical vine as one among 100 of the “world’s worst” invaders as it affects
ecosystem and causes biodiversity loss to a greater extent. The Central and South
America and the Caribbean with geographical expansion between 35 °N and 35 °S
(Day et al. 2003) is its native range. It was originally a hybrid species developed in
sixteenth century at Europe for ornamental purposes. Since then, it has become
prevalent in different forms, colours and varieties throughout the world (Parsons
and Cuthbertson 1992). Flower colour variation enlightens the genetic diversity of
L. camara and has been proved by Scott (1998) through RAPD analysis.
L. camara is one of the major invasive species in Indian forests (Murali and Setty
2001; Hiremath and Sundaram 2005; Sharma and Raghubanshi 2006; Prasad et al.
2006; Sahu and Singh 2008; Love et al. 2009). It was introduced by the British in
Calcutta Botanical Garden in 1809 as an ornamental plant (Brandis 1882; Aravind
and Rao 2001; Nanjappa et al. 2005). After two centuries of colonization and
establishment, this aggressive weed has spread throughout India in farms, pastures,
fallow lands, roadside, railway tracks, canals and forests except the Thar Desert
238 SM. Sundarapandian et al.

(Kannan et al. 2008, 2009; Aravind et al. 2010; Kimothi and Dasari 2010; Surampalli
2010; Dobhal et al. 2011; Sharma et al. 2005; Kohli et al. 2006; Dogra et al. 2010).
The species eliminates native flora, by forming dense thickets (GISIN 2011). Global
Invasive Species Information Network now categorizes Lantana as one among the
top ten invasive species in the world (GISIN 2011). As of now, Lantana has invaded
more than 13 million ha in India, 5 million ha in Australia, and 2 million ha in South
Africa (Wells and Stirton 1988; Simelane 2005; Sharma and Raghubanshi 2011).
A large-scale (55 ha) liana diversity inventory (≥1.5 cm dbh) was carried out on
six major hill complex of Indian Eastern Ghats yielded a total of 32,033 liana
individuals (Muthumperumal and Parthasarathy 2010) of which native lianas
formed 76 % (24,319 individuals) and 24.08 % (7,714 individuals) were invasive
lianas: Lantana camara alone contributed 24.04 % (7,703 stems). The seeds of
Lantana are dispersed by frugivorous birds and several wild and domestic animals.
Furthermore, many species have ecological amplitude to grow in a wide range of
forest types, from tropical thorn to tropical wet evergreen, and Lantana is an aggres-
sive colonizer practically at forest edges and disturbed and denuded areas (Singh
et al. 2006). There was a strong negative non-linear relationship between native
species richness and Lantana cover in south-eastern Australia (Gooden et al. 2009).
Lantana camara seeds are dispersed through birds and mammals (Khoshoo and
Mahal 1967). In addition, they are also propagated vegetatively by layering and
reshooting. Wijayabandara et al. (2013) found that seeds are viable from 2–5 years.
In dry deciduous forests, Lantana alters the spatial pattern of herbaceous vegetation
as well as the microhabitat (Sharma and Raghubanshi 2010).
Lantana cover increases water run-off, thereby leading to soil erosion, although
some studies have also proven that it prevents soil erosion on barren areas and
mountain slopes (Greathead 1968; Ghisalberti 2000). Simba et al. (2013) studied the
impact on soil properties in Nairobi National Park, Kenya. The results indicated
that significant differences in soil pH, concentrations of magnesium, calcium and
potassium between the invaded and non-invaded sites and their levels were higher
in invaded sites. More over magnesium, calcium and potassium varied significantly
across the forest, riverine and shrub-grassland. These results suggest that L. camara
can improve the nutrient levels of soil and therefore influence nutrient cycling
resulting to making the ground better for its growth and this might explain the
capacity of the invasive species to outcompete the native ones. Jain et al. (1989)
reported that Lantana inhibits the germination and growth of other plants by the
presence of 14 phenolic compounds. Lantana leaves, stems and fruits contain toxic
compounds, namely pentacyclic triterpenes (Kellerman et al. 1996), lantadene A
and B (Morton 1994). A series of aromatic alkaloids and phenolic substances
can be identified from the extracts of Lantana Khan et al. (2003). Allelochemical
concentration in root, stem and leaf (Chaudhary and Bhansali 2002), makes it toxic
to grazers (Ambika et al. 2003).
Though Lantana camara is noxious, it has some potential medicinal properties.
Some of research work have conducted on chemical constituents in different parts
of this plants: Leaves (Singh et al. 1996; Noble et al. 1998; Nagao et al. 2002);
Whole Plant (Ghisalberti 2000; Day et al. 2003); Flower (Mohan Ram and Mathur
12 Biological Invasion of Vines, Their Impacts and Management 239

1984; Day et al. 2003); Bark (Priyanka and Joshi 2013) etc. L. camara has several
therapeutic uses such as anti-inflammatory, anti-tumour, and anti-AIDS, antimi-
crobial, fungicidal, insecticidal, nematicidal, biocidal activities (Sharma et al.
1988, 1999, 2007; Sharma and Sharma 1989; Begum et al. 2000; Saxena 2000).
Lantana oil is used externally for leprosy and scabies (Ghisalberti 2000). Plant
extracts are used as medicine for the treatment of cancers, chicken pox, measles,
asthma, ulcers, swellings, eczema, tumors, high blood pressure, bilious fevers,
catarrhal infections, tetanus, rheumatism, malaria and atoxy of abdominal viscera
(Begum et al. 2000). Lantana is now a naturalized plant in India that plays a vital
role in ecosystem functions viz., by providing cover to carnivores, food for birds
and wild herbivores as well as benefitting the local communities (Soni et al. 2006).
Some of domestic/economic utilities have been developed such as paper pulp
which is used for wrapping, writing and printing paper (Ray and Puri 2006;
Naithani and Pande 2009); baskets and temporary shelters (Kannan et al. 2008)
and also biofuels (Sharma et al. 1988). The Male Madeshwara Hills of Karnataka,
India launched a project to control Lantana camara by effectively using the stems
to make furniture which is cheaper than cane, but equally sturdy (Aravind et al.
2006; Kannan et al. 2008).
Mechanical and chemical control of Lantana was reviewed by Cilliers and Neser
(1991) who concluded that, even it is very effective, these methods are often labour-
intensive and expensive. Using fires is not always a suitable option as infestations
are often closer or inside the indigenous forests, grazing lands and plantations. The
biological control programme, including previously introduced natural enemies on
L. camara in South Africa, was reviewed by Cilliers and Neser (1991) suggested a
possible shift in the selection of new natural enemies, in which stem borers, root
feeders and flower feeders should take precedence over leaf feeders. The abundance
of potential agents suggests that, despite the problems associated with the program,
L. camara remains a candidate for biological control in South Africa (Baars and
Neser 1999). Lantana camara have many negative impacts including to interrupt
the succession cycle, displace the native biota resulting in decreased biodiversity
(Murali and Setty 2001; Sharma and Raghubanshi 2010). The allelopathic effect
reducing the species richness, because of its drastic increment of density in wild
forest (Day et al. 2003).There are several controlling measures have been practiced.
Mechanical clearing is suitable for small areas, while fire is preferred in larger areas.
Chemical control is the most effective method, while little success has been obtained
through bio-control. Therefore planners and managers has developed a new strate-
gies, aimed to utilize the species viz., craft making, herbal medicine and biofuel
through involvement of local community.
Priyanka et al. (2013) suggested that sustainable management of invasive species
is challenging but inevitable given the increasing range of alteration caused by inva-
sion which has little prospect of irreversibility. To be sustainable, Invasive Species
Management Framework (ISMF) strategies must include environmental, social,
economic and political factors that influence the causes, impacts, and control of invasive
species across spatio-temporal scales. Although these elemental management
240 SM. Sundarapandian et al.

strategies are easy to document and comprehend but their implementation is often
limited by insufficient control measures, funds, research, socio-economic pressures
and political constraints. Based on ISMF strategies, if incorporated into a manage-
ment plan, will lead to effective management through increased coordination, com-
munication, transparency, accountability and help avert potential risks posed by
accidental and/or intentional introduction of L. camara. Due to its wide distribution
and significant impact on biodiversity, forestry and agriculture, Lantana has been
recognized to be an important weed worldwide. Impacts of Lantana invasion are
minimized to develop a strategy through best-practice guidelines and increasing in
community awareness (Sharma et al. 2005).

12.6 Resource Values of Invasive Vines

Several non-native plants can provide various ecosystem services such as food,
medicine, shelter, and aesthetic values as well (Ewel et al. 1999; Mack 2001). These
apart, native plants are also known to be benefitted from the invasive plants as
the introduction of the latter is known to attract more pollinators, which facilitate
the pollination of native plants too (Bartomeus et al. 2008). Most of the alien plants
have been introduced out of their native ranges for some captivating resource value
such as ornamental, edible or ethnomedicinal characteristics. Pysek et al. (2009)
stated that the planting intensity, residence time and species traits determine the
invasion success of alien, woody species. The probability of invasion success
increases with residence time (Rejmanek 2000) in the area introduced. Of the
52 invasive vines, 36 % were introduced for ornamental, 25 % for edible, 21 % for
medicinal values (modern as well as folk medicine) and 10 % for economic
purposes (Table 12.2). The threats posed by the introduced invasive vines are often
realized, only after the colonization had become too rampant.

12.7 Impact of Invasive Vines in Different Ecosystems

Invasive plant species pose a major threat to local biodiversity, habitat quality,
and ecosystem processes. Invasive vines have economic impacts through loss of
production, utilization of resources and expensive implementation of control mea-
sures (Harrington et al. 2003). Many invasive vines are known to act as disturbing
agents, particularly in soil-disturbed areas (Mack and D’Antonio 1998). According
to Charles and Dukes (2007), invasive vines alter the community structure and
influence natural cycles, which in turn, affect ecosystem services directly and indi-
rectly. Direct effects include decline in the abundance of economically important
species. Moreover, some invasive vines disturb the mutualistic relationship between
the native flora and fauna, affecting pollination and pest control services. On the
12 Biological Invasion of Vines, Their Impacts and Management 241

other hand, invasive vines also decrease an ecosystem’s ability to resist and with-
stand changes, thereby making them more vulnerable to further invasion (Hooper
et al. 2005; Simberloff and Von Holle 1999). Therefore, it can be precisely stated
that rapid spreading of invasive vines drives loss of genetic diversity and species
extinctions (Ellstrand and Schierenbeck 2000; McGeoch et al. 2010).

12.8 Successive Controlling Measures of Invasive Vines

Eradicating invasive vines and vines can be a herculean task and there is no standard
control measure that would be universally most successful for all the invasive vines.
Principally, control measures could be classified into physical, chemical, biological
or cultural control measures (Table 12.1). In some cases, a single approach might
prove successful, but in many cases, an integrated approach is chosen for the best
outcome. It is also important to ensure that the control method followed shouldn’t
exert adverse impact on the ecosystem and affect the native biodiversity. Control
measures is challenging task which requires huge economy investment that unable
to meet out by most developing countries.

12.8.1 Physical Control

Physical control measures include activities such as hand-pulling, smothering,


flooding, cutting, mowing, prescribed burning, etc. These techniques are most
appropriate when the area to be cleared of invasive vines is small and also during
early phases of invasion. However, if the area is extensive or if the invasion has been
widespread, then physical control measures may be labour-intensive and might not
be economically feasible. In such circumstances, an alternative control measure has
to be chosen (Anonymous 2006).

12.8.2 Chemical Control

Chemical control refers to the use of pesticides/herbicides to get rid of invasive


vines. Chemical control could be achieved by foliar sprays or basal bark application
or cut-stem treatment. While choosing an herbicide it is important to take into
consideration the following: target population to be controlled, stage of invasion,
proximity of the area to water resources, soil characteristics and other environmen-
tal conditions. Although this method is notably successful, it is not so eco-friendly
(http://dnr.wi.gov/topic/Invasives/control.html).
242 SM. Sundarapandian et al.

12.8.3 Biological Control

Biological control refers to the use of animals, fungi, etc. to induce diseases and
thereby kill/suppress the invasion of invasive vines. Biological control measures do
not abruptly eliminate the invasive species and it may take years to achieve the
desired outcome. It is important to make sure that the chosen biological control
agent attacks only the invasive plants and not the native ones. Also, to introduce a
potential biological control agent in an area, it is necessary to obtain permits from
Central and State Governments. Understanding the effects of different types of
herbivory on invasive plants is essential to plan biological control programs (Inouye
1982; Maron 1998; Blossey and Hunt-Joshi 2003; Hunt-Joshi et al. 2004; Weed and
Casagrande 2010; Weed et al. 2010). The selection of biological control agents
involves not only recognizing host-specific herbivores, but also to identify herbi-
vores that reduces the population of the targeted weed (Sheppard 2003; van Klinken
and Raghu 2006). For example, air potato vine (Dioscorea bulbifera) has been bio-
logically controlled in Florida by the United States Department of Agriculture by
releasing an Asian leaf beetle (Lilioceris cheni). Cat’s claw creeper Dolichandra
unguis-cati is a woody vine that is controlled in Australia by the release of a leaf-
mining jewel beetle, Hylaeogena jureceki (Dhileepan et al. 2013). The rust fungus
Maravalia cryptostegiae is known to profoundly control the highly invasive rubber
vine, Cryptostegia grandiflora in Queensland, Australia (Tomley and Evans 2004).

12.8.4 Cultural Control

Cultural control involves the manipulation of the forest structure and composition to
control the invasive species or the alteration of the nature of the stand so that the
effects of the invasion would be minimal even if it occurs. Cultural control includes
techniques such as mulching, prescribed burning, soil solarization, canopy closure
(to obstruct the growth of shade-intolerant species), etc. (http://dnr.wi.gov/topic/
Invasives/control.html).

12.9 Future Prospects of Invasive Vines

According to McGeoch et al. (2010), the Red List Index has shown that invasive
alien species impel decline in species diversity and although the policy response trend
had proven to have had positive impact in the recent past, only half of the countries,
that are signatory to Convention on Biological Diversity have adopted national
legislations relevant to invasive alien species. The authors have also opined that
although the pressure posed by invasive alien species has triggered a positive, policy
response, the implementation with the motive of reducing adverse impact on
12 Biological Invasion of Vines, Their Impacts and Management 243

biodiversity has so far been inadequate. The green industry and other NGOs could
produce both state-wise and regional lists on problematic plants. Moreover, land-
scape managers and horticulturists should take special care and adopt new guide-
lines, while designing landscapes. This could be done by educating and spreading
awareness to the general public. If desired, some of these popular invasive vines
could be replaced with sterile hybrids or other similar, attractive plant (Harrington
et al. 2003). Also, new innovative ways by which invasive vines could be put to use
economically and on a large scale should be developed, so that the existing large
population of the invasive vines could be minimized.

References

Ambika SR, Poornima S, Palaniraj R, Sati SC, Narwal SS (2003) Allelopathic plants: 10. Lantana
camara L. Allelopath J 12:147–162
Anderson H (2012) Invasive dog-strangling vine (Cynanchum rossicum) best management
practices in Ontario. Ontario Invasive Plant Council, Peterborough
Anning AK, Yeboah-Gyan K (2007) Diversity and distribution of invasive weeds in Ashanti
Region, Ghana. Afr J Ecol 45:355–360
Anonymous (2006) Identifying and controlling invasive plants of the forest. Northern Kentucky
Urban and Community Forestry Council. Center for Applied Ecology, Northern Kentucky
University, USA. Online at: http://www.nkyurbanforestry.org/SitePages/download_files/
documents/Invasive_Plant_Control.pdf
Aravind NA, Rao D (2001) Biodiversity: an introduction. In: Hosetti BB, Venkateshwaralu M
(eds) Current trends in wildlife biodiversity, conservation and management, vol II. Daya
Publishing House, New Delhi, pp 1–26
Aravind NA, Rao D, Vanaraj G, Ganeshaiah KN, Shaanker RU, Poulsen JG (2006) Impact of
Lantana camara on plant communities at Male Mahadeshwara reserve forest, South India. In:
Rai LC, Gaur JP (eds) Invasive alien species and biodiversity in India. Banaras Hindu
University, Banaras, pp 68–154
Aravind NA, Rao D, Ganeshaiah KN, Uma Shaanker R, Poulsen JG (2010) Impact of Lantana
camaraon bird assemblage at Male Mahadeshwara reserve forest, South India. Trop Ecol
51:325–338
Arnold MA (2013) Landscape plants for Texas and environs, 4th edn. http://aggie-horticulture.
tamu.edu/syllabi/308/Lists/Fourth%20Edition/Epipremnumaureum.pdf
Austin DF (2007) Water spinach (Ipomoea aquatica, Convolvulaceae): a food gone wild. Ethnobot
Res Appl 5:123–146
Averill KM, DiTommaso A, Mohler CM, Milbrath LR (2010) Establishment of the invasive
perennial Vincetoxicum rossicum across a disturbance gradient in New York State, USA. Plant
Ecol 211:65–77
Baars J-R, Neser S (1999) Past and present initiatives on the biological control of Lantana camara
(Verbenaceae) in South Africa. Afr Entomol Mem 1:21–33
Banko PC, Oboyski PT, Slotterback JW, Dougill SJ, Goltz DM, Johnson L, Laut ME, Murray TC
(2002) Availability of food resources, distribution of invasive species, and conservation of a
Hawaiian bird along a gradient of elevation. J Biogeogr 29:789–808
Barden LS, Matthews JF (1980) Change in abundance of honeysuckle (Lonicera japonica) and
other ground flora after prescribed burning of a piedmont pine forest. Castanea
45:257–260
Baret S, Rouget M, Richardson DM, Lavergne C, Egoh B, Dupont J, Strasberg D (2006) Current
distribution and potential extent of the most invasive alien plant species on La Reunion (Indian
Ocean, Mascarene islands). Austral Ecol 31:747–758
244 SM. Sundarapandian et al.

Bartomeus I, Vila M, Santamaria L (2008) Contrasting effects of invasive plants in plant-pollinator


networks. Oecologia 155:761–770
Begum S, Wahab A, Siddiqui BS (2000) Pentacyclic triterpenoids from the aerial parts of Lantana
camara. Chem Pharm Bull 51:134–137
Bhagwat SA, Breman E, Thekaekara T, Thornton TF, Willis KJ (2012) A battle lost? Report on two
centuries of invasion and management of Lantana camara L. in Australia, India and South
Africa. PLoS One 7(3):e32407
Binggeli P (1996) A taxonomic, biogeographical and ecological overview of invasive woody
plants. J Veg Sci 7:121–124
Binggeli P, Hall JB, Healey JR (1998) An overview of invasive woody plants in the tropics,
pp 1–83. http://pages.bangor.ac.uk/~afs101/iwpt/web1-99.pdf. Accessed 10 Sept 2014
Blossey B, Hunt-Joshi TR (2003) Belowground herbivory by insects: influence on plants and
aboveground herbivores. Annu Rev Entomol 48:521–547
Bohn KK, Minogue PJ, Pieterson EC (2011) Control of invasive Japanese climbing fern (Lygodium
japonicuni) and response of native ground cover during restoration of a disturbed longleaf pine
ecosystem. Ecol Restor 29(4):346–356
Brandis D (1882) The forests of South India. Indian Forester 7:363–369
Bungard RA, Daly GT, McNeil DL, Jones AV, Morton JD (1997) Clematis vitalba in a New Zealand
native forest remnant: does seed germination explain distribution? N Z J Bot 35:525–534
Carroll SP, Mathieson M, Loye JE (2005) Invasion history and ecology of the environmental weed
balloon vine, Cardiospermum grandiflorum Swartz, in Australia. Plant Prot Q 20:140–144
Charles H, Dukes JS (2007) Impacts of invasive species on ecosystem services. In: Nentwig W (ed)
Biological invasions. Springer, Berlin, pp 217–237
Chaudhary BL, Bhansali E (2002) Effect of different concentration of Lantana camara Linn.
extract on spore germination of Physcomitrium japonicum Hedw. in half Knop’s liquid medium
and double distilled water. Res Bull Punjab Univ (Sci) 52:161–165
Chen BM, Ni GY, Ren WT, Peng SL (2007) Effects of aqueous extracts of the invasive plant
Mikania micrantha on litter decomposition of native plants in South China. Allelopath J
20:307–314
Chen BM, Peng SL, Ni GY (2009) Effect of the invasive plant Mikania micrantha H.B.K. and soil
nitrogen availability through allelopathy in South China. Biol Invasions 11:1291–1299
Choudhury AK (1972) Controversial Mikania (vine) – a threat to forests and agriculture. Indian
Forester 98:178–186
Cilliers CJ, Neser S (1991) Biological control of Lantana camara (Verbenaceae) in South Africa.
Agric Ecosyst Environ 37:57–75
Cronk QCB, Fuller JL (1995) Plant invaders: the threat to natural ecosystems. Chapman and Hall,
London
Day MD, Wiley CJ, Playford J, Zalucki MP (2003) Lantana: current management, status and
future prospects. Aust Cent Int Agric Res 5:1–20
De Araujo LD, Quirino ZG, Machado IC (2014) High specialisation in the pollination system
of Mandevilla tenuifolia (J.C. Mikan) Woodson (Apocynaceae) drives the effectiveness of
butterflies as pollinators. Plant Biol 16:947–955
Deng X, Ye WH, Feng HL, Yang QH, Cao HL, Xu KY, Zhang Y (2004) Gas exchange characteristics
of the invasive species Mikania micrantha and its indigenous congener M. cordata (Asteraceae)
in South China. Bot Bull Acad Sin 45:213–220
Dhileepan K, Taylor DBJ, Lockett C, Treviño M (2013) Cat’s claw creeper leaf-mining jewel
beetle Hylaeogena jureceki Obenberger (Coleoptera: Buprestidae), a host-specific biological
control agent for Dolichandra unguis-cati (Bignoniaceae) in Australia. Australian Weed Risk
Assessment for Dolichandra unguis-cati. J Entomol 52:175–181
Dietz H, Kueffer C, Parks CG (2006) MIREN: a new research network concerned with plant inva-
sion into mountain areas. Mt Res Dev 26:80–81
Ding J, Reardon R, Wu Y, Zheng H, Fu W (2006) Biological control of invasive plants through
collaboration between China and the United States of America: a perspective. Biol Invasions
8:1439–1450
12 Biological Invasion of Vines, Their Impacts and Management 245

DiTommaso A, Lawlor FM, Darbyshire SJ (2005) The biology of invasive alien plants in Canada.
2. Cynanchum rossicum (KLEOPOW) BORHIDI [=Vincetoxicum rossicum (KLEOPOW)
BARBAR.] and Cynanchum louiseae (L.) KARTESZ & GANDHI [=Vincetoxicum nigrum (L.)
MOENCH]. Can J Plant Sci 85:243–263
Dobhal PK, Kohli RK, Batish DR (2011) Impact of Lantana camara L. invasion on riparian
vegetation of Nayar region in Garhwal Himalayas (Uttarakhand, India). J Ecol Nat Environ
3:11–22
Dogra KS, Sood SK, Dobhal PK, Sharma S (2010) Alien plant invasion and their impact on
indigenous species diversity at global scale: a review. J Ecol Nat Environ 2:175–186
Dreyer G, Baird L, Fickler C (1987) Celastrus scandens and Celastrus orbiculatus: comparisons
of reproductive potential between a native and an introduced woody vine. Bull Torrey Bot Club
114:260–264
Du F, Yang YM, Li JQ, Yin WY (2006) A review of Mikania and the impact of M. micrantha
(Asteraceae) in Yunnan. Acta Bot Yunnanica 28(5):505–508
Duxbury C, Glasscock S, Staniszewska I (2003) Control of regrowth from air potato (Disoscorea
bulbifera) bulbils. Wildland Weed 6:14–15
Ellstrand NC, Schierenbeck KA (2000) Hybridization as a stimulus for the evolution of invasive-
ness in plants? Proc Natl Acad Sci U S A 97:7043–7050
Ellsworth JW (2003) Controls on the establishment and early growth of oriental bittersweet
(Celastrus Orbiculatus Thunb.), an invasive woody vine, University of Massachusetts at
Amherst, p 118
Emms J, Virtue JG, Preston C, Bellotti WD (2004) Do all legumes pose the same weed risk?
Development of a method to evaluate the risk of introduced legumes to temperate Australia,
pp 105–108. In: Sindel BM, Johnson SB (eds) 14th Australian Weeds Conference proceedings:
weed management – balancing people, planet, profit. p 718
Ernst CM, Cappuccino N (2005) The effect of an invasive alien vine, Vincetoxicum rossicum
(Asclepiadaceae) on arthropod populations in Ontario old fields. Biol Invasions 7:417–425
Ewel JJ, O’Dowd DJ, Bergelson J, Daehler CC, D’Antonio CM, Gomez LD, Gordon DR, Hobbs
RJ, Holt A, Hopper KR, Hughes CE, La Hart M, Leakey RRB, Lee WG, Loope LL, Lorence
DH, Louda SM, Lugo AE, McEvoy PB, Richardson DM, Vitousek PM (1999) Deliberate intro-
ductions of species: research needs. Bio Sci 49:619–630
Floyd AG (1989) The vine weeds of coastal rainforests. In: Proceedings of the 5th Biennial
Noxious Plants Conference. New South Wales Department of Agriculture and Fisheries,
Sydney, pp 1109–1115
Ghisalberti EL (2000) Lantana camara Linn. (review). Fitoterapia 71:467–485
GISIN (2011) Global invasive species information network. Available: http://www.gisin.org/
cwis438/Websites/GISINDirectory/SpeciesStatus_TopInvasives.php?WebSiteID4. Accessed
10 Aug 2014
Global Invasive Species Database (2013). http://www.issg.org/database/
Gooden B, French K, Turner P (2009) Invasion and management of a woody plant, Lantana
camara L., alters vegetation diversity within wet sclerophyll forest in south-eastern Australia.
For Ecol Manage 257:960–967
Gordon DR, Riddle B, Pheloung PC, Ansari S, Buddenhagen C, Chimera C, Daehler CC, Dawson
W, Denslow JS, Tshidada NJ, La Rosa A, Nishida T, Onderdonk DA, Panetta FD, Pysek P,
Randall RP, Richardson DM, Virtue JG, Williams PA (2010) Guidance for addressing the
Australian Weed Risk Assessment questions. Plant Prot Q 25:56–74
Granados J, Koerner C (2002) In deep shade, elevated CO2 increases the vigor of tropical climbing
plants. Glob Chang Biol 8:1109–1117
Greathead DJ (1968) Biological control of Lantana – a review and discussion of recent developments
in East Africa. PANS 14:167–175
Greenberg CH, Smith LM, Levey DJ (2001) Fruit fate, seed germination and growth of an invasive
vine – an experimental test of ‘sit and wait’ strategy. Biol Invasions 3:363–372
Groves RH, Willis AJ (1999) Environmental weeds and loss of native plant biodiversity: some
Australian examples. Aust J Environ Manag 6:164–171
246 SM. Sundarapandian et al.

Harden GJ, Fox MD (1988) Wingham Brush regeneration assessment. Report 9, Royal Botanic
Gardens Sydney
Harden GJ, Fox MD, Fox BJ (2004) Monitoring and assessment of restoration of a rain forest
remnant at Wingham Brush, NSW. Austral Ecol 29:489–507
Harrington RA, Kujawski R, Ryan HDP (2003) Invasive plants and the green industry. J Arboric
29:42–48
Harwood E, Sytsma M (2003) Risk assessment for Chinese Water Spinach (Ipomoea aquatica) in
Oregon. Portland State University, Portland. http://www.oregon.egov.com/oisc/docs/pdf/ipaq_ra.pdf
Hills L (1999) Mile-a-minute. Agnote, 535. Online at http://www.nt.gov.au/dpif/pubcat/agntes/535.
htrn
Hiremath J, Sundaram B (2005) The fire-Lantana cycle hypothesis in Indian forests. Conserv Soc
3:26–42
Holm LG, Plucknett DL, Pancho JV, Herberger JP (1977) The world’s worst weeds. Distribution
and biology. University Press of Hawaii, Honolulu
Hooper DU, Chapin FS, Ewel JJ, Hector A, Inchausti P, Lavorel S, Lawton JH, Lodge DM,
Loreau M, Naeem S, Schmid B, Setala H, Symstad AJ, Vandermeer J, Wardle DA (2005)
Effects of biodiversity on ecosystem functioning: a consensus of current knowledge. Ecol
Monogr 75:3–35
Huang ZL, Cao HL, Liang XD, Ye WH, Feng HL, Cai CX (2000) The growth of damaging effects
by Mikania micrantha in different habitats. J Trop Subtrop Bot 8:131–138
Hulme PE (2007) Biological invasions in Europe: drivers, pressures, states, impacts and responses.
In: Hester R, Harrison RM (eds) Biodiversity under threat. Cambridge University Press,
Cambridge, pp 56–80
Hunt-Joshi TR, Blossey B, Root RB (2004) Root and leaf herbivory on Lythrum salicaria: implica-
tions for plant performance and communities. Ecol Appl 14:1574–1589
Inouye DW (1982) The consequences of herbivory: a mixed blessing for Jurinea mollis
(Asteraceae). Oikos 39:269–272
Ismail MI (2006) Inhibitory effects of Na-hypochlorite and heating on the mycobiota associated
with fruits or juice of passion (Passiflora edulis Sims) in Uganda. Mycobiology 34:92–98
Ismail BS, Mah LS (1993) Effects of Mikania micrantha H.B.K. on germination and growth of
weed species. Plant Soil 157:107–113
Jain R, Singh M, Dezman DJ (1989) Qualitative and quantitative characterization of phenolic
compounds from Lantana (Lantana camara) leaves. Weed Sci 37:302–307
Johnson K (2011) Cat’s claw creeper (Dolichandra unguis-cati) – weed management guide.
Biosecurity Queensland DEEDI. Online at: http://www.weeds.org.au/WoNS/catsclawcreeper/
docs/WMG_CATS_CLAW_CREEPER-final.pdf
Kannan R. Uma Shaanker R, Joseph G. (2009) Putting an invasive alien species to good use. In:
Business.2010, Business & Biodiversity June 2009, vol 4, no 1, Convention on Biological
Diversity, pp 18–19
Kannan R, Aravind NA, Joseph G, Ganeshaiah KN, Shaanker RU (2008) Lantana craft: a weed for
a need. Biotech News 3:9–11
Kay FM, Marina Oleiro M, Fourie A, Simelane D (2010) Natural enemies of balloon vine
Cardiospermum grandiflorum (Sapindaceae) in Argentina and their potential use as biological
control agents in South Africa. Int J Trop Insect Sci 30:67–76
Keighery GJ, Gibson N, Kenneally KF, Mitchell AA (1995) Biological inventory of Koolan Island,
Western Australia 1. Flora and vegetation. Rec West Aust Mus 17:237–248
Kellerman TS, Naude TW, Fourie N (1996) The distribution, diagnoses and estimated economic
impact of plant poisonings and mycotoxicoses in South Africa. Onderstepoort J Vet Res
63:65–90
Khan M, Srivastava SK, Jain N, Syamasundar KV, Yadav AK (2003) A chemical composition of
fruit and stem essential oil of Lantana camara from northern India. Flavour Fragrance J
18:376–379
Khoshoo TN, Mahal C (1967) Versatile reproduction in Lantana camara. Curr Sci 36:201–203
12 Biological Invasion of Vines, Their Impacts and Management 247

Kim KD (2012) An exotic invasive Liana, Wisteria in Korea. International conference on biological
and life sciences, IPCBEE, vol 40. IACSIT Press, Singapore
Kimothi MM, Dasari A (2010) Methodology to map the spread of an invasive plant (Lantana
camara L.) in forest ecosystems using Indian remote sensing satellite data. Int J Remote Sens
31:3273–3289
Kohli RK, Batish DR, Singh HP, Dogra KS (2006) Status, invasiveness and environmental threats
of three tropical American invasive weed (Parthenium hysterophorus L., Ageratinum conyzoides
L., Lantana camara L.) in India. Biol Invasions 8:1501–1510
Kricsfalusy V, Miller GC (2010) Community ecology and invasion of natural vegetation by
Cynanchum rossicum (Asclepiadaceae) in the Toronto region, Canada. Thaiszia J Bot 20:53–70
Kriticos DJ, Sutherst RW, Brown JR, Adkins SW, Maywald GF (2003) Climate change and biotic
invasions: a case history of a tropical woody vine. Biol Invasions 5:145–165
Kueffer C, Daehler CC (2009) A habitat-classification framework and typology for understanding,
valuing, and managing invasive species impacts. In: Inderjit (ed) Management of invasive
weeds. Springer, Dordrecht, pp 77–101. doi:10.1007/978-1-4020-9202-2_5
Lake E, Cutting K, Hough-Goldstein J (2011) Integrating biological control and native plantings
to restore sites invaded by mile-a-minute weed, Persicaria perfoliata, in the Mid-Atlantic USA.
In: Proceedings of the XIII international symposium on biological control of weeds – 2011. pp
254–261. Online at http://www.invasive.org/proceedings/pdfs/Lake.pdf
Lake EC, Hough-Goldstein J, Amico V (2014) Integrating management techniques to restore sites
invaded by mile-a-minute weed, Persicaria perfoliata. Restor Ecol 22:127–133
Leatherman AD (1955) Ecological life history of Lonicera japonica Thunb. Dissertation,
University of Tennessee, Knoxville
Leicht SA, Silander JA (2006) Differential responses of invasive Celastrus orbiculatus (celastra-
ceae) and native C. scandens to changes in light quality. Am J Bot 93:972–977
Leicht-Young SA, Silander JA, Latimer AM (2007) Comparative performance of invasive and
native Celastrus species across environmental gradients. Oecologia 154:273–282
Li MG, Zhang WY, Liao WB, Wang BS, Zan QJ (2000) The history and status of the study on
Mikania micrantha. Ecol Sci 19:41–45
Li LY, Peng TX, Liu WH (2002) Actinote anteas D & H (Lepidoptera; Nymphalidae; Acraeinae) a
new biological agent for controlling the weed Mikania micrantha. Nat Enemies Insect 24:49–52
Li WH, Zhang CB, Jiang HB, Xin GR, Yang ZY (2006) Changes in soil microbial community
associated with invasion of the exotic weed, Mikania micrantha H.B.K. Plant Soil
281:309–324
Loewenstein NJ, Loewenstein EF (2005) Non-native plants in the understory of riparian forests
across a land use gradient in the Southeast. Urban Ecosyst 8:79–91
Lott MS, Volin JC, Pemberton RW, Austin DF (2003) The reproductive biology of the invasive
ferns Lygodium microphyllum and L. japonicum (Schizaeaceae): implications for invasive
potential. Am J Bot 90:1144–1152
Love A, Babu S, Babu CR (2009) Management of Lantana, an invasive alien weed, in forest
ecosystems of India. Curr Sci 97:1421–1429
Lowe S, Browne M, Boudjelas S, Depoorter M (2001) 100 of the world’s worst invasive alien
species. A selection from the Global Invasive Species Database. IUCN/SSC Invasive Species
Specialist Group (ISSG), Auckland
Mack RN (2001) Motivations and consequences of the human dispersal of plants. In: McNeely JA
(ed) The great reshuffling: human dimensions of invasive alien species. ICUN, Gland/
Cambridge, UK, pp 23–34
Mack MC, D’Antonio CM (1998) Impacts of biological invasions on disturbance regimes. Trends
Ecol Evol 13:195–198
Maron JL (1998) Insect herbivory above-and belowground: individual and joint effects on plant
fitness. Ecology 79:1281–1293
McGeoch MA, Butchart SHM, Spear D, Marais E, Kleynhans EJ, Symes A, Chanson J, Hoffmann
M (2010) Global indicators of biological invasion: species numbers, biodiversity impact and
policy responses. Divers Distrib 16:95–108
248 SM. Sundarapandian et al.

McNab WH, Meeker M (1987) Oriental bittersweet: a growing threat to hardwood silviculture in
the Appalachians. North J Appl For 4:174–177
Michigan Department of Natural Resources (2012) Michigan Natural Features Inventory 2/2012.
http://www.michigan.gov/documents/dnr/Oriental_Bittersweet_389123_7.pdf
Miller JH (2003) Nonnative invasive plants of southern forests: a field guide for identification and
control. General Technical Report SRS-62, USDA Forest Service Southern Research Station,
Asheville
Mohan Ram HY, Mathur G (1984) Flower colour changes in Lantana camara. J Exp Bot
35:1656–1662
Morton JF (1994) Lantana, or red sage (Lantana camara L., [Verbenaceae]), notorious weed and
popular garden flower; some cases of poisoning in Florida. Econ Bot 48:259–270
Murali KS, Setty RS (2001) Effect of weeds Lantana camara and Chromelina odarata growth on
the species diversity, regeneration and stem density of tree and shrub layer in BRT sanctuary.
Curr Sci 80:675–678
Muthumperumal C, Parthasarathy N (2010) A large-scale inventory of vine diversity in tropical
forests of South Eastern Ghats. India Syst Biodivers 8:289–300
Nagao T, Abe F, Kinjo J (2002) Antiproliferative constituents in plants: flavones from the leaves of
Lantana montevidensis Briq. and consideration of structural relationship. Biol Pharm Bull
25:875–879
Naithani S, Pande PK (2009) Evaluation of Lantana camara Linn. stem for pulp and paper mak-
ing. Indian Forester 135:1081–1087
Nanjappa HV, Saravanane P, Ramachandrappa BK (2005) Biology and management of Lantana
camara L. – a review. Agric Rev 26:272–280
Ni GY, Song LY, Zhang JL, Peng SL (2006) Effects of root extracts of Mikania micrantha H.B.K.
on soil microbial community. Allelopath J 17:247–254
Noble D, Lane SJ, Sidebottom PJ, Lynn SM (1998) Isolation of translactone-containing triterpenes
with thrombin inhibitory activities from the leaves of Lantana camara. J Nat Prod
61:1328–1331
Ogle CC, La Cock GD, Arnold G, Mickleson N (2000) Impact of an exotic vine Clematis vitalba
(F. Ranunculaceae) and of control measures on plant biodiversity in indigenous forest, Taihape,
New Zealand. Austral Ecol 25:539–551
Olden JD, Poff NL (2003) Toward a mechanistic understanding and prediction of biotic homogeni-
zation. Am Nat 162:442–460
Palazzola E, Burnham RJ (2013) Climbers: censusing lianas in mesic biomes of eastern regions
Hedera helix L. University of Michigan. http://climbers.lsa.umich.edu/?p=172
Parker C (1972) The Mikania problem. Pest Artic News Summ 18:312–315
Parsons WT, Cuthbertson EG (1992) Noxious weeds of Australia. Inkata Press, Melbourne
Patterson DT (1974) The ecology of oriental bittersweet, Celastrus orbiculatus, a weedy
introduced ornamental vine. Ph.D. dissertation, Duke University, Durham
Patterson DT (1975) Photosynthetic acclimation to irradiance in Celastrus orbiculatus Thunb.
Photosynthetica 9:140–144
Pauchard A, Kueffer C, Dietz H, Daehler CC, Alexander J, Edwards PJ, Arevalo JR, Cavieres L,
Guisan A, Haider S, Jakobs G, McDougall K, Millar CI, Naylor BJ, Parks CG, Rew LJ, Seipel
T (2009) Ain’t no mountain high enough: plant invasions reaching new elevations. Front Ecol
Environ 7:479–486
Paynter Q, Harman H, Waipara N (2006) Prospects for biological control of Merremia peltata.
Landcare Research, Auckland, p 34
Pejchar L, Mooney HA (2009) Invasive species, ecosystem services and human wellbeing. Trends
Ecol Evol 24:497–504
Pemberton RW, Pratt PD (2002) Skunk vine. In: Van Driesche R, Blossey B, Hoddle M, Lyon S,
Reardon R (eds) Biological control of invasive plants in the eastern United States, USDA
Forest Service Publication FHTET-2002-04, pp 343–351, 413p. http://dnr.state.il.us/steward-
ship/cd/biocontrol/pdf/27SkunkVine.pdf
12 Biological Invasion of Vines, Their Impacts and Management 249

Pisula NL, Meiners SJ (2010) Relative allelopathic potential of invasive plant species in a young
disturbed woodland. J Torrey Bot Soc 137:81–87
Prabu NR, Stalin N, Swamy PS (2014) Ecophysiological attributes of Mikania micrantha, an
exotic invasive weed, at two different elevations in the tropical forest regions of the Western
Ghats, South India. Weed Biol Manag 14:59–67
Prasad AM, Iverson LR, Liaw A (2006) Newer classification and regression techniques: bagging
and random forests for ecological prediction. Ecosystems 9:181–199
Priyanka N, Joshi PK (2013) A review of Lantana camara studies in India. Int J Scic Res Public
3:1–11
Priyanka N, Shiju MV, Joshi PK (2013) A framework for management of Lantana camara in India.
Proc Int Acad Ecol Environ Sci 3:306–323
Putz FE (1991) Silvicultural effects of lianas. In: Putz FE, Mooney HA (eds) The biology of vines.
Cambridge University Press, Cambridge, pp 493–501
Pysek P (1998) Is there a taxonomic pattern to plant invasions? Oikos 82:282–294
Pysek P, Richardson DM (2010) Invasive species, environmental change and management, and
health. Annu Rev Environ Resour 35:25–55
Pysek P, Krivanek M, Jarosik V (2009) Planting intensity, residence time, and species traits
determine invasion success of alien woody species. Ecology 90:2734–2744
Pysek P, Jarosik V, Hulme PE, Pergl J, Hejda M, Schaffner U, Vila M (2012) A global assessment
of invasive plant impacts on resident species, communities and ecosystems: the interaction of
impact measures, invading species’ traits and environment. Glob Chang Biol 18:1725–1737
Raghu S, Dhileepan K, Trevino M (2006) Response of an invasive liana to simulated herbivory:
implications for its biological control. Acta Oecol 29:335–345
Ram AK (2008) Impact of Mikania micrantha on Rhinoceros habitat in Chitwan National Park,
Central Nepal. PhD thesis (unpublished), submitted to Tribhuvan University, Pokhara
Ramachandran K, Subramaniam B (1983) Scarlet gourd, Coccinia grandis, little-known tropical
drug plant. Econ Bot 37:380–383
Ramakrishnan PS (1984) The science behind rotational bush fallow agricultural systems (jhum).
Proc Indian Acad Sci Plant Sci 93:379–400
Ray AK, Puri MK (2006) Modeling H factor-kappa number for kraft pulping of Lantana camara
plant- an experimental investigation. Adv Biocatal Protein Eng 15:1–62
Raymond KL (1996) Geophytes as weeds: bridal creeper (Asparagus asparagoides) as a case
study. In: Shepherd RCH (ed) Proceedings of the eleventh Australian weeds conference. Weed
Science Society of Victoria, Frankston, pp 420–423
Reardon R (2006) Overview of the forest health technology enterprise team biological control
program for invasive species – 1995 through 2007. Forest Health Technology Enterprise Team.
Online at http://www.fs.fed.us/foresthealth/technology/pdfs/BiocontrolsOverview.pdf
Rejmanek M (2000) Invasive plants: approaches and predictions. Austral Ecol 25:497–506
Rejmanek M, Richardson DM, Higgins SI, Pitcairn MJ, Grotkopp E (2005) Ecology of invasive
plants – state of the art. In: Mooney HA, Mack RN, Mc Neely JA, Neville L, Schei PJ, Waage
J (eds) Invasive alien species: a new synthesis. Island Press, Washington, DC, pp 104–161
Remaley T (2005) Weeds gone wild: alien plant invaders of natural areas. http://www.nps.gov/
plants/alien/
Richardson DM (1998) Forestry trees as invasive aliens. Conserv Biol 12:18–26
Richardson DM (2011a) Forestry and agroforestry. In: Simberloff D, Rejmanek M (eds)
Encyclopedia of biological invasions. University of California Press, Berkeley, pp 241–248
Richardson DM (2011b) Trees and shrubs. In: Simberloff D, Rejmanek M (eds) Encyclopedia of
biological invasions. University of California Press, Berkeley, pp 670–677
Richardson DM, Rejmanek M (2004) Invasive conifers: a global survey and predictive framework.
Divers Distrib 10:321–331
Richardson DM, Pyusek P, Rejmanek M, Barbour MG, Panetta FD, West CJ (2000) Naturalisation
and invasion of alien plants: concepts and definitions. Divers Distrib 6:93–107
Robertson DJ, Robertson MC, Tague T (1994) Colonization dynamics of four exotic plants in a
northern Piedmont natural area. Bull Torrey Bot Club 121:107–118
250 SM. Sundarapandian et al.

Sahu PK, Singh JS (2008) Structural attributes of lantana-invaded forest plots in Achanakmar–
Amarkantak Biosphere Reserve, Central India. Curr Sci 94:494–500
Sapkota L (2007) Ecology and management issues of Mikania micrantha in Chitwan National
Park, Nepal. Banko Janakari 17:27–39
Satheesh LS, Murugan K (2011) Antimicrobial activity of protease inhibitor from leaves of
Coccinia grandis (L.) Voigt. Indian J Exp Biol 49:366–374
Sax DF, Gaines SD (2003) Species diversity: from global decreases to local increases. Trends Ecol
Evol 18:561–566
Saxena MK (2000) Aqueous leachate of Lantana camara kills water hyacinth. J Chem Ecol
26:2435–2447
Schierenbeck KA (2004) Japanese honeysuckle (Lonicera japonica) as an invasive species;
history, ecology, and context. Crit Rev Plant Sci 23:391–400
Schierenbeck KA, Mack RN, Sharitz RR (1994) Effects of herbivory on growth and biomass
allocation in native and introduced species of Lonicera. Ecology 75:1661–1672
Scientific Committee, New South Wales Government. (2011). http://www.environment.nsw.gov.
au/determinations/ExoticVinesKtp.htm
Scott L (1998) Identification of Lantana spp. Taxa in Australia and the South Pacific. Final report.
Australian Centre International Agricultural Research, Canberra
Shao H, Peng S, Wei X, Zhang D, Zhang C (2005) Potential allelochemicals from an invasive weed
Mikania micrantha H.B.K. J Chem Ecol 31:1657–1668
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 (2010) How lantana invades dry deciduous forest: a case study from
Vindhyan highlands, India. Trop Ecol 51(2S):305–316
Sharma GP, Raghubanshi AS (2011) How lantana invaded India. Curr Conserv 4:21–22
Sharma OP, Sharma PD (1989) Natural products of the lantana plant – the present and prospects.
J Sci Ind Res 48:471–478
Sharma OP, Makkar HPS, Dawra RK (1988) A review of the noxious plant Lantana camara.
Toxicon 26:975–987
Sharma S, Singh A, Sharma OP (1999) An improved procedure for isolation and purification of
lantadene A, the bioactive pentacyclic triterpenoid from Lantana camara leaves. J Med Aromat
Plant Sci 21:686–688
Sharma GP, Raghubanshi AS, Singh JS (2005) Lantana invasion: an overview. Weed Biol Manag
5:157–167
Sharma OP, Sharma S, Pattabhi V, Mahato SB, Sharma PD (2007) A review of the hepatotoxic
plant. Lantana camara. J Sci Ind Res 37:313–352
Shen S, Xu G, Zhang F, Jin G, Liu S, Liu M, Chen A, Zhang Y (2013) Harmful effects and chemi-
cal control study of Mikania micrantha H.B.K in Yunnan, Southwest China. Afr J Agric Res
8:5554–5561
Sheppard AW (2003) Prioritising agents based on predicted efficacy: beyond the lottery approach.
In: Jacob HS, Briese DT (eds) Improving the selection, testing and evaluation of weed biologi-
cal control agents. Cooperative Research Centre for Australian Weed Management, Adelaide,
pp 11–21
Simba YR, Kamweya AM, Mwangi PN, Ochora JM (2013) Impact of the invasive shrub, Lantana
camara L. on soil properties in Nairobi National Park, Kenya. Int J Biodivers Conserv
5:803–809
Simberloff D, Von Holle B (1999) Positive interactions of nonindigenous species: invasional
meltdown? Biol Invasions 1:21–32
Simelane DO (2005) Biological control of Lantana camara in South Africa: targeting a different
niche with a root-feeding agent, Longitarsus sp. BioControl 50:375–387
Singh U, Wadhwani AM, Johri BM (1996) Dictionary of economic plants of India. Pbl. Indian
Council of Agricultural Research, New Delhi, p 118
12 Biological Invasion of Vines, Their Impacts and Management 251

Singh JS, Singh SP, Gupta SR (2006) Ecology environment and resources conservation. Anamaya
Publication, New Delhi, p 668
Soni PL, Naithani S, Gupta PK, Bhatt A, Khullar R (2006) Utilization of economic potential of
Lantana camara. Indian Forester 132:1625–1630
Sorrie BA, Somers P (1999) The vascular plants of Massachusetts: a county checklist. Massachusetts
Division of Fisheries and Wildlife Natural Heritage & Endangered Species Program, Westborough
Stansbury CD (2001) Dispersal of the environmental weed Bridal creeper, Asparagus asparagoi-
des by Silvereyes, Zosterops lateralis in south-western Australia. Emu 101:39–45
Starr F, Starr K, Loope L (2003a) Buddleja madagascariensis, Scrophulariaceae. United States.
Geological Survey – Biological Resources Division, Haleakala Field Station, Maui
Starr F, Starr K, Loope L (2003b) Delairea odorata Cape ivy, Asteraceae. United States. Geological
Survey – Biological Resources Division, Haleakala Field Station, Maui
Starr F, Starr K, Loope L (2003c) Hiptage benghalensis Hiptage, Moraceae. United States.
Geological Survey – Biological Resources Division, Haleakala Field Station, Maui
Stone KR (2009) Wisteria floribunda, W. sinensis. In: Fire Effects Information System, [Online].
U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station, Fire
Sciences Laboratory (Producer). Available: http://www.fs.fed.us/database/feis/
Stratton L (1996) The impact and spread of Rubus ellipticus in ‘Ola’a Forest Tract Hawaii
Volcanoes National Park. Technical report 107, Cooperative National Park Resources Studies
Unit, University of Hawaii at Manoa; Western Region, National Park Service, Honolulu, p 35
Surampalli D (2010) Abundance of Lantana camara in open canopy, partial canopy, closed
canopy areas in forest trails, Karnataka. https://mail.google.com/mail/u/0/?tab=wm&pli=
1#inbox/1492482fd06906c0?projector=1. Accessed 10 Oct 2014
Swamy PS (1986) Eco-physiological and demographic studies of weeds of secondary successional
environments after slash and burn agriculture in north-eastern India. PhD thesis, North-Eastern
Hill University, Shillong, 270 pp
Swamy PS, Ramakrishnan PS (1987a) Weed potential of Mikania micrantha H.B.K. and its control
in fallows after shifting agriculture (Jhum) in north-east India. Agric Ecosyst Environ
18:195–204
Swamy PS, Ramakrishnan PS (1987b) Effect of fire on population dynamics of Mikania micrantha
H.B.K. during early succession after slash and burn agriculture (jhum) in north-east India.
Weed Res 27:397–403
Swamy PS, Ramakrishnan PS (1987c) Contribution of Mikania micrantha during secondary
succession following slash-and-burn agriculture (Jhum) in north-east India I. Biomass, litterfall
and productivity. For Ecol Manage 22:229–237
Swamy PS, Ramakrishnan PS (1987d) Contribution of Mikania micrantha during secondary
succession following slash-and-burn agriculture (Jhum) in north-east India II. Nutrient cycling.
For Ecol Manage 22:239–249
Swamy PS, Ramakrishnan PS (1988a) Growth and allocation patterns of Mikania micrantha in
successional environments after slash and burn agriculture. Can J Bot 66:1465–1469
Swamy PS, Ramakrishnan PS (1988b) Effect of fire on growth and allocation strategies of Mikania
micrantha under early successional environments. J Appl Ecol 25:653–658
Sydney Weeds Committees (2012) http://sydneyweeds.org.au/wp-cms/wp-content/uploads/Weed-
Fact-Sheet-Cape-Ivy.pdf
Tassin J, Riviere JN, Cazanove M, Bruzzese E (2006) Ranking of invasive woody plant species for
management on Reunion Island. Weed Res 46:388–403
Thomas JR, Middleton B, Gibson DJ (2006) A landscape perspective of the stream corridor inva-
sion and habitat characteristics of an exotic (Dioscorea oppositifolia) in a Pristine watershed in
Illinois. Biol Invasions 8:1103–1113
Thunbergia – Thunbergia spp. Department of Agriculture, Fisheries and Forestry (2013) Fact
sheet, declared class 1 and 2 Pest Plant , p 23 Dec 2013
Timmins SM, Reid V (2000) Climbing asparagus, Asparagus scandens Thunb.: a South African in
your forest patch. Austral Ecol 25:533–538
252 SM. Sundarapandian et al.

Tiwari S, Adhikari B, Siwakoti M, Subedi K (2005) An inventory and assessment of invasive alien
plant species in Nepal. IUCN The World Conservation Union, Kathmandu
Tomley AJ, Evans HC (2004) Establishment of, and preliminary impact studies on, the rust,
Maravalia cryptostegiae, of the invasive alien weed, Cryptostegia grandiflora in Queensland,
Australia. Plant Pathol 53:475–484
Tripathi RS, Khan ML, Yadav AS (2012) Biology of Mikania micrantha H.B.K.: a review. In: Bhatt
JR, Singh JS, Singh SP, Tripathi RS, Kohli RK (eds) Invasive alien plants: an ecological
appraisal for the Indian subcontinent. CAB International, Oxfordshire, pp 99–107
Van Klinken RD, Raghu S (2006) A scientific approach to agent selection. Aust J Entomol
45:253–258
Vila M, Basnou C, Pysek P, Josefsson M, Genovesi P, Gollasch S, Nentwig W, Olenin S, Roques
A, Roy D, Hulme PE, DAISIE Partners (2009) How well do we understand the impacts of alien
species on ecological services? A pan-European cross-taxa assessment. Front Ecol Environ
8:135–144
Vila M, Espinar JL, Hejda M, Hulme PE, Jarosik V, Maron JL, Pergl J, Schaffner U, Sun Y, Pysek
P (2011) Ecological impacts of invasive alien plants: a meta-analysis of their effects on species,
communities and ecosystems. Ecol Lett 14:702–708
Vitelli JS, Madigan BA, Van Haaren PE, setter S, logan P (2009) Control of the invasive liana,
Hiptage benghalensis. Weed Biol Manag 9:54–62
Waggy MA (2010) Hedera helix. In: Fire Effects Information System, [Online]. U.S. Department
of Agriculture, Forest Service, Rocky Mountain Research Station, Fire Sciences Laboratory
(Producer). Available: http://www.fs.fed.us/database/feis
Wakjira M (2011) An invasive alien weed giant sensitive plant (Mimosa diplotricha Sauvalle)
invading Southwestern Ethiopia. Afr J Agric Res 6:127–131
Wang BS, Liao WB, Miao RH (2001) Revision of Mikania from China and the key of four relative
species. Acta Sci Nat Univ Sunyatseni 40:72–75
Wang BS, Wang YJ, Liao WB, Zhan QJ, Li MG, Peng SL (2004) The invasion ecology and
management of alien weed Mikania micrantha H.B.K. Science Press, Beijing, pp 76–85
Wang RL, Zeng RS, Peng SL, Chen BM, Liang XT, Xin XW (2011) Elevated temperature may
accelerate invasive expansion of the liana plant Ipomoea cairica. Weed Res 51:574–580
Waterhouse DF (1994) Biological control of weeds: Southeast Asian prospects, vol 125. ACIAR,
Canberra
Waterhouse BM (2003) Know your enemy: records of potentially serious weeds in northern
Australia, Papua New Guinea and Papua (Indonesia). Telopea 10:477–484
Weed AS, Casagrande RA (2010) Biology and larval feeding impact of Hypena opulenta
(Christoph) (Lepidoptera: Noctuidae): a potential biological control agent for Vincetoxicum
nigrum and V. rossicum. Biol Control 53:214–222
Weed AS, Gassmann A, Leroux AM, Casagrande RA (2010) Performance of potential European
biological control agents of Vincetoxicum spp. with notes on their distribution. J Appl Entomol
135:700–713
Wells MJ, Stirton CH (1988) Lantana camara: a poisonous declared weed. Farming in South
Africa. Weeds A-27. Department of Agriculture and Water Supply, Pretoria
Wijayabandara SMKH, Jayasuriya KMGG, Jayasinghe JLDHC (2013) Seed dormancy, storage
behavior and germination of an exotic invasive species, Lantana camara L. (Verbenaceae). Int
Res J Biol Sci 2:7–14
Williams PA, Cameron EK (2006) Creating gardens: the diversity and progression of European
plant introductions. In: Allen RB, Lee WG (eds) Biological invasions in New Zealand. Springer,
Berlin, pp 33–47
Williamson M, Fitter A (1996) The varying success of invaders. Ecology 77:1661–1666
Willis AJ, McKay R, Vranjic JA, Kilby MJ, Groves RH (2003) Comparative seed ecology of the
endangered shrub, Pimelea spicata and a threatening weed, Bridal Creeper: smoke, heat and
other fire-related germination cues. Ecol Manage Restor 4:55–65
Winter M, Schweiger O, Klotz S, Nentwig W, Andriopoulosd P, Arianoutsoud M, Basnoue C,
Delipetrouf P, Didziulisg V, Hejdah M, Hulmei PE, Lambdonj PW, Perglh J, Pyšekh kP, Royl
12 Biological Invasion of Vines, Their Impacts and Management 253

DB, Kuhn I (2009) Plant extinctions and introductions lead to phylogenetic and taxonomic
homogenization of the European flora. Proc Natl Acad Sci U S A 106:21721–21725
Wu SH, Chaw SM, Rejmanek M (2003) Naturalized Fabaceae (Leguminosae) species in Taiwan:
the first approximation. Bot Bull Acad Sin 44:59–66
www.wvnps.org/FightingInvasives.pdf
Xie Y, Li Z, William PG, Li D (2000) Invasive species in China – an overview. Biodivers Conserv
10:1317–1341
Ye WH, Zhou X (2001) The plant killer Mikania micrantha in South China. Aliens 13:7
Ye UH, Lin SS, Yu DY, Liang YX (2013) Is the overspreading of paederia scandens in highly
disturbed areas just occasional? Pak J Bot 45:1149–1158
Yu XM, Yang FJ (2011) Ecological characters and invasion route of Mikania micrantha in
Shenzhen Bay. J Northeast For Univ 39:51–52
Zhang WJ, Chen B (2011) Environment patterns and influential factors of biological invasions: a
worldwide survey. Proc Int Acad Ecol Environ Sci 1:1–14
Zhang LY, Ye WH, Cao HL, Feng HL (2004) Mikania micrantha H.B.K. in China – an overview.
Weed Res 44:42–49
Zheng H, Wu Y, Ding J, Binion D, Fu W, Reardon R (2006) Invasive plants of Asian origin estab-
lished in the United States and their natural enemies, vol 1. USDA, Morgantown, 147p. Online
at http://www.fs.fed.us/foresthealth/technology/pdfs/IPAOv1ed2.pdf
Chapter 13
Liana Diversity and the Future of Tropical
Forests

Mason Campbell, Ainhoa Magrach, and William F. Laurance

Abstract Lianas contribute substantially to the total species richness of tropical


forests, accounting for up to a quarter of the woody plant diversity. However, liana
diversity is intrinsically linked with forest condition and consequently is altered by
human-induced forest modifications. Multiple environmental drivers including
forest fragmentation, logging and climate change are impacting tropical forests; the
extent and intensity of their effects will likely define future global liana diversity.

Keywords Climate change • Disturbance • Fragmentation • Hunting • Lianas •


Logging • Vines

13.1 Introduction

Global liana diversity is both spatially and temporally heterogeneous (Gentry 1991;
Schnitzer and Bongers 2002). Liana diversity peaks in tropical climes where it is
intrinsically linked with the distribution of closed-canopy forests (Gentry 1991).
However, within the closed-canopy forests of the tropics, there is considerable
regional variation in liana diversity in response to local environmental or ecological
conditions and the evolutionary history of a region’s flora (Gentry 1991; Schnitzer
and Bongers 2002; Durigon et al. 2013; Schnitzer 2005; Gianoli 2004; Hegarty and
Clifford 1991).
Regional liana diversity also varies over time, as both the lianas and the forests
they inhabit respond to changes in the prevailing environmental conditions and,
more recently, human impacts (Laurance et al. 2014a; Phillips et al. 2002; Schnitzer

M. Campbell (*) • W.F. Laurance


Centre for Tropical Environmental and Sustainability Science (TESS) and College of Marine
and Environmental Sciences, James Cook University, Cairns, QLD 4878, Australia
e-mail: mason.campbell@my.jcu.edu.au; bill.laurance@jcu.edu.au
A. Magrach
Ecosystem Management Group, Institute of Terrestrial Ecosystems, ETH Zurich,
Universitätstrasse 16, 8092 Zurich, Switzerland
e-mail: ainhoamagrach@hotmail.com

© Springer International Publishing Switzerland 2015 255


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1_13
256 M. Campbell et al.

et al. 2011; Addo-Fordjour et al. 2012a, b, 2013; Laurance et al. 2001a). Human
impacts on forests will likely shape future patterns of regional liana diversity due
to their broad spatial extent and increasing intensity as the human population
continues to expand (Gerland et al. 2014).
Here we focus on some of the more influential impacts that humans are currently
having on tropical forests and how these affect local liana diversity. The impacts we
will assess include deforestation, forest fragmentation, logging and other silvicul-
tural practices, forest disturbance, climate change and hunting. Finally, we conclude
by examining the implications that these human impacts will likely have on the
liana diversity of future tropical forests.

13.2 Liana Diversity and Anthropogenic Forest


Modifications

13.2.1 Deforestation

Any discussion on tropical closed-canopy forests and the retention of species diver-
sity therein must begin with deforestation, the single biggest threat to tropical bio-
diversity (Dirzo and Raven 2003; Gibson et al. 2011). In fact, approximately half of
the original tropical closed-canopy forest and its constituent biodiversity has already
been lost due to deforestation (Wright 2005). Moreover, tropical forests continue to
be lost at an alarming rate (Asner et al. 2009; Hansen et al. 2008, 2013).
Lianas are highly vulnerable to deforestation as they rely on the trees of the tropi-
cal closed-canopy forests for structural support (Gentry 1991). Additionally, lianas
are vulnerable to deforestation as many liana species are sparsely distributed
(Laurance et al. 2001a; Gentry 1991; Parthasarathy et al. 2004; Santos et al. 2009).
For such rare species, localized extirpation may occur in the heavily degraded land-
scapes and small forest remnants created in the aftermath of deforestation (Laurance
et al. 1999). Deforestation has surely resulted in the local extirpation of many liana
species and any tropical region facing the loss of closed-canopy forests on a broad-
scale are likely to exhibit a concurrent decline in their resident liana diversity.

13.2.2 Forest Fragmentation

Tropical deforestation rarely results in the removal of all pre-existing vegetation in


a given area (Laurance and Bierregaard 1997). The resulting landscape is often a
matrix of isolated forest fragments surrounded by new habitat types (Wilcove et al.
1986). Forest fragmentation is occurring on an immense scale throughout the
tropical regions of the globe with nearly 9,000 fragments of <100 km2 in area
generated from 1999 to 2002 in the Brazilian Amazon alone (Broadbent et al. 2008).
With continuing deforestation, the extent of remnant fragments will continue to
increase (Broadbent et al. 2008; Achard et al. 2002; Wright 2005).
13 Liana Diversity and the Future of Tropical Forests 257

Fig. 13.1 Lianas are significantly more diverse on the edge of forest fragments than in their
interior (Laurance et al. 2001a)

One by-product of forest fragmentation is that it greatly increases the area of


forest edge, which exposes the surviving biota to numerous environmental changes
associated with these edges, such as increased light penetration and desiccation
(Laurance and Yensen 1991; Wilcove et al. 1986; Laurance 2008; Williams-Linera
1990; Briant et al. 2010). Additionally, the forest edge offers lianas an increased
availability of climbing trellises (Chittibabu and Parthasarathy 2001; Putz 1984;
Balfour and Bond 1993; Londre and Schnitzer 2006; Williams-Linera 1990). The
increased desiccation, light and climbing trellises within forest fragments and in
particular at forest edges is often a result of the higher level of disturbance and
tree-turnover found there (Laurance et al. 2011; Laurance and Curran 2008;
Laurance et al. 1997, 2000; Briant et al. 2010). The juxtaposition of ecological and
environmental traits in forest fragments and in particular forest edge favours liana
growth requirements and as a result these areas often have high liana diversity
(Fig. 13.1, Laurance et al. 2001a; Zhu et al. 2004; Mohandass et al. 2014).
Although fragmentation increases liana abundance and diversity on forest edges
it can also lead to a decline in overall species diversity at regional scales. This
decline in regional diversity may occur as a direct consequence of the fragmentation
process itself through mechanisms such as genetic drift, isolation of breeding
populations, propagule-dispersal limitation, and pollination limitation (Hernandez-
Stefanoni 2005; Lienert 2004; Young et al. 1996; Aizen and Feinsinger 1994). These
fragmentation processes are especially likely to impact lianas that are rare (Laurance
et al. 1999), although the extent of the impact of each particular process on lianas
specifically is still poorly understood. However, given that the percentage of rare
liana species within any tropical forest region is not insignificant (Laurance et al.
2001a; Gentry 1991; Parthasarathy et al. 2004; Santos et al. 2009), it is important to
better understand how such rare species are affected by fragmentation.
Liana diversity within forest fragments may also decline as the result of addi-
tional anthropogenic impacts after fragmentation. For instance, heavy disturbance
258 M. Campbell et al.

(see below) within fragments often leads to a considerable decline in intra-fragment


and thus regional liana diversity (Addo-Fordjour et al. 2012b; Muthuramkumar
et al. 2006; Arroyo-Rodriguez and Toledo-Aceves 2009; Mohandass et al. 2014). It
is believed that the decline of liana diversity in heavily disturbed fragments may be
due to a decrease in the availability of structural hosts (trees) for lianas to climb
(Addo-Fordjour et al. 2012b; Arroyo-Rodriguez and Toledo-Aceves 2009; van der
Heijden and Phillips 2008). This loss of potential tree hosts can result from logging,
from tree damage via other anthropogenic resource extraction, or from a progressive
decay in fragment condition over time (Oliveira et al. 1997; Laurance and
Curran 2008; Laurance et al. 2000). Fragment decay can be driven by elevated tree
mortality and turnover from increased wind damage and unfavourable microcli-
matic changes near forest edges (Kapos 1989; Laurance and Curran 2008; Laurance
et al. 1998, 2000, 2006; Williams-Linera 1990) and by edge-related fires (Cochrane
and Laurance 2002, 2008). Consequently, regional liana diversity in areas with a
low to moderate level of fragment disturbance are more likely to retain their liana
diversity than are those suffering heavy disturbance (Mohandass et al. 2014; Addo-
Fordjour et al. 2012b; Arroyo-Rodriguez and Toledo-Aceves 2009; Muthuramkumar
et al. 2006).

13.2.3 Logging and Other Silvicultural Practices

Selective logging has varied impacts on regional liana diversity. Selective logging is
a major driver of forest degradation across the tropics, with ~20 % of all tropical
forests logged between 2000 and 2005 (Asner et al. 2009). Globally, more than 400
million hectares of tropical forest are in logging estates (Blaser et al. 2011). Selective
logging, via the felling of canopy trees, collateral killing of many other trees, and
the creation of logging roads and skid trails (Laporte et al. 2007), modifies the local
microclimate by reducing canopy cover. A loss in canopy cover causes a decline of
many forest-interior specialists and an increase in abundance of disturbance-tolerant
and edge-adapted taxa, including most species of lianas (Arroyo-Rodriguez and
Toledo-Aceves 2009; Ding and Zang 2009; Laurance et al. 2001a; Schnitzer and
Bongers 2011; Schnitzer et al. 2004; Parren and Doumbia 2005). However, the
overall response of regional liana species richness to logging varies considerably
depending upon the intensity and type of logging applied (Gerwing 2006; Fox 1968;
Gerwing and Vidal 2002). Additionally, selective logging of forests has been
suggested to unequally impact individual liana species by favoring lianas of certain
climbing guilds via alteration of the available climbing trellises (Ding and Zang 2009).
The increase in liana abundance in logged forests has received considerable
attention given their effects on important timber trees. Timber trees may suffer
increased tree mortality and stem deformation in forests with high liana abundance
(Putz and Mooney 1991; Ingwell et al. 2010). Additionally, lianas compete with
timber trees for light and nutrients, lowering tree growth rates (Wright et al. 2005;
Schnitzer et al. 2005; Tobin et al. 2012). Consequently, it has been suggested that
cutting lianas might be a cost-effective (but expensive, Perez-Salicrup et al. 2001)
13 Liana Diversity and the Future of Tropical Forests 259

Fig. 13.2 Liana cutting during logging operations may threaten liana diversity and with them their
reliant faunal counterparts

way of improving the value of production forests (Alvira et al. 2004; Schnitzer et al.
2004; Gerwing and Uhl 2002). Additionally, liana cutting has been suggested to
improve seed production by certain timber tree species by 50–100 % and hence
enhance their recruitment (Nabe-Nielsen et al. 2009). Nevertheless, this practice
might considerably reduce the conservation value of logged forests by diminishing
liana diversity and their valuable role in ecosystem functioning (Schnitzer and
Bongers 2002). In particular, the loss of localised liana diversity through deliberate
cutting may impact faunal inhabitants by reducing food resources, arboreal walk-
ways and nesting sites (Fig. 13.2, Ansell et al. 2011; Thorpe et al. 2009; Yanoviak
260 M. Campbell et al.

and Schnitzer 2013; Asensio et al. 2007; Gentry 1991). Such practices can have
long-term effects on liana communities. For example, after pre-logging liana
cutting, the local liana community of one logged forest was found to display a
reduction in animal-dispersed liana species (Gerwing and Vidal 2002) and an
increase in vegetatively reproducing liana species (Gerwing 2006).
Another intervention prescribed by silviculturalists to decrease liana abun-
dance in logged forests is controlled burning. Burning leads to the increased
mortality of lianas compared with trees and consequently is likely to cause a
decrease in liana diversity (Balch et al. 2011; Gerwing 2001). Burning also
causes changes in liana community composition, because burn-induced mortal-
ity varies among different species and size classes of lianas (Balch et al. 2011).
Additionally, burning favors liana species that are able to coppice after fires
(Gerwing 2001). Finally, previously burnt stands have been found to become
more susceptible to further burns (Gerwing 2001; Cochrane et al. 1999) and sub-
sequent burn events may further impact liana diversity and result in a sustained
degradation of the liana community.
A better understanding of both the direct and indirect implications of selec-
tive logging on liana communities is needed in order to determine the direction
and extent of changes to local and regional liana diversity. However, it appears
likely that liana diversity will be maintained within selectively logged forests
(Ding and Zang 2009) but decline in logged forests where lianas are cut or burnt
(Balch et al. 2011; Gerwing and Vidal 2002). Consequently, a balanced perspective
between the negative impacts of lianas on timber production and the positive
impacts of liana diversity on associated ecosystem functioning and faunal support
must be weighed when determining regional forest-usage strategies (Gerwing
and Vidal 2002).

13.2.4 Forest Disturbance

Disturbance of the forest canopy, soil and micro-climatic conditions can occur
through both human-induced and natural environmental processes. Spatially,
these disturbances can range in size from a single treefall to forest-wide distur-
bances such as canopy defoliation following cyclones or hurricanes (Turton
and Siegenthaler 2004). Small-scale forest disturbance often occurs through
treefalls, which usually disturb an area of forest only tens of metres in extent.
However, even though a single treefall may be spatially small, the amount of
treefall disturbance throughout a forest is closely linked with the maintenance of
forest-wide liana diversity (Ledo and Schnitzer 2014; Schnitzer and Bongers
2005; Schnitzer and Carson 2001).
It is believed that abundant treefall gaps maintain liana diversity in forests, as
lianas are proportionately more diverse in treefall gaps than other interior forest
locations (Schnitzer and Carson 2001). Lianas are successful within treefall gaps as
13 Liana Diversity and the Future of Tropical Forests 261

Fig. 13.3 Liana tangles in treefall gaps limit tree regeneration during forest succession and enable
the maintenance of localized liana diversity

they can both colonize gaps in high numbers and survive in gaps for a long period
of time (Schnitzer and Bongers 2005). Lianas take possession of treefall gaps so
successfully because they can colonize the gap through seed and seedling banks
as do trees, survive the treefall event itself and subsequently regrow (Putz 1984) and
colonize the gap via long-distance clonal growth (Yorke et al. 2013; Penalosa 1984).
Once lianas have colonized a treefall gap they can survive there for long periods of
time, which again enhances their localized diversity (Schnitzer and Bongers 2005;
Schnitzer et al. 2000; Oliveira et al. 1997). They are also successful at retaining pos-
session of forest gaps because they often form dense “tangles” in the gap that may
prevent tree-seedling establishment and damage tree saplings (Fig. 13.3, Schnitzer
and Carson 2010). Liana proliferation in treefall gaps may stall tree succession by
increasing tree mortality, which may even change the successional trajectory of the
262 M. Campbell et al.

gap vegetation (Schnitzer and Bongers 2005; Schnitzer et al. 2000). Consequently,
small-scale disturbance of tropical forests through treefalls often leads to the main-
tenance of high regional liana diversity (Dalling et al. 2012; Schnitzer and Carson
2001; Ledo and Schnitzer 2014; Anbarashan and Parthasarathy 2013).
Large-scale disturbance of tropical closed-canopy forests can occur through
anthropogenic processes such as forest fragmentation, climate change (such as
an increase in intensity and frequency of cyclones or droughts) and selective log-
ging. Following the large-scale disturbance of a forest, liana diversity peaks as
the forest reaches an intermediate successional age and then may slowly decline
in mature forest (DeWalt et al. 2000; Letcher and Chazdon 2009). It has been
suggested that the peak in liana diversity in intermediate-aged forest corresponds
with the availability of liana-limiting resources such as light and climbing
trellises within the forest (DeWalt et al. 2000; Putz 1984). In addition, liana
recruitment into forests may increase by up to 500 % after disturbance (Benitez-
Malvido and Martinez-Ramos 2003). However, such recruitment is not even
across all liana species and thus the disturbance may alter local liana-community
composition and possibly reduce overall local liana diversity (Benitez-Malvido
and Martinez-Ramos 2003). Finally, recent studies suggest that excessive distur-
bance of a forest may result in a collapse of local liana diversity (Addo-Fordjour
et al. 2009, 2012a, 2013).
Globally, there is a prodigious amount of tropical forest currently suffering from
myriad anthropogenic disturbances (Achard et al. 2002; Hansen et al. 2008). This is
likely to increase in future as humanity places more demands on the resources of
tropical forests (e.g. Laurance et al. 2014b; Seto et al. 2012; Gerland et al. 2014).
However, disturbed future tropical forests are likely to maintain high regional liana
diversity if these disturbances can be managed so that they are both spatially and
temporally variable and of low to medium intensity.

13.2.5 Climate Change

Long-term changes in liana communities within undisturbed forests are currently


occurring (Laurance et al. 2014a; Phillips et al. 2002; Wright et al. 2004). In particu-
lar, researchers have observed that liana abundance, recruitment, productivity and
biomass have been increasing, at least in the New World tropics (Laurance et al.
2014a; Wright et al. 2004). Three main hypotheses have been proposed to explain
the increase in liana abundance in undisturbed forests:
1. Elevated atmospheric CO2 concentration is enabling increased liana growth and
fecundity (Granados and Korner 2002; Körner 2009);
2. Seasonal droughts (Fu et al. 2013) disproportionately benefit lianas as they are often
more resilient to dry conditions than trees (Schnitzer 2005; Cai et al. 2009), and;
3. Increasing tree mortality and turnover in forests is creating additional treefall
gaps, which favors lianas (Schnitzer et al. 2000; Dalling et al. 2012; Putz 1984).
13 Liana Diversity and the Future of Tropical Forests 263

These mechanisms are not mutually exclusive and it is possible that a combina-
tion of all three is leading to the observed changes (Laurance et al. 2014). However,
recent research comparing leaf-chemical traits in trees and lianas across the tropics
(Asner and Martin 2012) suggests that the systematic differences found between
trees and lianas in leaf nutrient concentrations (in particular N, P and Ca) may act
to favor the growth of lianas under elevated CO2. This finding gives more impor-
tance to CO2 fertilization as a potential major driver of rising liana productivity
(Korner 2004; Granados and Korner 2002; Laurance et al. 2014a). Although the
findings mentioned above specifically relate to liana abundance and productivity
increases, this long-term trend could influence regional liana diversity as well.
If the increase in liana abundance and productivity in response to increasing
atmospheric CO2 concentration is species specific (Condon et al. 1992), the liana
community composition within forests may be altered. As such, studies to deter-
mine, whether and how, long-term changes in liana diversity in undisturbed forests
are occurring in apparent response to climate change and rising atmospheric CO2
concentration should be a priority.
In addition to increasing atmospheric CO2 concentration, a current gradual
increase in El Niño Southern Oscillation (ENSO) intensity and the amplitude of its
oscillations might intensify droughts across many tropical forests, especially those
of the western Pacific (Malhi and Wright 2004; Tudhope et al. 2001; Power et al.
2013). As mentioned above, intensified droughts or extended dry seasons may give
lianas a competitive advantage over trees in tropical forests (Schnitzer 2005;
although see van der Sande et al. 2013; DeWalt et al. 2010). Continued drought
might also contribute to increased tree mortality and turnover, facilitating forest
disturbance and favoring lianas (Ledo and Schnitzer 2014). However, if these
droughts result in excessively depleted deep-soil moisture, liana mortality may
increase (Nepstad et al. 2007), potentially leading to a localized diversity decline.
As well as increased drought-induced liana mortality, future climatic conditions
may lead to the demise of suitable host trees (Laurance et al. 2001b; Nepstad et al.
2007), potentially resulting in negative, indirect consequences for liana diversity, as
host-tree availability has been identified as a major determinant of liana success
(van der Heijden and Phillips 2008; Arroyo-Rodriguez and Toledo-Aceves 2009;
Addo-Fordjour et al. 2012b).
A final future impact of climate change that may result in increased tree turnover
in tropical closed-canopy forests is a predicted increase in the frequency and inten-
sity of cyclones (also known as hurricanes or typhoons) (Elsner et al. 2008; Emanuel
2005; Mendelsohn et al. 2012). These intense events often destroy much of the
canopy cover of affected forests, leading to high-disturbance and high-light
conditions (Turton and Siegenthaler 2004). In addition, forests in cyclonic and
hurricane zones may not reach successional maturity due to high levels of distur-
bance and tree turnover (Whigham et al. 1991). As a consequence of these cyclone
impacts and their associated environmental conditions, liana abundance and diver-
sity may increase at the expense of local tree abundance and diversity (Webb 1958;
Putz 1984; Allen et al. 2005).
264 M. Campbell et al.

13.2.6 Hunting

The unsustainable hunting of vertebrates is one of the more insidious anthropogenic


impacts influencing modern tropical forests, with important indirect impacts on their
resident liana diversity. Unsustainable hunting occurs on a vast scale throughout the
tropics with very few regions not touched by its influence (Corlett 2007; Bennett
et al. 2002; Peres and Palacios 2007; Abernethy et al. 2013). Moreover, hunting
often occurs simultaneously with, and is promoted by, other forest-degrading pro-
cesses such as logging and road construction (Laurance et al. 2009, 2014b). Hunting
is a major threat to targeted faunal communities, whose localized extirpation may
lead to the demise of seed predators and dispersers and thus substantially impact
many plant species (Muller-Landau 2007; Wright et al. 2007). For instance, the
competitive ability of large-seeded liana species may be improved (Wright et al.
2007) due to decreased predation by hunted fauna (however see Peres and Palacios
2007). Additionally, the localized extirpation of animal seed dispersers may favor
wind-dispersed liana species (Fig. 13.4, Wright et al. 2007).

Fig. 13.4 An empty seed pod of the wind-dispersed liana Aristolochia acuminata Lam. after a
successful dispersal event. Hunting is suggested to competitively advantage wind-dispersed liana
species over those dispersed by animals, as wide-scale, unsustainable hunting results in the localized
extirpation of many bird and mammal species throughout the tropics
13 Liana Diversity and the Future of Tropical Forests 265

Neotropical lianas have a disproportionate representation of wind-dispersed


species (Gentry 1991) and appear to be regenerating increasingly well in forests
that have been hunted (Wright et al. 2007). Thus, wind-dispersed lianas are likely to
maintain their diversity and potentially increase their abundance in heavily hunted
Neotropical forest regions. However, the competitive ability of animal-dispersed
lianas in these regions may decline, altering liana community composition. Outside
the Neotropics, in some tropical regions, such as north Queensland (Australia), the
majority of liana species possess fleshy fruits (Research 2010). The lianas in these
regions require animal dispersers and thus may suffer much of the negative conse-
quences of hunting that their fleshy-fruited tree counterparts do (Muller-Landau
2007; Wright et al. 2007). Of course, the degree to which hunting influences seed
dispersal or predation depends on the prevalence of hunting within the particular
region (and is minimal in tropical Australia). As a result, the impact of hunting on
regional liana diversity and community composition will likely be region-specific
and determined by the combination of hunting intensity and the proportion of
animal- and wind-dispersed lianas within the local flora.

13.3 Conclusion

Human impacts on tropical closed-canopy forests are unequivocally modifying


regional liana diversity patterns. Besides substantial liana-diversity declines in
tropical regions experiencing deforestation, human impacts are also modifying
liana diversity in the remaining forests in an often complex manner. To continue
to unravel these patterns we suggest that researchers focus specifically on assessing
liana-diversity responses to each human impact. It appears that the often-overarching
changes in liana abundance and biomass found in many studies to date may
‘drown out’ any fine-scale changes in liana community composition. In fact,
when examined closely, it seems that many of the human impacts on tropical
forests studied to date may be unequally favouring certain guilds or species of
lianas at the expense of others (Table 13.1). Consequently, although liana abun-
dance may remain high in the human-impacted forests of the future, it appears
likely that their community compositions will be altered (Table 13.2). Given the
complexity of ecological interactions within tropical forests, any alteration to
liana community composition is likely to have a variety of flow-on effects for
ecosystem functioning and for fauna that rely on liana resources (Yanoviak and
Schnitzer 2013; Gentry 1991).
266 M. Campbell et al.

Table 13.1 Liana traits influenced by human impacts on tropical closed-canopy forests
Human impact on Liana traits
tropical forest Liana traits favoured disadvantaged Source(s)
Deforestation and Spatially common Spatially rare Laurance et al.
Forest (1999)
Fragmentation
Forest Forest edge specialists/ Forest interior Arroyo-Rodriguez
Fragmentation full sun-tolerant specialists/full and Toledo-Aceves
sun-intolerant (2009) Laurance
et al.(2001a)
Mohandass et al.
(2014)
Zhu et al. (2004)
Forest Desiccation tolerant Desiccation intolerant Cai et al. (2009)
Fragmentation species species Schnitzer (2005)
and climate
change
Burning (logging Fire tolerant Fire intolerant Balch et al. (2011)
management Gerwing (2001)
practice)
Logging Persistent seed bank Non-persistent seed Gerwing (2006)
bank (unless reduced
impact logging),
Re-sprouting/coppicing No re-sprouting/ Gerwing and Vidal
from fallen and coppicing from fallen (2002)
prostrate stems and prostrate stems
Logging (Mid Stem-twining climbers Tendril climbers Ding and Zang
successional (2009)
stage)
Logging (Early Tendril climbers Stem-twining climbers DeWalt et al. (2000)
successional Muthuramkumar and
stage) Parthasarathy (2000)
Forest disturbance Species using stolon/ Species not using Penalosa (1984)
clonal treefall gap stolon/clonal treefall Yorke et al. (2013)
colonization gap colonization
Forest disturbance Disturbance induced Non-disturbance Benitez-Malvido and
germination induced germination Martinez-Ramos
(2003)
Climate change Species more receptive Species less receptive Condon et al. (1992)
to elevated CO2 to elevated CO2 Granados and
Korner (2002)
Körner (2009)
Hunting (seed Large seeded species Small seeded species Wright et al. (2007)
predation)
Hunting (seed Wind dispersal Animal dispersal Wright et al. (2007)
dispersal)
13 Liana Diversity and the Future of Tropical Forests 267

Table 13.2 Human impacts and their effect on regional liana diversity
Human impact on
tropical forest Impact Reason Source(s)
Deforestation Regional liana Direct loss of lianas Laurance et al. (1999)
diversity decline and loss of liana host
trees
Forest fragmentation Maintenance or Increased availability Arroyo-Rodriguez and
proportional increase of forest edge habitat Toledo-Aceves (2009)
in regional liana Laurance et al.
diversity (2001a)
Mohandass et al.
(2014)
Zhu et al. (2004)
Forest fragmentation Regional liana Differential species Benitez-Malvido and
diversity decline and recruitment following Martinez-Ramos
community disturbance (2003)
composition change
Forest fragmentation Regional liana Direct loss of lianas Addo-Fordjour et al.
diversity decline or loss of suitable (2012b)
host trees Arroyo-Rodriguez and
Toledo-Aceves (2009)
Laurance et al. (1999)
van der Heijden and
Phillips (2008)
Logging and Liana community Alteration in Ding and Zang (2009)
associated forest composition change availability of
management climbing trellises
practices
Logging and Regional liana Forest burning and Balch et al. (2011)
associated forest diversity decline and liana cutting Chittibabu and
management community Parthasarathy (2001)
practices composition change Gerwing (2001, 2006)
Gerwing and Vidal
(2002)
Forest disturbance Maintenance or Enhanced resource Anbarashan and
proportional increase availability and Parthasarathy (2013)
in regional liana suitable growing Laurance et al.
diversity conditions (2001a)
Ledo and Schnitzer
(2014)
Malizia and Grau
(2008)
Schnitzer and Bongers
(2002, 2013)
Schnitzer and Carson
(2001)
(continued)
268 M. Campbell et al.

Table 13.2 (continued)


Human impact on
tropical forest Impact Reason Source(s)
Forest disturbance Regional liana Heavy disturbance Addo-Fordjour et al.
diversity decline may result in the (2012b, 2013)
direct damage of
lianas and loss of host
trees
Climate change Liana community Species specific Condon et al. (1992)
composition change response to increased Granados and Korner
atmospheric CO2 (2002)
Körner (2009)
Hunting Liana community Loss of pollinators Wright et al. (2007)
composition change and seed predators
and dispersers

Acknowledgements This research was supported by an ARC Discovery Grant awarded to


WL. AM was funded by an ETH fellowship.

References

Abernethy KA, Coad L, Taylor G, Lee ME, Maisels F (2013) Extent and ecological consequences
of hunting in Central African rainforests in the twenty-first century. Philos Trans R Soc B Biol
Sci 368:1625. doi:10.1098/rstb.2012.0303
Achard F, Eva HD, Stibig H-J, Mayaux P, Gallego J, Richards T, Malingreau J-P (2002)
Determination of deforestation rates of the world’s humid tropical forests. Science
297(5583):999–1002
Addo-Fordjour P, Obeng S, Addo MG, Akyeampong S (2009) Effects of human disturbances and
plant invasion on liana community structure and relationship with trees in the Tinte Bepo forest
reserve, Ghana. For Ecol Manage 258:728–734. doi:10.1016/j.foreco.2009.05.010
Addo-Fordjour P, Rahmad ZB, Amui J, Pinto C, Dwomoh M (2012a) Patterns of liana community
diversity and structure in a tropical rainforest reserve, Ghana: effects of human disturbance. Afr
J Ecol 51:217–227. doi:10.1111/aje.12025
Addo-Fordjour P, Rahmad ZB, Shahrul AMS (2012b) Effects of human disturbance on liana com-
munity diversity and structure in a tropical rainforest, Malaysia: implication for conservation.
J Plant Ecol 5(4):391–399. doi:10.1093/jpe/rts012
Addo-Fordjour P, El Duah P, Agbesi DKK (2013) Factors influencing liana species richness and
structure following anthropogenic disturbance in a tropical forest, Ghana. ISRN Forestry
2013:11. doi:10.1155/2013/920370
Aizen MA, Feinsinger P (1994) Forest fragmentation, pollination, and plant reproduction in a
Chaco Dry Forest, Argentina. Ecology 75(2):330–351. doi:10.2307/1939538
Allen BP, Sharitz RR, Goebel PC (2005) Twelve years post-hurricane liana dynamics in an old-
growth southeastern floodplain forest. For Ecol Manage 218(1–3):259–269. doi:10.1016/j.
foreco.2005.08.021
Alvira D, Putz FE, Fredericksen TS (2004) Liana loads and post-logging liana densities after liana
cutting in a lowland forest in Bolivia. For Ecol Manage 190(1):73–86. doi:10.1016/j.
foreco.2003.10.007
Anbarashan M, Parthasarathy N (2013) Diversity and ecology of lianas in tropical dry evergreen
forests on the Coromandel Coast of India under various disturbance regimes. Flora – morphol-
ogy, distribution, functional ecology of plants 208(1):22–32. doi:10.1016/j.flora.2012.12.004
13 Liana Diversity and the Future of Tropical Forests 269

Ansell FA, Edwards DP, Hamer KC (2011) Rehabilitation of logged rain forests: avifaunal
composition, habitat structure, and implications for biodiversity-friendly REDD+. Biotropica
43(4):504–511. doi:10.1111/j.1744-7429.2010.00725.x
Arroyo-Rodriguez V, Toledo-Aceves T (2009) Impact of landscape spatial pattern on liana
communities in tropical rainforests at Los Tuxtlas, Mexico. Appl Veg Sci 12(3):340–349
Asensio N, Cristobal-Azkarate J, Dias PAD, Vea JJ, Rodriguez-Luna E (2007) Foraging habits of
Alouatta palliata mexicana in three forest fragments. Folia Primatol 78(3):141–153.
doi:10.1159/000099136
Asner GP, Martin RE (2012) Contrasting leaf chemical traits in tropical lianas and trees: implica-
tions for future forest composition. Ecol Lett 15(9):1001–1007
Asner GP, Rudel TK, Aide TM, Defries R, Emerson R (2009) A contemporary assessment of change
in humid tropical forests. Conserv Biol 23(6):1386–1395. doi:10.1111/j.1523-1739.2009.01333.x
Australian Tropical Rainforest Plants Edition 6.1 [online version] (2010) Centre for Australian
National Biodiversity Research. http://www.anbg.gov.au/cpbr/cd-keys/rfk/index.html
Balch JK, Nepstad DC, Curran LM, Brando PM, Portela O, Guilherme P, Reuning-Scherer JD, de
Carvalho O (2011) Size, species, and fire behavior predict tree and liana mortality from experi-
mental burns in the Brazilian Amazon. For Ecol Manage 261(1):68–77. doi:10.1016/j.
foreco.2010.09.029
Balfour DA, Bond WJ (1993) Factors limiting climber distribution and abundance in a southern.
Afr For J Ecol 81(1):93–100
Benitez-Malvido J, Martinez-Ramos M (2003) Impact of forest fragmentation on understory plant
species richness inAmazonia.ConservBiol17(2):389–400.doi:10.1046/j.1523-1739.2003.01120.x
Bennett E, Eves H, Robinson J, Wilkie D (2002) Why is eating bush meat a biodiversity crisis.
Conserv Biol Pract 3:28–29
Blaser J, Sarre A, Poore D, Johnson S (2011) Status of tropical forest management 2011, ITTO
technical series no 38. ITTO, Yokohama
Briant G, Gond V, Laurance SGW (2010) Habitat fragmentation and the desiccation of forest cano-
pies: a case study from eastern Amazonia. Biol Conserv 143(11):2763–2769. doi:http://dx.doi.
org/10.1016/j.biocon.2010.07.024
Broadbent EN, Asner GP, Keller M, Knapp DE, Oliveira PJC, Silva JN (2008) Forest fragmenta-
tion and edge effects from deforestation and selective logging in the Brazilian Amazon. Biol
Conserv 141(7):1745–1757. doi:10.1016/j.biocon.2008.04.024
Cai ZQ, Schnitzer SA, Bongers F (2009) Seasonal differences in leaf-level physiology give lianas
a competitive advantage over trees in a tropical seasonal forest. Oecologia 161(1):25–33.
doi:10.1007/s00442-009-1355-4
Chittibabu CV, Parthasarathy N (2001) Liana diversity and host relationships in a tropical
evergreen forest in the Indian Eastern Ghats. Ecol Res 16(3):519–529
Cochrane MA, Laurance WF (2002) Fire as a large-scale edge effect in Amazonian forests. J Trop
Ecol 18(3):311–325
Cochrane MA, Laurance WF (2008) Synergisms among fire, land use, and climate change in the
Amazon. Ambio 37(7/8):522–527
Cochrane MA, Alencar A, Schulze MD, Souza CM, Nepstad DC, Lefebvre P, Davidson EA (1999)
Positive feedbacks in the fire dynamic of closed canopy tropical forests. Science
284(5421):1832–1835. doi:10.1126/science.284.5421.1832
Condon MA, Sasek TW, Strain BR (1992) Allocation patterns in 2 tropical vines in response to
increased atmospheric CO2. Funct Ecol 6(6):680–685
Corlett RT (2007) The impact of hunting on the Mammalian Fauna of tropical Asian forests.
Biotropica 39(3):292–303. doi:10.1111/j.1744-7429.2007.00271.x
Dalling JW, Schnitzer SA, Baldeck C, Harms KE, John R, Mangan SA, Lobo E, Yavitt JB, Hubbell
SP (2012) Resource-based habitat associations in a neotropical liana community. J Ecol
100(5):1174–1182. doi:10.1111/j.1365-2745.2012.01989.x
DeWalt SJ, Schnitzer S, Denslow JS (2000) Density and diversity of lianas along a chronose-
quence in a central Panamanian lowland forest. J Trop Ecol 16(1):1–19
270 M. Campbell et al.

DeWalt SJ, Schnitzer SA, Chave J, Bongers F, Burnham RJ, Cai ZQ, Chuyong G, Clark DB,
Ewango CEN, Gerwing JJ, Gortaire E, Hart T, Ibarra-Manriquez G, Ickes K, Kenfack D, Macia
MJ, Makana JR, Martinez-Ramos M, Mascaro J, Moses S, Muller-Landau HC, Parren MPE,
Parthasarathy N, Perez-Salicrup DR, Putz FE, Romero-Saltos H, Thomas D (2010) Annual
rainfall and seasonality predict pan-tropical patterns of liana density and basal area. Biotropica
42(3):309–317. doi:10.1111/j.1744-7429.2009.00589.x
Ding Y, Zang R (2009) Effects of logging on the diversity of lianas in a lowland tropical rain forest
in Hainan Island, South China. Biotropica 41(5):618–624. doi:10.1111/j.1744-7429.2009.00515.x
Dirzo R, Raven P (2003) Global state of biodiversity and loss. Ann Rev Environ Resour 28:137–167
Durigon J, Durán SM, Gianoli E (2013) Global distribution of root climbers is positively associated
with precipitation and negatively associated with seasonality. J Trop Ecol 29(4):357–360.
doi:10.1017/S0266467413000308
Elsner JB, Kossin JP, Jagger TH (2008) The increasing intensity of the strongest tropical
cyclones. Nature 455(7209):92–95. doi:http://www.nature.com/nature/journal/v455/n7209/
suppinfo/nature07234_S1.html
Emanuel K (2005) Increasing destructiveness of tropical cyclones over the past 30 years.
Nature 436(7051):686–688. doi:http://www.nature.com/nature/journal/v436/n7051/sup-
pinfo/nature03906_S1.html
Fox JED (1968) Logging damage and the influence of climber cutting prior to logging in the low-
land Dipterocarp forest of Sabah. Malays For 31:326–347
Fu R, Yin L, Li W, Arias PA, Dickinson RE, Huang L, Chakraborty S, Fernandes K, Liebmann B,
Fisher R, Myneni RB (2013) Increased dry-season length over southern Amazonia in recent
decades and its implication for future climate projection. Proc Natl Acad Sci 110(45):18110–
18115. doi:10.1073/pnas.1302584110
Gentry AH (1991) The distribution and evolution of climbing plants. In: Putz FE, Mooney HA
(eds) The biology of vines. Cambridge University Press, Cambridge
Gerland P, Raftery AE, Ševčíková H, Li N, Gu D, Spoorenberg T, Alkema L, Fosdick BK, Chunn
J, Lalic N, Bay G, Buettner T, Heilig GK, Wilmoth J (2014) World population stabilization
unlikely this century. Science. doi:10.1126/science.1257469
Gerwing JJ (2001) Testing liana cutting and controlled burning as silvicultural treatments for a
logged forest in the eastern Amazon. J Appl Ecol 38(6):1264–1276
Gerwing JJ (2006) The influence of reproductive traits on liana abundance 10 years after conven-
tional and reduced-impacts logging in the eastern Brazilian Amazon. For Ecol Manage 221(1–3):
83–90. doi:10.1016/j.foreco.2005.09.008
Gerwing JJ, Uhl C (2002) Pre-logging liana cutting reduces liana regeneration in logging gaps in
the eastern Brazilian Amazon. Ecol Appl 12(6):1642–1651
Gerwing JJ, Vidal E (2002) Changes in liana abundance and species diversity eight years after liana
cutting and logging in an eastern Amazonian forest. Conserv Biol 16(2):544–548
Gianoli E (2004) Evolution of a climbing habit promotes diversification in flowering plants. Proc
R Soc B Biol Sci 271(1552):2011–2015. doi:10.1098/rspb.2004.2827
Gibson L, Lee TM, Koh LP, Brook BW, Gardner TA, Barlow J, Peres CA, Bradshaw CJA, Laurance
WF, Lovejoy TE, Sodhi NS (2011) Primary forests are irreplaceable for sustaining tropical
biodiversity. Nature 478(7369):378–381. doi:http://www.nature.com/nature/journal/v478/
n7369/abs/nature10425.html#supplementary-information
Granados J, Korner C (2002) In deep shade, elevated CO2 increases the vigor of tropical climbing
plants. Glob Chang Biol 8(11):1109–1117
Hansen MC, Stehman SV, Potapov PV, Loveland TR, Townshend JRG, DeFries RS, Pittman KW,
Arunawati B, Stolle F, Steininger MK, Carroll M, DiMiceli C (2008) Humid tropical forest
clearing from 2000 to 2005 quantified by using multitemporal and multiresolution remotely
sensed data. Proc Natl Acad Sci U S A 105(27):9439–9444
Hansen MC, Potapov PV, Moore R, Hancher M, Turubanova SA, Tyukavina A, Thau D, Stehman
SV, Goetz SJ, Loveland TR, Kommareddy A, Egorov A, Chini L, Justice CO, Townshend
JRG (2013) High-resolution global maps of 21st-century forest cover change. Science
342(6160):850–853. doi:10.1126/science.1244693
13 Liana Diversity and the Future of Tropical Forests 271

Hegarty EE, Clifford HT (1991) Climbing angiosperms in the Australian Flora. In: Werren G,
Kershaw P (eds) The rainforest legacy; Australian national rainforests study, vol 2, Flora and
fauna of the rainforests. Australian Government Publishing Services, Canberra
Hernandez-Stefanoni JL (2005) Relationships between landscape patterns and species richness
of trees, shrubs and vines in a tropical forest. Plant Ecol 179(1):53–65. doi:10.1007/
s11258-004-5776-1
Ingwell LL, Wright SJ, Becklund KK, Hubbell SP, Schnitzer SA (2010) The impact of lianas on
10 years of tree growth and mortality on Barro Colorado Island, Panama. J Ecol 98(4):879–
887. doi:10.1111/j.1365-2745.2010.01676.x
Kapos V (1989) Effects of isolation on the water status of forest patches in the Brazilian Amazon.
J Trop Ecol 5(02):173–185. doi:10.1017/S0266467400003448
Korner C (2004) Through enhanced tree dynamics carbon dioxide enrichment may cause tropical
forests to lose carbon. Philos Trans R Soc Lond Ser B Biol Sci 359(1443):493–498. doi:10.1098/
rstb.2003.1429
Körner C (2009) Responses of humid tropical trees to rising CO2. Ann Rev Ecol Evol Syst
40(1):61–79. doi:10.1146/annurev.ecolsys.110308.120217
Laporte NT, Stabach JA, Grosch R, Lin TS, Goetz SJ (2007) Expansion of industrial logging in
Central Africa. Science 316(5830):1451. doi:10.1126/science.1141057
Laurance WF (2008) Theory meets reality: how habitat fragmentation research has transcended island
biogeographic theory. Biol Conserv 141(7):1731–1744. doi:10.1016/j.biocon.2008.05.011
Laurance WF, Bierregaard RO Jr (eds) (1997) Tropical forest remnants: ecology, management, and
conservation of fragmented communities. The University of Chicago Press, Chicago
Laurance WF, Curran TJ (2008) Impacts of wind disturbance on fragmented tropical forests: a
review and synthesis. Austral Ecol 33(4):399–408. doi:10.1111/j.1442-9993.2008.01895.x
Laurance WF, Yensen E (1991) Predicting the impacts of edge effects in fragmented habitats. Biol
Conserv 55(1):77–92. doi:10.1016/0006-3207(91)90006-u
Laurance WF, Laurance SG, Ferreira LV, Rankin-de Merona JM, Gascon C, Lovejoy TE (1997)
Biomass collapse in Amazonian forest fragments. Science 278(5340):1117–1118. doi:10.1126/
science.278.5340.1117
Laurance WF, Ferreira LV, Rankin-de Merona JM, Laurance SG (1998) Rain forest fragmentation
and the dynamics of Amazonian tree communities. Ecology 79(6):2032–2040
Laurance WF, Gascon C, Rankin-de Merona JM (1999) Predicting effects of habitat destruction on
plant communities: a test of a model using. Amazon Trees Ecol Appl 9(2):548–554.
doi:10.1890/1051-0761(1999)009[0548:peohdo]2.0.co;2
Laurance WF, Delamonica P, Laurance SG, Vasconcelos HL, Lovejoy TE (2000) Conservation:
rainforest fragmentation kills big trees. Nature 404(6780):836
Laurance WF, Perez-Salicrup D, Delamonica P, Fearnside PM, D’Angelo S, Jerozolinski A, Pohl
L, Lovejoy TE (2001a) Rain forest fragmentation and the structure of Amazonian liana
communities. Ecology 82(1):105–116
Laurance WF, Williamson GB, Delamonica P, Oliveira A, Lovejoy TE, Gascon C, Pohl L (2001b)
Effects of a strong drought on Amazonian forest fragments and edges. J Trop Ecol 17(6):771–
785. doi:10.1017/S0266467401001596
Laurance WF, Nascimento HEM, Laurance SG, Andrade A, Ribeiro JELS, Giraldo JP, Lovejoy
TE, Condit R, Chave J, Harms KE, D’Angelo S (2006) Rapid decay of tree-community com-
position in Amazonian forest fragments. Proc Natl Acad Sci U S A 103(50):19010–19014
Laurance WF, Goosem M, Laurance SGW (2009) Impacts of roads and linear clearings on tropical
forests. Trends Ecol Evol 24(12):659–669. doi:http://dx.doi.org/10.1016/j.tree.2009.06.009
Laurance WF, Camargo JLC, Luizão RCC, Laurance SG, Pimm SL, Bruna EM, Stouffer PC,
Bruce Williamson G, Benítez-Malvido J, Vasconcelos HL, Van Houtan KS, Zartman CE, Boyle
SA, Didham RK, Andrade A, Lovejoy TE (2011) The fate of Amazonian forest fragments: a
32-year investigation. Biol Conserv 144(1):56–67. doi:10.1016/j.biocon.2010.09.021
Laurance WF, Andrade AS, Magrach A, Camargo J, Campbell M, Fearnside PM, Edwards W,
Valsko JJ, Lovejoy TE, Laurance SG (2014) Apparent environmental synergism drives the
dynamics of Amazonian forest fragments. Ecology 95(11):3018–3026. doi:10.1890/14-0330.1
272 M. Campbell et al.

Laurance WF, Andrade AS, Magrach A, Camargo JLC, Valsko JJ, Campbell M, Fearnside PM,
Edwards W, Lovejoy TE, Laurance SG (2014a) Long-term changes in liana abundance and forest
dynamics in undisturbed Amazonian forests. Ecology 95(6):1604–1611. doi:10.1890/13-1571.1
Laurance WF, Clements GR, Sloan S, O’Connell CS, Mueller ND, Goosem M, Venter O, Edwards
DP, Phalan B, Balmford A, Van Der Ree R, Arrea IB (2014b) A global strategy for road build-
ing. Nature 513(7517):229–232. doi:10.1038/nature13717, http://www.nature.com/nature/
journal/v513/n7517/abs/nature13717.html#supplementary-information
Ledo A, Schnitzer SA (2014) Disturbance and clonal reproduction determine liana distribution and
maintain liana diversity in a tropical forest. Ecology 95(8):2169–2178. doi:10.1890/13-1775.1
Letcher SG, Chazdon RL (2009) Lianas and self-supporting plants during tropical forest succes-
sion. For Ecol Manage 257(10):2150–2156. doi:10.1016/j.foreco.2009.02.028
Lienert J (2004) Habitat fragmentation effects on fitness of plant populations – a review. J Nature
Conserv 12(1):53–72. doi:http://dx.doi.org/10.1016/j.jnc.2003.07.002
Londre RA, Schnitzer SA (2006) The distribution of lianas and their change in abundance in tem-
perate forests over the past 45 years. Ecology 87(12):2973–2978. doi:10.1890/0012-
9658(2006)87[2973:tdolat]2.0.co;2
Malhi Y, Wright J (2004) Spatial patterns and recent trends in the climate of tropical rainforest
regions. Philos Trans R Soc B Biol Sci 359(1443):311–329
Malizia A, Grau HR (2008) Landscape context and microenvironment influences on liana com-
munities within treefall gaps. J Veg Sci 19(5):597–604. doi:10.3170/2008-8-18413
Mendelsohn R, Emanuel K, Chonabayashi S, Bakkensen L (2012) The impact of climate change
on global tropical cyclone damage. Nat Clim Change 2(3):205–209. doi:http://www.nature.
com/nclimate/journal/v2/n3/abs/nclimate1357.html#supplementary-information
Mohandass D, Hughes AC, Campbell M, Davidar P (2014) Effects of patch size on liana diversity
and distributions in the tropical montane evergreen forests of the Nilgiri mountains, southern
India. J Trop Ecol 30:579–590
Muller-Landau HC (2007) Predicting the long-term effects of hunting on plant species composition
and diversity in tropical forests. Biotropica 39(3):372–384. doi:10.1111/j.1744-7429.2007.00290.x
Muthuramkumar S, Parthasarathy N (2000) Alpha diversity of lianas in a tropical evergreen forest
in the Anamalais, Western Ghats, India. Divers Distrib 6(1):1–14. doi:10.2307/2673371
Muthuramkumar S, Ayyappan N, Parthasarathy N, Mudappa D, Raman TRS, Selwyn MA,
Pragasan LA (2006) Plant community structure in tropical rain forest fragments of the Western
Ghats, India. Biotropica 38(2):143–160. doi:10.1111/j.1744-7429.2006.00118.x
Nabe-Nielsen J, Kollmann J, Pena-Claros M (2009) Effects of liana load, tree diameter and distances
between conspecifics on seed production in tropical timber trees. For Ecol Manage
257(3):987– 993. doi:10.1016/j.foreco.2008.10.033
Nepstad DC, Tohver IM, Ray D, Moutinho P, Cardinot G (2007) Mortality of large trees and lianas
following experimental drought in an Amazon forest. Ecology 88(9):2259–2269
Oliveira AT, deMello JM, Scolforo JRS (1997) Effects of past disturbance and edges on tree
community structure and dynamics within a fragment of tropical semideciduous forest in
south-eastern Brazil over a five-year period (1987–1992). Plant Ecol 131(1):45–66
Parren MPE, Doumbia F (2005) Logging and lianas in West Africa. CABI Publishing, Wallingford,
pp 183–201
Parthasarathy N, Muthuranikumar S, Reddy MS (2004) Patterns of liana diversity in tropical evergreen
forests of peninsular India. For Ecol Manage 190(1):15–31. doi:10.1016/j.foreco.2003.10.003
Penalosa J (1984) Basal branching and vegetative spread in two tropical rain forest lianas.
Biotropica 16(1):1–9. doi:10.2307/2387886
Peres CA, Palacios E (2007) Basin-wide effects of game harvest on vertebrate population densities
in Amazonian forests: implications for animal-mediated seed dispersal. Biotropica 39(3):304–
315. doi:10.1111/j.1744-7429.2007.00272.x
Perez-Salicrup DR, Claros A, Guzman R, Licona JC, Ledezma F, Pinard MA, Putz FE (2001) Cost
and efficiency of cutting lianas in a lowland liana forest of Bolivia. Biotropica 33(2):324–329
13 Liana Diversity and the Future of Tropical Forests 273

Phillips OL, Martinez RV, Arroyo L, Baker TR, Killeen T, Lewis SL, Malhi Y, Mendoza AM, Neill
D, Vargas PN, Alexiades M, Ceron C, Di Fiore A, Erwin T, Jardim A, Palacios W, Saldias M,
Vinceti B (2002) Increasing dominance of large lianas in Amazonian forests. Nature
418(6899):770–774. doi:10.1038/nature00926
Power S, Delage F, Chung C, Kociuba G, Keay K (2013) Robust twenty-first-century projections
of El[thinsp]Nino and related precipitation variability. Nature 502(7472):541–545. doi:10.1038/
nature12580
Putz FE (1984) The natural history of lianas on Barro-Colorado island, Panama. Ecology
65(6):1713–1724
Putz FE, Mooney HA (1991) The biology of vines. Cambridge University Press, Cambridge, UK
Santos KD, Kinoshita LS, Rezende AA (2009) Species composition of climbers in seasonal semi-
deciduous forest fragments of Southeastern Brazil. Biota Neotropica 9(4):175–188
Schnitzer SA (2005) A mechanistic explanation for global patterns of liana abundance and distri-
bution. Am Nat 166(2):262–276
Schnitzer SA, Bongers F (2002) The ecology of lianas and their role in forests. Trends Ecol Evol
17(5):223–230. doi:http://dx.doi.org/10.1016/S0169-5347(02)02491-6
Schnitzer SA, Bongers F (2005) Lianas and gap-phase regeneration: implications for forest
dynamics and species diversity. CABI Publishing, Wallingford, pp 59–72
Schnitzer SA, Bongers F (2011) Increasing liana abundance and biomass in tropical forests: emerging
patterns and putative mechanisms. Ecol Lett 14(4):397–406. doi:10.1111/j.1461-0248.2011.01590.x
Schnitzer SA, Carson WP (2001) Treefall gaps and the maintenance of species diversity in a tropi-
cal forest. Ecology 82(4):913–919
Schnitzer SA, Carson WP (2010) Lianas suppress tree regeneration and diversity in treefall gaps.
Ecol Lett 13(7):849–857. doi:10.1111/j.1461-0248.2010.01480.x
Schnitzer SA, Dalling JW, Carson WP (2000) The impact of lianas on tree regeneration in tropical
forest canopy gaps: evidence for an alternative pathway of gap-phase regeneration. J Ecol
88(4):655–666
Schnitzer SA, Parren MPE, Bongers F (2004) Recruitment of lianas into logging gaps and the
effects of pre-harvest climber cutting in a lowland forest in Cameroon. For Ecol Manage
190(1):87–98. doi:10.1016/j.foreco.2003.10.008
Schnitzer SA, Kuzee ME, Bongers F (2005) Disentangling above- and below-ground competition between
lianas and trees in a tropical forest. J Ecol 93(6):1115–1125. doi:10.1111/j.1365-2745.2005.01056.x
Schnitzer S, Bongers F, Wright SJ (2011) Community and ecosystem ramifications of increasing
lianas in neotropical forests. Plant Signal Behav 6(4):598–600
Seto KC, Güneralp B, Hutyra LR (2012) Global forecasts of urban expansion to 2030 and direct
impacts on biodiversity and carbon pools. Proc Natl Acad Sci 109(40):16083–16088.
doi:10.1073/pnas.1211658109
Thorpe SKS, Holder R, Crompton RH (2009) Orangutans employ unique strategies to control
branch flexibility. Proc Natl Acad Sci 106(31):12646–12651. doi:10.1073/pnas.0811537106
Tobin MF, Wright AJ, Mangan SA, Schnitzer SA (2012) Lianas have a greater competitive effect
than trees of similar biomass on tropical canopy trees. Biological Science Faculty Publications
(Ecosphere) 3(2):Publication 2. doi:10.1890/ES11-00322.1
Tudhope AW, Chilcott CP, McCulloch MT, Cook ER, Chappell J, Ellam RM, Lea DW, Lough JM,
Shimmield GB (2001) Variability in the El Niño-southern oscillation through a glacial-
interglacial cycle. Science 291(5508):1511–1517. doi:10.1126/science.1057969
Turton SM, Siegenthaler DT (2004) Immediate impacts of a severe tropical cyclone on the
microclimate of a rain forest canopy in north-east Australia. J Trop Ecol 20(5):583–586
van der Heijden GMF, Phillips OL (2008) What controls liana success in Neotropical forests? Glob
Ecol Biogeogr 17(3):372–383. doi:10.1111/j.1466-8238.2007.00376.x
van der Sande M, Poorter L, Schnitzer S, Markesteijn L (2013) Are lianas more drought-tolerant
than trees? A test for the role of hydraulic architecture and other stem and leaf traits. Oecologia
172(4):961–972. doi:10.1007/s00442-012-2563-x
274 M. Campbell et al.

Webb LJ (1958) Cyclones as an ecological factor in tropical lowland rainforest, North Queensland.
Aust J Bot 6(3):220–230
Whigham DF, Olmsted I, Cano EC, Harmon ME (1991) The impact of Hurricane Gilbert on trees,
litterfall, and woody debris in a dry tropical forest in the Northeastern Yucatan Peninsula.
Biotropica 23(4):434–441. doi:10.2307/2388263
Wilcove DS, McLellan CH, Dobson AP (1986) Habitat fragmentation in the temperate zone. In:
Soule ME (ed) Conservation biology: the science of scarcity and diversity. Sinauer Associates,
Sunderland, pp 237–256
Williams-Linera G (1990) Vegetation structure and environmental conditions of forest edges in
Panama. J Ecol 78(2):356–373
Wright SJ (2005) Tropical forests in a changing environment. Trends Ecol Evol 20(10):553–560.
doi:10.1016/j.tree.2005.07.009
Wright SJ, Calderon O, Hernandez A, Paton S (2004) Are lianas increasing in importance in tropical
forests? A 17-year record from Panama. Ecology 85(2):484–489
Wright SJ, Jaramillo MA, Pavon J, Condit R, Hubbell SP, Foster RB (2005) Reproductive size
thresholds in tropical trees: variation among individuals, species and forests. J Trop Ecol
21(3):307–315. doi:10.2307/4092035
Wright SJ, Hernandez A, Condit R (2007) The bushmeat harvest alters seedling banks by favoring
lianas, large seeds, and seeds dispersed by bats, birds, and wind. Biotropica 39(3):363–371.
doi:10.1111/j.1744-7429.2007.00289.x
Yanoviak SP, Schnitzer S (2013) Functional roles of lianas for forest canopy animals. In: Lowman
M, Devy S, Ganesh T (eds) Tree tops at risk. Springer, New York
Yorke SR, Schnitzer SA, Mascaro J, Letcher SG, Carson WP (2013) Increasing liana abundance
and basal area in a tropical forest: the contribution of long-distance clonal colonization.
Biotropica 45(3):317–324. doi:10.1111/btp.12015
Young A, Boyle T, Brown T (1996) The population genetic consequences of habitat fragmentation
for plants. Trends Ecol Evol 11(10):413–418. doi:10.1016/0169-5347(96)10045-8
Zhu H, Xu ZF, Wang H, Li BG (2004) Tropical rain forest fragmentation and its ecological and
species diversity changes in southern Yunnan. Biodivers Conserv 13(7):1355–1372.
doi:10.1023/B:BIOC.0000019397.98407.c3
Index

A Biotechnological tool, 3
Above ground biomass (AGB), 3, 45, 124, Broken stick model, 115, 119
128, 129, 162, 164, 165, 167, 168,
170, 236
Abrus, 104, 136, 165, 167, 175, 181, 183, 188, C
189, 214, 225, 226 Calamus, 65, 69, 75, 105, 114, 116, 117,
Actinidia, 68, 75, 104, 114 137, 138
Afro-montane forest, 86, 87 Cameroon, 86, 91
Alien species, 212, 242 Capparis, 105, 138, 168, 175
Animal diet, 162 Carbon
Apocynaceae, 20, 22–24, 30, 62–65, accumulation, 49, 51, 153
68–72, 76, 77, 83, 103, 105, 106, cycling, 3, 44–48, 52, 153, 157
108–110, 112, 128, 134, 136–141, dynamics, 49, 152–153
143, 167–169, 186, 215, 216, 225, storage, 3, 28, 43–52, 150
231, 232 Carbon sink, 45, 48, 52
Asian bittersweet, 8 Carbon stock, 3, 44, 46, 47, 49, 50, 52, 176
Asparagus, 73, 75, 104, 114, 116, 137, 167, Cell culture, 188, 189, 191–197
213, 226 Chaco forest, 18–20, 22–26
Atlantic forest, 18–27, 29, 30, 32, 34, Chemical control, 228, 232, 233, 239, 241
36, 38 Chronosequence, 51
Climate change, 3, 124, 212, 227, 256,
262–263, 266, 268
B Climbing mechanism, 1, 2, 11, 23, 30, 32,
Barro Colorado Island (BCI), 151, 154 34, 36, 38, 84, 85, 99, 102, 104–110,
Bauhinia, 33, 65, 75, 104, 105, 112, 119, 125, 128, 130–132, 135,
114–117, 137 136, 180
Bignoniaceae, 22, 23, 26, 31, 64, 65, 78, Climbing plants, 18, 57–78, 82, 126, 151
217, 234 CO2 Concentration, 10, 45, 48, 124,
Bioactive molecules, 183, 184, 186, 188, 262, 263
196, 201 Coleoptera, 176
Biodiversity, 3, 20–22, 29, 46, 93–94, 100, Combretum, 101, 106, 116, 128, 139,
123–145, 162–165, 167, 168, 170, 176, 163–165, 169, 175, 184
237, 239–241, 243, 256 Community composition, 2, 11, 28,
Biological activity, 182, 183, 201 46, 153–154, 260, 263, 265,
Biological control, 217, 227–230, 236, 267, 268
239, 242 Competitive effects, 152–154

© Springer International Publishing Switzerland 2015 275


N. Parthasarathy (ed.), Biodiversity of Lianas, Sustainable Development
and Biodiversity 5, DOI 10.1007/978-3-319-14592-1
276 Index

Conservation, 3, 28–38, 101, 119, 162, 177, F


179–202, 235, 236, 259 Fabaceae, 20, 22–24, 26, 33, 83, 103–107,
Conservation status, 117–118, 120 109–111, 134, 184, 214, 217, 220, 221,
Control measures, 212, 214–230, 234, 224, 225, 231, 233, 234
240–242 Faunal assemblage, 135, 176
Convolvulaceae, 63–65, 69–74, 76, 77, 83, Florivores, 165, 167, 168, 170
104, 107–110, 137, 140, 141, 143, 169, Foliar herbivores, 164, 165, 167, 168, 170, 176
170, 218–220, 225, 231 Forest
Coromandel Coast, 126, 127, 129, 161–177 biomass, 150, 153, 180
canopy, 44, 48, 82, 88, 89, 92, 99, 124,
151, 156, 157, 260
D disturbance, 3, 9, 28, 87–89, 124, 126, 134,
Deforestation, 3, 256, 265–267 256, 260–263, 266–268
Dendrochronology, 7, 12 dynamics, 2, 3, 9, 44, 45, 47, 82, 100, 154
Diaspore, 2, 125, 128, 131 management, 10, 91–93, 267
Dioscorea, 2, 65, 68–70, 72, 106, 114, 116, recovery, 29, 49
117, 140, 165, 169, 175, 217, 226, 228, structure, 7, 18, 125, 149, 150, 162, 176,
233, 242 180, 242
Distribution, 3, 7–13, 18–22, 26, 27, 29, Fragmentation, 28, 44, 45, 49, 87, 256–258,
57–78, 85, 100, 101, 103, 114, 118, 262, 266, 267
124, 125, 134, 136, 150, 154–157, Frugivores, 165, 167, 168, 170
212–231, 240, 255 Functional ecology, 3, 177
Diversity, 1–3, 7–13, 18, 25, 26, 28, 29, Functional traits, 123–145
57–78, 81–94, 99–120, 124–126,
128–129, 135, 149–157, 161–177,
180, 201, 202, 213, 226, 231–233, G
237, 238, 241, 242, Gap dynamics, 28
255–268 Geographical aspects, 17–38
Diversity patterns, 265 Global change, 124
Dominance pattern, 102, 115–117
Dominance-diversity curve, 115, 117, 119
Dominant taxa, 65–66, 68–74 H
Dry season length, 9, 24, 26 Himalaya, 59, 65–70, 76, 100, 118
Human disturbance, 10, 86–89, 94
Human health, 176, 212
E Hunting, 87, 256, 264–266, 268
Eastern Ghats, 118, 126–128, 238 Hymenopterans, 165, 172
Eastern Himalayas, 99–120
Ecological aspects, 17–38
Ecological services, 162, 164–174 I
Economic services, 175 Indo-Malaya, 118
Ecosystem Infestation, 47, 48, 89, 91, 92, 213, 225, 232,
functioning, 125, 126, 150, 212, 239, 259, 234, 236, 239
260, 265 Invasive vines, 212–243
processes, 2, 125, 240 Ipomoea, 2, 61, 69, 71, 108, 141, 169, 218,
services, 3, 161–177, 240 225–228, 231
Endemic taxa, 66, 68–74
English ivy, 8, 10, 12, 234
Environmental factors, 76, 89–91, 100, 134 J
Eurasia, 57–78, 234 Japanese honeysuckle, 8, 10, 232
Evopotranspirational demand, 9 Jasminum, 66, 69, 108, 114, 141, 142,
Explants, 189–192, 196–201 165, 169
Index 277

L O
Lantana camara, 135, 142, 174, 213, 219, Old World forests, 63–64, 77, 126
228, 237–240 Oligarchy, 26, 27
Large vessel elements, 9 Orthoptera, 176
Leaf traits, 125 Over exploitation, 2, 3, 180
Lepidopterans, 2, 135, 162, 165, 172, 176
Liana(s)
abundance, 3, 7–13, 27, 28, 44, 49, 85–94, P
100, 119, 124, 130–132, 153, 157, 164, Pan tropical, 9, 155–156
165, 257, 258, 260, 262, 263, 265 Permanent plots, 11, 24, 46, 47
assemblage, 81–94 Phenology, 135, 151
colonisation, 27 Piper, 65, 69, 143
communities, 7, 10–12, 26, 28, 44, 83, 84, Pollination, 176, 177, 231, 240, 257
103, 124, 162, 175, 177, 260, 262, 263, Pueraria tuberosa, 110, 116, 181, 186,
265, 267, 268 188–190, 194, 195, 198, 201
cutting, 29, 91–94, 152, 153, 157, 259,
260, 267
diversity, 3, 11–12, 25, 26, 58, 72, 82–85, R
87–91, 94, 100, 102–115, 124, 126, Rainfall, 9, 10, 20, 25, 28, 44, 52, 83, 86,
128, 135, 154–156, 161–177, 255–268 90, 94, 100, 124, 126, 134, 135,
Litter production, 236 154–156, 163
Livelihood, 94, 175–177 RAPD markers, 201
Logging, 24, 27–29, 44, 50, 87, 91–93, 232, Recruitment, 2, 28, 44, 46, 48, 51, 82, 88, 89,
256, 258–260, 262, 264, 266, 267 152, 157, 259, 262, 267
Low land tropical forest, 150, 151, 155 Red Data Book, 102, 117, 118
Regeneration, 2, 7, 28, 29, 49–52, 82, 87, 88,
91, 93, 100, 119, 120, 124, 150, 162,
M 180, 188–191, 196–200, 213, 221, 231,
Malay Archipelago, 58, 60, 63, 65, 66, 68, 233, 237, 261
75, 77 Resource values, 175, 214–224, 240
Mean annual precipitation (MAP), 9, 25, Resources, 2, 3, 13, 44, 47, 48, 82, 88, 119,
26, 155 150, 151, 162, 165, 172, 175–177, 180,
Medicinal lianas, 179–202 188, 213, 227, 237, 240, 241, 259,
Mediterranean, 59, 66, 71, 72, 75 262, 265
Micro environment, 213 Rubiaceae, 36, 63–65, 69, 71–73, 77, 109,
Micropropagation, 188–190, 196–202 112, 138, 142, 143, 220
Mikania micrantha, 2, 109, 116, 213, 220,
235–237
Montane sub tropical forest, 100, 101, 113, S
115, 117, 118 Seasonal drought, 3, 9, 47, 157, 262
Seasonally dry tropical forest, 20
Secondary forest, 44, 49–52, 84, 85, 88
N Secondary metabolites, 189, 192–196
Natural disturbance, 10, 124, 154 Seed dispersal, 125, 135, 176, 177, 236,
Negative impacts, 150, 212, 236, 239, 260 265, 266
Neo tropical forest, 26, 126, 150, 153, Seedling bank, 261
157, 265 SEF. See Semi-evergreen forest (SEF)
Niche partitioning, 152, 155 Semi deciduous forest, 18–21, 24–27, 29
Nitrogen deposition, 28, 29 Semi-evergreen forest (SEF), 118, 126–132,
Non timber forest product, 126, 162 134–136, 138, 140, 142, 144
North Africa, 57–78 Sequestration, 2, 52, 100
North-eastern India, 99–120, 236 Shannon index, 102, 103, 113, 114
278 Index

Shoot culture, 190–192, 194 Treefall gaps, 49, 51, 124, 152, 154, 156, 157,
Silviculture, 231 260–262, 266
Simpson index, 102, 103, 113 Tropical dry evergreen forest, 161–177
Southeast Asia, 3, 81–94, 235 Tropical forests, 2, 7, 18, 43–52, 66,
Species 81–94, 100, 123–145, 150, 162,
biology, 235–240 180, 255–268
dominance, 26, 115–117
survival, 177, 180
Strychnos, 34, 111, 144, 151, 163, 164, V
171, 174 Venn diagram, 23
Sub Sahara, 60, 63–65 Vitaceae, 22, 38, 63–65, 68, 69, 72,
Subtropical forest, 17–38, 103, 114, 115, 77, 103–105, 109, 111, 128,
118, 119 134, 136–139, 144, 167, 168,
Succession, 44, 49–52, 236, 237, 239, 261 183, 187, 214
Sustainable development, 3, 176 Vitis, 2, 8, 61, 72, 75, 151, 181, 187, 226
Sustainable utilization, 28–38, 179–202

W
T WEF. See Wet-evergreen forest (WEF)
Taxonomical aspects, 17–38 Western Ghats, 118, 126–128
Taxonomic diversity, 128–129 Wet-evergreen forest (WEF), 126–132,
Temperate forest, 7–13, 18, 82, 100, 101, 134–136, 138, 140, 142, 144
103–115, 117–119, 150 Wild life, 150, 151, 233
Tetrastigma, 65, 111, 114, 116, 144 Wood density, 3, 45, 46, 52, 180
Timber resource, 91, 162, 258, 259 Wood productivity, 45, 47–48
Topography, 90 Woody vines, 7, 11, 13, 99, 112,
Transformed cultures, 195–196, 202 231–234, 242
Tree growth, 3, 44, 45, 47, 49, 51, 52, 150,
152, 258
Tree mortality, 3, 45, 48, 52, 150, 152, 258, Z
261–263 Ziziphus, 128, 145, 171, 174, 175

Potrebbero piacerti anche