Download as pdf or txt
Download as pdf or txt
You are on page 1of 369

MANGROVE ECOLOGY, SILVICULTURE AND CONSERVATION

Mangrove Ecology, Silviculture


and Conservation

by

Peter Saenger
Southern Cross University,
Lismore, Australia

,~

Springer-Science+Business Media, B.V.
A C.I.P. Catalogue record for this book is available from the Library of Congress.

ISBN 978-90-481-6050-1 ISBN 978-94-015-9962-7 (eBook)


DOI 10.1007/978-94-015-9962-7

Printed on acid-free paper

All Rights Reserved


© 2002 Springer Science+Business Media Dordrecht
Originally published by Kluwer Academic Publishers in 2002.
Softcover reprint ofthe hardcover I st edition 2002
No part of this work may be reproduced, stored in a retrieval system, or transmitted
in any form or by any means, electronic, mechanical, photocopying, microfilming, recording
or otherwise, without written permission from the Publisher, with the exception
of any material supplied specifically for the purpose of being entered
and executed on a computer system, for exclusive use by the purchaser of the work.
CONTENTS

Preface IX

1. Introduction: The Mangrove Environment 1

2. The Mangrove Flora 11

2.1 Introduction 11
2.2 The Mangrove Flora 11
2.3 Contemporary Biogeographical Processes 21
2.3.1 Floristic Attenuation with Latitude 21
2.3.2 Floristic Attenuation with Aridity 22
2.3.3 Floristic Changes induced by Natural Upheavals and Human Activities 23
2.4 Contemporary Biogeographic Patterns 25
2.4.1 Vicariants 26
2.4.2 Hybridizations 26
2.4.3 Discontinuities and Endemism 27
2.4.4 AtLantic and Eastern Pacific Biogeogrqaphy 28
2.4.5 Indian Ocean and East African Biogeography 32
2.4.6 Australasian Biogeography 37
2.4.7 Western Pacific Biogeography 39
2.4.8 North-East Asian Biogeography 41
2.5 Towards a Global Biogeographic History 43

3. Adapting to the Mangrove Environment 49

3.1 Introduction 49
3.2 Dealing with High Salt Concentrations 49
3.2.1 Salt Exclusion 50
3.2.2 Salt Extrusion 53
3.2.3 Storage of Salt 57
3.2.4 Succulence 58
3.2.5 Compartmentalization 60
3.2.6 Osmocompensation 62
3.3 Conserving Desalinated Water 63
3.3.1 Xeromorphic Features 63
3.3.2 Transpiration 68
3.3.3 Optimizing Leaf Temperatures 73
3.4 Root Aeration in Waterlogged Soils 75
3.4.1 Morphological Adaptations 75
3.4.2 Physiological Adaptations 81
3.4.3 Metabolic Adaptations 83
3.5 Surviving Seawater Dispersal 84
3.5.1 Vivipary and Cryptovivipary 84
3.5.2 Propagule Production 87
3.5.3 Propagule Dispersal and Establishment 91
VI

4. Physico-chemical Factors and Mangrove Performance 101

4.1 Introduction 101


4.2 Temperature 101
4.2.1 Mangroves and Low Temperatures 102
4.2.2 Responses to High Temperatures 103
4.2.3 Inter- and Intra-specific Differences 104
4.3 Light and Photosynthesis 105
4.3.1 Photosynthetic Efficiency of Mangroves 105
4.3.2. Canopy Gaps: Gap-enhancement of Reproduction and Growth 110
4.3.3 The Mangrove Understorey and the Role of Sulfides 113
4.4 Wind and other Atmospheric Phenomena 114
4.4.1 Ambient Winds 114
4.4.2 Extreme Winds and Mangrove Vegetation 116
4.4.3 Hail 121
4.4.4 Lightning 121
4.5 Sedimentary Conditions and Processes 122
4.5.1 Mangrove Soil Development 122
4.5.2 Aeration and Drainage Properties of Mangrove Soils 128
4.5.3 Nutrients in Mangrove Soils 131
4.6 Salinity of Soil Water 137

5. Biotic Interactions and Mangrove Performance 147

5.1 Plant-Plant Interactions 147


5.1.1 Parasitism 147
5.1.2 Antagonism (Ammensalism) 150
5.1.3 Mutualism 152
5.1.4 Competition 154
5.2 Plant-Animal Interactions 160
5.2.1 Flowering and Pollination 160
5.2.2 Bioturbation of Sediments 166
5.2.3 Grazing and Trampling (Leaf Herbivory) 168
5.2.4 Other Mutualistic Interactions 176
5.2.5 Trophic Pathways 179

6. Mangrove Structure and Classification 183

6.1 Classification of Mangrove Communities 183


6.1.1 Phytosociological Classification 183
6.1.2 Classification Using Structural Attributes 184
6.1.3 Classification Using Geomorphological Settings 187
6.1.4 Classification Using Physiographic and Structural Attributes 190
6.2 Zonation of Mangroves 194
6.2.1 Shoreline Zonation 194
6.2.2 Longitudinal Upriver Zonation 201
6.2.3 Similarities and Differences in Shoreline and Upriver Zonation 204
Vll

7. The Value of Mangroves 207

7.1 Introduction 207


7.2 Components 209
7.2.1 Plant Resources 209
7.2.2 Fisheries Resources 213
7.2.3 Wildlife Resources 21S
7.2.4 Water Supply Resources 21S
7.2.S Agricultural Resources (including Salt Production and Aquaculture) 21S
7.2.6 Forage Resources 21S
7.2.7 Water Transport Resources 216
7.2.8 Recreational Resources 217
7.2.9 Energy Resources 217
7.2.10 Pharmaceutical Resources 218
7.3 Functions 219
7.3.1 Shoreline Protection 219
7.3.2 Windbreak and Stonn Protection 219
7.3.3 Sediment Regulation 219
7.3.4 Nutrient Retention 220
7.3.S Water Quality Maintenance 220
7.3.6 External Support 221
7.3.7 Groundwater Discharge and Recharge 222
7.3.8 Local Microclimatic Stabilization 223
7.4 Attributes 223
7.4.1 Biodiversity 223
7.4.2 Uniqueness and Heritage 224
7.5 An Economic Perspective 225

8. Mangrove Silviculture and Restoration 229

8.1 Introduction 229


8.2 Objectives for Mangrove Planting 229
8.2.1 Setting the Objectives 230
8.2.2 Timber Production 231
8.2.3 Shoreline Protection. Channel Stabilization and Stonn Protection 237
8.2.4 Fisheries and Wildlife Enhancement 238
8.2.S Legislative Compliance 239
8.2.6 Social Enrichment 240
8.2.7 Ecological Restoration 241
8.3 Macropropagation of Mangroves 243
8.3.1 Direct Planting of Propagules collected from the Wild 243
8.3.2 Outplanting of up to One Year old Nursery-raised Propagules 244
8.3.3 Outplanting of small Seedlings after Nursery-raising 244
8.3.4 Direct Transplanting of Seedlings and Shrubs 244
8.3.S Raising ofAir-layered Material 24S
8.3.6 Use of Stem Cuttings 24S
8.3.7 Use of Propagule Segments 246
8.4 Tissue Culture or Micropropagation of Mangroves 247
Vl11

8.5 Site·Species Matching 249


8.5.1 Some General Principles 249
8.5.2 Nursery and Planting Techniques 250
8.5.3 Field Trials 252
8.6 Silvicultural Management of Planted Areas 252
8.6.1 Objectivesfor Management 252
8.6.2 Site Management 253
8.6.3 Natural and Assisted Regeneration 253
8.6.4 Cost of Assisted Regeneration 255
8.7 Plantation Performance 256
8.7.1 Optimal Planting Season 256
8.7.2 Optimal initial Spacing 257
8.7.3 Survival 257
8.7.4 Standing Stock 257
8.7.5 Mean Annual increment 265
8.7.6 Rotation and Thinning Schedules 268
8.8 Indices of 'Health' in Mangrove Communities 269

9. Conservation and Management of Mangroves 271

9.1 The Need for Mangrove Management 271


9.2 Management Frameworks 275
9.3 Some Specific Management Issues 277
9.3.1 Excessive Extractive Use 277
9.3.2 Discharge of Wastes 278
9.3.3 Hydrocarbon Contamination 282
9.3.4 Reclamation and Foreshore Development 286
9.3.5 Mangroves and Global Climate Change 291
9.4 Management Tools 294
9.4.1 The Reserves 294
9.4.2 Zoning of Mangrove Areas 297
9.4.3 Ecosystem Modelling 298
9.4.4 Raising Public Awareness 298
9.5 Future Prospects of Mangroves 299

10. References 303

Index 351
Preface

Historically, there has not been a greater appreciation for natural environments than
exists today. Indeed, most of the emphasis of today's world is on sustainable
development and the re-establishment of natural systems! And, mangroves are no
exception. Considering that mangroves occupy less than I % of the world's surface,
they appear from the publication record to have occupied a 'figure-head' place in the
scientific literature. This trend seems to be continuing.
Personally, I have advocated the use and conservation of mangroves for many
years, and I have often been dismayed by the thoughtless destruction of mangroves.
Believing, as I do, that raising public awareness of mangroves is a major step for
their conservation, this book is intended to inform resource managers and decision-
makers on the capabilities and values of mangroves, and the processes governing
their continued existence. I have included a chapter on silviculture and restoration to
assist with restoring damaged systems, and to provide information on the sustainable
use of mangrove forests.
While preparing the book has been enjoyable, I have experienced some
difficulties which I have dealt with by arbitrary, but consistent, decisions. The first
is the problems of 'units': the sheer multiplicity of units has, at times, been
overwhelming. Even salt concentrations, one of the fundamental parameters of
mangrove ecology, can be expressed in a range of units depending on the particular
interest; salinity (in %0 or without units as required by the Practical Salinity Scale)
is commonly used although it is only an approximate, density-based measure of salt
concentrations. Similarly, conductivity (in S m-I, IlS em-lor mmhos cm- 1) is often
used for easy field measurements and, while it is correlated with salt concentration
and salinity, the linear relationship breaks down at high salinities such as are found
on saltflats. Elsewhere, salt concentration (in g L- 1, mM NaCl, mmol m- 3, or meq L-
1) or osmolality (in mmol kg-1) may be more appropriate. For osmotic pressure
equivalence, salt concentrations are in MPa. An equally diverse array of units is used
for virtually all other measurements. For the most part, I have retained the units of
the original source but have provided more common equivalents for rarely used
units.
Second, I have also had some difficulties with spelling and the use of accents;
in the reference section, all citations are given as they were in the original. In the
body of the text, I have adopted mostly the English form of spelling rather than the
American - hence, centre rather than center, and kilometre rather than kilometer.
Where quotes have been used, their original spelling have been retained except where
the incorrect spelling of taxonomic binomials might lead to confusion. In terms of
place names, I have usually given the original version as reported in the research
report I have drawn on, but where different names are now in use, the altered names
have been given in brackets as close to the first use of the name as possible.
Third, wherever possible, I have tried to summarize information into tables.
Even though this practice breaks the continuity of the text, I have persevered as
many readers may not have access to the necessary library resources to locate the
individual publications. I have also freely cited 'references' to enable the reader to
follow up on any specific points.
Last but not least, I sifted through huge numbers of papers and databases in an
attempt to provide an overall synthesis of the material, only to find that many state
the obvious, repeat what has long been known, or simply rephrase, sometimes with
IX
x

little, novel philosophical twists, what has been previously reported by someone
else. Personally, I know of few measures of information redundancy, but I would be
happy to wager a beer or two, that by most measures, information redundancy in the
mangrove literature is extremely high compared with most other areas of the natural
sciences.
All of these factors have made my task all the harder. Nevertheless, I have tried
to be impartial and I have always tried to give credit for any specific ideas or
advances where that credit was due. I have been conscientious in terms of
acknowledging all authors of ideas, points of view or approaches - and, if I have
not, then, it is entirely accidental. I should also l¥kl that several of my colleagues
have read and commented on earlier versions of the various chapters (see
acknowledgements), but, at the end of the day, the views and comments expressed
here are my responsibility.
Let me make the final comment, that the mangrove literature today is so vast,
and it is becoming increasingly difficult to provide the sort of overview that I have
attempted. More importantly, if I have managed to introduce some questions that
give rise to further detailed investigation or critical review, I will have fulfilled my
role. That has been my approach! I hope that it stimulates and enhances your interest
in mangroves; if it does, my objectives have been met.
Acknowledgements

This work is a compilation of numerous individual investigations carried out


over about three decades. Numerous persons provided individual assistance during
that time and, in particular, I would like to acknowledge Carole Hartman for the
painstaking thin-sectioning of mangrove leaves; Greg Luker for preparing the maps,
Heather Glyde for collecting leaf dimensional data; Max Egan for photographic
microscopy and assistance with image digitizing and processing; Tarek Youssef for
photosynthetic data analysis; Rob Fleetwood for field assistance with the IRGA; and
John Moverley, Don Norman and Rodney Hamilton for their general assistance with
field work, often under trying conditions.
Several colleagues provided useful information, and commented on ideas or
earlier versions of the manuscript, helping to improve it substantially. They include
Marilyn Ball, Fran~ois Blasco, Barry Clough, Colin Field, Fran~ois Fromard, Hal
Heatwole, Nicholas Holmes, David McConchie, Eong Jin Ong, Marianne Popp,
Alison Specht, Brett Stubbs, Leigh Sullivan, and Phil West. I gratefully
acknowledge the editorial assistance provided by Brett Stubbs, and the assistance
with the preparation of the Index from Sonia Weiss. I thank Brian Hutchinson for
his assistance in guiding me through IT problems.
Photographs kindly made available have been acknowledged in the text but I
would particularly like to thank Fran~ois Blasco, John Clark, Norm Duke, J.C.
Lefeuvre, Leigh Sullivan, and Tarek Youssef for allowing me to use their
photographs.
I had financial assistance to pursue my mangrove studies from a variety of
sources as follows (listed in approximate chronological order): Mozambique and
South African mangrove ecology - Rhodes University, Internal Research Grant;
Gulf of Carpentaria mangrove ecology - Broken Hill South Pty. Ltd., Southern
Cross University, Internal Research Grant; Gladstone leafing study - Queensland
Electricity Generating Board; Proserpine leafing study - Condor Oil Shale
Developments; Bangladesh afforestation study - World Bank and FAO; Ballina
leafing study - Southern Cross University, Internal Research Grant; West African
mangrove ecology - UNIDO; leaf heavy metals - Australian Research Council
Research Grant, Brisbane Port Authority, NABALCO and Queensland Aluminium
Pty. Ltd.; mangrove leaf morphology - Southern Cross University, Internal
Research Grant; Arabian Gulf mangrove ecology - UNIDO, Kuwait Institute of
Scientific Research, Saudi Arabian Ministry of Fisheries and Agriculture, and the
Heritage Club of the United Arab Emirates; photosynthetic efficiency of mangroves
- Southern Cross University, Internal Research Grant; mangrove palaeogeography -
France Ministry of Higher Education Research Grant; Kimberley coast mangrove
ecology - W.A. Museum; mangrove genetics and propagation - Australian Centre
for International Agricultural Research.
I would particularly like to acknowledge the assistance of my wife, Helen, who
took notes for me at conferences, spent laborious hours on xerox machines on my
behalf, proof-read the entire manuscript and checked the references, and put up with
my absences but made my presences enjoyable.
I would also like to acknowledge Southern Cross University'S Studies Leave
Program which enable me to pursue my studies in some of the foremost centres for
mangrove research.

Xl
1. Introduction: The Mangrove Environment
In every ... direction you will see the
apparently endless walls of mangrove,
unvarying in colour, unvarying inform,
unvarying in height, save from perspective.
Beneath and between you and them
lie the rotting mud waters of Bonny River, and away
up and down river, miles of rotting mud waters
fringed with walls of rotting mud mangrove-swamp.

Mary Kingsley (1897:96)

Mangroves are the characteristic intertidal plant fonnations of sheltered tropical and
subtropical coastlines. They have been variously described as 'coastal woodlands',
'mangals', 'tidal forests' or 'mangrove forests'. Where conditions are optimal,
mangroves do indeed fonn extensive and productive forests (fig. 1.1). Where
conditions are not optimal, however, dwarfed and scattered shrubs of mangroves may
occur, which fall short of being either 'woodlands' or 'forests'.

Fig. 1.1 Tidal forests at Grand Lahou, Cote d'Ivoire, probably not very different from
those noted by Mary Kingsley in the Bonny River, Nigeria, fringing the waterways
with their proproots and overarching foliage.

Given suitable conditions for growth, propagules of mangrove species colonize


intertidal areas and become established. Species interact among themselves and
respond to environmental conditions, with the result that a characteristic grouping of
species, tenned a 'community', is fonned. Such a community, in combination with
the physico-chemical environment with which it interacts, forms the subject of this
book.
The etymology of the English word 'mangrove' is much disputed. Numerous
suggestions have been made including the old Malay manggi-manggi, the
Senegalese mangue, or the Portuguese and Spanish mangue, manguezal, mangle and
manglares. Others have related the tenn to a composite word, either of native
2 Mangrove Ecology, Silviculture and Conservation

American or African and Portuguese or Spanish origin, or a combination of old


Malay and the Arabic el gurm to mang-gurm, or a combination of the Portuguese
mangle with the English word grove. Unfortunately, we can only speculate on the
etymology of the word, as its origin is obscured by the passage of time.
Moreover, in the English language, 'mangrove' is used in two different ways: it
can refer either to an individual plant or to an assemblage of plants that contains
many species. These two meanings are traditionally used interchangeably; that
tradition, although perhaps initially confusing, is maintained in this book. Where
appropriate, the term 'mangrove community' is used to distinguish the plant
assemblage from the individual mangrove species.
While the term 'mangrove' is well understood generally, it is difficult to define
precisely what constitutes a mangrove species. Mangrove communities comprise
plants belonging to many different genera and families, many of which are not
related closely to one another phylogenetically. Thus, various definitions have been
applied to what does or does not constitute a mangrove plant (Mepham and Mepham
1985). The difficulty arises because mangroves ('mangrove communities') are an
ecological assemblage rather than a taxonomic or a morphological grouping. In
other words, representatives of various plant families have devised means, often
novel, whereby they are able to survive the changing and demanding conditions of
the mangrove habitat and, thus, have become a member of the mangrove
assemblage. Some mangrove species may be able to survive also in other habitats,
whilst other more specialized ones 'have put all of their adaptive eggs into the
mangrove basket'. Some mangrove species do not occur in all mangrove habitats but
are regionally confined. Others (such as Hibiscus tiliaceus and Thespesia populnea)
may be widely present in mangrove habitats yet appear to show no obvious
morphological adaptations for dealing with the demands of the mangrove habitat.
What mangrove species have in common, is a variety of morphological,
physiological, biochemical and reproductive adaptations that enable them to grow in
the particular kind of rather unstable, difficult environment that comprises the
mangrove habitat. On the basis of the common possession of some of these various
adaptations and on their fidelity to the mangrove habitat, approximately 84 species
of plants belonging to about 39 genera in 26 families are recognized throughout the
world as being mangroves (Saenger et al. 1983, Mepham 1983, Blasco 1984,
Tomlinson 1986, Duke 1992, Field 1995, Mastaller 1997, Duke et al. 1998, Naskar
and Mandal1999, Kathiresan and Bingham 2(01).
Different species vary in their dependence on the littoral habitat. Nevertheless,
of the total number of species accepted world-wide as mangrove plants, 63 are only
found in mangrove communities and 21 are important but non-exclusive, extending
beyond the immediate upper tide levels (Saenger et al. 1983). The latter have been
labelled as 'non-exclusive', or 'back' or 'associate' mangroves (Mepham and Mepham
1985, Mastaller 1997). Additionally, Tomlinson (1986) has divided the 'exclusive'
mangroves into 'major' and 'minor' mangroves, depending on the structural
contribution the individual species make; this distinction has not been used here as it
is geographically too variable.
A major difficulty in delimiting mangrove communities is that, because they
lie at the land-sea interface, many of the processes that regulate them have their
origin elsewhere. These external processes, governing water availability, the pool of
available nutrients and the stability of the habitat, often are not seen as part of the
mangroves - and if they are, then the physical boundaries of mangrove communities
Introduction 3

become virtually impossible to define. In view of this, it seems preferable to leave


the delimitation of mangrove communities rather loose.
Another characteristic of mangrove· communities is that tides and coastal
currents bring unremitting variation, where plants and animals must adapt
continuously to the changing chemical, physical and biological characteristics of
their environment. Thus, whilst we speak of mangrove communities we need to
recognize that they differ in space and time from each other. Nevertheless, the
existence of extensive mangrove communities appears to depend on a number of
basic environmental requirements (Walsh 1974, Chapman 1975, 1977), including:
(1) air temperature within a certain range, (2) muddy substrate, (3) protection from
wave action, (4) access to tidal waters, (5) tidal range, (6) ocean currents and (7)
shallow shores. These will be reviewed in turn.
According to Walsh (1974) and Chapman (1975, 1977), extensive mangrove
development occurs only when the average air temperature of the coldest month is
higher than 20°C and where the seasonal range does not exceed 10 degrees. It should
be noted that the world distribution of mangroves (fig. 1.2), particularly at the
northern and southern limits, appears to correlate reasonably well with the 16°C
isotherm for the air temperature of the coldest month (Chapman 1977).

(5)

Fig. 1.2 World distribution of mangroves showing the relationship to the range of 20°C
water temperature isotherms in summer and winter and the bioregions 1 to 6 used to
describe the distributions of individual species (see Table 2.1).

Barth (1982), on the other hand, has shown that an equally good correlation can
be obtained using water temperatures; the presence of mangroves seems to be
correlated with those areas where the water temperature of the warmest month
exceeds 24°C, and their latitudinal limits occur in those waters that never exceed
24°C throughout the year. Despite these generalizations, the occurrences of
mangroves in south-eastern Africa, South America, Australia (south-western
Western Australia and Victoria) and in the North Island of New Zealand appear to be
exceptions, regardless of whether air or sea temperatures are used; these mangroves
are discussed in more detail below (see section 4.2).
Although mangrove communities are able to grow on sand, peat, rock and
coral, the most extensive and luxuriant mangroves are invariably associated with
mud and muddy soils, as noted by Mary Kingsley in her 19th century travels around
tropical West Africa. Such soils are usually found along deltaic coasts, in lagoons,
4 Mangrove Ecology, Silviculture and Conservation

and along estuarine shorelines. Once established on a shoreline, mangroves


themselves may reduce wave action and tidal currents (Brinkman et al. 1997, Mazda
et al. 1997a, b), thereby influencing the rate of deposition and sediment
composition, even accelerating mud accretion on coral islands (Steers 1977) and in
deltas (Furukawa and Wolanski 1996).
Walsh (1974) and Chapman (1975, 1977) argued that protected coastlines are
essential as mangrove communities cannot develop on exposed coasts where wave
action prevents establishment of seedlings. Bays, lagoons, estuaries and shores
behind barrier islands and spits are suitable localities. Nevertheless, mangroves
modify the local wave climate (Mazda et al. 1997a, b) and it is conceivable that in
growing out from a sheltered environment, mangroves can progressively grow into
medium-high energy environments.
Mangroves are not obligate halophytes, although there is evidence that some
species have their optimal growth in the presence of some sodium chloride. Some
species, such as Rhizophora mangle (Werner and Stelzer 1990) and Sonneratia alba
(Ball and Pidsley 1988, 1995) cannot develop fully in freshwater. However, the real
importance of salt lies in the fact that mangroves are generally slow-growing and
therefore they cannot compete with faster-growing species unless those species are
suppressed or eliminated by salt. In this sense, salt is an essential ecological
requirement for mangrove development. Increasingly, high levels of sodium chloride
are viewed as an 'ecological' rather than a 'physiological' requirement.
Coupled with local topography, the tidal range primarily influences the lateral
extent of mangrove development. The greater the tidal range, the more potential
habitat becomes available for mangrove communities. For a given tidal range, steep
shores tend to have narrower mangrove zones than do gently sloping ones. Although
Walsh (1974) and Chapman (1975, 1977) considered tidal range to be important,
there are exceptions. For example, considerable mangrove development occurs on the
microtidal coasts (mean spring range <2 m) in the Caribbean, West Africa and
northern Australia (fig. 1.3).
Exceptionally, natural mangroves have been reported from landlocked, tideless
areas in Australia (Beard 1967), in Madagascar (Weiss 1966), in the Caribbean
(Stoddart et al. 1973, Lugo 1981, Vegas Vilam1bia 2(00) and on some Pacific
islands (Woodroffe 1987, Mueller-Dombois and Fosberg 1998), and as remnants
occurring considerably above tidal limits. For example, Sonneratia caseolaris occurs
in Papua New Guinea 75 m above sealevel (van Steenis 1963). Christmas Island in
the Indian Ocean, with its stepped topography and inland cliffs, rises to a central
plateau at about 250-300 m. The lowest, most recent terrace, was formed around
120,000 y BP (Woodroffe 1988) and has a mangrove community consisting of
Bruguiera gymnorhiza and B. sexangula at its base (van Steenis 1984, Woodroffe
1988). This community occurs about 100 m inland from the present shoreline, and
lies between 24-37 m above sealevel, yet trees reach a height of around 40 m (Du
Puy 1993). A Heritiera littoralis stand also occurs at the base of the first terrace but,
even more remarkably, a stand of the mangrove Cynometra ramijlora occurs on the
upper terrace at an altitude of around 300 m (Du Puy 1993). Similarly, the inland
occurrences of Avicennia germinans (accompanied by such non-halophytes as CmJaba
farinosa, Dichrostachys glomerata and Adansonia digitata) in Senegal, West Africa,
are also apparently relicts of much more widespread mangrove communities dating
back to the Nouakchottian transgression of around 5,700 to 5,300 Y BP (Trochain
and Dulau 1942, Tastet 1981, Lezines 1988, 1996, Gowthorpe and Lamarche 1993).
Introduction 5

Fig. 1.3 Aerial view of mangroves in the south-eastern Gulf of Carpentaria, Australia,
showing extensive mangrove development in an area of microtides «2 m) but with an
extremely flat topography (mean slope of I: 10,000).

For mangroves occurring or planted in atypical habitats, van Steenis (1958,


1984) has summarized information on those apparently growing successfully in non-
tidal, freshwater areas. These include Bogor Botanic Gardens, Indonesia, with an
altitude of around 265 m above sealevel (fig. 1.4).
Clearly, tides play an important role in the functioning of mangrove
communities although they are not an absolute ecophysiological requirement.
However, regular inundation with seawater is probably the means whereby mangrove
soils do not reach excessive salt levels. Salt exclusion by mangrove roots (see
section 3.2.1) is very effective and Passioura et al. (1992) have calculated that this
process could lead to salt build-up in the soil with an assumed salt diffusion
coefficient of 7 x 10-5 m2 dol. From experimental studies, Hollins et al. (2000) found
that the diffusion coefficient is even lower (4.6 x 10-5 m 2 dol) than the assumed
value, thereby strengthening the argument that mangroves may salinize the root
zone. Consequently, the presence of macropores and animal burrows in intertidal
sediments, together with regular tidal flushing (Wolanski and Gardner 1981, Ridd
1996), become critical habitat features to prevent the excessive build-up of salt to
levels at which water uptake could be severely limited.
Favourable currents have also been viewed as essential as they disperse
mangrove propagules and distribute them along coasts. Chapman (1975) noted that
the southern limit of mangroves on the western coast of Africa coincides with the
boundary between a southern cold-water upwelling and warm currents. He noted a
similar situation occurring on the western coasts of Australia and South America.
However, the northern mangrove limit on the western coast of Africa (fig. 1.5) does
not correlate with upwelling but is set by aridity (Saenger and Bellan 1995). The
temperature of currents, particularly their role in influencing the aridity of adjacent
coasts, is undoubtedly of primary importance to mangrove distribution.
Apart from the temperature of cold currents, their direction of flow also affects
mangrove distribution. Chapman (1975) argued that in all cases of western shores in
6 Mangrove Ecology, Silviculture and Conservation

the Southern Hemisphere such currents flow northwards, thereby inhibiting the
southerly drift of floating propagules. Whether such dispersal characteristics of ocean
currents are of significance, however, remains doubtful; the southern limits of
mangroves on the western coast of Africa and South America are certainly limited by
the aridity induced by cold northward-flowing currents (Cliisener and Breckle 1987,
Saenger and Bellan 1995). Even on the west coast of Australia, aridity rather than
dispersal limitations seems to play the major biogeographic role (Saenger and
Moverley 1985, Duke 1992). In any case, ocean currents are likely to be ineffective
in dispersing propagules. Many coastal currents are variable, with some even
reversing direction from time to time, and most mangrove propagules have limited
viability. This, in turn, suggests that oceanic dispersal may not be as significant as
is often assumed (see also 3.5.3). In New Zealand, as elsewhere where coastal current
velocities are low, it is the large distances between suitable habitats that limit
coastal mangrove dispersal (Lange and Lange 1994). It seems that wind-induced drift,
littoral drift and nearshore tidal currents are likely to play major roles.

Fig. 1.4 The mangrove Sonneratia caseolaris was planted in a non-tidal freshwater
pond in the Bogor Botanic Gardens, Indonesia, in the late I 950s. It continues to
survive, grow, flower and fruit although no seedlings are evident.

Mangroves grow in relatively shallow water as seedlings cannot become


anchored in water deeper than 20-30 cm at mean low water. The physical size of
mangroves and their need to have a great proportion of their body above the water,
Introduction 7

but at the same time to be anchored in the soil, makes occupancy of deeper water
impossible. In fact, the relevant water depth and the size of propagules has been used
to explain the commonly seen zonation of mangroves (Rabinowitz 1975,
Tomlinson 1986, Van Spreybroeck 1992), where zones paralleling the shoreline
have been described (see section 6.2.1).
Chapman (1975) maintained that the gentler the shoreline slopes and the more
extensive the shallows, the greater the extent of mangrove development; on steeply
shelving shores, where the zone of shallow water is narrow, only fringe
communities develop.
The features described above seem to be the general prerequisites for mangrove
establishment and development and, although detailed information on the
prerequisites for all the individual mangrove species is lacking, the statement can be
made that if certain conditions prevail, such as a protected shoreline with suitable
climate, muddy substrate and suitable tidal regime, then a mangrove community is
likely to develop, provided, of course, that there is a proximal source of propagules.
Furthermore, this mangrove community will consist of some combination of
characteristic but limited plant species - as relatively few species are able to grow in
the 'mangrove environment' of tropical and subtropical sheltered shorelines.

Fig. 1.5 Sand flats with clumps of Avicennia germinans on the arid coast of
Mauritania, close to the northernmost occurrence of this species in West Africa.
(Photo. J.e. Lefeuvre)

Despite the limitations on species richness, mangrove communities are one of


the most widely distributed and structurally and floristically uniform higher plant
communities around the world (Table 1.1).
The reason for the limited number of species in mangrove communities is
fairly transparent. Any plant capable of surviving and growing in areas of shallow,
tidally influenced areas must be able to tolerate salt at variable concentrations,
withstand any tidal buffeting that it may be subjected to from time to time, and have
seeds that can survive seawater dispersal if the species is to grow and reproduce itself
successfully. Furthermore, if relatively sheltered conditions prevail, the sediments
8 Mangrove Ecology, Silviculture and Conservation

will be enriched by organic matter which, because of waterlogging, will become


anoxic. In tum, anoxia will cause certain breakdown products of the organic matter
to be in a reduced state, resulting in relatively high concentrations of sulfide, ferrous
and manganous ions. These reduced ions are toxic to plants and any plants that grow
in such situations must either be able to avoid these phytotoxins or have some
means to deal with them (Youssef and Saenger 1998a, 1999).
Table 1.1 Continental mangrove areas, together with percentage figures of global totals.

Region Mangrove area (km2)


Ref. 1 Ref. 2 Ref. 3
Asia 75.173 (42%) 76,226 (38%) 51,766 (31%)
Oceania 18,789 (10%) 15,145 (8%) 16,980 (10%)
America 49,096 (27%) 51,286 (26%) 67,446 (40%)
West Africa 27,995 (16%) 49,500 (25%) 27,110 (16%)
East Africa 10,024 (6%) 6,661 (3%) 5,508 (3%)
Total 181,077 198,818 168,810
ISpalding et al. 1997; 2Pisher and Spalding 1993; 3Saenger et al. 1983.

The inherent capacity of higher plants to deal with these adverse characteristics
of the 'mangrove environment' is limited and requires the evolution of salt and
waterlogging tolerance and of seawater seed survival and dispersal. Not surprisingly,
since the Late Cretaceous (-60 my BP) few species have successfully developed
these features to the extent that they can occupy this habitat. Equally, however, we
should not consider mangroves as 'masterpieces' of adaptation: other plants have
developed some, but not all, of the required features. The tolerances of higher plants
are cross-tabulated in fig. 1.6 with the 'nominal' communities in which they occur.
Clearly, mangroves are at one extreme in a range of angiosperm strategies to occupy
available habitats.
To illustrate this point, it is instructive to examine one mangrove, Osbomia
octodonta, a member of the Myrtaceae. This plant family contains numerous genera,
including Eucalyptus with around 600 species and Melaleuca with around 100
species. Based on glasshouse trials of 38 species of Eucalyptus and 20 species of
Melaleuca grown under salinity conditions ranging from around 25 to 45 ppt
(equivalent to 430-770 mol m- 3) with and without waterlogging (van der Moezel et
al. 1991), a relative degree of salt tolerance was exhibited by many of the species in
these genera, including E. angustissima, E. calycogona, E. kondininensis, E.
micranthera, E. yilgamensis, E. halophila, E. sideroxylon, E. tereticomis and
Melaleuca eleuterostachya, M. glome rata, M. halmatuorum, M. lateriflora, M.
stenostachya and M. viridiflora. However, waterlogging in combination with saline
waters resulted in generally low sUrvival rates and very poor growth. Nevertheless,
some species showed a degree of saline waterlogging tolerance including E.
intertexta, E. microtheca, E. occidentalis, E. raveretiana, E. sargentii, E. spathulata,
E. striaticalyx, E. tereticomis, M. lateriflora, M. thyoides, M. halmatuorum, M.
quinquenervia and M. acuminata.
Despite this degree of salinity and waterlogging tolerances in some species of
the Myrtaceae (van der Moezel et al. 1991), only the monospecific genus Osbomia
has developed both sufficiently high salt and waterlogging tolerances to enable it to
grow within the 'mangrove habitat'. Other species, such as Melaleuca quinquenervia
with high waterlogging tolerance and the ability to modify its leaf form depending
on the extent of waterlogging (Lockhart 1996), but with relatively low salinity
Introduction 9

tolerance, frequently occurs in the fringing swamps around mangroves in high


rainfall areas from Australia to as far north as Vietnam. Similarly, Melaleuca
halmatuorum forms the landward fringe of mangroves in many areas of southern
Australia with winter rainfall.

SALT TOLERANCE

HIGH MEDIUM LOW


I
I

,
I
I
I

::c
I
I
I

0
I
Humid-zone I Swamp
Most I

!Il
I
I
IBack Forests &
Mangroves I
I
I
Mangroves Streambanks
Ej I
I
I
,
Z I
I

<: I
I
------------~--------------~-------------
~
,
I
W I
I
....l I
0 I
I
E-<
~
I
Arid-zone I
IMost Swamp
0 ..... Back
I
I
Dune/Strand Margin
Z
...... 0 I
I
I
0 W Mangroves
I
I
IPlants Plants
0 ~ I
I

0 I
I
I
I

~ I
I
------------~--------------~-------------

~ I
I
I
<: I
I
I

~
g
I
I
I
I
Most I
I

Headland
, Most
....l Saltmarsh
I
I
I
Upland
I
I
Plants Plants
Plants I
I
I
I
I
I
I
I
I
I

Fig. 1.6 Conceptual model of the relationship of mangroves to other saline or non-
saline swamp communities.

In other regions, similar examples of 'almost mangroves' can be identified.


Thus, Pterocarpus ojJicinalis (Fabaceae) with its large buttress roots and its abundant
aerenchyma and lenticels (Saur et al. 1998) displays various adaptations to extreme
waterlogging, but does not tolerate salinity above 12%0 (Alvarez-Lopez 1990). It,
together with Raphia taedigera, occupies a similar marginal position to that of
Melaleuca throughout Central America and the Caribbean (Imbert et al. 2000b).
Similarly, Pterocarpus indicus, together with MeIaleuca spp., occupies the
freshwater lowland forests behind the mangroves in south-east Asia (Naskar and
Mandai 1999).
These examples of 'almost mangroves' bring us back to the problem of defining
a mangrove. A plant with low salinity tolerance may occur in the 'mangrove habitat'
in a particular region of high rainfall where salinity never becomes a limiting factor.
For example, Bruguiera gymnorhiza is a dominant mangrove throughout the
10 Mangrove Ecology, Silviculture and Conservation

intertidal zone in Kosrae, Micronesia, which has a mean annual rainfall of around
5,000-6,000 mm (Ewel et al. 1998a). In another region, this same species may be
confined to the landward fringe of the mangrove community because the low or
seasonal rainfall causes soil salinity to build up at times to the point where this
plant is unable to permanently occupy the 'mangrove habitat'.

Fig. 1.7 Repulse Creek. near Proserpine. Queensland. is a subtropical deltaic system in
north-eastern Australia with extensive mangroves abutting 'almost mangrove'
(Me/aleuca quinquenervia) forests. Despite the 1.800 rom mean annual rainfall.
incipient salt flat formation can be seen in the more tidally elevated areas.

Bruguiera gymnorhiza - a 'rr' or just the one 'r'?

In his Botany of Mangroves, Tomlinson (1986) advocated the use of the 'rr' for this
species. However. the International Rules for Botanical Nomenclature do not allow
such stylistic changes simply because they are grammatically or etymologically
preferable. Linnaeus first described Rhizophora gymnorhiza in his Species Plantarum
Vol. I. p. 443. published in 1753. on a type specimen from India. Thus. 'gymnorhiza'
with one 'r' is the basionym. This was changed to Bruguiera gymnorhiza by Savigny in
Lamarck's Encyclopaedia in 1798. again with a single 'r'. Quite separately. Bentham in
his Flora Australiensis Vol. 2. pp. 495. published in 1864. described a Bruguiera
gymnorrhiza from New Guinea - with a 'rr'. This taxon was renamed by Ding Hou in
1957 as Bruguiera exaristata - a distinctive and different species from B. gymnorhiza.
Thus. the use of B. gymnorrhiza (L.) Savigny in Lamk. is incorrect and even B.
gymnorrhiza without its author citation should not be used as it confounds the identity
with B. exaristata - as it did Tomlinson. who states that 'it seems remarkable that this
distinctive species should have remained overlooked until so recently'!

This ongoing dilemma can never be resolved in its entirety and, at some point,
we need to set some arbitrary limits. If a definition must be in place, the following
operational definition, which has been used in this book, might be as complex but
as good as any: 'A mangrove is a higher plant (tree, shrub, palm, herb or fern) which
(1) predominantly grows in the intertidal areas of tropical and subtropical shorelines,
which (2) exhibits a marked degree of tolerance to high salt concentrations and soil
anoxia and, which (3) has propagules able to survive dispersal by seawater.' Bearing
this definition in mind, chapter 2 examines the actual species that have managed to
occupy many of the sheltered tropical and SUbtropical coastal areas.
2. The Mangrove Flora
Dec. 14. 1846:
We crossed a great number of small creeks,
coming from the eastward,
and draining the ridges of the neck of the Peninsula.
Scattered Pandanus and drooping tea-trees grew on their banks
as far as the fresh water extended;
when they were succeeded by the salt-water
tea-tree and the mangrove,
covering andJringing their beds,
which enlarged into stiffplains,
without vegetation,
or into mangrove swamps.
The latter were composed of iEgiceras, Bruguiera, and Pemphis.

Ludwig Leichhardt (1847:528)

2.1 Introdnction
Given the adverse characteristics of the 'mangrove environment', it is not surprising
that relatively few transitions from terrestrial to shallow marine ecosystems have
occurred (Vermeij and Dudley 2(00) and that a numerically small flora has occupied
this habitat. Nevertheless, around 15 transitions have occurred (Ricklefs and Latham
1993, Schwarzbach and Ricklefs 2001) and taxonomic isolation of mangrove species
from non-mangrove sister taxa varies from the level of species within genera
(Heritiera, Excoecaria, Xylocarpus) to the level of tribe (Rhizophoreae in the
Rhizophoraceae) or family, if such monotypic families as Aegialitidaceae,
Avicenniaceae, Pellicieraceae and Nypaceae are accepted.

2.2 The Mangrove Flora


Despite the problems with definitions, there is general agreement that the mangrove
flora is reasonably well delimited. Lists of what are considered to be mangrove
species (classified in some lists into such categories as minor, major, exclusive,
associated and so on) have been widely disseminated. However, some of the earlier
ones (e.g. Watson 1928, Fernando and Pancho 1980) were clearly of a regional
nature. Many later lists were arrived at consensually. Thus, the world-wide listing in
Saenger et al. (1983) was reached after a consensus of 23 mangrove specialists.
Nevertheless, it contained errors and omissions, some of which were corrected by
Tomlinson (1986) who, unfortunately, introduced additional errors and omissions.
Mepham and Mepham (1985), using a broad definition, produced an alternate list
which has not been widely accepted. Duke (1992) provided an updated list which
addressed many of the omissions and this was further augmented by the consensual
list published by Field (1995). The recent listing by Duke et al. (1998) presented an
even more detailed list which has been adopted here, augmented by additional data
only where needed.
A listing of mangroves of the world is presented in Table 2.1. It shows that the
mangrove flora consists of around 39 genera in 26 families. Each of the mangrove-
containing families is reviewed briefly below. More detailed descriptions can be
found in Tomlinson (1986) and Naskar and MandaI (1999). It should be noted,
however, that with the development of molecular genetic techniques, new light is
being shed both on the ancestry and interrelationships of the mangrove species,
genera and their families. For example, large studies based on various genes
12 Mangrove Ecology, Silviculture and Conservation

(Savolainen et al. 2000, Soltis et al. 2000) have proved extremely useful In
depicting interfamilial relationships while others have necessitated redefining family
limits (Steane et aI. 1999, Schwarzbach and Ricklefs 2001). These interfamiliaI
relationships are briefly discussed under the individual mangrove families.
Table 2.1 World distribution of mangrove species in six global biogeographic regions (see fig. 1.2).
together with common upriver location and intertidal positions (based on Fernando and Pancho 1980.
Barth 1982. Saenger et al. 1983. Blasco 1984. Tomlinson 1986. Duke 1992. Field 1995. Mastaller 1997,
Duke et al. 1998. Naskar and Mandai 1999 and Kathiresan and Bingham 2001). Full binomial listings
with author citations are given under the individual mangrove family descriptions in the text. Life form:
F - Fern. S - Shrub. T - Tree. P - Palm; Upriver location: D - downstream. I - intermediate. U -
upstream location; Tidal position: L - low intertidal. M - mid-intertidal. H - high intertidal; Species
occurrence: + present in region. (+) introduced to region.

Genera and Species Biogeographic Regions Upriver Tidal


Location position
2 3 4 5 6 DIU L M H
ACANTHUS (Acanthaceae)
A. ebreeteatus (S) + + + + +
A. ilieifolius (S) + + + + + +
A. volubilis (S) + + + + + +
A. xiamenensis (S) + + +
ACROSTICHUM (Pteridaceae)
A. aureum(F) + + + + + + + +
A. danaeifolium (F) + + + + +
A. speeiosum (F) + + + +
AEGIAUTIS (Plumbaginaceae)
A. annulata (S) + + + + + + +
A. rotundifolia (S) + + + + +
AEGICERAS (Myrsinaceae)
A. eomieulatum (S) + + + + + +
A·floridum(S) + + + + +
AGlAIA «Meliaceae)
A. eueullata (SfT) + + +
ATALANTIA (Rutaceae)
A. eorrea (S) + + +
AVICENNIA Verbenaceae)
A. alba (T) + + + + +
A. bieolor (SfT) + + + +
A. germinans (T) + + + + + + +
A. integra (SfT) + + +
A. marina (T) + + + + + + + +
A. officinalis (T) + + + +
A. rumphiana (T) + + + + +
A. sehaueriana (T) + + + +
BARRINGTONIA (Lecythidaceae)
B. raeemasa (T) + + + + + +
BROWNLOWIA (Tiliaceae)
B. argentea (T) + + + + +
B. tersa (S) + + + +
BRUGUJERA (Rhizophoraceae)
B. cylindriea (T) + + + + +
B. exaristata (SfT) + + + +
B. gymnorhiza (T) + + + + + + +
B. hainesii (T) + + + +
B. parvijlora (T) + + + + +
B. sexangula (T) ( +) + + + + + +
CAMPTOSTEMON «Bombaceae)
c. philippinensis (T) + + +
C. sehultzii (T) + + + + + +
CERBERA (Apocynaceae)
C. manghas (T) + + + +
C. odollam (SfT) + + + +
C.floribunda (S) + + +
The Mangrove Flora 13

Table 2.1 continued


Genera and Species Biogeographic Regions Upriver Tidal
Location position
2 3 4 5 6 D I U L M H
CERIOPS (Rhizophoraceae)
C. australis (Sm + + + +
C. decandra (Srr) + + + + +
C. tagal(Sm + + + + + + +
CLERODENDRUM (Verbenaceae)
c. inerme (S) + + + + + +
CONOCARPUS (Combretaceae)
C. erectus (Srr) + + + + +
CYNOMEI'RA (Caesalpiniaceae)
C. iripa (S) + + + + +
C. ramiflora (S) + + + +
DALBERGIA (Fabaceae)
D. ecastophyllum (S) + + + + +
D. amerimnion (S) + + + + + +
DERRIS (Fabaceae)
D. triJoliata (S) + + + + + + +
D. littoralis (T) + + + + +
DIOSPYROS (Ebenaceae)
D. littoralis (T) + + + + +
DOUCHANDRONE (Bignoniaceae)
D. spathacea (T) + + + +
EXCOECARIA (Euphorbiaceae)
E. agallocha (Srr) + + + + + + + +
E. ovalis (Srr) + + + + + +
HERITlERA (Sterculiaceae)
H.Jomes(n + + + +
H. littoralis (n + + + + + +
HIBISCUS (Malvaceae)
H. tiliaceus (T) + + + + + + + + +
KANDEUA (Rhizophoraceae)
K. candel (Sm + + + +
LAGUNCULARIA (Combretaceae)
La. racemosa (Srr) + + + + + + + +
LUMNI1ZERA (Combretaceae)
L littorea (Sm + + + + +
L racemosa (Srr) + + + + + + + +
MORA (Caesalpiniaceae)
M. oleifera (T) + + + +
NYPA (Arecaceae)
N. fruticans (P) (+) (+) + + + + + +
OSBORNIA (Myrtaceae)
O. octodonta (Srr) + + + + + +
PAVONIA (Malvaceae)
Pa. rhizophorae (S) + + +
Pa. spicata (S) + + + +
PELUCIERA (Pellicieraceae)
Pe. rhizophorae (T) + + + + + +
PEMPHIS (Lythraceae)
P. acidula (Srr) + + + + +
PREMNA (Verbenaceae)
P. obtusifolia (S) + +
RHIZOPHORA (Rhizophoraceae)
R. apiculata (T) + + + + + +
R. harrisonii (T) + + + + + +
R. mangle (Srr) + + + (+) + + +
R. mucronata (T) + + + + + + +
R. racemosa (T) + + + + + +
R. stylosa (Srr) + + + + + +
SCYPHIPHORA (Rubiaceae)
S. hydrophyllacea (S) + + + +
14 Mangrove Ecology, Silviculture and Conservation

Table 2.1 continued

Genera and Species Biogeographic Regions Upriver Tidal


Location position
2 3 4 5 6 D I U L M H
SONNERATIA (Sonneratiaceae)
S. alba (T) + + + + +
S. apetala (T) + + + + +
S. caseolaris (T) + + + + +
S. griffithii (T) + + +
S. hainanensis (T) +
S. lanceolata (T) + + + +
S. ovata (T) + + + +
S. paracaseolaris (T) + + +
TABEBUJA (Bignoniaceae)
T. palustris (S) + + + +
THESPESIA (Malvaceae)
T. papulnea (T) + + + + + + + + +
XYWCARPUS (Meliaceae)
X. granatum (T) + + + + + +
X. moluccensis (T) + + + + + +

ACANTHACEAE
A large family with around 4,000 tropical and subtropical species; the Acanthaceae
is monophyletic with characteristic fruits (Scotland et al. 1995, McDade and Moody
1999). It differs from the Scrophulariaceae by obscure botanical characters such as
the hardening of the stalks of the ovules, which act as a catapult, forcibly ejecting
the four seeds from the dried capsule. Only one genus is represented in the
mangroves; the genus Acanthus has around 30 species, most of which are terrestrial
but with four mangrove species, Acanthus ilicifolius L., A. volubilis Wall., A.
ebrecteatus Vahl. and the recently described Chinese endemic, A. xiamenensis R.T.
Zhang (1985).

APOCYNACEAE
A large pantropical family with around 220 genera and 2,100 species. One genus,
Cerbera contains three mangroves which, although most abundant on beaches and
dunes, are commonly encountered on the landward margins of mangrove
communities. They are Cerbera manghas L. ranging from tropical eastern Africa
through south-east Asia to Australia and the Pacific Islands, and C. odollam Gaertner
with a similar range but which does not include Australia. The third coastal species,
C. floribunda K. Schumann, is endemic in New Guinea (Tomlinson 1986).

ARECACEAE(Palmace~)
This large family of 212 genera and 2,800 species contains the monospecific
mangrove palm, Nypa fruticans (Thunb.) Van Wurmb. A separate family, the
Nypac~, or subfamily, the Nypoideae, has been used for this genus which is quite
distinct from other members of this family (Tomlinson 1986, Gee 2001).

BIGNONIACEAE
A tropical family of trees and lianes with large tubular flowers. Two species are
associated with mangrove communities, including Tabebuia palustris Hemsl., a tree
of the neotropics, and Dolichandrone spathacea (L.) K. Schumann, a tree ranging
from India to Australia, Papua New Guinea and New Caledonia.
The Mangrove Flora 15

BOMBACEAE
Only one mangrove genus, Camptostemon, is recognized, with two species: C.
schultzii Masters from northern Australia and Papua New Guinea and C.
philippinensis (Vidal) Becc. from the Philippines to Borneo.

CAESALPINIACEAE
This family is found throughout the tropical and temperate regions of the world and
differs from the Fabaceae in that the five petals are not joined, the posterior petal is
neither enlarged nor more showy than the rest; and the ten stamens are not fused. It
contains two mangrove genera. The Indo-Australian Cynometra contains two
species, C. iripa Kostel from India to Australia and C. ramijlora L. from India to the
Pacific Islands but not including Australia. The neotropical Mora oleifera (Triana)
Ducke is recorded from mangroves on the Pacific coast of Colombia (Gentry 1982).

COMBRETACEAE
Combretaceae has around 50 genera and 1,000 species of tropical and subtropical
distribution in America, Africa and Australasia. Three mangrove genera are known,
two from the neotropics and one from the Indo-West Pacific. The mangrove species
are Laguncularia racemosa (L.) Gaert., Cononcarpus erectus L., Lumnitzera racemosa
Willd. and Lumnitzera littorea (Jack) Voigt. Hybrids of L. littorea and L. racemosa
have been identified by Tomlinson et al. (1978) as Lumnitzera rosea.

EBENACEAE
This family contains three genera with slightly more than 500 species from the
tropics and sUbtropics of the world. Two genera have limited distributions and few
species in North Africa and South America respectively, and only the genus
Diospyros, with around 500 species, occurs in all continents. One species is
considered to be a mangrove; although there is nomenclatural uncertainty about this
species (Duke 1992), the name Diospyros littoralis (R. Br.) Bakh. is used here.

EUPHORBIACEAE
A cosmopolitan family of small trees and herbs well represented in south-east Asia
and Australia. Although containing around 320 genera with around 8,000 species,
only one large genus, Excoecaria, has mangrove species (fig. 2.1). These include E.
agallocha and E. ovalis, the latter recently elevated to species status on the basis of
morphological (Wightman 1989) and DNA (Maguire and Saenger 2000) data.
Another species, E. indica, now usually referred to as Sapium indica, occurs
from tidal marshes and rivers to around 250 m in evergreen forests from Sri Lanka to
Papua New Guinea. As it is not confined to the mangrove environment, it is not
included here as a mangrove.

FABACEAE
A cosmopolitan family that includes several mangrove shrubs and climbers,
including Dalbergia ecastophyllum (L.) Taub. from the Atlantic coast of Africa, and
Derris trifoliata Lour. from Asia and northern Australia. Another species, D. indica
Bennett, which was formerly known as Pongamia pinnata, is not considered a
mangrove as it occurs at inland locations to an altitude of 500 m.
16 Mangrove Ecology, Silviculture and Conservation

Fig. 2.1 The wide-ranging Excoecaria agallocha (A), with its crenulated leaf margins
and obovate leaves, is easily distinguished from Excoecaria ovalis (B), with its entire
margins and elliptic leaves, which appears to be restricted to northern Australia.
Genetically, they are also quite distinct.

LECYTHIDACEAE
A pantropical family with 20 genera and 450 species. Ba"ingtonia racemosa (L.)
Sprengel occurs in the landward mangrove fringe of the Indo-West Pacific.

LYTHRACEAE
A mainly tropical family with 28 genera and around 660 species. The monotypic
genus Pemphis has one species, Pemphis acidula Forster and G. Forster which is
common on sheltered beaches and in mangroves on sandy substrates. A second but
upland species, Pemphis madagascariensis, endemic to the semi-arid savannas of the
southern half of Madagascar, was known (Tomlinson 1986). Although P.
madagascariensis has been invoked as a sister taxon to P. acidula (Ricklefs and
Latham 1993), recent detailed investigations (Graham et al. 1986: 795) led to the
conclusion that 'all evidence suggests there is no direct relationship between P.
madagascariensis and P. acidula'. This taxon has subsequently been placed into a new
monotypic genus, Koehneria, allied most closely with the neotropical genera
Adenaria and Pehria and the old world genus Woodfordia (Graham et al. 1986).

MALVACEAE
A cosmopolitan family with around 120 genera containing more than 1,500 species
throughout the world. Two large genera, Hibiscus and Thespesia, have some species
which are considered to be commonly associated with the mangrove habitat and
sometimes referred to as 'almost mangroves'. They include Hibiscus tiliaceus L. and
Thespesia populnea (L.) Soland. Ex Correa. Another species commonly associated
with mangroves is Thespesia populneoides (Roxb.) Kostel., now generally reduced
to synonymy with T. populnea. In the neotropics, two species of Pavonia, P.
spicata Cav. and P. rhizophorae Killip., are closely associated with mangroves
throughout tropical America and the coast of Colombia respectively.

MELIACEAE
This large tropical family of 50 genera and around 1,200 species, contains the
mangrove genus Xylocarpus with two mangrove species; according to Mabberley
(1995), following a detailed nomenclatural analysis, the correct names of the two
mangrove species are X. granatum Koenig and X. moluccensis (Lam.) Roemer. In
addition, the genus Aglaia with A. cucullata (Roxb.) Pellegrin can be considered as a
The Mangrove Flora 17

mangrove, occurring on estuarine muds with its abundant aerial roots from the
Sundarbans to Thailand. A second species, A. e/aeagnoidea, has been excluded as it
also occurs in coastal forests to 1,000 m.

MYRSINACEAE
The family is represented by 35 genera and 1,000 species throughout the tropics and
sUbtropics of the world, particularly in South Africa. Australia and New Zealand.
The Myrsinaceae was included within the Ericales based on the chloroplast gene rbcL
(Savolainen et al. 2(00). One genus with two species is represented in the Indo-West
Pacific mangroves: Aegiceras comiculatum (L.) Blanco and Aegiceras floridum
Roem. & Schultes.

MYRTACEAE
This large family of around 155 genera and over 3,000 species occurs mainly in the
southern hemisphere and tropics. Only one species, Osbomia octodonta F. Muell.,
is considered to belong to the mangrove flora of northern Australia and eastern
Malesia. Its position in the Myrtaceae is problematical and it is probably an
isolated, monogeneric group close to the ancestral Myrtoideae stock (Johnson and
Briggs 1981).

PELLICIERACEAE
Formerly included within the Theacaceae, this family contains the monotypic
Pelliciera rhizophorae Triana & Planchon (fig. 2.2).

Fig. 2.2 The monotypic Pelliciera rhizophorae, with its fluted trunks, occurs widely in
high rainfall sites on the Pacific coast of Central America but is less common on the
Caribbean coast.

Although presently only found in abundance from Buenaventura Bay in western


Colombia to Costa Rica, occasional occurrences on the Atlantic coasts may be
relicts of its earlier greater distribution, which included the wider Caribbean as well
as Africa and Europe during the Eocene (Graham 1995, Saenger 1998).
18 Mangrove Ecology, Silviculture and Conservation

PLUMBAGINACEAE
A small family of about 22 genera of perennial herbs and shrubs distributed
throughout the world, mostly in salty or littoral areas. Only one genus, Aegialitis,
is considered to be a mangrove, although the family contains another genus,
Limonium, of widespread occurrence in salt marshes. Aegialitis contains two
species, A. rotundifoZia Roxb. and A. annuZata R. Br. The genus Aegialitis does not
fit well in the Plumbaginaceae and a separate family Aegialitidaceae has been
proposed (Prakasah and Lim 1995).

PTERIDACEAE
A small tropical and warm temperate fern family with 5 genera and around 300
species. Only one genus with three species occurs as an understorey plant in
mangroves of somewhat reduced salinity. Acrostichum aureum L. has a pantropical
distribution while Acrostichum speciosum Willd. occurs in the Indo-West Pacific
and Acrostichum danaeifolium Langdorf & Fischer is confined to the Neotropics.

RHIZOPHORACEAE
This family of 16 genera and 120 species is the richest mangrove family with four
genera and 16 species, exclusive of the recently described Rhizophora annomaIayana
(Kathiresan 1995), now recognized as a probable hybrid of R. mucronata and R.
apicuZata (see 2.4.2).
Based on morphological and anatomical features, the Rhizophoraceae has been
included in the order Myrtales (Takhtajan 1980) or assigned to its own separate order
Rhizophorales (Thome 1992). However, a parsimony analysis of the DNA sequence
data from the chloroplast gene rbcL (Conti et aZ. 1996), strongly suggests that the
Rhizophoraceae does not belong to the Myrtales. A more recent compilation of the
phylogeny of the eudicots by Savolainen et al. (2000) reviewed the phylogeny of
589 rbcL sequences from 308 families including some of those containing mangrove
species. The study revealed that Rhizophoraceae and Erythroxylaceae form a distinct
subclade within the Malpighiales (this order also includes Humiaceae, Euphorbiaceae
and Salicaceae to name a few). The Malpighiales are a sister group to the Oxalidales
which includes families such as Elaeocarpaceae and Cunoniaceae (Savolainen et al.
2000).
More specific studies of the Rhizophoraceae, based on chloroplast DNA
(Setogushi et al. 1999, Schwarzbach and Ricklefs 2000) has confirmed that the
Erythroxylaceae within the Malpighiales is the sister group of the Rhizophoraceae.
Using a combination of data from chloroplast DNA, nuclear ribosome DNA, and
morphology, the Rhizophoraceae was shown to consist of three monophyletic tribes
with the Gynotrocheae being the sister group of the highly derived mangrove clade,
the Rhizophoreae (Schwarzbach and Ricklefs 2000). Interestingly, within the tribe
Rhizophoreae, the genus Bruguiera occupies the most isolated position in
comparison to the other genera (Kandelia, Ceriops and Rhizophora) in this mangrove
group. This suggests a possible ancestral origin for this genus, and supports
morphological observations on dispersal mechanisms. In Bruguiera the seedling
disperses initially with the fruit, whereas in the other more advanced genera only the
seedling disperses.
The mangrove species of this family include Bruguiera cy/indrica (L.) Blume,
Bruguiera exaristata Ding Hou, Bruguiera gymnorhiza (L.) Sav., Bruguiera hainesii
The Mangrove Flora 19

C.G. Rogers, Bruguiera parviflora (Roxb.) Wight & Arn., Bruguiera sexangula
(Lour.) Poir., Ceriops australis (C.T. White) E.R. BaIlment, T.J. Smith & J.A.
Stoddart, Ceriops decandra (Griff.) Ding Hou, Ceriops tagal (Perr.) C.B. Robinson,
Kandelia candel (L.) Druce, Rhizophora apiculata Blume, Rhizophora harrisonii
Leechman, Rhizophora mangle L., Rhizophora mucronata Lam., Rhizophora
racemosa Meyer and Rhizophora stylosa Griff. Tomlinson (1986) has regarded
Rhizophora samoensis (Hochr.) Salvoza as a geographical outlier of R. mangle.

RUBIACEAE
This is one of the largest plant families, comprising some 500 genera and 6,000
species of tropical and subtropical plants. Only one old world mangrove species,
Scyphiphora hydrophyllacea Gaertn., is recognized, occurring from India. through
south-east Asia to the Philippines, south to Australia and east as far as the Solomon
Islands and Palau.

Manila, the 'Distinguished and Ever Loyal City'

The national capital of the Republic of the Philippines spreads along the eastern shore
of Manila Bay, a large inlet with ocean access through a channel to the southwest. It
occupies the narrow deltaic plain of the Pasig River along whose banks the mangrove
Scyphiphora hydrophyllacea was once common. Known as the nilad plant, the locality
was named Maynilad which was subsequently shortened to Maynila and then to its
present form. Although originally a walled Muslim settlement, Spanish conquistadors
under Miguel LOpez de Legazpi entered the mouth of the Pasig River in 1571,
destroyed the existing settlement and founded a fortress city. The Augustinian friars
who accompanied the expedition, established Roman Catholicism and, by 1574,
Manila was decreed jointly by the Vatican and Spain as the 'Distinguished and Ever
Loyal City'. Interestingly, the Vatican through Pope Alexander VI had by the Treaty
of Tordesillas of 1494 divided the world into Spanish and Portuguese hemispheres and,
even by the revised line of demarcation agreed to by Pope Julius II in 1506, Manila
was well within the Portuguese realm! More interestingly, that line of demarcation still
forms the eastern boundary of the Australian state of Western Australia.

RUTACEAE
The monotypic genus Atalantia contains the mangrove A. correa Roem.
(synonymous with Merope angulata and Atalantia monophylla) which was first
collected and described by G.E. Rumphius in the East Indies (Ambon) as Limonellus
angulosus. Ranging from Papua New Guinea, Indonesia including Irian Jaya,
Borneo, Malaysia, Thailand and Vietnam, westwards to Burma and the Sundarbans,
this salt-tolerant plant is found in small populations in mangrove swamp forests and
in related brackish water habitats, often associated with Nypa fruticans (Macintosh
1987, Naskar and MandalI999).

SONNERATIACEAE
This family of two genera contains the mangrove genus Sonneratia whose
Australian species have recently been revised by Duke and Jackes (1987). In addition,
two endemic species have been described from Hainan Island, China (Ko 1985,
1993). Thus, the mangroves of this family include Sonneratia alba J. Smith,
Sonneratia apetala Buch.-Ham., Sonneratia caseolaris (L.) Engler, Sonneratia
griffithii Kurz, Sonneratia hainanensis Ko, E.Y. Chen and W.Y. Chen, Sonneratia
lanceolata Blume, Sonneratia ovata Backer and Sonneratia paracaseolaris Ko, E.Y.
Chen and W.Y. Chen.
20 Mangrove Ecology, Silviculture and Conservation

STERCULIACEAE
This large family of 65 genera and about 1,100 species is found throughout the
tropical and subtropical zones of the world. The large genus Heritiera has two
mangrove species, H. littoralis Dryand. and H. fomes Buch.-Ham.; a third species,
Heritiera globosa Kostermans is an 'almost mangrove' which is only known from
the upper reaches of rivers in Borneo, where the water is always fresh (Kostermans
1959).
H. littoralis is widely distributed from eastern tropical Africa to south-east
Asia, Australia and to the Pacific while H. fomes occurs from India and Bangladesh
to Myanmar (Burma). Random amplified polymorphic DNA (RAPD) studies of
these two mangrove species have shown them to be closely related to each other,
with a high level of intraspecific polymorphism (Das et al. 2001).

TILIACEAE
Containing around 50 genera and 400 species, this family is chiefly tropical. The
genus Brownlowia contains about 30 species which are widely distributed in south-
east Asia. Two species are considered as mangroves as they occur on the soft mud of
estuarine rivers and creeks. They include Brownlowia tersa (L.) Kosterman which
ranges from Malaysia to the Solomon Islands, and B. argentea Kurz which is found
from Malaysia north to the Philippines (Tomlinson 1986, Aksornkoae et al. 1992).

VERBENACEAE
This large family has 75 genera and 3,000 species, mainly in the tropics and
sUbtropics. The genus Avieennia has been placed in the Verbenaceae or, more
recently, within a monotypic family, Avicenniaceae. However, on the basis of a
study of the chloroplast gene rbeL (Hanagata et al. 1999, Savolainen et al. 2000),
the Avicenniaceae was found to form a distinct subclade with the Acanthaceae,
Pedaliaceae and Gesneriaceae within the Lamiales. The taxonomic placement within
the Acanthaceae has been confirmed by the analysis of chloroplast and nuclear
ribosomal DNA sequences (Schwarzbach and Ricklefs 2001, Schwarzbach and
McDade in press).
The genus Avieennia contains 8 species, including Avieennia alba Blume,
Avieennia bieolor Standley, Avieennia germinans (L.) Steam, Avieennia integra
N.C. Duke, Avieennia marina (Forsk.) Vierh., Avieennia officinalis L., Avieennia
rumphiana Hallierf. and Avieennia sehaueriana Stapf and Leechman. Duke (1991b)
has recently revised the Australasian species while Everett (1994) has formalized the
subspecies within A vieennia marina.
The genus Clerodendrum is a large, pantropical genus with 400-500 species
which, until recently, was also considered to be part of the Verbenaceae, but which
is now classified within the Larniaceae. In a study of ribosomal and chloroplast
DNA sequences, the genus was shown to be paraphyletic and divided into four
clades. C. inerme fell within Clade ill (Steane et al. 1999), which included several
other species that inhabit coastal areas of central America (C. aeuleatum),
Australasia (c. inerme) and Africa (c. aeerbianum, C. eriophylium, C. glabrum and
C. aff. humbertii). All the species of Clade III might be worth investigating for their
affinity to the mangrove environment. At present, however, only Clerodendrum
inerme (L.) Gaertn. is considered as a mangrove.
The genus Premna contains one species, P. obtusifolia R. Br., which may be
considered to be a mangrove of the landward margins, occupying extensive stands
The Mangrove Flora 21

near the high tide levels in mangroves from China in the north, to northern
Australia, extending westwards as far as India.

Fig. 2.3 The scrambling shrub mangrove, Clerodendrum inerme, occurs from the
coastal areas of India to north-eastern Australia and to the Pacific Islands. Throughout
its range it has been used by coastal people as a pOUltice for skin disease and spear
wounds.

2.3 Contemporary Biogeographical Processes


A much debated question concerning mangroves is where they first arose and how
they got where they are today (Duke 1995, Ellison et al. 1999). Whatever the exact
historical origin(s) and dispersal routes of mangroves, the present distributions of
mangroves show several features reflecting some of the more modern processes
described below (Saenger 1998). These modern constraints are manifested as a
reduction of species with increasing latitude and/or aridity on the one hand, and
selective removal or loss of species by natural upheavals or human activity on the
other.

2.3.1 Floristic Attenuation with Latitude


The latitudinal limits of mangrove vegetation on each of the major landmasses
(Table 2.2 and fig. 1.2) show that these limits are quite variable. They are, however,
broadly related to temperature and aridity.
Table 2.2 Latitudinal limits of mangrove vegetation on major land masses. (Adapted from Saenger
1998)

Continental land mass Northern limit Southern limit


Atlantic America 32° 20' 28° 56'
Pacific America 30° 15' 5°30'
Atlantic Africa 19° 50' 12° 20'
Eastern Africa/Red Sea 28° 24' 32° 59'
Western Australia 33° 16'
Eastern Australia 38° 45'
Pacific Asia 31 ° 21'
22 Mangrove Ecology, Silviculture and Conservation

It should also be noted that there is a gradual reduction of mangrove


development (e.g. height and extent of mangrove vegetation) with latitude, and that
an attenuation of species with increasing latitude precedes these latitudinal limits
(Lot et al. 1975, Saenger and Snedaker 1993). By way of example, the southward
attenuation of species on the non-arid east coasts of Africa and Australia is shown in
Table 2.3.
Table 2.3 Southern latitudinal limits of shared mangrove species on the eastern coasts of Australia and
Africa. (Adapted from Saenger 1998)

Species E. African E. Australian


Southern Southern
Limit Limit
Sormeratia alba 23° 55' 22° 30'
Pemphis acidula 24° 18' 15° 42'
Thespesia populnea 26° 01' 27° 30'
Xylocarpus granatum 26° 01' 25° 30'
Xylocarpus moluccensis 26° 33' 23° 50'
Heritiera littoralis 26° 33' 22° 15'
Ceriops tagal 26° 50' 28° 11'
Lumnitzera racemosa 26° 50' 27° 30'
Acrostichum aureum 28° 58' 26° 05'
Barringtonia racemosa 30° 44' 17° 21'
Rhizophora mucronata 31° 42' 18° 25'
Bruguiera gymnorhiza 32° 14' 29° 25'
Hibiscus tiliaceus 32° 16' 31 ° 27'
Avicennia morina 32° 59' 38° 45'

This comparison indicates that on coasts where mangroves are not limited by
aridity, temperature is the major factor in reducing number of species with latitude
(Saenger and Moverley 1985, Smith and Duke 1987). As temperature is more or less
correlated with latitude, a similar sequence of species loss with increasing latitude
should be discernible. While there is some intra-specific variability between the two
continents, the data support the suggestion that, in the presence of adequate rainfall,
latitude as a proxy for temperature is related to the southern limits of mangrove
species distribution.

2.3.2 Floristic Attenuation with Aridity


The structure and diversity of mangroves decline with increasing salinity and aridity
which affect the availability of water at the roots and the evaporative demand at the
leaves (Ball and Sobrado 1998). As shown in Table 2.2, the latitudinal limits of
mangroves on the West African and South American Pacific coasts are unexpectedly
curtailed. These limits coincide with the limits of arid regions (UNESCO 1979)
defined as: summer rainfall and winter drought, 12 months/year with <30mm
rainfall, and a precipitation to potential evapotranspiration ratio (PIPEt) <0.03. This
suggests that the mangrove distribution on western coasts of southern continents is
generally more limited by aridity than by temperature (Saenger and Moverley 1985,
CHisener and Breckle 1987, Smith and Duke 1987).
Similar considerations are also relevant to parts of western Asia and the Middle
East. Another interesting example of aridity-induced species attenuation (but with
decreasing latitude) is provided by the eastern African-Arabian Peninsula region
(Saenger 1998). The extreme aridity of the 'Hom of Africa' (northern Somalia)
causes an abrupt loss of species and only Avicennia, Rhizophora, Bruguiera and
Ceriops extend north to the southern Red Sea.
The Mangrove Flora 23

2.3.3 Floristic Changes induced by Natural Upheavals and Human Activities


Enormous natural changes have affected the mangroves through historical time.
Probably the most dramatic changes have occurred in the distribution of the
mangrove genera Pelliciera and Nypa. both of which have undergone severe
contractions since the early Miocene (Graham 1995. Saenger 1998, Ellison 1999,
Gee 2001). Equally, however, there has been the sudden formation and subsequent
decline of 'giant mangrove swamps' in northern Australia due to continued estuarine
infilling about 6,000 y BP (Woodroffe et al. 1985) until their transition to
freshwater wetlands around 1,300 y BP (Clark and Guppy 1988). Similarly, over the
past 12,000 years there has been a large reduction of Rhizophora in the swamps of
the Panama isthmus as a result of changing sealevels, with Rhizophora disappearing
from Gatun Lake 4,200 y BP after dominating the area for the previous 30,000 years
(Bartlett and Barghoorn 1973). In French Guiana, a similarly large reduction in
Rhizophora occurred approximately 3,500 y BP with stabilizing sealevels (Tissot et
al. 1988).
By way of contrast, in the northern part of Tribuga Gulf, on the Pacific coast of
Colombia, local subsidence due to tectonic faulting (-2,600 y BP) has allowed
mangroves (Pelliciera rhizophorae and Rhizophora mangle) to invade local tropical
lowland rainforests (Jaramillo and Bayona 2000). Similarly, Lezine (1986) presents
data from two cores 7.5 m deep from Lake Guiers in Senegal, at 16°N, showing
extensive mangrove development around the lake area at around 6,000 y BP. The
palaeogeographic changes are contemporaneous with an increased flow of a tributary
of the Senegal River (the Ferlo River) between 6,000 and 2,000 y BP. After 2,000 y
BP, there was a sudden onset, without any transitional stages, of modem semi-arid
conditions and a concomitant decrease in mangrove areas.
In addition to natural changes in mangrove vegetation, human activities have
significantly modified the mangrove vegetation in various parts of the world on a
scale at least equivalent to some of the natural changes (Ellison and Farnsworth
1996, Blasco et al. 2001). In particular, these human activities include species
introductions and habitat modification, as well as the more recent phenomena of
mangrove afforestation, often accompanied by species introductions.
As with other plant groups, mangroves have been deliberately introduced into
areas beyond their natural distributional ranges to meet human needs or aspirations.
While some of these introductions are well documented, others are not and are likely
to confound apparent distributional ranges.
By way of example, the present West African populations of Nypa fruticans
were introduced to Calabar in 1906 and Oron in 1912 (Wilcox 1985) from the
Singapore Botanic Gardens, even though this species had occurred throughout the
Niger delta until 25 my BP (Sowunmi 1986). To date, this species has spread into
the Niger, Imo, Bonny and Cross Rivers and its rate of spread is perceiVed to have
been accelerating over recent years. It has now reached the Wouri Estuary in
Cameroon where local villagers, who value its thatching properties (Din 1991),
facilitate its dispersal. In Nigeria, various non-governmental organizations (NGOs)
are currently trying to reduce its spread while in other areas, such as the Volta River
delta, Ghana, it has been recommended for afforestation purposes. Introduced stands
of Nypa jruticans have also recently been recorded from the Atlantic coast of
Panama, where a similar spread of the species is predicted (Duke 1991a).
Even though Pelliciera fossils indicate its presence in the Caribbean, Europe
and West Africa during the Eocene to Oligocene (Fuchs 1970, Graham 1977), the
24 Mangrove Ecology, Silviculture and Conservation

present populations of Pelliciera rhizophorae on the Atlantic coasts of Colombia


have a mixed origin. This species was introduced from the Pacific to the Atlantic
coast of Colombia, around the Canal del Dique y Covenas, in the last twenty years
(Paez 1994), although some apparently ancient stands occur on the Atlantic coast of
Panama, Colombia and Nicaragua (Jimenez 1984, Roth and Grijalva 1991, Duke
pers. comm.).

Fig. 2.4 Traditional fish ponds on Oahu, Hawaii, showing the dense growth of the
introduced 'old world-new world' hybrid mangrove community.

An extreme example of the extension of the distribution of species is the


introduction of Rhizophora mangLe to Molokai in 1905 (MacCaughey 1917), and of
Rhizophora mucronata, Bruguiera parvijIora, Bruguiera sexanguLa and Ceriops tagaL
to Oahu, Hawaii in 1922. Of these introductions, only R. mangLe and B. sexanguLa
have persisted (fig. 2.4), forming an unusual 'old world-new world' mangrove hybrid
community (Wester 1981).

The Mangroves of Mo'orea, French Polynesia


As recently as 1998, uncertainty was expressed as to whether the occurrence of
Rhizophora stylosa on Mo'orea in French Polynesia was natural and whether it
represented the most westerly point this species had reached naturally. Palynological
studies on Lake Temae deposits by Parkes (1997) showed that there was no
Rhizophora represented in the pollen profile covering the last 1,572 years BP.
Moreover, that author explained that a small stand of Rhizophora mucronata (sic) had
been planted at Vaiare 'in an attempt to stabilize the sediments.' Similarly, Mueller-
Dombois and Fosberg (1998:389) report that 'there is no real mangrove swamp in
Eastern Polynesia, except for an introduced occurrence of Rhizophora on Mo'orea,
which is spreading in the Society Islands.' In 1996 while going through French
research archives in Toulouse, I came across the following research grant report
(Contribution de I'URA CNRS 1453) which noted that the Rhizophora stylosa stand on
Mo'orea had been planted using propagules from New Caledonia between 1930-1935
and that 'a la suite d'une etude realisee en 1987, la popUlation de paletuviers a ete
estimee a pres de 3200 pieds.'

Other introductions in the Pacific include Rhizophora mangLe on Enewetak


Atoll (Taylor 1950, Duke et al. 1998), Bruguiera gymnorhiza on Bikini (St. John
The Mangrove Flora 25

1960) and Rhizophora stylosa as far east as Mo'orea, Tahiti and Bora Bora (Taylor
1979, Mueller-Dombois and Fosberg 1998). Similarly, the occurrence of
Rhizophora mangle in Fiji is almost certainly due to introductions, most probably
by the American Missionary Society which has been active on the island since the
1840s.
Human activity (such as pollution, water diversion and selective clearing) can
also significantly change mangrove distributional ranges. Thus, the northernmost
stands of R. racemosa in West Africa, reported by Adam (1965) from the ile re
Thiong (16° 03' N), Mauritania, have been cleared by the local inhabitants in the last
two decades (Gowthorpe and Lamarche 1993). Similarly, significant stands of
Bruguiera gymnorhiza have been selectively felled for boat building in the Sudan and
Yemen (Zahran 1977, Ormond et al. 1988) while Rhizophora mucronata has
disappeared from the Arabian Gulf in historical time for similar reasons.
In parts of eastern Africa, tanbark extraction from mangroves commenced in
1900 with Germany being the major importer. Tanbark export commenced in
Kenya, with Tanzania, Mozambique and Madagascar soon following (Grewe 1941).
The tanbark export from Kenya reached record values around the start of World War
I. This activity had collapsed by the end of the war. Even with some regrowth during
WW I, the tanbark resources of Madagascar were exhausted by 1927, while those of
Mozambique managed to survive until 1935, largely supplying the American leather
industry. Unsustainably high tanbark extraction took place in Tanzania from 1921
until that resource also was exhausted by 1937. Given that tanbark was derived only
from Rhizophora mucronata, Ceriops tagal and Bruguiera gymnorhiza, the volume
of bark exported represents a huge and selective loss of these particular species.
The large-scale replacement of mangroves by aquaculture ponds is a widespread
modem phenomenon that has been responsible for major losses of mangroves in, for
example, the Philippines, Indonesia and Ecuador. Elsewhere, the construction of
dams and barrages has altered prevailing salinity regimes to the detriment of some
species. For example, Heritiera fomes appears to be increasingly confined in its
distribution to the wetter eastern Sundarbans in Bangladesh as a result of tectonic
changes in combination with dam construction and excessive freshwater extraction
(Blasco et al. 1996).
There are now several examples of major mangrove afforestation programs from
around the world as well as numerous less extensive ones (Saenger and Siddiqi 1993,
Field 1996, Ellison 2000). However, in terms of how mangrove afforestation
programs can change mangrove distributions or relative abundance, we need only to
examine the Bangladesh mangrove afforestation program, where, since 1966,
approximately 120,000 ha of mangrove plantations have been established (Saenger
and Siddiqi 1993, Siddiqi 2001). As a consequence of the ease of transplanting
seedlings of Sonneratia apetala, the greatest part of the plantations consist of
monocuItures of this species which, prior to this program, was a relatively obscure
species. More recently, Sonneratia apetala has been successfully introduced to
Guangdong Province, China (Li et al. 1999), several thousand kilometres beyond its
natural range.

2.4 Contemporary Biogeographic Patterns


Patterns in contemporary mangrove biogeography are the outcome of the various
processes outlined above (Duke 1995, Saenger 1998, Ellison et al. 1999). Features
of these patterns are briefly reviewed.
26 Mangrove Ecology, Silviculture and Conservation

2.4.1 Vicariants
There are a few examples of mangrove species within one genus with non-
overlapping ranges. Such vicariant species are found in Aegialitis arxl
Camptostemon (Tomlinson 1986), two genera whose earliest fossils are known
from Borneo 23 my BP (Muller 1964). Aegialitis annulaJa is distributed from
northern Australia to Papua New Guinea while Aegialitis rotundifolia is restricted to
shorelines of the Bay of Bengal and the Andaman Sea (Saenger and Luker 1997).
Similarly, Camptostemon schultzii occurs in northern Australia, Papua New Guinea
and, possibly, as far north as Borneo, while Camptostemon philippinense occurs in
the Philippines, Borneo and Sulawesi. Whether there is any overlap between these
species in Borneo is not presently known.
By way of contrast, the genus Aegiceras contains one species (Aegiceras
comiculatum) with a wide Australasian distribution that completely encompasses the
more restricted Philippine distribution of Aegiceras floridum. Similarly the genus
Heritiera contains at least two mangrove species with one of limited distribution (H.
Jomes) being totally included within the range of the other wide-ranging species (H.
littoralis). Similarly, Excoecaria has one very wide-ranging species (Excoecaria
agallocha) which totally encloses the range of the recently confirmed species E.
ovalis, which appears to be confined to the Australian distributional range of E.
agallocha (Maguire et al. 2000).
Both mangrove species in the genus Cynometra (C. iripa and C. ramiflora)
have largely overlapping distributions in south-east Asia, although only C. iripa
extends southwards to the Australian mainland. C. ramiflora occurs as close as the
Australian territory of Christmas Island (Du Puy 1993).

2.4.2 Hybridizations
Hybrids are known between several species of mangroves including Sonneratia,
Rhizophora, Xylocarpus and Lumnitzera (Duke and Bunt 1979, Duke 1984,
Tomlinson 1986) which suggests that the genetic isolation between species of some
genera is not complete. On the other hand, molecular studies of a limited number of
mangroves from India and Sri Lanka suggest that between closely related genera,
genetic isolation between genera and most species is complete (Ballment et al. 1988,
Parani et al. 1998, Abeysinghe et al. 1999).
Molecular studies have also been used to shed more light on the parent species
of hybrids where these occur. Thus, based on morphological data, the sterile hybrid
Rhizophora lamarckii was interpreted as the crossing result of R. apiculata and R.
stylosa (Duke and Bunt 1979, Tomlinson et al. 1979). In 1995, Kathiresan described
the new species, Rhizophora annamalayana, from southern India; it had first been
identified as R. lamarckii, before being described as a new species. However, on the
basis of genomic DNA analysis, Parani et al. (1997a) have suggested that it is a
hydrid between R. apicu/ata and R. mucronata. The mitochondrial DNA of the
hybrid, inherited from the maternal parent, was identical to that of R. apiculata,
indicating that this species is the maternal parent. Consequently, R. annamalayana is
now recognized as a sterile FI hybrid of R. mucronata and R. apiculata.
By way of contrast, some species that have been considered as possible hybrids
are apparently not hybrids on the basis of biochemical data. For example, as Breteler
(1969) had suggested for the American Rhizophoras, Wilcox (1985) has indicated
that R. harrison;; might be a hybrid of R. racemosa and R. mangle in West Africa.
This is supported by intermediate morphological features (such as lengths of
The Mangrove Flora 27

peduncle and pedicels and the number of flowers in the inflorescence), the low
viability of R. harrisonii pollen (Wilcox 1985, Breteler, 1969), and the poor fruit set
of this species (Keay 1953, Savory 1953, Breteler, 1969). However, the hybrid
status of R. harrisonii could not be confirmed on the basis of foliar wax chemistry
by Dodd et al. (1995) and Rafii et al. (1996, 1999).

2.4.3 Discontinuities and Endemism


Given their seawater dispersal, most mangrove species were assumed to have a more
or less continuous distribution around various coastlines but there are exceptions and
the basis for this assumption needs to be critically re-assessed. Existing examples of
discontinuities from among mangrove genera include:
(1) All the neotropical species of Rhizophora show disjunct distributions with
West African, Atlantic American and Pacific American occurrences. In the case of
Rhizophora mangle, even the West African distribution appears to be disjunct with a
Senegal to Liberia and a Nigeria to Angola distribution (Saenger and Bellan 1995).
(2) Bruguiera hainesii is discontinuous between western Malaysia and New
Guinea (Tomlinson 1986) while Ceriops tagal and Bruguiera gymnorhiza show
major discontinuities between eastern Africa and southwest Asia, being absent from
Pakistan to Djibouti (Dodd et al. 1999).
(3) Sonneratia alba also appears to have several disjunct populations centred on
East Africa/Madagascar, India/Sri Lanka and Australasia. Even the northern
Australian distribution appears to be separated by a significant discontinuity in the
Gulf of Carpentaria. Similarly, Sonneratia ovata has a major disjunction between its
Thai-Indonesian and Papua New Guinea occurrences while Sonneratia lanceolata
appears to have three discontinuous distributions in southeast Asia and northern
Australia (Duke and lackes 1987).
(4) Avicennia rumphiana appears to be the only species of Avicennia with three
disjunct distributional ranges while A. integra is to date only known from a small
area of northern Australia (Duke 1988, 1991a, Wightman 1989).
(5) Pelliciera rhizophorae is of very limited distribution and is endemic to the
tropical Pacific coast of America. Its presence on the Caribbean coasts, first noted in
the early 1980s, is of uncertain origin.
Other locally restricted species include Acanthus xiamenensis, Aegialitis
rotundijolia, A. bicolor, A. schaueriana, Excoecaria ovalis, Atalantia correa,
Sonneratia griffithii, S. hainanensis, S. paracaseolaris and S. apetala.
The explanations for many of these disjunctions are undoubtedly to be found in
historical processes (Duke 1995, Saenger 1998, Ellison et al. 1999) in combination
with the ecological requirements of the individual species. The influence of historical
processes on the species composition of the mangroves of the northern and southern
coasts of New Guinea, for example, is discussed in 2.4.6. In a broad biogeographical
sense, however, temperature and rainfall are probably the main determinants.
Nevertheless, the high endemicity of the east Asian and northern Australian areas
may reflect their proximity to one of the main centres of mangrove origin and
subsequent dispersal routes. Obversely, there are no endemic mangroves on the
coasts of Africa, in contrast to all other major mangrove regions that have at least
one endemic species. This absence of endemic species suggests that the African
mangrove flora may be a relict flora. This is supported by the fossil record which
indicates that during the late Eocene to early Miocene, a richer mangrove flora was
present, including, amongst others, Nypa, Mauritia and Pelliciera (Saenger 1998,
28 Mangrove Ecology, Silviculture and Conservation

Ellison et al. 1999, Gee 2(01). To that must be 00ded that the mangrove vegetation
of Africa is in retreat from relatively recent climatic and sealevel changes (Rossi
1981, Saenger and Bellan 1995). It also suggests that the conditions along this
coastline were not suitable to trigger speciation in the taxa present.

2.4.4 Atlantic and Eastern Pacific Biogeography


Several major ocean currents influence the western coast of Africa (Longhurst 1962).
Two cold currents, the Canary Current in the northern hemisphere and the Benguela
Current in the southern hemisphere, move towards the equator 'squeezing' the region
of warm water on the coast into a narrower belt than would be expected from a
consideration of latitude alone (Lawson 1966). As they approach the equator, these
two currents tum westward and are then known as the North and South Equatorial
currents respectively. Between them, the warmer Counter Equatorial current flows in
the opposite direction towards the coast, moving northwards during the northern
winter. Like the Counter Equatorial current. the Guinea current flows in an easterly
direction travelling along the length of the Guinea coast and turning south in the
Bay of Bonny, joining the westward flowing South Equatorial current.
Another feature of importance in the Gulf of Guinea is the phenomenon of
'upwelling' of cold water that takes place along its coastline. This causes short-term
temperature drops and brings increased nutrients to the surface in June-August. Its
influence is pronounced from western Nigeria to Cote d'Ivoire but it is much reduced
west of there.
Tides on the Atlantic coast of Africa are of the semi-diurnal type with ranges
that are generally small although locally variable. Atlantic swells are ubiquitous
along most of the unprotected western coast of Africa and, consequently, it is only
in semi-enclosed bays and in lagoons and estuaries that conditions can be described
as sheltered. As the amplitude of the waves often exceeds the range of the tides, wave
action is relatively more important in determining the presence and vertical extent of
mangrove vegetation on the open shoreline. In estuaries and lagoons, highly variable
hydrodynamic conditions prevail and it is difficult to generalize.
Much of the western coast of Africa is uniformly mild to warm throughout the
year but there are large differences in total annual rainfall. In the north (fig. 2.5),
coastal Mauritania and Senegal are arid (9-11 dry months) while similar conditions
prevail in the south, largely as a result of the cold offshore watermasses. In between
these arid zones are the humid tropical zones with very high annual rainfalls (2,000-
3,000 mm) in Guinea, Guinea-Bissau and Liberia as well as the head of the Gulf of
Guinea (Nigeria, Cameroon, Equatorial Guinea and Gabon). The areas between the
humid zones (Cote d'Ivoire, Ghana, Togo and Benin) have an annual rainfall between
1,000-2,000 mm punctuated by two short dry periods. It should be noted, however,
that along the entire coastline, maximal rainfall generally occurs in the warmer
months. Unlike many mangrove coastlines elsewhere, cyclone activity is absent on
the Atlantic coast of Africa (Anthony 1989).
According to Lang and Paradis (1977, 1984), the last glaciation around 18,000
y BP resulted in a lowering of sealevels of about 100-150 m along the entire Gulf of
Guinea coastline. Beginning around 10,000 y BP, sealevel rose again due to the
melting of the glaciers, and reached a new maximum between 6,500 and 4,500 y BP
during the Nouakchottian transgression (Anthony 1989). All valleys and river
mouths were flooded in Senegal, Gambia, Guinea, Cote d'Ivoire and Gabon, forming
rias still common today. Many interdunal areas in Mauritania and elsewhere were
The Mangrove Flora 29

also flooded (up to 250 kIn inland) and occupied by dense vegetation as a result of
the humid conditions which accompanied this transgression (Faure and Hebrard
1977, Elouard 1968, Gowthorpe and Lamarche 1993).

Fig. 2.5 Satellite image of the arid Senegalese mangroves of the Sine Saloum delta
complex, with extensive development salt flats or 'tannes'. (Part of SPOT Kj 022/333,
dated 30/04197. ©cnes 1997 - distribution Spot Image. Reprinted with permission.)

Towards the end of the Nouakchottian transgression, between 4,800-4,200 y


BP, a north-south (along the Senegal to Guinea coast) and an east-west (along the
Gulf of Guinea) littoral drift became established which gradually caused littoral sand-
barriers to close the gulfs and rias that had been formed. This phenomenon was
progressive, starting in Mauritania (4,000 y BP), then Senegal (3,900-3,200 y BP)
and subsequently in Guinea, Cote d'Ivoire and Benin. By 3,000 y BP, for example,
the mouth of the Casamance in Senegal was transformed into a large lagoon where
mangroves had become established. This sand-barrier formation was further
consolidated by the small Taffolian regression of about 1 m at around 3,000 y BP.
By about 2,000-1,500 y BP the Atlantic coastline of Africa had largely developed its
present configuration and the climate had changed abruptly from a humid type to the
modem semi-arid climate without any transitional stages (Delibrias et al. 1976,
Kalck 1978, Barusseau et al. 1986, Lezines 1988). Over the last 1,500 years,
sealevel fluctuations have been small and difficult to establish with certainty,
30 Mangrove Ecology, Silviculture and Conservation

although rises of between 1-3.4 mm y-l have been reported from this region
(Verstraete 1989, Emery and Aubrey 1991).
Based on pollen analysis of sediments from the Niger delta (Sowunmi 1981,
1986), offshore cores (Rossignol-Strick and Duzer 1979) and peat deposits (Michel
and Assemien 1970, Assemien 1971), the extent and floristic composition of the
mangrove vegetation of western Africa has altered dramatically in conjunction with
changing climatic conditions and palaeo-sealevels, contracting significantly along
the entire west coast of Africa compared to the much greater extent around 5,000 y
BP. The remnants along parts of the coastline (e.g. Senegal, Cote d'Ivoire, Ghana,
Togo, Benin and western Nigeria) are characteristically confined to 'physiographic
refugia' i.e. semi-enclosed coastal lagoons or embayments, generally with
constrained tidal exchange and with limited (and markedly seasonal) freshwater input.
Additional 'physiographic refugia' can be recognized where offshore sandbanks (e.g.
Guinea), coral communities (e.g. Limbe, Cameroon) or coral rubble banks (e.g.
Liberia, Nigeria) may protect the shoreline from wave action that. in tum, allows
frontal mangroves to develop.
In addition, however, there are also some large deltas on this coastline which
constitute 'salinity refugia'. Mangroves are associated with these river deltas,
particularly the larger ones such as the Gambia, Volta, Niger, Rio Muni, Ogooue,
and Zaire; these settings are characterized by more-or-Iess continuous freshwater
input. free tidal exchange, and the abundant deposition of alluvial sediments. These
areas can be further subdivided into estuarine (polyhaline) or fluvial (mesohaline)
areas depending on the prevailing salinity regime.
The early literature concerning the mangrove vegetation of the Atlantic coast of
Africa failed to distinguish the three species of Rhizophora (e.g. Pellegrin 1952) and
it was not until Keay (1953) clarified this situation that reliable floristic data became
available for analysis. Avicennia germinans has a similarly long synonymy
(including a period when it was considered as an endemic species - A. ajrica:num)
before it became generally accepted that there were no significant morphological
differences between the American and African populations (Compere 1963). Given
the current taxonomic status, there are six indigenous and one introduced· (Nypa
jruticans) species of mangroves on the Atlantic coast of Africa (Blasco 1985,
Saenger and Bellan 1995). This contrasts with the Atlantic coast of the Americas
where there is an additional species of A vicennia. This suggests that since the late
Eocene (38 my BP), when the proto-Atlantic was being invaded by mangrove
species from the central and eastern Tethys Sea, there has been complete and
differential segregation between the Atlantic African and the East AfricanlIndo-
Pacific mangrove flora. However, since the mid-Miocene (20-14 my BP), when the
Tethys Sea dispersal route was no longer open (although it still contained extensive
Avicennia communities with Indo-Pacific affinity; Bessedik, 1981, 1985) and direct
dispersal between America and Africa was apparently no longer possible, there has
been speciation only in the genus Avicennia in America, but apparently not in
Rhizophora. IAguncularia or Conocarpus on either coast (Keay 1953, Tomlinson
1986, Jimenez 1987, Dodd et al. 1998).
Even when associated species are added to this list, only 12 families constitute
this widespread vegetation type on the Atlantic coast of Africa (Saenger and Bellan
1995). All of these families have counterpart species or genera on the American
Atlantic coasts and the majority have counterparts along the eastern coast of Africa.
The Mangrove Flora 31

The indigenous species of the Atlantic coast of Africa are shared with the Atlantic
and Pacific coasts of America.
The northern limits of mangroves (fig. 2.6) is formed by Avicennia, which is
recorded as far north as the lIe Tidra (19° 50' N) in Mauritania (Naurois and Roux
1965), in an archipelago now included in the 11,700 km 2 Le Parc National du Banc
d'Arguin, established in June 1976 (and as a World Heritage Site in 1989). The
genus Rhizophora is reported to extend to a little north of Saint Louis in the delta of
the Tiallakh River (16° 11' N) in southern Mauritania (Naurois and Roux 1965),
although it should be noted that the stands of R. racemosa reported by Adam (1965)
from the lIe de Thiong have been cleared by the local inhabitants in the last two
decades (Gowthorpe and Lamarche 1993). The northern limit of Conocarpus occurs
at 16° 03' N, on the Mauritanian banks of the Senegal River (Adam 1965). The
northern limit of Laguncularia appears to be from the delta of the Senegal River
(Nicou 1956).
The southern limit of extensive mangrove vegetation is found in the Angolan
estuary of the Rio Longa (10° 18' S), although isolated occurrences of Avicennia
occur slightly south of Lobito (12° 20' S) (Gossweiler and Mendon~a 1939, Airy-
Shaw 1947). Gossweiler and Mendon~a (1939) also indicate that Rhizophora does
not occur south of Lobito where its niche appears to be occupied by Laguncularia.
The most southerly record of Conocarpus is from the northern banks of the Rio
Zaire (6° 10' S) (Pynaert 1933). The southern limit of Laguncularia lies between
Lobito and M~amedes (15° 10' S).
· BctmudI
ATlANTlC
OCEAN

Fig. 2.6 Distributional gradients of maximal numbers of mangrove species on the


American and African Atlantic coasts together with the East Pacific coast of
America.

Both the northern and southern limits of mangroves in West Africa coincide
with the limits of arid regions (UNESCO 1979). This suggests that, as found on
other western coasts of southern continents (e.g. Australia; Saenger and Moverley
1985), mangrove distribution in western Africa is more limited by aridity rather than
by temperature.
Along the western shorelines of Africa, mangrove species distributions are not
uniform but show a degree of patchiness. For example, Rhizophora mangle is absent
32 Mangrove Ecology, Silviculture and Conservation

in Cote d1voire, Ghana and Benin, and is very rare with only a small stand of R.
mangle known from the lagoon at Owendo in Gabon (Lebigre 1983). In addition to
their variable distribution along the coastline, variable distances have been recorded
for mangrove penetration up river systems, ranging from around 100 kIn in Guinea-
Bissau to less than 20 kIn in Equatorial Guinea. These upriver distances probably
reflect seasonal differences in hydrology, with greater salinity intrusion during the
dry season.
On the other side of the Atlantic (fig. 2.6), mangroves extend from Laguna,
Brazil (28 0 30' S) in the south to their northernmost limit at Bermuda (32 0 20' N).
Brazil, with a 7,408 kIn coastline, possesses the largest mangrove area (13,400
km2), with clear latitudinal variation of the mangrove structure (Spalding et al.
1997). In northern Brazil there are well developed mangrove forests while there is a
southward decrease in the structural characteristics (Schaeffer-Novelli et al. 1985).
Northwards, too, the structure of mangroves is reduced along with their plant species
richness. Thus, in Bermuda only two species, Rhizophora mangle and Avicennia
germinans, occur (Ellison 1997). Unlike those of West Africa, the mangroves on the
American coasts of the Atlantic are limited by cold stress rather than aridity (Lugo
and Patterson-Zucca 1977, Johnstone 1983, Kangas and Lugo 1990, Snedaker 1995).
The most structurally developed and species rich mangroves along the American
coast occur in the high rainfall areas of both the Pacific and Caribbean coasts of
central America. Because of the recent closure of the Panamanian isthmus, merely
some 3 my BP, no differences in species composition occur between the Pacific and
Caribbean coasts of central America, although in some high rainfall areas such as
the Pacific coast of Colombia,. several rainforest species have been recorded as
mangrove associates (Gentry 1982).

2.4.5 Indian Ocean and East African Biogeography


The coastline of eastern Africa, the Arabian Peninsula and West Asia is extremely
variable. The Arabian shores of the Arabo-Persian Gulf include classical carbonate
and evaporite systems with essentially calcareous sediments (Purser 1982, Baltzer
and Purser, 1990) while the west Asian coastline is characterized by tectonic
movement and the resultant high frequency of earthquakes and changes in coastal
geomorphology (Blasco and Aizpuru 1997). The east African coast shows the signs
of the numerous major transgressions that have been recorded from the region
(Rossi, 1981) with sealevels well above present levels (Table 2.4), leaving extensive
supratidal fossil coral reefs as well as variously aged dunal systems. With each
transgression, the mangrove systems would have been relocated, as with each of the
subsequent regressions. In addition, these transgressions have been accompanied by
wet phases as a result of which rainfall, runoff and sediment supply would have
varied enormously.
Plaziat (1995) has identified raised beds of potamiid molluscs (Terebralia
palustris) in Egypt (near Hurgada), on the Saudi Gulf Coast (near Ras al Khaimah),
and around Abu Dhabi, United Arab Emirates, as indicating a former raised sealevel
associated with a Holocene transgression 3,800 y BP. At this time, he suggested,
Rhizophora was abundant in the Red Sea and the Arabo-Persian Gulf, favoured by
the somewhat wetter, more humid climatic conditions. This was accompanied by
such characteristic mangrove molluscs as Terebralia palustris, which made a brief
excursion into the Persian Gulf, at least to Abu Dhabi, where it does not occur today
(Plaziat 1995).
The Mangrove Flora 33

Table 2.4 Dates of transgressions and palaeoclimates of East Africa and Madagascar (Madagascan
names, where different, are given in brackets). (Adapted from Rossi 1981 and P1aziat 1987)

Transgression Time Climate Sealevel


(y BP) change (m)
Dry Phase
Nakurian 2,200-3,800 Humid Phase 3 to 4
(Fiandrian)
Dry Phase
Makalian 11,000-18,000 Wet Phase ?

Dry Phase
Upper Gamblian 25,000-30,000 Wet Phase ?
(lrodian)
Dry Phase
Lower Gamblian 80,000-150,000 Wet Phase 50 to 55
(Karimbolian)
Dry Phase -40 to -45
Kanjerian 240,000 Wet Phase 90
(Ivovonian)
Dry Phase
Kamasian 400,000 Wet Phase 80 to 85
(Antsiranian)
Dry Phase
Kagurian 2,200,000 Wet Phase 70 to 80
(Tatsimian)
Dry Phase -50to-70
Early Pliocene -3,000,000 Wet Phase

Two types of mangrove settings can be recognized on much of the Indian


OceanlEast African coasts. The dominant setting in the more humid areas consists of
deltas formed around the river mouths of the major rivers. Where streams are
intermittent, wadis are a much smaller analogue. The second setting, predominant in
the more arid regions of this coastline, consists of living or raised coral reefs along
which there are series of small bays that cut into the raised beach. These bays are
partly landlocked, often blocked by living coral which reduces wave action to a
minimum and allows mangroves to become established. Gazi Bay in Kenya is an
excellent example (Slim et al. 1996) where extensive mangroves occur, sheltered by
raised reefs and sand spits.
Grewe (1941) recognized four mangrove landscapes in eastern Africa and the
Middle East as follows: (1) Avicennia marina mangroves in the Sinai to Massaua;
(2) Avicennia marina and Rhizophora mucronata mangroves in southern Red Sea to
Somalia (although Ceriops tagal occurs in Djibouti; Faye 1993); (3) Full
complement of species from Lamu Archipelago, Kenya to the mouth of the
Zambesi, Mozambique; and (4) species poor scattered mangroves of southern
Mozambique to Natal. Similar landscape units have been recognized in Pakistan and
India (Blasco and Aizpuru 1997, Saifullah 1997).
The flora of eastern Africa, the Arabian peninsula and western Asia is a sub-set
of the Indo-Pacific mangrove flora. Each of the mangrove species has a wide Indo-
Pacific distribution and there are no endemic species in this region. While several
endemic species have been described (e.g. Ceriops somalensis Chiov., Ceriops
boiviniana Tul., Sonneratia mossambicensis Klotzsch ex Peters and Xylocarpus
benadirensis Mattei), these have been shown on closer examination to be conspecific
with their Indo-Pacific counterparts.
The coasts of the Arabian Red Sea and Arabian Gulf comprise the northern
latitudinal limits of the Indo-Pacific-East African mangrove realm and, as elsewhere
34 Mangrove Ecology, Silviculture and Conservation

toward the latitudinal limits, the mangroves are species-poor (Dodd et al. 1999).
Only two species of mangroves are recorded from the Red Sea coast of Saudi Arabia,
with A vicennia marina the most widespread. This species is tolerant of relatively
high salinity together with low rainfall and high temperature conditions, and
characterizes the high latitude limits of mangroves in the Arabian peninsula.

\
/
!
,

Fig. 2.7 Distributional gradients of maximal numbers of mangrove species on the


coasts of the Arabian Peninsula and East African.

In the Red Sea, the northerly limit of this species is reportedly at Shura EI-
Monqata associated with the Wadi Kid, near Nabq (28 0 24' N) on the Sinai Peninsula
(Por et al. 1977, Ish-Shalom-Gordon and Dubinsky 1993), where this species occurs
'in a dense stand ... rooted in heavy silt' sheltered by fossilized coral reefs. Other
northern occurrences are at Ras Mohammed (27 0 40' N) on the southern tip of the
Sinai Peninsula. On the Egyptian mainland coast, the northern limits occur in Hyos
Hormos Bay, about 22 km north of Hurghada (27 0 IT N), where Avicennia marina
forms low stands (Zahran 1977, 1982). On the Red Sea coast of Saudi Arabia, the
northernmost stands occur at Sharrn Zubeir at 27 0 34' N (Ormond et al. 1988).
In the Arabian Gulf, the northernmost natural occurrence of A vicennia marina
is around Qurma (Gurmah) Island near Al Jubail (27 0 08' N) although the largest
stands occur in Tarut Bay near Damman. Experimental plantations of Avicennia
marina exist further north. At Khor AI-Maqta near Khafji (28 0 25' N) in Saudi
Arabia, a plantation was established by the Japanese under contract to the Arabian
The Mangrove Flora 35

Oil Company in 1981 (Kogo 1988). Similarly, small experimental plantations were
established by researchers from the Kuwait Institute for Scientific Research in 1991
on the southern shores of Kuwait Bay (29° 25' N). Within 7 years, these plantations
had reached heights of around 2.5 m (AboEI-Nil 2(01).

Fig. 2.8 Mangroves of Tarot Bay near Damman on the Arabian Gulf coast of Saudi
Arabia, close to the northern limit of mangroves in the Arabian Gulf. While
Rhizophora mucronata is recorded from this region in historical time, today only
Avicennia marina occurs. Garbage dumping and/or landfilling is a widespread threat
to mangroves in the Arabian Gulf.

A second species, Rhizophora mucronata, occurs at a few sites in the Red Sea
but is absent from the Arabian Gulf. Its most northerly localities are on Umm
Rumah Island on the Wajh Bank (25° 44' N) in the Arabian Red Sea coast (AI-
Wetaid et al. 2(00) and it is particularly common in the Farasan Islands (16° 40' N),
in Egypt as far north as Marsa el-Madfa (22 19' N) near Marsa Halaib on the
0

Sudanese-Egyptian border, where it forms an almost pure stand (Zahran 1977), in the
Dahlak Archipelago off Mits'iwa (Massaua) on the Eritrean Red Sea coast (Kassas
1957), and at a few sites in Yemen (Ormond et al. 1988).
According to Zahran (1977), Bruguiera gyrrmorhiza has also been reported from
near Suakin (19° 15' N) in Sudan and from offshore islands near Hudaydah in Yemen
(Draz 1956). However, it has disappeared from Suakin over the last few decades
because of intensive cutting (Zahran 1977), while the occurrence in Yemen could not
be confirmed during a recent survey (Ormond et al. 1988). Today, its most northerly
occurrence is at Aquiq Bay (18° 10' N) in southern Sudan (Hughes and Hughes
1992).
The southern limit of mangroves appears to be in the vicinity of the Nahoon
River (32° 59' S) just north of East London in Natal which marks the latitudinal
limit of Avicennia marina (Hughes and Hughes 1992, Steinke 1998). The southern
limits of the other species are shown in Table 2.5 together with their northern
limits.
36 Mangrove Ecology, Silviculture and Conservation

Fig.2.9 At AI-Qandal in the Farasan Islands in the southern Red Sea, extensive stands
of Rhizophora mucronata still occur, protected as a nature reserve since 1989.

As noted earlier, mangrove communities are generally characterized by


decreasing plant height and by increasingly disjunct (patchy) distribution towards
their latitudinal limits. Both of these characteristics are apparent in the mangroves of
Saudi Arabia where heights rarely exceed 5 m, reflecting the generally limiting
environment (high salinity, low rainfall, extreme temperatures) in which they occur.
Similarly, the distribution of mangroves in the southern Arabian Red Sea is more
continuous than in the northern Red Sea where mangroves are confined to restricted,
favourable habitats. In the northern Arabian Gulf, only two widely separated
mangrove communities still remain, both confined to highly sheltered areas. In
southern Africa, the same picture emerges in reverse (Steinke 1998) while from Iran
to Pakistan and India, a similar attenuation of structure and species has been
described (Blasco 1975, Baltzer and Purser 1990, Blasco and Aizpuru 1997, Dodd et
al. 1999).
Table 2.5 Latitudinal limits of East African mangroves, compiled from various sources including Day
1981, Hughes and Hughes 1992, Steinke 1998, Saenger I99S, and AI-Wetaid et al. 2000.

Species Southern Northern


Limit Limit
Sonneratia alba 23° 55' S oool7'S
Ceriops tagal 26° 50' S 15° 50' N
Lumnitzera racemosa 26° 50' S 00° 51' S
Acrostichum aureum 2S0 5S' S 02° OS'S
Barringtonia racemosa 30°44' S oool7'S
Rhizophora mucronata 31°42'S 25°44' N
Hibiscus tiliaceus 32° 16' S 04°04' S
Bruguiera gymnorhiza 32° 14' S ISO 10' N
A vicennia marina 32° 59' S 2SO 24' N
Heritiera littoralis 26° 33' S 02°05' S
Xylocarpus granatum 26° 01' S 00° 17' S
Xylocarpus moluccensis 26° 33' S 02°05' S
Pemphis acidula 24° IS'S 03° 10' S
Thespesia populnea 26° 01' S 02° 32' S
The Mangrove Flora 37

2.4.6 Australasian Biogeography


Using floristic and environmental data, Saenger et al. (1977) divided the mangrove
coastlines of Australia into twelve biogeographic regions. Semeniuk et al. (1978)
subsequently subdivided one of the Western Australian regions into two, resulting in
a total of thirteen. More recently, Bridgewater and Cresswell (1999) re-analyzed
phytosociological mangrove and salt marsh data, recognizing eleven biogeographic
regions. Although it is expected that a further refinement of boundaries will occur,
these biogeographic regions correlate closely with certain environmental (particularly
meteorological and tidal) as well as physiognomic features for both the mangrove
and saltmarsh vegetation (Saenger 1995, Digby et at. 1999). This correlation
suggests that meteorological and tidal features of the coastline are involved in the
distribution and the physiognomy of the mangrove vegetation and its constituent
species.
Fossilized mangrove pollen and wood from the southern Australian mainland
(Churchill 1973, Stover and Evans 1973) and Tasmania (Cookson and Eisenack
1967, Pole and Macphail 1996) indicates that tropical coastal waters extended along
these shores during the Early to Mid Eocene (50 my BP), before the formation of the
major ice sheet on Antarctica around 37 my BP. Several of the species recorded as
fossils in southern Australia do not occur there today, and none occur in Tasmania.
It appears, therefore, that there has been a loss of these elements from southern
Australia since the Eocene. In view of the more restricted distribution of these
species today, past changes in climate and coastal conditions appear to have had a
sifting effect on the Australian mangrove flora. This sifting had undoubtedly
contributed to the existing species gradients, not only of the mangroves themselves
but also of associated plants and animals.
In Australia and Papua New Guinea, approximately thirty species of trees and
shrubs, belonging to fourteen families of angiosperms, are generally considered to be
part of the mangrove flora. With the exception of Avicennia integra and Excoecaria
ovalis, none of these species is endemic to the Australian region. Macnae (1966)
stated that three species appear to be purely Australian, Aegialitis annulata,
Bruguiera exaristata (fig. 2.8) and Osbomia octodonta. However, A. annulata occurs
throughout the Indo-West Pacific region, while B. exaristata extends to Papua New
Guinea and Timor, and O. octodonta extends as far north as the northern Philippines
(van Steenis 1979).
Duke (1992) has described a sharp floristic discontinuity in mangrove species
between the southern and northern shores of New Guinea. Similar floristic
discontinuities have also been described for upland plants of New Guinea (Heads
2001). These discontinuities reflect tectonic events associated with the collision of
the Australian and Pacific plates. The southern shores of New Guinea form part of
the stable Australian plate and have been subjected to alternate periods of emergence
and submergence asssociated with periods of glaciation, with the most recent around
18,000 y BP when sealevels were about 100-150 m lower than at present. The
northern shores of New Guinea lie on the northern edge of the Australian plate
which, although tectonically active, has remained submerged. The mangroves there
probably represent more ancient communities than the relatively young
communities along the southern shores. Probably of equal significance is that the
northern flora is continuous with, and derived from, the Indo-Malesian mangroves,
while the southern flora is largely derived from north-eastern and north-western
Australia.
38 Mangrove Ecology, Silviculture and Conservation

In Australia, an additional ten species from eight families have been noted as
associated lianes, epiphytes, or understorey species (Saenger et al. 1977), and a
further ten to fifteen species, although occasionally occurring in the mangrove
community, find their greatest development away from it. A large number of other
plants such as algae and seagrasses, bryophytes, fungi and lichens also have been
recorded from mangrove communities (Saenger et al. 1977, Stevens and Rogers
1979, Cribb 1979, Stevens 1979, Windolf 1989), but most of these species are not
confined to mangrove environments.

Fig. 2.10 Bruguiera exaristata is common on the coast of Queensland and extends
north to the southern shores of Papua New Guinea and Timor.

The distributions of the abundant species of mangroves around the Australian


coastline show that the largest number of species occurs on the northern and north-
eastern coasts (fig. 2.10). This diversity of mangrove species and associated plants in
the north-eastern area of Australia may be attributed to three main factors. First, this
region was one of the centres of origin of mangroves (see section 2.5) and the point
of their secondary dispersal into and out of Australia by virtue of its land
connections with south-eastern Asia (Walker 1972) during the various changes in
palaeo-sealevels. This interpretation accords with that for other Australian floristic
elements (Burbidge 1960). Second, the climatic regime of this area is similar to that
under which mangrove vegetation first developed and, consequently, little or no loss
of species has occurred there. In fact, Mepham (1983) argued that the north-eastern
coastline provides refuges for the once widespread and diverse Australian mangrove
flora as it withdrew northwards with the onset of arid conditions in the Oligocene.
Consequently, these north-eastern mangrove forests may be regarded as relicts
maintained by suitable climatic regimes (Smith and Duke 1987). Third, the coastline
configuration in this region, with its numerous estuaries generally sheltered by the
offshore Great Barrier Reef, provides large areas of low-energy coastline suitable for
mangrove colonization and development, and because of its climatic suitability, has
provided refugia during periods when climatic conditions were not optimal.
The Mangrove Flora 39

The distributions of many individual species of mangroves are patchy within


their general range. For example. several species are absent along the southern
shores of the Gulf of Carpentaria (Saenger and Hopkins 1975. Wells 1982. 1983)
but occur in the Northern Territory (Wells 1982. Wightman 1989) as well as in
north-eastern Queensland. Similarly. Xylocarpus granatum is absent from the
mainland coast south of the mouth of Raglan Creek (23 0 40' S) but small stands
occur on Fraser Island (25 0 20' S). In addition. some species are relatively
unimportant (in terms of numbers and size) along one part of the coast but are
among the more important constituents of the mangrove vegetation in other areas.
These patchy distributions around the Australian coastline are most probably due to
regional temperature and rainfall regimes (Saenger and Moverley 1985. Smith and
Duke 1987. Duke 1992).
Along the Queensland coastline. the most highly developed mangrove
vegetation. in terms of species and structural complexity. is found in those areas
where the annual rainfall exceeds 1.250 mm; these areas are generally where
elevations greater than 700 m occur in proximity to the coast. With increasing
latitude. both on the eastern and western coastlines. the number of species declines
rapidly. Lower water and air temperatures as well as the predominance of winter
rainfall effectively limit the southward extension of many species. reducing the
southern mangrove flora to a solitary species. A vicennia marina. which survives as
relict populations in disjunct pockets (for example. Abrolhos Islands in Western
Australia; Ceduna. Spencer Gulf. and Gulf St Vincent in South Australia; and
Barwon Heads. Port Phillip Bay. Westernport Bay and Comer Inlet in Victoria).
Macnae (1966) suggested two explanations for the present-day distribution of
the southerly mangrove populations: transmission by ocean currents or persistence
of relicts of previously warmer seas. He preferred the latter explanation. From the
work of Ludbrook (1963) and others. it is clear that during the Tertiary (including
the Pliocene) the seas around southern Australia were warmer than they are today.
The occurrence of other mangrove fossils from the Early to Middle Eocene in
southern Australia (Churchill 1973. Pole and Macphail 1996) suggests that the
present-day mangrove vegetation on the southern Australian coastline is a relict of
these earlier. warmer conditions. having managed to maintain itself in a few
favourable localities.
In New Zealand. where only a single species of mangrove occurs. it also seems
likely that present day mangroves are a relict of the more widespread Eocene
mangroves. On the west coast of New Zealand. A vicennia marina naturally occurs as
far south as Kawhia Harbour (3r 47' S) although planted mangroves occur slightly
further south (Crisp et al. 1990). On the east coast. mangroves extend as far south as
Ohiwa Harbour (38 0 00' S). although Mildenhall and Brown (1987) and Mildenhall
(1994) have shown Early Holocene mangrove remains further south in Poverty Bay
(38 0 43' S).

2.4.7 Western Pacific Biogeography


Mangroves do not cover a large area on islands in the Pacific Ocean and there is
a rapid attenuation of species richness and a general decline in mangrove forest
structure eastwards from south-east Asia (Woodroffe 1987. Woodroffe and Grindrod
1991. Mueller-Dumbois and Fosberg 1998. Devoe and Cole 1998). Mangroves are
absent from the eastern central Pacific despite their occurrence on the west coast of
Central and South America and in the Galapagos Islands. a feature ascribed to the
40 Mangrove Ecology, Silviculture and Conservation

difticulties of dispersal from an Indo-West Pacific centre of diversity (van Steenis


1962, Saenger 1998, Dodd et al. 1998, Ellison et al. 1999).
East of Samoa, natural mangroves are either absent or, as on the Cook Islands,
consist of only scattered plants of such species as Pemphis acidula and Hibiscus
tiliaceus (Mueller-Dumbois and Fosberg 1998), although Ellison (1994) recorded
Rhizophora pollen in cores dated around 2,500 y BP from Mangaia, Cook Islands.
Other islands east of Samoa (e.g. the Society and Hawaiian Islands) have only
planted mangroves (fig. 2.11) and Holocene core samples have failed to provide any
evidence of earlier, natural occurrences (Parkes 1997).
This large area of the eastern central Pacific without natural mangroves has
given rise to considerable controversy between those who favour a western Pacific
crossing of mangroves and those who support the opposite view (Saenger 1998).
The picture is further confused by the occurrence of Rhizophora mangle (R.
samoensis according to some authors) in Samoa, Tonga, Fiji and New Caledonia.
Although this species is probably a historical introduction, Ellison (1991) claimed
that the occurrence of fossil R. mangle pollen in a Holocene mangrove peat from
Tonga suggests that this species arrived here by island hopping during periods of
lower sealevel during the Eocene. However, R. mangle type pollen is
indistinguishable from pollen of R. apiculata and the hybrid R. lamarckii, two taxa
widely distributed in this area. The alternative interpretation is that R. apiculata
occurred in pre-historic times in Tonga but became extinct during some adverse
period, only to be ecologically replaced by the introduced R. mangle during historic
time. This possibility was not considered by Ellison (1991) but cannot be ruled out.
Ellison's (1991) simple assumption that the pre-historic apiculata-lamarckii-mangle
type pollen represents R. mangle needs further investigation (Dodd et al. 1998).

Mia~ ~I
\ - -..
• t> PAOPIC OCEAN

IlTOR

T<I)P
<

Fig. 2.11 Distributional gradients of maximal numbers of mangrove species across the
western Pacific.

Further west, the mangrove species richness and structural complexity increases
rapidly. Thus, three species (Rhizophora stylosa, Lumnitzera littorea and Pemphis
The Mangrove Flora 41

acidula) are recorded from Tuvalu (Woodroffe and Moss 1984). Mangroves are absent
in the eastern Marshall Islands but four species occur on the western islands
(Mueller-Dumbois and Fosberg 1998). Similarly, seven species of mangroves are
recorded eastern Micronesian islands of Trok, Ponape, Kosrae and Guam but ten
species are known from the western islands of Palau and Yap (Mueller-Dumbois and
Fosberg 1998). In Vanuatu, mangrove areas can be found on nine islands, namely
Hiou Vanua Lava, Mota Lava and Ureparapara in the north, Malekula, Epi, Ernae
and Efate in the central region, and Aniwa in the south. The largest mangrove stand,
at Port Stanley on Malekula, consists of 17 species of mangroves (Marshall and
Medway 1976) and illustrates the rapid increase in mangrove species diversity
towards the Solomon Islands and Papua New Guinea.

2.4.8 North-East Asian Biogeography


The mangroves of north-eastern Asia show a northward attenuation both in species
richness, productivity and structural development (Hong and San 1993, Li and Lee
1997). While mangroves extended further north in China (Zhang 1991, Zhang and
Wang 1994, Zhang et al. 1997) and Japan (Tanai 1972, Tsuchi 1992) during the last
10 my BP, today mangroves are found in the provinces of Hainan, Guangdong
(including Hong Kong and Macau), Guangxi, Fujian, and in Taiwan. The northern
limit of natural mangrove distribution on the mainland is at Fuding (27° 20' N,
Fujian Province, China) while in Japan, Kandelia candel extends north to Kiire,
Kagoshima at 31 ° 22' N (Suzuki and Saenger 1996).

o
,

;
JO ~/

Fig. 2.12 Distributional gradients of maximal numbers of mangrove species in north-


eastern Asia.

The northern limit on the mainland is probably temperature related as the


minimum monthly mean temperature at Fuding is 8.4°C in January (Lin and Xin-
Men 1983); at this locality, only one species (K. candel) is present and the average
tree height is ::;; 2 m. It is worth noting, however, that there is a remnant stand of
KaruJelia candel further north in Zhejian province. This was planted in the 1950s
42 Mangrove Ecology, Silviculture and Conservation

and, although largely destroyed by human disturbance with only 8 ha remaining at


present (Li and Lee 1997), it seems to suggest that the present thermal limit lies
somewhat beyond the natural geographic limit.
In China, there are 37 species of mangroves and mangrove associates with the
greatest species richness on Hainan Island. It contains 35 species and the best
developed mangrove forests, with a maximum tree height of 15 m (Li and Lee
1997). With increasing latitude, species richness decreases and the mangroves
become less well-developed. Guangdong, Guangxi, Taiwan and Fujian have species
numbers of 20, 15, 19 and 9 respectively (Li and Lee 1997, Tam et al. 1997).
In Taiwan, mangroves were once more widely distributed on the west coast of
the island, with particularly large areas at Keelung and Kaohsiung Bays. However, in
1929 Keelung Harbour was developed, destroying the most northerly K. candel stand
on the island. Similarly, the development and expansion during 1908-1944 of
Kaohsiung Harbour greatly reduced the largest and most diverse mangrove
communities on Taiwan. In the process, two indigenous species (c. tagal and B.
gymnorhiza) were entirely eliminated from Taiwan, and the local population of R.
stylosa was decimated (Hsueh and Lee 2000). At present, there are 32 mangrove
stands remaining in Taiwan's western shorelines with a total area of 289 ha (Hseu
and Chen in press).

Fig. 2.13 Shown here at Hainan Island, China, Kandelia candel is the characteristic
species of the north-eastern Asian mangroves, occurring from Malaysia to the
northern limits of mangroves at Kiire, Kagoshima (3 I 0 22' N), southern Japan.

The floristic composition of the five Chinese provinces reveals that Guangdong
and Guangxi are most closely tied floristically, whereas Hainan Island, with the
highest mangrove species richness in China, is most floristically distant from
Fujian, the northern limit of Chinese mangrove distribution. Taiwan's mangrove
species assemblage exhibits higher similarity to those of Guangdong and Guangxi
than to Fujian, despite the latter's proximal geographical position to Taiwan.
This deviation from the latitudinal trend is a result of the influence of the warm
Kuroshio Current on Taiwan (Hsueh and Lee 2000) but not Fujian (Li and Lee
The Mangrove Flora 43

1997). Hsueh and Lee (2000) examined species diversity against three indicators of
winter coldness and found that the number of species increased logarithmically with
increase in annual mean air temperature and winter mean temperature and decreased
with increase in the number of months at monthly mean air temperature below
20°C.
Winter air temperatures are, in turn, dependent on the warm Kuroshio Current
which influences the east and south-west coast of Taiwan as well as the Ryukyu
Islands of southern Japan, while the cold North China Coastal Current influences air
temperatures in Fujian Province and the north-west coast of Taiwan (Hsueh and Lee
2000, Taira 2001).
The warm Kuroshio Current also affects water and air temperatures in the
Ruyuku Islands. Despite their occurrence north of Fujian Province, several additional
species of mangroves occur; on Iriomoto Island, for example, 10 species have been
recorded, including Kandelia candel, Avicennia marina, Sonneratia alba, Rhizophora
stylosa, Bruguiera gymnorhiza, Lumnitzera racemosa, Heritiera littoralis, Excoecaria
agallocha, Nypa fruticans and Acrostichum aureum (Suzuki 1979). The Nypa
fruticans stands on Iriomote Island were declared a 'national monument' in 1989.

2.5 Towards a Global Biogeographic History


Land plants probably invaded shallow marine environments as early as the Late
Carboniferous, when the gymnospermous cordaitean, Amyelon, possibly formed
'mangroves' (Cridland 1964, Raymond and Phillips 1983). Additional mangrove-
forming plants such as the fern, Weichselia, are known from the Late Jurassic, and
several conifers from the Cretaceous (Shinaq and Bandel 1998, Vermeij and Dudley
2000, Plaziat et al. 2001). Ricklefs and Latham (1993) have described a number of
further invasions from the Late Cretaceous to the Miocene. Nevertheless, it remains
a puzzle why only three clades of flowering plants became fully marine (seagrasses)
and a mere 15 other clades entered what are now mangrove environments (Vermeij
and Dudley 2000, Schwarzbach and Ricklefs 2001). Most of these transitions to the
mangrove environment appear to have occurred in the old world tropics (Ricklefs and
Latham 1993) although there is some evidence that a few transitions took place in
the American and the European tropics of the Eocene (Saenger 1998, Gee 2001).
What prompted these transitions must largely remain speCUlative. However,
Proches (2001) has suggested that tropical transitions from the terrestrial to the
marine environment in mangroves (and turtles and sea snakes) were promoted by
such biotic interactions as competition.
The world-wide occurrence of mangrove vegetation with its floristic divergence
between the 'old' and the 'new' world mangroves, can only be explained by historical
processes (Duke 1995, Saenger 1998, Ellison et al. 1999, Plaziat et al. 2001). The
composition of the modern mangrove flora at anyone location, although subject to
present-day ecological processes, is essentially the end result of historical processes
(Tomlinson 1986, Duke 1995). While some authorities (Thome 1973, Smith 1973)
have suggested that Angiosperms arrived on the scene too late for their modem
distribution to have been seriously affected by continental drift, being water-
dispersed, mangroves could only be dispersed where suitable passages of water
existed between continents. Consequently, the present distribution of mangroves and
their constituent species must be seen within the context of plate tectonics and
continental drift (Saenger 1998). Although several interpretations have been offered
to relate mangrove distributions to past continental movements (van Steenis 1962,
44 Mangrove Ecology, Silviculture and Conservation

McCoy and Heck 1976, Specht 1981b, Mepham 1983, Rico-Gray 1993, Duke
1995, Plaziat 1995), none has been universally accepted (Saenger 1998, Plaziat et al.
2001).
Each of these various interpretations is based on the existence during the
Cretaceous of an extensive tropical sea, the Tethys Sea, separating the northern
supercontinent of Laurasia from the southern Gondwanaland. Mangroves evolved in
the late Cretaceous within the Tethys Sea and dispersed outwards via the available
coastal waterways. Ultimately, around 18 my BP, the western Tethys Sea became
isolated with the enclosure of the Mediterranean by the collision of Africa and Asia
Minor (Table 2.6). At that time, the pantropical mangrove flora became disjunct and
developed as two isolated floras. This assumes that the southerly extent of Africa and
South America formed impassable barriers to mangrove dispersal; given that the
latest ice-core data suggest that Antarctica formed a permanent icesheet around 37 my
BP, this assumption seems realistic. Nevertheless, given the rapidity with which
Nypa colonized southern Australia during the Eocene (Pole and Macphail 1996), this
assumption may yet need revision.
Table 2.6 Significant geological events and mangrove fossil records. (Adapted from Saenger 1998 and
references therein)

Period Approx. Fossil Record


myBP
Pliocene 3 Panama route closed
Miocene 10 Avicennia and Sonneratia pollen in Borneo; Aegialitis and Sonneratia
pollen in India; Avicennia pollen in Nigeria
16 Acanthus, Aegialitis and Camptostemon pollen in Australia;
Bruguiera. Sonneratia. Avicennia. Ceriops. Rhizophora. Scyphiphora.
Nypa and Excoecaria pollen in Japan; Tethys Sea route closed;
A vicennia still present in Mediterranean
23 Aegiaiitis. Avicennia and Camptostemon pollen in Borneo;
Scyphiphora. Lumnitzera and Avicennia pollen in the Marshall
Islands
24 Rhizophora in Nigeria. Senegal and South America
Oligocene 36 Pelliciera pollen in Nigeria. Cameroon and Puerto Rico
38 Sonneratia in south-east Asia; Rhizophora in South America
Eocene 40 Rhizophora. Avicennia and Sonneratia in south-west Australia;
Rhizophora and PelLiciera pollen in Panama; Rhizophora,
Conocarpus and Laguncularia in North America
42 Fossils of Nypa. Ceriops, Paiaeobruguiera. Rhizophora and
Acrostichum-like ferns in London Clay Flora; Bruguiera. Ceriops.
Avicennia, Pelliciera and Nypa pollen in Paris basin; Heritiera and
Aegiceras pollen near Pyrenees; Nypa in Egypt and Sonneratia wood
in Libya
50 Pelliciera pollen in Jamaica; Rhizophora in Borneo and India; Nypa
and Rhizophoraceae near Pyrenees
54 Nypa pollen and/or fruit in Europe, North America, Asia and
Australia; Acrostichum spores and fronds in Japan; Pelliciera pollen
near Pyrenees
Palaeocene 57 Sonneratiaceae pollen near Pyrenees
60 Rhizophora pollen in Australia
63 Nypa pollen in Nigeria
69 Nypa fruit in Brazil
Cretaceous 110 First flowering plants

Finally, around 3 my BP, the Panama gap closed due to the collision of North
and South America. Today, therefore, there are three disjunct mangrove floras of
which that of the eastern Pacific is too recent in origin to markedly differ
morphologically from its Atlantic progenitor.
The Mangrove Flora 45

The known fossil record of mangroves is summarized in Table 2.6. It suggests


that mangroves evolved soon after the appearance of flowering plants, with the
earliest records of Nypa in the late Cretaceous to early Palaeocene (Dolianiti 1955),
closely followed by Sonneratia (Gruas-Cavagnetto et al. 1988). Other genera evolved
at various times and in apparently differing regions.
Given this evolutionary history, four broad views have been suggested for the
centre of origin and the dispersal of the mangrove flora: (1) an eastern Tethys Sea
origin with eastward dispersal across the Pacific and via the Panama gap eastwards
into the Atlantic (e.g. van Steenis 1962, Muller and Caratini 1977, Ellison 1991);
(2) an eastern Tethys Sea origin with dispersal north-westwards into the Atlantic and
then via the Panama gap into the eastern Pacific (e.g. Specht 1981b, Rico-Gray
1993); (3) a western Tethys Sea origin with dispersal south via southern Africa to
the eastern Tethys Sea (e.g. Duke 1995); and (4) a two-way dispersal from a central
Tethys Sea origin (Por 1984, Plaziat and Cavagnetto 1996) or an eastern Tethys Sea
origin (Bousquet-Melou 1996).
None of these various interpretations can be ruled out entirely as mangrove
fossil records (Table 2.6) are somewhat ambiguous and certain genera (e.g.
Avicennia) are not well represented (Plaziat 1995, Plaziat et al. 2001). However, the
vicariance hypothesis that different mangrove plants and animals arose in various
parts of the Tethys Sea from where they migrated and diversified as a result of
continental drift is strongly supported by the mangrove fossil record and the modem
distributions of mangroves and mangrove-associated gastropods (Plaziat 1995,
Saenger 1998, Ellison et al. 1999, Plaziat et al. 2001).
However, the fossil record generally supports an eastern Tethys Sea origin for
RhizopJwra and perhaps Avicennia (Cookson and Eisenack 1967, Stover and Evans
1973, Churchill 1973, Martin 1982, Schwarzbach and Ricklefs 2001), with a
westward dispersal via the Mediterranean route which until about 18 my BP,
contained extensive Avicennia communities with Indo-Pacific affinity (Bessedik
1981, 1985). For Avicennia, a cladistic analysis of the species of this genus
(Bousquet-Melou 1996) indicated that the Australasian species (A. marina and A.
alba) are the ancestral species while the three neotropical species (A. germinans, A.
schaueriana and A. bicolor) are of later origin. Similarly, based on leaf aliphatic
hydrocarbon and triterpenoid fractions, Dodd et al. (1998) concluded that, in
comparison with West African provenances, the American provenances showed
derived characteristics; the African material consistently showed a greater diversity in
lipid composition and a tendency for longer carbon chain lengths.
Fossil remains of several other mangrove genera (including Aegialitis,
Camptostemon and Scyphiphora) have only been found in the Australasian region
and are clearly of somewhat later, eastern Tethys Sea origin. However, it seems
equally clear that the genera Nypa and perhaps Laguncularia and Conocarpus of the
Combretaceae, have a western Tethys Sea origin (Berry 1924, 1930, Gee 2001).
Finally, on the basis of presently known fossil location and ages, genera such as
Sonneratia, Heritiera, Pelliciera and Aegiceras made their earliest appearances in the
central Tethys Sea in the late Palaeocene to mid Eocene (Gruas-Cavagnetto et al.
1988, Cavagnetto and Anad6n 1995, Plaziat et al. 2001).
Given these various centres of origin, the following dispersal patterns seem
plausible (Saenger 1998). During the late Cretaceous to Palaeocene (63-55 my BP),
when the western Tethys Sea (proto-Atlantic) was still a relatively narrow waterway
and western Africa was experiencing a 'wet' phase, there was no mangrove vegetation
46 Mangrove Ecology, Silviculture and Conservation

per se in the Niger delta, but extensive estuarine swamp communities dominated by
Nypa (Sowunmi 1981, 1986). Nypa, which is first known from the late Cretaceous
of Brazil (Dolianiti 1955), apparently evolved in, and spread throughout, the proto-
Atlantic at this time (Moore 1973, Gee 2(01). This Nypa-dominated swamp
community remained in West Africa during the entire Eocene (Sowunmi 1981,
1986) and extended as far north as Europe (Wilkinson 1981).

Fig. 2.14 Acrostichum speciosum is a hardy fern, forming abundant sporangia over the
entire undersides of fertile fronds. As a result, it is an effective disperser and
opportunist, rapidly able to colonize most canopy gaps in tropical mangrove forests.

At much the same time (i.e. Palaeocene), Sonneratia appeared in the central
Tethys Sea (Gruas-Cavagnetto et al. 1988), followed in the mid-Eocene by Heritiera,
Aegiceras and Pelliciera (Cavagnetto and Anad6n 1995). Three of these genera
apparently dispersed towards the eastern Tethys Sea while Pelliciera dispersed first
westwards to Jamaica (Graham 1975), then southwards to Cameroon (Salard-
Cheboldaeff 1976). Genera of the Rhizophoraceae together with Avicennia rapidly
dispersed westwards from the eastern Tethys Sea during the early Eocene. Thus, by
the mid-Eocene, Europe appears to have been the first tropical meeting point of
mangrove genera from the eastern, central and western Tethys Sea. It is interesting
to note that, in the fossil beds of France and England during this tropical era, pollen
or spores of Sonneratia, Avicennia, Rhizophora, Palaeobruguiera, Ceriops, Nypa,
Aegiceras, Heritiera, Acrostichum and Pelliciera have been recorded (Chandler 1951,
The Mangrove Flora 47

Wilkinson 1981, Gruas-Cavagnetto 1987, Ollivier-Pierre et al. 1987, Gruas-


Cavagnetto et al. 1988, Plaziat and Cavagnetto 1996).
By the late Eocene, the genera Conocarpus and Laguncularia appeared in the
western Tethys Sea (Berry 1924, 1930) and became widely dispersed around the
American and West African shorelines. As seasonally dry conditions were widespread
in West Africa at the end of the Eocene, Nypa declined in abundance and finally
disappeared altogether in the early Miocene (24 my BP) - incidentally, at the same
time Nypa disappeared from the fossil record of Venezuela. These disappearances
coincided with the sudden and predominating appearance of Rhizophora throughout
the region (Sowunmi 1981, 1986) which, given the earlier records of this genus
from the eastern Tethys Sea, must have reached West Africa via the Mediterranean.
This is further supported by the earlier appearance of Rhizophora in southern France
(50 my BP; Gruas-Cavagnetto et al. 1988) and in the London Clay (45 my BP;
Chandler 1951) than in North (40 my BP; Berry 1924, 1930) and South America (38
my BP; Germeraad et al. 1968, Van der Hammen 1972).
Plaziat and Cavagnetto (1996) and Plaziat et al. (2001) have suggested that the
'climatic cooling' of the late Eocene not only changed the species composition and
distribution of the mangroves but also started the process of differentiation of the
'old' and 'new' world mangroves during the Oligocene, rather than in the mid-
Miocene when the Tethys Sea route closed.
Thus, while some mangroves appear to have evolved in the central Tethys Sea
(Mediterranean), several other mangroves (Nypa and an ancestral Combretaceae)
apparently evolved in the western Tethys Sea (the proto-Atlantic) and dispersed out
of the region (Exell and Stace 1972, Moore 1973, Saenger and Bellan 1995, Ellison
et al. 1999). Given this western Tethys Sea origin of an ancestral Combretaceae, the
occurrence of a non-mangrove species of Conocarpus, C. lancifolius, on the coastal
plains of the Red Sea, Sudan (van Vliet 1979, Tomlinson 1986), raises some
interesting problems, and emphasizes how speCUlative much of our information
about past dispersal and speciation events appears to be.
Other members (Rhizophora and Avicennia) of the eastern Tethys Sea (Indo-
Pacific) mangrove flora were able to enter and survive in the proto-Atlantic before
the Tethys Sea became unavailable as a migration route. Acrostichum may have had
a north-eastern Tethys Sea origin where the earliest remains of Acrostichum have
been found in Japan (Tanai 1972). However, Acrostichum, which has been world-
wide in its distribution since the mid-Eocene and extra-tropical in many parts of its
distribution (fig. 2.14), is an effective disperser and opportunist.
If, as indicated by the cladistic analysis of Avicennia (Bousquet-Melou 1996),
the West African-American species are derived from Australasian ancestral stock,
then the process of differentiation between the Indo-Pacific and African-American
species has occurred relatively recently, during the 'climatic cooling' of the late
Eocene, or after the closure of the Mediterranean around 18 my BP. Similarly, the
differentiation between the African-American genus Laguncularia and the closely
related Indo-Pacific genus Lumnitzera would also date from this period. On the other
hand, given the apparently derived nature of the American Avicennia. Rhizophora
and Laguncularia from African stock (Dodd et al. 1998, 2000), the process of
differentiation between the Neotropical and West African species and between
American Atlantic and Pacific species is even more recent, and is as yet incomplete
(Saenger 1998, Dodd et al. in press).
3. Adapting to the 'Mangrove Environment'
When the coral structure reaches low-water sea-level
the life of the coral-insect ends,
and the winds and tides, and the mangrove finish the work.
The mangroves live just at the edge of the shore,
throwing outward into the salt water their straight, forked roots,
thus forming an interlocking net-work
into which the drift is carried and finds lodgement.
The storms bring in solid material,
and the sea is slowly but surely encroached upon,
and the naked reef in time covered with vegetation.

C.R. Dodge (1894:23)

3.1 Introduction
As we have already noted, the mangrove environment is highly variable and stressful
owing to a combination of periodic fluctuations and extremes in physico-chemical
parameters. Despite this variability, the mangrove flora has successfully colonized
this environment, apparently aided by the development of numerous morphological,
reproductive and physiological adaptations (Saenger 1982, Clough et al. 1982,
Tomlinson 1986, Hutchings and Saenger 1987, Naskar and Mandai 1999). Many of
these adaptations are inferred; that is, adaptations of mangrove species have generally
been identified simply by comparing the characteristics of mangroves with those of
species from non-mangrove environments. Experimental investigation of the
efficiency of many of these adaptations remain to be carried out; nevertheless,
mangroves exhibit several specific features which allow them to occupy the
mangrove environment.

3.2 Dealing with High Salt Concentrations


Salt is the single most important and abundant characteristic of the mangrove
environment, and mangroves inevitably absorb some sodium and chloride ions from
their surroundings. Seawater, containing about 35 g L· t of dissolved salts (483 mM
Na+ and 558 mM en, has an osmotic potential of approximately -2.5 MPa. That of
soil water may be even lower. The fact that mangroves are able to grow in such
highly saline substrates and even grow better in the presence of some salt (Connor
1969, Downton 1982, Ball and Pidsley 1995, Khan and Aziz 2001) suggests that
they are able to control the intake of salt and maintain a water balance which is
physiologically acceptable. Even though these processes are understood in general
terms, reliable data are lacking on many details.
Although the term 'salt tolerance' was used earlier, Popp (1995) has discussed
the semantics of the term 'tolerance' and 'resistance' in relation to salt and other
stresses. In her comprehensive review, she has pointed out that in dealing with salt,
mangroves have developed resistance which consists of salt avoidance and regulation
on the one hand, and salt tolerance or accommodation on the other. All mangroves
show some, or all, of the features of salt resistance, which include salt exclusion,
extrusion, storage, succulence, compartmentalization and osmocompensation. Each
of these aspects of salt resistance are considered below.
50 Mangrove Ecology, Silviculture and Conservation

3.2.1 Salt Exclusion


Controlling the intake of salt in an environment high in ambient salt is the first line
of resistance; mangroves provide a striking example of salt exclusion from the
xylem sap. They possess an effective mechanism in the roots (Rains and Epstein
1967, Scholander 1968, Werner and Stelzer 1990, Popp et al. 1993), whereby water
is taken up and salt is largely excluded. Genera which are particularly effective in
excluding salt are Rhizophora, Ceriops, Sonneratia, A vicennia, Osbomia, Bruguiera,
Excoecaria and Aegiceras.

Fig. 3.1 The effectiveness of salt exclusion in mangroves is demonstrated by


Rhizophora apiculata growing in seawater at Taklong Island, Philippines.

Early measurements of the osmotic potential of xylem sap in species which


lack salt glands gave values of less than -0.2 MPa (Scholander et. al. 1962),
indicating that the concentration of soluble salts in the xylem is similar to that of
many plants from non-saline environments. The osmotic potential of xylem sap in
salt-extruding species was reported to range from -0.4 to -0.7 MPa (Scholander et al.
1962), showing that they are somewhat less efficient in excluding salt than those
species without salt glands.
However, when Scholander and his colleagues developed the pressure chamber
technique (Scholander et al. 1965, 1966, Scholander 1968) more negative values,
ranging from -3.8 to -5.2 MPa, were found. While this technique has been criticized
(Zimmermann et al. 1994) for a variety of reasons (see 3.3.2), recent laboratory
studies have shown that the osmotic potential of the xylem sap of the mangroves
Avicennia marina (Downton 1982, Tuan et al. 1995, Khan and Aziz 2001),
Aegiceras comiculatum and Conocarpus erectus (Popp et al. 1993), and Rhizophora
mucronata and Ceriops tagal (Khan and Aziz 2001) is correlated with the salinity of
the growth medium, ranging from -1.6 MPa at zero salinity to around -6.1 MPa at
full seawater.
Takemura et al. (2000) investigated the response of Bruguiera gymnorhiza to
increasing salinities. When transferred from low to high salinity, an immediate
increase in transpiration was observed. This, they suggested, was a response to
Adapting to the 'Mangrove Environment' 51

osmotic shock whereby the plant increased the internal salt concentration to balance
the increased external osmoticum. The time needed to re-establish a steady-state rate
of transpiration was of the same order as that required by the internal salt
concentrations to stabilize, suggesting that these two processes are closely coupled.
Such a response provides a great ecological advantage under the fluctuating salinity
conditions common in mangroves in the field.
Typical values of leaf water potential in mangroves in the field range from -2.5
to -6 MPa (Rada et al. 1989, Sternberg et al. 1991, Popp et al. 1993, Tuan et al.
1995, Weiper 1995, Sobrado 2000). Nevertheless, most mangroves, whether they
can extrude salt or not, apparently still exclude 80-95% of the salt in seawater at the
roots (Scholander 1968, Waisel et al. 1986, Werner and Stelzer 1990, Popp et al.
1993, Weiper 1995).
The mechanism by which this occurs is not clear; Scholander (1968) found that
neither chilling nor metabolic inhibitors caused any change in the capacity of the
roots to exclude salt, and he concluded that the process was simply a passive
function of the differential permeability of membranes in the root. This was
supported by the absence of any obvious diurnal variation in the salt concentration
ofthe xylem sap (Scholander et al. 1966), which suggests that the flux of salt into
the root is tied closely to water uptake. Weiper (1995) investigated the ultra-structure
of a range of mangrove absorption roots to identify any possible sites where such
exclusion might be operating; no such sites were found. What she did find, however,
was that (with the exception of Aegialitis annulata) the apoplastic pathway by which
salt may enter developing roots was regulated by the very early (within 2-5 mm of
the root tip) initiation of a suberized hypodermis. In other words, except for the
extreme root tip, ionic uptake is regulated via the symplasm (Weiper 1995).
Moreover, Popp et al. (1993) have questioned whether a passive ultrafiltration
is involved, arguing that a passive process would result in unchanged proportions of
Na+ and Cl- both in the nutrient solution and in the xylem sap. These relative ionic
proportions do not remain unchanged in Conocarpus erectus; with a highly saline
nutrient solution, concentrations in the xylem sap of Cl- merely doubled while Na+
increased ten-fold (Popp et al. 1993). Similarly in Avicennia marina, in increasing
saline solutions from 0% to 150% seawater, there was only a five-fold increase in
cr concentration but an 8-fold increase in Na+ (Tuan et al. 1995). Popp et al. (1993)
also questioned the chilling techniques and the metabolic inhibitors used by
Scholander (1968), advising the repetition of those experiments with modern
techniques. Nevertheless, it is clear that avoidance of salt intake is achieved by
special mechanisms in the root systems. Once in the xylem, however, no evidence
was found for any exclusion of salt from the leaves (Popp et al. 1993, Popp 1995).
It follows that if salt exclusion is so effective, then the localized build-up of
soil salinity will lead to such high levels that localized inhibitory salinization is
likely around the absorption roots of mangroves (Passioura et al. 1992, Hollins et
al. 2000). It seems clear that to avoid such excessive salinization, continued flushing
with tidal waters becomes an ecophysiological necessity for mangroves under highly
saline conditions. In other words, regular tidal inundation is necessary to remove the
build-up of salt in the mangrove root zone.
Despite the effective strategy of excluding 80-95% of the salt during water
uptake, the regulation of the remaining 5-20% of the salt that enters the roots is of
utmost importance for the survival of mangroves. Given that the mean lifespan of
leaves of mangroves (Table 3.1) ranges from around 3 months (in Excoecaria
52 Mangrove Ecology, Silviculture and Conservation

agallocha) to nearly 75 months (in Cynometra iripa), a range of efficient


mechanisms must exist in the mangrove flora to deal with the continuing intake of
salt and to protect the tissues from excessive salt build-up.
Table 3.1 Leaf turnover and mean leaf lifespans of various mangrove species.

Species Mean daily Mean daily Turnover Mean Latitude


leafing shedding lifespan
(per 1000) (per 1000) (per year) (months)
A. ilicifolius 3.67 2.93 1.07 11.2 22°S
A. annulata 3.37 4.52 1.65 7.3 22°S
5.32 4.47 1.63 7.4 24°S
A. comiculatum 1.67 1.68 0.61 19.6 22°S
1.62 1.42 0.52 23.2 24°S
2.08 1.63 0.59 20.2 29°S
A. marina 4.60 3.00 1.09 11.0 8°N
2.79 3.81 1.39 8.6 22°S
4.65 2.34 0.85 14.0 24°S
3.41 1.41 0.51 23.3 29°S
B. cyclindrica 5.10 2.60 0.95 12.6 8°N
B. exaristata 1.95 1.70 0.62 19.3 22°S
B. parvij/ora 1.67 1.13 0.41 29.1 22°S
B. gymnorhiza 0.93 12.9 ION
0.43 28.0 18°S
1.65 1.30 0.47 25.3 22°S
C. tagal 2.20 1.80 0.66 18.3 8°N
0.90 0.33 36.5 18°S
2.12 1.23 0.45 26.7 22°S
2.05 1.57 0.57 20.9 24°S
C. iripa 0.36 0.44 0.16 74.7 22°S
E.agallocha 30.06 9.82 3.58 3.3 22°S
6.85 6.44 2.35 5.1 24°S
18.60 9.75 3.56 3.4 29°S
H. littoralis 1.04 0.69 0.25 47.6 22°S
H. tiliaceus 3.43 3.5 ION
4.63 5.60 2.04 5.9 22°S
K. candel 1.04 11.5 22°N
L racemosa 8.49 4.89 1.78 6.7 22°S
4.08 4.65 1.70 7.1 24°S
L littorea 7.30 3.50 1.27 9.4 8°N
O. octodonta 3.09 1.64 0.60 20.0 22°S
3.60 2.03 0.74 16.2 240 5
R. apiculata 0.97 12.4 ION
1.80 2.20 0.80 14.9 8°N
0.57 21.0 18°S
R. lamarckii 0.52 23.0 18°S
R. mucronata 3.30 3.00 1.09 11.0 8°N
23.2 300S
R. stylosa 1.20 10.0 60 S
0.63 19.0 18°S
3.09 2.93 1.07 11.2 22°S
2.78 1.86 0.68 17.7 24°S
S. hydrophyl/acea 5.50 3.20 1.17 10.3 8°N
S. alba 1.82 6.6 ION
X. moluccensis 4.11 5.83 2.13 5.6 22°S
5.12 3.64 1.33 9.0 24°S
X. granatum 3.38 1.52 0.55 21.6 22°S

ION: Halmahera Island, Indonesia. Moriya et al. 1988; 6 0 S: Pari Island. Indonesia. 5ukardjo et al. 1987;
8°N: Phuket Island. Thailand. Christensen and Wium-Andersen 1977. Wium-Andersen and Christensen
1978. and Wium-Anderson 1981; 18°S: Duke et aI. 1984; 22°S: Proserpine. data collected by author
during 1982-1984; 22°N: Lee 1991; 24°S: Gladstone. data collected by author during 1974-1983; 300 5:
Mgeni Estuary. South Africa. Steinke 1988; 29°S: BaIlina, data collected by author during 1986-88.
Adapting to the 'Mangrove Environment' 53

3.2.2 Salt Extrusion


Salt extrusion occurs through salt glands (figs. 3.2 and 3.3) in the leaves of
Avicennia (Baylis 1940, Fitzgerald et al. 1992), Aegiceras (Cardale and Field 1971),
Aegialitis (Atkinson et al. 1967), Acanthus (Mullen 1931b), Clerodendrum inerme
(fig. 3.4), and possibly via cork warts in the leaves of Rhizophora (Baijnath and
Charles 1980). In A vicennia (Tuan et al. 1995) and Aegiceras (Cardale and Field
1971), salt glands are formed whether or not salt is present in the medium, whereas
they are reported to be entirely absent from Acanthus when grown in freshwater
(Mullan 1931 b).

Fig. 3.2 Salt glands on the adaxial leaf surface of (A) Acanthus ilicijolius and (B)
A vicennia marina. Scale bars 10 11m.

Joshi et al. (1975) concluded that among salt-extruding species Avicennia is the
most efficient and, consequently, is able to grow in highly saline conditions,
whereas the less efficient Acanthus and Aegiceras are restricted to less salty habitats.
The extrusion rates (Table 3.2) do not entirely support this contention; at high
salinities, the salt glands of Aegialitis annuJata can extrude around 30% of the
internal salt content of leaves in 12 hours (Popp et al. 1993, Weiper 1995) while
Avicennia marina and A. germinans can extrude 23% and 33% respectively of their
internal salt content in 24 hours (Tuan et al. 1995, Weiper 1995). Aegiceras
comiculatum was the least efficient, extruding a maximum of 2.1 % of its internal
NaCI content in 24 hours (Wei per 1995).
Boon and Allaway (1986) found that excised leaves of A vicennia marina
extruded chloride at markedly higher rates (up to 3.5 ~mol m-2 S-I) when their
petioles stood in concentrated solutions of sodium chloride, reaching an order of
magnitude greater than that commonly found for intact leaves on mangroves under
field conditions (Table 3.2). Maximal extrusion rates occurred when the excised
leaves were exposed to 1,000 mol m- 3 sodium chloride, and minimal rates when they
were exposed to dilute (<100 mol m- 3) or very concentrated (>2,000 mol m- 3)
solutions. Over the first 24 hours, the excised leaves maintained an approximate salt
balance by extruding as much Cl- as they took up in the transpiration stream from
NaCI concentrations of 100-2,000 mol m-3 (0.10-2.0 mol L-I ).
The salt glands of excised leaves of Avicennia marina also maintained their ion
extrusion specificity at low concentrations. Thus, when they were exposed to 10
mol m-3 solutions of NaCl, KCl or MgCI 2, they extruded mainly Na+, and very little
K+ or Mg2+ (Boon and Allaway 1986). In contrast, when excised leaves were exposed
54 Mangrove Ecology, Silviculture and Conservation

to 1,000 mol m- 3 solutions of NaCl, KCI or MgCI 2 , the cation they extruded was
mainly that specific to the treatment solution.
Table 3.2 Extrusion rates (J.1mol m- 2 leaf area S-I) of salt glands of various mangrove species.

Species NaClmol m-3 Extrusion Rate Ref.


root medium Na+ CI" NaCI
A. annulata 0.1-0_6 I
A. annulata 0.93 2
A. annulata
Young leaves 50 0.18 0.20 3
Old leaves 50 0_08 0.12 3
Young leaves 500 0.44 0.42 3
Old leaves 500 0.28 0.25 3
A. comiculatum 1.1-2.2 I
A. comiculatum
Young leaves 50 0.02 0.02 9
Old leaves 50 0.02 0.02 9
Young leaves 500 0_03 0_04 9
Old leaves 500 0.04 0.04 9
A. marina 804 0.12 4
A. marina 0.40 0.53 5
A. marina
Young leaves 154 0.16 0.22 6
Young leaves 599 0.53 0.59 6
A. germinans 50 0.05 7
400 0.33 0.30 7
A. germinans
Old leaves 50 0.18 0.15 9
Old leaves 500 0.40 0.45 9
Young leaves 500 0.45 0.45 9
A. marina
Excised leaves 1,000 3.5 8

IScholander et at. 1962; 2Atkinson et at. 1967; 3pOpp et at. 1993; 'Ish-Shalom-Gordon and Dubinsky
1993; 5800n and Allaway 1982; &ruan et at. 1995; 7Stewart and Popp 1987; 8800n and Allaway 1986;
9Weiper 1995.

Table 3.3 Na+ concentrations in young leaves and in the xylem sap of salt-extruding mangroves grown
in 100% seawater, together with the salt gland frequencies, extrusion rates during daylight, and the
efficiency of extrusion (as % extrusion of the leaf content). (Data from Weiper 1995).

Parameter A. annulata A. germinans A. comiculatum


Na+ conc. (mol m-2) 0.10 0.12 0.12
Na+ conc. in xylem (mol m-3) 25.4 17.4 5.9
Salt glands mm-2 12 60 8
Min. extrusi~n (J.1mol m-2 S-I~ 0.65 0.32 0_033
Max. extrusIOn (J.1mol m-2 s- ) 1.34 0.62 0.044
Extrusion per salt gland (nmol 12h- l ) 4.8 0.45 0.24
Efficiency (%) 34 17 1.4

As shown in Table 3.3, however, the extrusion rates and efficiencies vary
considerably in the three species examined in detail, not only in terms of the salinity
of the external medium and leaf age (Weiper 1995). Thus, Aegialitis annulata has the
least efficient salt exclusion in the roots while Aegiceras comiculatum has the most
efficient in 100% seawater; on the other hand, the extrusion rates of A. annulata are
very high when compared with those of A. comiculatum. Avicennia genninans
occupies an intermediate position in both instances. Given its high salt gland
frequency, when extrusion is expressed per salt gland, A. genninans and A.
comiculatum have extrusion rates an order of magnitude lower than those of A.
annulata. Combining all of these features shows that the amount of Na+ extruded, as
a percentage of the leaf content, is highest in the species (A. annulata) with least salt
Adapting to the 'Mangrove Environment' 55

exclusion, and is lowest in the species (A. corniculatum) with the highest salt
exclusion.
Ultrastructural studies of the glands of Aegiceras (Cardale and Field 1971,
Bostrom and Field 1973) have shown that they consist of 24-40 secretory cells
situated over a single large, basal cell. The secretory cells are densely packed with
mitochondria and other organelles, suggesting some metabolically active function.
The living contents of the basal cell and the secretory cells are linked by fine
cytoplasmic threads (plasmodesmata) that pass through the cell walls. On the other
hand, the junction between the basal cell and the sub-basal cells, which form a layer
above the palisade mesophyll, seems to be partially cutinized. Field et al. (1984)
showed, however, that there is a small slit-like opening between the cuticle of the
gland and that of the leaf; it is through this slit that salt extrusion occurs. The
mesophyll cells contain two types of vacuoles: one type contains large amounts of
an organic solute and little or no chloride whereas the other is free of organic solute
but rich in chloride (Van Steveninck et al. 1976). The fluxes of Na+, K+ and Cl- have
been measured using radioisotopes (Cardale and Field 1975), and all of these ions are
actively transported out of the parenchyma by the gland cells.

Fig. 3.3 Salt glands on the adaxial leaf surface of (A) Aegiceras corniculatum and (B)
Aegialitis annulata. Scale bars 10 J.1m.

A mechanism similar to that of Aegiceras operates in Aegialitis annulata; the


flow-path of salt was traced using CI, and it was found to pass directly from the leaf
veins via the palisade mesophyll to the salt glands (Atkinson et al. 1967). The salt
glands of Acanthus appear to have a similar ultrastructure; the vacuoles appear to
contain a fine precipitate, but nearer the epidermis this seems to be replaced by
round, dark vesicles (Wong and Ong 1984). It seems likely that these two vacuolar
inclusions correspond to the two types of vacuoles found in Aegiceras by Van
Steveninck et al. (1976).
The ultrastructure of the salt glands of A vicennia has been described as being
similar to those of Aegiceras, Aegialitis and Acanthus (Balsamo and Thomson
1993); unlike the other genera, however, Avicennia has smaller but no less active
glands on the abaxial leaf surface (Drennan and BeIjak 1979, Fitzgerald et al. 1992).
Petiolar glands of Laguncularia racemosa have been interpreted as extrafloral
nectaries, hydathodes, bacterial leaf nodules, salt glands, or simply as glands (Rumm
1944, Biebl and Kinzel 1965). Field-fixed leaves of Laguncularia racemosa show a
pair of flask-shaped bodies embedded at the distal end of each petiole (Kemis and
56 Mangrove Ecology, Silviculture and Conservation

Lersten 1984). Each of these bodies consists of a few layers of slightly elongate
cells lining a central cavity, which is filled with a secretory product consisting of
mannitol (Popp 1995). Externally, except for a couple of dimples, there is no
evidence of an opening. The zone of elongate cells is bounded by 2-3 layers of
small, flattened cells. The mode of secretion has not yet been determined but these
glands are not involved in salt extrusion.
Leaf glands also occur in the mesophyll of this species (fig. 3.4) and their
frequency appears to depend on the salinity of the medium in which the plant is
growing (Biebl and Kinzel 1965, Roth 1992).

Fig. 3.4 Leaf sections showing salt glands (A) on the leaf surface of Clerodendrum
inerme and (B) salt glands in the upper mesophylJ layer of Laguncularia racemosa.
Although no obvious openings can be seen, the glands of L racemosa are formed only
in the presence of salt and can be assumed to have a role in salt extrusion. Scale bars
JOllm.

Similarly in Conocarpus erectus, bilaterally positioned petiolar glands occur


which have previously been interpreted as hydathodes, salt glands or nectaries.
Kemis (1984) was able to show that these conspicuous glands are extrafloral
nectaries. Glands on a living plant secreted a clear sweet-tasting nectar consisting of
glucose, fructose, and sucrose. Each nectary consisted of a uniseriate palisade
epidermis, below which were several cell layers of dense cytoplasmic nectariferous
tissue. Phloem connections via small vascular bundles terminated in a thin-walled
subsecretory region, 2-5 cell layers beneath the actual secretory tissue.
The actual mechanism of salt extrusion is still not entirely clear; adenosine
triphosphate (ATP) localization in the stalk and secretory cells of Avicennia marina
confirms the involvement of active transport processes (Drennan et al. 1992). Young
leaves of Avicennia marina have a diurnal pattern of salt extrusion, peaking at noon
(Ish-Shalom-Gordon and Dubinsky 1993). Mature leaves have the same pattern of
extrusion, but do so at a slower rate. Extrusion appears to have a light-dependent
component which causes it markedly to decline at night in Aegialitis annulata
(Weiper 1995) or to continue, but at a slow rate during the night, in Avicennia
marina (Ish-Shalom-Gordon and Dubinsky 1993). In other species examined both in
the glasshouse and field, including Aegiceras comiculatum, Avicennia germinans and
A. marina (Weiper 1995), no differences in extrusion rates between day and night
were found.
Surprisingly, no significant temperature effects on extrusion rates could be
identified (Ish-Shalom-Gordon and Dubinsky 1993). The highest correlation between
salt extrusion and irradiance was found when a 2-hour lag between the data sets was
introduced, suggesting either that time is needed to build up salt concentration to a
Adapting to the 'Mangrove Environment' 57

threshold needed for salt extrusion, or that the lag represents the time needed to build
up photosynthate pools to generate the ATP to drive the extrusion process.
Salt is carried via the xylem to the leaf mesophyll from where it appears to be
temporarily relocated to the hypodermis before being moved to the secretory cells by
symplastic pathways (Fitzgerald and Allaway 1991). From the secretory cells, the
salt is released via a vesicle-like structure through pores in the cuticle (Ish-Shalom-
Gordon and Dubinsky 1990). When the vesicle-like structure ruptures, the contents
are ejected onto the surrounding leaf area and the vesicle-like structure disintegrates.
The salt glands of Avicennia marina then appear to move into an inactive phase
until a new extrusion cycle commences (Ish-Shalom-Gordon and Dubinsky 1990).
As a result, this is an asynchronous process with a large proportion of salt glands
inactive at anyone time.

3.2.3 Storage of Salt


Although Popp et al. (1993) and Popp (1995) found no evidence that salt, once in
the xylem. is excluded from the leaf, there is evidence that internal salt movement is
regulated. Some mangroves (Excoecaria, Lumnitzera, A vicennia, Osbomia,
Bruguiera, Rhizophora, Sonneratia and Xylocarpus) often deposit sodium and
chloride in the bark of stems and roots, and in older leaves (Atkinson et al. 1967,
Joshi et al. 1975, Clough and Attiwill 1975, Wang and Lin 1999). Joshi et al.
(1975) showed that prior to leaf fall in Sonneratia, Excoecaria and Lumnitzera, Na
and CI ions are deposited in senescent leaves. Senescing leaves in Rhizophora
mangle become depleted in K+ and enriched in Na+ (Werner and Stelzer 1990) while
in both Ceriops tagal and Rhizophora mucronata from Gazi Bay, Kenya, senescing
leaves were enriched in Cl- while N was depleted by around 50% in both species
(Slim et al. 1996). In Bruguiera gymnorhiza, Na and CI ion concentrations increased
by 45 and 31 % respectively during leaf senescence at the same time that 60% of its
N, 48% of its P, and 46% of its K was transferred out of the senescing leaf (Wang
and Lin 1999). In this way, excess salt is removed from metabolic tissue while
nutrients are conserved.
It is noteworthy that deciduous mangrove species (Excoecaria agal/Ocha,
Lumnitzera racemosa and Xylocarpus moluccensis) have relatively short leaf
lifespans (-4, 7 and 8 months respectively). Lumnitzera and Xylocarpus have a
congeneric species which is not deciduous and which has a longer leaf lifespan; thus,
for example, Lumnitzera littorea and Xylocarpus granatum have leaf lifespans of 9
and 21 months respectively. Deciduousness may be viewed as an adaptation whereby
leaf lifespans are shortened and co-ordinated with hydrological seasons so as to rid
the entire canopy of salt. For deciduous species of the genera Xylocarpus,
Lumnitzera and Excoecaria, annual leaf fall may be an effective mechanism for the
removal of excess salt prior to the onset of a new growing and fruiting season. On
the other hand, Cram et al. (2001) have argued that leaf fall is not a major
mechanism for a reduction in salt from non-deciduous mangroves in moist tropical
environments.
Interestingly, while Popp et al. (1984, 1993) were unable to show any build-up
in Na+ and ct- in senescent leaves of Ceriops tagal, Osbomia octodonta, Lumnitzera
racemosa and Conocarpus erectus, they did show a build up in S04- with leaf age in
Ceriops tagal. As S04 is less mobile than Na, CI and K, it may be that the build up
of S04 is the ultimate trigger for leaf senescence.
58 Mangrove Ecology, Silviculture and Conservation

The movement of salt into viviparous and cryptoviviparous seedlings while


still attached to the parent tree. is also regulated in Rhizophora. Ceriops. Bruguiera.
Aegiceras and Avicennia (Chapman 1944a, b. LOtschert and Liemann 1967. Joshi et
al. 1975). In Rhizophora mangle seedlings. there is an effective barrier between the
cotyledonary body and the peripheral tissues (L6tschert and Liemann 1967),
characterized by small, nearly spherical cells with papillose protuberances in the cell
walls (termed 'transfer cells' by Wise and Juncosa 1989), and by an increased
phosphatase activity (Pannier 1962), resulting in marked concentration gradients of
Cl, Na, K, Ca and N. LOtschert and Liemann (1967) concluded that the reduced salt
uptake by seedlings of R. mangle is accomplished by the activity of this glandular
tissue.
Smith and Snedaker (1995) planted propagules of R. mangle from high and low
salinity areas under both high and low salinity culture conditions to investigate
whether the salinity conditions of the seedling mother trees affected the subsequent
growth performance of the seedings. They found no evidence of pre-conditioning,
concluding that propagules did not appear to possess any differential capacity for
growth in salinity conditions similar to those in which they were produced.
Glandular tissue ('transfer cells') similar to that of Rhizophora mangle is also
present on the outside of the cotyledonary body in the Australian R. stylosa (Saenger
1982), and may be identical to the papillose layer described from R. stylosa and
Ceriops tagal by Carey (1934). Highly vacuolated, metabolically active cells also
have been described from the outer cotyledons of the propagules of A vicennia marirul
(Butler and Steinke 1976); these cells may have a similar regulatory role. Seedlings
taken from A vicennia marina growing on tidal mudflats had osmotic potentials more
negative than seawater, yet they contained little sodium or chloride (Downton 1982).
It appears that while still attached to the tree, developing seedlings can control the
uptake of sodium and chloride, and adjust osmotically by the accumulation of
organic rather than inorganic solutes (Downton 1982) but that, after falling, they
rapidly increase their salt content until their root system is capable of ultra-filtering
seawater (Chapman 1944. Field 1984).

3.2.4 Succulence
Leaf storage of salt (and Na+ in particular) is generally accompanied by succulence;
dilution of the accumulated salt by increasing water content per leaf area or shoot
volume is a common strategy of halophytes. While succulence has been interpreted
as a xeromorphic feature common in mangrove leaves, studies of Rhizophora
(Bowman 1921, Camilleri and Ribi 1983) and Sonneratia (Walter and Steiner 1936)
growing in saline and freshwater conditions, indicate that succulence is a response
primarily to the presence of Na+ and cr, although increases in leaf succulence with
leaf age have also been described (Walter and Steiner 1936).
The potential importance of changes in leaf succulence in regulating salt
concentrations was first shown for the non-extruding mangrove 1o,guncularia
racemosa (Biebl and Kinzel 1965). The chloride content of leaves of increasing age
was found to rise when calculated on a leaf area basis but remained more or less
constant when calculated on the basis of plant water per leaf area. In other non-
extruding species, succulence was shown to be essential for salt balance (Popp et al.
1993). These investigators found that the cr content per leaf area was highly
correlated with the water content per leaf area in three non-extruding species
Adapting to the 'Mangrove Environment' 59

(Rhizophora mangle, Laguncularla raceTIWsa and Conocarpus erectus), while no


correlation at all was found in Avicennia germinans.
Ball et al. (1988) showed that in members of the Rhizophoraceae, salinity
tolerance was positively correlated with succulence; the least salt-tolerant Bruguiera
gymnorhiza showed a mean succulence of 262.5 g plant water m-2 leaf area, followed
by Rhizophora apiculata (348.4 g water m-2), Rhizophora stylosa (387.9 g water m-
2) and Ceriops tagal (463.2 g water m-2).

Table 3.4 Degree of succulence in mangrove leaves. (Data compiled from Biebl and Kinzel 1965.
Karmarkar 1982. Camilleri and Ribi 1983. Kuraishi et aI. 1985a, Lacerda et aI. 1986b. Stewart and
Popp 1987. Smith et aI. 1989. Sobrado 2000, and unpublished data).

Species g I1lant water g plant water Mean leaf Moisture


m·2 leaf area g. dry weight area (mm2 ) (% F. Wt.)
A. ilicifolius 496.1 2.78 2.237 73.7
A. annulata 239.6 2.66 1.248
23L1 1.39 1.709
A. comiculatum 215.8 2.20 373
314.4 1.35 1.463 55.7
220.7 Lll 1.628
A. germinans 253.0 1.65 1.590
62.8
A. marina 336.8 2.21 502
308.3 1.15 1.330 57.4
370.3 1.36 1.704
A. schaueriana 71.7
B. exaristata 264.1 1.83 1.007
510.2 2.30 1.583
B. gymnorhiza 334.7 1.97 4.187 65.0
C. erectus 67.7
C. tagal 265.8 1.69 285
425.6 2.51 1.713
C. iripa 102.7 1.25 487 61.9
H. littoralis 128.6 0.98 13.060 49.2
H. tiliaceus 113.9 1.40 17.733 76.7
La. racemosa 1077.5 1.380
368.2 1.635
1140.4 1.845
70.3
321.0 2.52 1.840
L. racemosa 471.5 3.05 1.230 74.1
O. octodonta 320.0 1.83 344 62.2
R. apiculata 338.8 2.08 6.020 67.4
444.9 2.70 7.102 72.9
R. mangle 69.1
2.34 70.0
382.0 2.27 2.110
R. mucronata 516.3 2.31 7.844 69.9
R. stylosa 296.4 2.09 939
679.1 2.42 2.638
484.1 2.06 2.974 67.1
S. alba 465.5 2.54 1.912 71.0
S. apetala 3.566 71.4
S. caseolaris 336.5 2.78 4.545 73.6
S.ovata 342.5 2.10 5.810 68.2
T. populnea 180.9 2.30 13.100 69.7
X. granatum 227.7 2.49 3.820 72.0
X. moluccensis 481.0 2.95 2.266 72.1

Data on water content of mangrove leaves for variable leaf sizes (an assumed
surrogate for leaf age) are presented in Table 3.4; compiled from various sources am
collected under heterogenous environmental conditions, these data are not convincing
in terms of increasing succulence with leaf size (age). Nevertheless, the data ch
60 Mangrove Ecology, Silviculture and Conservation

indicate that in mangroves in general, water content is relatively high compared to


terrestrial species. It should also be noted that, where data are available, leaf
succulence appears to vary less with leaf size (leaf age) in salt-extruding species
(Avicennia marina) than in non-extruding ones (Ceriops tagal).
When comparing the water relations of Avicennia germinans and Conocarpus
erectus at high and low salinities during the wet and dry seasons in Venezuela,
Smith et al. (1989) found that leaf succulence barely changed in A. germinans (a
moisture content of around 63% on a fresh weight basis). In Conocarpus erectus,
however, leaf succulence was generally greater in old leaves than young leaves,
reflected in mean moisture contents of 65% in young leaves to around 75% in older
leaves. Leaf sections showed that the differences in succulence in these iosbilateral
leaves were attributable to an increase in cell lengths of the central mesophylliayers.

Fig. 3.5 Leaf section of Conocarpus erectus showing rnesophyll cells between the
abaxial (lower) surface and the central vascular bundles; these cells can enlarge
enormously when grown under highly saline conditions.

Anatomical factors contributing to succulence in leaves include the presence of


a well-developed, large-celled, water-storing hypodermis, a strongly developed
palisade mesophyll and generally small intercellular volumes (Saenger 1982, Rao
and Tan 1984, Das et al. 1995). Those species with isobilateral leaves (Lumnitzera,
Laguncularia, Conocarpus, Sonneratia and Pemphis) do not possess a hypodermis,
but in several of these, large undifferentiated mesophyll cells form a central water-
storing tissue (Saenger 1982, Roth 1992). Spongy mesophyll is absent from species
with isobilateralleaves and it generally forms less than 40% of the cross-sectional
area of those species with dorsi ventral leaves.

3.2.5 Compartmentalization
Although it is clear that the internal salt concentration in mangroves must be
maintained if turgor potential is to be constant, the metabolic effects of salt at the
cytoplasmic level are inadequately known. The salt concentrations found in leaves of
mangroves from the field are usually at levels where a range of enzymes from
halophytic plants are markedly inhibited. Thus, salt may potentially influence the
functioning of metabolic enzymes and therefore affect such vital processes as
respiration, photosynthesis and protein synthesis (Akatsu et al. 1996). For example,
Joshi et al. (1974, 1975) suggested that high salt concentrations in the cell inhibit
ribulose bisphosphate carboxylase, an enzyme of the carboxylation process. In
addition, activity of the enzyme malic dehydrogenase was significantly lower in
Adapting to the 'Mangrove Environment' 61

mangroves than in other plants, and this was attributed to salt inhibition and/or the
unavailability of calcium to the metabolic tissues.
Mizrachi et al. (1980) tested the response of seedlings of Avicennia germinans
and Rhizophora mangle to different salt concentrations and simultaneously
determined the chloride and nitrogen content (total N, protein N and amino N) ani
the rate of uptake of the labelled amino acid, leucine, in both leaves and roots. The
two species responded differently in some respects, but both showed a reduction in
leucine uptake with increasing soil salinity, indicating reduction in protein
synthesis. In R. mangle the amino N increased with increasing salinity, whereas in
A. marina there was an initial increase to a salinity of 9.6%0 followed by a rapid
decline. Amino N accumulated at all salt concentrations in the roots of both species.
In Avicennia marina. Ashihara et al. (1997) studied various enzymes related to
carbohydrate metabolism and their responses to increased salt concentrations. They
found that six out of seven of the enzymes studied decreased in activity as NaCl
concentrations increased and only aldolase (fructose-l,6-bisphosphate aldolase) was
stimulated to peak activity at about half seawater salt levels (250 mM salt). At 500
mM salt, all enzymes were strongly inhibited. Takemura et al. (2000) studied the
effects of increasing salt concentrations on the antioxidant enzymes, superoxide
dismutase (SOD) and catalase in Bruguiera gymnorhiza. These enzymes are
important to protect the plant from 'active' oxygen produced in response to stressors.
Both of these enzymes showed an immmediate increase after the plants were
transferred from water to high salinity, and retained full activity at least up to
seawater salt concentrations (5oomM). At 1,000 mM salt, the activities of both
enzymes were strongly inhibited.
These effects of salt on enzyme activity suggest that the enzymes of mangroves
and saltmarsh plants do not differ from those of other plants; the enzymes probably
would not function if they were directly in contact with the salt levels implied by
the overall salt content of the plant. Ball and Anderson (1986) tested the sensitivity
of the enzymes associated with photosystem II (PSII) of A vicennia marina and were
unable to detect any differences in their sensitivity from those from the glycophyte,
Pisum sativum.
Interestingly, when isolated chloroplasts were exposed to high concentrations of
NaCl, there appeared to be a stimulation of PS I and II in A vicennia marina
(Critchley 1982). Avicennia chloroplasts were found to require chloride for maximal
production of oxygen during photosynthesis. Similar requirements were
subsequently found in Avicennia germinans (Critchley et al. 1982) and in two
saltmarsh species (Critchley et al. 1982, Critchley 1983), and it was suggested that
these halophytes might preferentially accumulate chloride in the chloroplasts.
However, the results obtained by Ball et al. (1984) provide no evidence that the
photosynthetic membranes of salt-tolerant species are adapted to function at the
elevated salt concentrations found in leaves of salt-grown plants. The requirement for
high chloride concentration for maximal rates of photosynthetic O2 evolution in
Avicennia marina was an artefact arising from the loss of the 23kD polypeptide
during isolation ofthe thylakoids (Anderson et al. 1984). Thus, there is no evidence
to suggest that enzymes from mangroves are any more salt-resistant than those of
other plants (Ball and Anderson 1986, Stewart and Popp 1987).
How are enzymes kept out of contact with unfavourable salt levels? The most
likely possibility is that most of the salt is contained in the vacuole, and that in the
cytoplasm, where enzymes are located, low concentrations of salt are maintained.
62 Mangrove Ecology, Silviculture and Conservation

However, the osmotic potential of the cytoplasm and the vacuole must be balanced.
Consequently, other solutes which do not adversely affect enzyme function must be
in the cytoplasm at concentrations sufficient to achieve an osmotic potential equal to
that of the vacuole. Such a mechanism is supported by the ultrastructural
observations of two types of vacuoles - one rich in chlorides and the other, rich in
organic solutes (Van Steveninck et al. 1976).

3.2.60smocompensation
In various halophytes, several organic compounds have been found to function as
cytoplasmic osmoregulators (Table 3.5). For example, glycinebetaine, a quarternary
ammonium compound, was detected in the leaves of Avicennia marina (Wyn Jones
and Storey 1981, Ashihara et al. 1997). Downton (1982) has calculated that if this
compound occupies a cytoplasmic volume of 5-10% of the cell, then its reported
concentration is sufficient to balance total leaf osmotic potential. This is consistent
with the current view that halophytes successfully compartmentalize inorganic ions
in a way that salt-sensitive species do not, utilizing ions from the environment to
maintain vacuolar osmotic potential at a lower level than that of the external
solution, while protecting the salt-sensitive cytoplasm from dehydration and ion
excess by the substitution of compatible organic solutes (Downton 1982). Some
evidence also suggests that these osmoregulators may prevent heat damage. to
enzymes.
Other compounds with osmoregulatory properties found in various halophytes
include choline-O-sulphate, choline-O-phosphate, cyclitols, the amino acid proline,
and the sugar alcohol sorbitol (Benson and Atkinson 1967, Balasubramaniam et al.
1992, Popp et al. 1993). The amino N accumulation reported with increasing salt
concentrations in Avicennia germinans and Rhizophora mangle (Mizrachi et al.
1980) indicates that proline and glycinebetaine also act as osmoregulators in these
species.
Table 3.5 Organic solutes in young leaves of mangroves and associated species from Australia,
Venezuela and Sri Lanka. Concentrations are given in mol m·l plant water or mmol kg· l dry weights
(*). (Data from Popp 1984, Popp et al. 1984, Medina et al. 1990, and Popp et al. 1993).

Species Proline Quaternary Mannitol Cyclitols


ammonium
compounds
A. ilicifolius 10.9 39.1
A. aureum 19
A. annulata 28.5 90
A. comiculatum 0.6 9.5 287
A. germinans 58.0
A. m. eucalyptifolia 68.7
A. m. marina 61.1
B. cylindrica 356*
B. exaristata 15.8 144
B. gymnorhiza 7.3 82
B. sexangula 215*
C. australis 182
C. tagal 162
C. erectus 220
E. agallocha 6.1 11.6 89
H. linoralis 71.5
H. tiliaceus 101.6
La. racemosa 325
L linorea 1.6 112
L racemosa 6.3 133
Adapting to the 'Mangrove Environment' 63

Table 3.5 continued

Species Proline Quaternary Mannitol Cyclitols


ammonium
compounds
O. octodonta 0.2 6.3
R. apiculata 223
R. lamarckii 0.6 238
R. mangle 208
R. mucronata 199*
R. stylosa 5.7 178
S. hydrophyUacea 0.7 233
S. alba 210
S. apetala 173*
S. caseolaris 93*
X. granatum 76.0 15.5
X. moluccensis 56.7 9.2

From the above discussion, it seems that mangroves have developed a number
of measures to resist the high salt concentrations of their environment, ranging from
exclusion, to extrusion in some species, and the ability to compartmentalize
inorganic ions in a way that salt-sensitive species cannot do. However, although the
generally adverse effects of salt on whole plants are well documented, the metabolic
bases for these effects are still speculative: nitrogen metabolism, protein synthesis
(particularly oxygen evolving enhancer proteins; Sugihara et al. 2(00), purine
nucleotide synthesis, carboxylation enzyme inhibition, and reduced stomatal
conductance all appear to be involved to some extent.

3.3 Conserving Desalinated Water


Mangroves have an abundant supply of water around them at most times. However,
because this water is saline compared with the internal sap concentration of the
mangrove, it must be taken up against an osmotic gradient. An energy cost is
coupled to this process, and the real availability to the plant of water of reduced
salinity is determined by the amount of metabolic energy the plant can make
available for desalination. In other words, it is the high metabolic cost of this water
that underlies the physiological dryness of the mangrove environment.
Having obtained desalinated water at considerable metabolic cost, many
mangroves display features which combine to conserve water, generally at the
expense of both carbon acquisition and light interception capacity of the leaf (Ball et
al. 1988); thus, transpiration in mangroves is generally low and many mangroves
show xeromorphic features which tend to conserve or retain water. These
physiological and morphological features are discussed below.

3.3.1 Xerorrwrphic Features


Xeromorphy (or sclerophylly) has been interpreted as an adaptive response of plants
for water or nutrient conservation, or for protection against herbivory or high light
intensities (Kienholz 1926, Wylie 1949, 1954, Shields 1950, Stace 1966, Lugo
1985, Choong et al. 1992, Gon~alves-Alvim et al. 2001). These hypotheses have
mostly been derived from correlating morphological and anatomical features of
plants with the environmental attributes of their habitat (fig. 3.6), rather than from
direct experimental data. Thus, the exact nature of the trigger for xeromorphy in a
plant or plant community remains unknown, and the possibility that several factors
in combination result in xeromorphy cannot be ruled out. Alternatively, xeromorphy
64 Mangrove Ecology, Silviculture and Conservation

may equally be the result of similar responses in plants to a number of different


selective forces.
Leaves of most mangroves exhibit a range of xeromorphic features (Areschoug
1902, Stace 1966, Saenger 1982, Camilleri and Ribi 1983, Rao and Tan 1984,
Vasil'eva and Vasil'ev 1988, Choong et al. 1992, Roth 1992), although their water
conservation function remains controversial (Uphof 1941, Feller 1996). There are
limited experimental studies in support of water-conserving functions of such leaf
characteristics. The major xeromorphic features are discussed below.
All species of mangroves have a thick-walled, often multi-layered epidermis
which, at least on the upper leaf surface, is covered by a thick, waxy, lamellar
cuticle that would seem to retard evaporative loss. On the lower surface, often there
is a dense layer of variously shaped hairs (Avicennia, Heritiera fornes, Pemphis,
Hibiscus tiliaceus) or scales (Heritiera littoralis, Camptostemon, Thespesia
populnea). These usually cover salt glands and stomata (when present) and,
presumably, reduce the loss of water through these apertures.
60

% 2f.
50
T
T
N' 40- i::
..
iii
~

~ 30-
.
';I
...:I
~
T
T
20-
~
10

0 I I I I I

2-
;;-
iii T
T
iii ~
.!! 1.5 ~
~
eo
0
iii
2 1-
~ ~
~
.
';I
...:I
~
0.5 -

0 I

0 20 30 35 40 42 67

Salinity (ppt)

Fig. 3.6 Changes in leaf area and leaf xeromorphy (expressed as leaf biomass per unit
leaf area) in Avicennia germinans under differing salinities. (Data from Gon\ialves-
Alvim et al. 2(01)
Adapting to the 'Mangrove Environment' 65

Stomata permit the passage of gases into and out of the leaf, and it is through
these that much evaporative loss occurs. Three types of stomata have been reported
from mangroves, but neither Sidhu (1975) nor Ramassamy and Kannabiran (1996)
were unable to attach any ecological significance to them. With few exceptions,
stomata are restricted to the lower leaf epidermis in mangroves. In terms of frequency
and dimension, mangrove stomata are similar to those of plants of other habitats,
but many species show stomata sunk beneath the level of the epidermis, for
example, Avicennia, Aegiceras, Bruguiera, Ceriops, Heritiera Lumnitzera, Nypa and
Rhizophora (Saenger 1982, Tomlinson 1986, Das et al. 1995).
Several mangroves have cuticular ledges (or ridges) on the stoma (fig. 3.7),
resulting in the formation of substomatal chambers found in Avicennia, Ceriops,
Rhizophora, Bruguiera, Scyphiphora, Laguncularia, Kandelia, Excoecaria, Pemphis
and Heritiera (Areschoug 1902, Sidhu 1975, Das et al. 1995 Ramassamy and
Kannabiran 1996). Members of the Rhizophorae show outer and inner cuticular
ledges, enclosing an outer antechamber and an inner chamber in the stomatal pore
(Ramassamy and Kannabiran (1996). The outer ledges are simple in species of
Rhizophora and in some species of Bruguiera (e.g. B. cylindrica) but they are two-
lipped in species of Ceriops and other species of Bruguiera (e.g. B. gymnorhiza).
While such cuticular ledges and lips appear to provide highly effective seals, their
adaptive significance is unknown.

Fig. 3.7 Stomatal pores of (A) Avicennia bicolor and (8) Rhizophora mucronata,
showing the sunken stoma and substomatal chamber together with the cuticular ledges
which provide highly effective seals. Scale bars 10 J.1m.

Both Stace (1966) and Sidhu (1975) concluded that the presence of a thick
cuticle, wax coatings, sunken stomata, and the distribution of cutinized and
sclerenchymatous cells throughout the leaf, including the epidermis, are xeric
characters which probably developed in response to the physiological dryness of the
mangrove environment.
66 Mangrove Ecology, Silviculture and Conservation

In several species, including Avicennia. Bruguiera and Ceriops, the ends of the
vascular bundles are surrounded by irregular groups of tracheid cells which are much
larger than the conducting elements (Areschoug 1902). Their walls bear spiral,
reticulated or pitted thickenings and, as they possess a flange-like connection to the
hypodermis, a water-storage function has been attributed to them (Baylis 1940).
A well-developed hypodermis is a conspicuous xeromorphic feature of many
mangrove leaves and various functions have been ascribed to it, including water
storage, salt accumulation or osmoregulation, mesophyll protection via light back-
scattering and/or heat dissipation, nutrient conservation, and preventing fungal
infestations via its tannin content (Mullan 1931a, Uphof 1941, Saenger 1982,
Camilleri and Ribi 1983, Roth 1992, Ramassamy and Kannabiran 1996, Feller
1996, Koizumi et al. 1998).
Camilleri and Ribi (1983) examined leaf thickness and fresh- to dry-weight
ratios in Rhizophora mangle from areas of constant high salinity, fluctuating
salinities, and low salinities. They found that leaf thickness and fresh- to dry-weight
ratios were consistently higher in plants growing at sites of constant high salinity
than at sites of fluctuating or lower salinities. As leaf thickness was primarily due to
greater development of hypodermal tissue, they concluded that the hypodermis played
a role in osmoregulation.
Based on freeze-fracture electron microscopy of the membrane ultrastructure in
Avicennia germinans, Balsamo and Thomson (1995) noted that significant changes
occurred in the membranes of both hypodermal and mesophyll cells when exposed to
high salt concentrations (300 mmol L- 1 NaCl). Salinity triggered an increase in
intramembranous particles (interpreted as ion transport, carrier proteins) in
hypodermal membranes, indicating that the hypodermis is an important site of salt
accumulation and storage away from the salt-sensitive photosynthetic apparatus of
mesophyll cells. This appears to confirm the existence of a storage capacity for salt
in Avicennia which had been postulated by Drennan and Pammenter (1982) to
account for the apparent lack of correlation between salt extrusion and transpiration,
and seems to indicate an active osmoregulatory role for the hypodermis.
In Rhizophora mangle, a multi-layered hypodermis occurs (fig. 3.8); usually
two layers of tannin-filled cells occur towards the upper epidermis, followed by two
layers of water-storing cells with a lower tannin content (Roth 1992). Finally, there
is a layer of funnel-shaped mucilage cells, whose stratified thickened walls fill the
cell lumen. Roth (1992) concluded that these cells are able to absorb large amounts
of water while, at the same time, collecting and distributing light for the palisade
cells which surround the slime cells. Koizumi et al. (1998) experimentally
demonstrated that the hypodermis can protect the palisade mesophyll from direct
light while enhancing the ability of the leaf to efficiently use low light intensities.
Hypodermal thickness in Rhizophora mangle is enhanced by high light
intensity, particularly in seedlings and saplings (Farnsworth and Ellison 1996). On
the other hand, Feller (1996) postulated a nutrient conserving role for the
hypodermis in Rhizophora mangle, based on fertilization trials. The hypodermis was
considerably reduced, particularly in older leaves, with P-treatment (but not with N-
treatment); in unfertilized controls, the hypodermis in older leaves was 60% thicker
than in leaves on P-treated plants although area per leaf did not change significantly.
While Feller (1996) cites these results as experimental support for the
oligotrophic xeromorphism hypothesis, this hardly seems conclusive as other
xeromorphic features (such as multi-layered epidermis, sunken stomata, etc)
Adapting to the 'Mangrove Environment' 67

remained unchanged. However, while it suggests that the hypodermis may have a
significant nutrient conserving role in addition to its osmoregulatory and light
capturing roles, the adaptive significance of a variably sized hypodermis in relation
to nutrient conservation remains unclear.

Fig. 3.8 Typical leaf section of Rhizophora racemosa showing the layer of elongated
mucilege eells (me), the adaxial hypodermis (h) and the lower spongy mesophyll (sm).

Other xeromorphic characters found in some mangroves include a dominance of


palisade mesophyll at the expense of spongy mesophyll; when present, the spongy
mesophyll often is compacted and lacking in lacunae (Areschough 1902, Roth
1992). Stone cells are often widely distributed among the palisade cells as, for
example, in Avicennia, Rhizophora, Sonneratia and Xylocarpus, while sclereids have
been reported from Rhizophora, Scyphiphora, Nypa and Bruguiera. Mucilege cells
are a conspicuous feature in Rhizophora, but are also found in Acanthus, Aegiceras,
Bruguiera, Excoecaria, Scyphiphora, Sonneratia and Xylocarpus. These cells
undoubtedly give toughness and rigidity to the leaves, reduce damage from wilting,
and may be involved in conserving water (Malaviya 1963, Saenger 1982, Choong et
al. 1992, Roth 1992, Das et al. 1995).
The anatomy of the wood of some mangroves also appears to be influenced by
the highly negative xylem pressures induced by the high salinities of the mangrove
environment in association with waterlogging. Generally, mangroves possess more
vessels per unit area, with a larger total cross-sectional area, and distinctly smaller
pores than do their nearest inland relatives (Panshin 1932, Janssonius 1950, van
Vliet 1976, 1979) or even the same mangrove species growing under less stressful
conditions (Yanez-Espinosa et al. 2001). Aegiceras comiculatum, however, has
fewer vessels per unit area, shorter vessel elements, and smaller fibre-tracheid
diameters and ray heights when grown under saline (23%0) than under low salinity
(3.5%0) conditions (Sun and Lin 1997). These various wood modifications have been
postulated to be related to resistance to water movement in the conducting tissue
(Jansonnius 1950), to a reduced likelihood of rupturing the water column in narrow
vessels (Reinders-Gouwentak 1953), or to enhanced vertical and horizontal water
transport, movement of photosynthates and gas exchange (Yanez-Espinosa et al.
2001).
In the Rhizophoraceae, the pitting and the perforation plates of the vessels also
are modified: perforation plates are exclusively scalariform, as are the inter-vessel
pits with thickened vessel walls; vessel pit parenchyma is simple; and rays are
mostly 3-5 seriate (Marco 1935, Carlquist 1975, van Vliet 1976). This combination
of characters is not found in the inland genera of this family. Carlquist (1975) noted
that the perforation plate arrangement seems 'ideal to resist collapse in vessels under
68 Mangrove Ecology, Silviculture and Conservation

tension'. These features seem to indicate a reduced vulnerability to xylem collapse or


drought-induced embolism in the mangrove members of the Rhizophoraceae
compared to the upland genera of this family.
The vulnerability of xylem conduits to cavitation and embolism was compared
between the mangrove Rhizophora mangle and the tropical forest species
Cassipourea elliptica by Sperry et al. (1988), using both acoustic and hydraulic
methods of detection. Rhizophora mangle was the less vulnerable to embolism of
the two species, showing minimal loss of xylem conductivity at xylem pressures up
to -5.9 MPa, then losing 80% of its hydraulic conductivity between -5.9 and -7.0
MPa. Cassipourea elliptica, on the other hand, lost conductivity linearly with
decreasing xylem pressure from -{).5 to -7.0 MPa. The vulnerability of the species
correlated closely with the physiological demands of their respective habitats am
could be accounted for by differences in the measured air permeability of intervessel
pit membranes. Embolism in the xylem conduits appears to occur when air enters a
water-filled vessel under tension from a neighbouring air-filled one via pores in
shared pit membranes. This led Sperry et al. (1988) to conclude that the intervessel
pit membrane appears to be the 'Achilles heel' of the xylem in relation to salinity-
induced xylem embolism.
Clearly, these anatomical differences between Rhizophora mangle am
Cassipourea elliptica contribute to the marked differences in their vulnerability to
embolism, enabling R. mangle to grow in more saline habitats. However, the
carbon cost incurred in constructing xylem to minimize the occurrence and severity
of embolism is likely to be significantly higher in the saline habitat where lower
soil water potentials are frequent.

3.3.2 Transpiration
As indicated above, many mangroves show xeromorphic characteristics which may
be regarded as being involved in conserving water. Hence, it is not surprising that
the transpiration rates of mangroves are generally low when compared with the rates
of non-saline plants (Gessner 1967, Moore et al. 1972, Lugo et al. 1975, Kuraishi
et al. 1985a) and that mangroves have low stomatal conductances and unusually high
water use efficiencies for C3 plants (Ball 1988a).
Transpiration rates in mangroves are not intrinsically low; rather, low
transpiration rates are imposed on mangroves by the high concentrations of salt in
their environment. For example, during the wet season in high rainfall areas, high
rates of transpiration and stomatal conductance were observed in Rhizophora
apiculata and Avicennia alba based on transpiration rates (Kuraishi et al. 1985a) am
on sap flow rates estimated from heat pulse velocity in entire, field-grown trees
(Becker et al. 1997). Similarly, Sobrado (1999b) has shown that when grown
without salt. Avicennia germinans has high transpiration rates and high stomatal
conductance. When grown at low salinity Bruguiera gymnorhiza also has high rates
of transpiration (Takemura et al. 2000). However, when transferred to higher
salinities, there is a short-term increase in transpiration followed by a marked
decline. Nevertheless, it seems that mangroves generally follow a very conservative
water-use strategy, becoming even more conservative with increased environmental
stress (Ball 1988a. Clough and Sim 1989). For example, during the rainy season in
northern Venezuela. transpiration rates in Avicennia germinans were around 4 mmol
m·2 S-l around noon, while during the dry season they remained at <1 mmol m-2 S-l at
all times (Smith et al. 1989). Conocarpus erectus followed a similar pattern, with
Adapting to the 'Mangrove Environment' 69

rates around 5 mmol m-2 S-l during the wet season and <0.5 mmol m- 2 S-l during the
dry period. Both stomatal conductance and transpirational water loss were more
reduced in the dry season for C. erectus than for A. genninans, with the result that
C. erectus showed a 3.4-fold increase in instantaneous transpiration efficiency
(Smith et al. 1989) during dry conditions.
Mangroves can maintain or even increase water-use efficiency with increasing
salinity, at least over the range of salinities at which growth is vigorous (Ball and
Farquhar 1984a, b, Ball 1988b). As Ball (l988a, 1996) has demonstrated, water-use
efficiency also increases with increase in the salt tolerance of the species. In other
words, for a wide range of environmental factors affecting photosynthesis over the
course of a day, stomatal conductance at a given assimilation rate is lower (and water
use is more conservative) the greater the salinity tolerance of the particular species
(Ball et al. 1988).
Some of the features that enable mangroves to operate with high water-use
efficiency remain speCUlative. For example, it has been suggested that the highly
negative xylem pressures (tension) in mangroves is an artefact of the pressure
chamber technique (Zimmermann et al. 1994, 1995) and that sub-atmospheric, but
still positive, xylem pressure values are found when measured by a xylem pressure
probe (Balling and Zimmermann 1990). However, the presence of viscous,
polymeric compounds (mucopolysaccharides) which line the xylem ducts of
mangroves may control flow rates under low xylem tension, thus limiting
transpiration (Zimmermann et al. 1994). Furthermore, Sobrado (2000) has shown
that the hydraulic architecture of mangrove shoots and leaves is generally limiting
('in the lowest end of the range reported for tropical trees') and contributes to low
transpiration rates of the neotropical species investigated.
We must recall, however, that transpiration is a trade-off; in order to maintain
intercellular CO 2 concentrations, the stomata must open, and while CO 2 enters the
leaf, water vapour is simultaneously lost. Thus, carbon gain must be accompanied
by some water loss. Just how much water the plant can lose for a given carbon gain
is dependent on both stomatal conductance and water vapour gradient between the
leaf and air.
There is a close correlation between the assimilation rate (A) and stomatal
conductance (g) in a wide range of plants, including mangroves (Ball and Farquhar
1984a, Andrews et al. 1984, Andrews and Muller 1985). Such a close relationship
has been associated with minimizing water expenditure while maintaining
assimilation rates at, or very close to, the capacity of the mesophyll for
photosynthesis (Cowan 1986). Clough and Sim (1989) found stomatal conductances
to range from 79-271 mmol m- 2 s-t, with corresponding assimilation rates ranging
from 5.8-19.1 jJmol CO 2 m- 2 S-l in mangroves in a low salinity environment. An
even lower range of stomatal conductances (18-84 mmol m- 2 S-l) and assimilation
rates (2.5-10.3 jJmol CO 2 m-2 S-l) occurred in plants from highly saline
environments. It has been also suggested that the slope of the relationship between
the assimilation rate and stomatal conductance is steeper for plants adapted to arid
environments than for unadapted individuals (Clough and Sim 1989, Youssef and
Saenger 1998b).
While low stomatal conductance restricts the loss of water, it also limits the
influx of CO 2 , causing a leaf to operate with low intercellular CO 2 concentrations
and correspondingly low assimilation rates, but with a high water-use efficiency
(Ball and Farquhar 1984a, b). These gas exchange characteristics are reflected in the
70 Mangrove Ecology, Silviculture and Conservation

carbon isotope composition of leaves because isotope fractionation occurs during


diffusion and carboxylation (Farquhar et al. 1982). Values of 013C in mangroves,
ranging from -19.6 to -26.5%0, are consistent with fractionation predicted from their
low intercellular CO 2 concentrations and high water-use efficiencies. These values
are less negative than ol3C values of -30.5 to -35.9%0 in less water-use efficient
leaves of tropical moist lowland rainforest species (Ball 1996).
When water is abundantly available, water loss becomes secondary to carbon
gain. However, in marginal habitats where water is scarce, all carbon gains come at
a high cost, and any means whereby carbon gain can be achieved with minimal water
loss is a significant advantage (Dudley 1996a, b). This trade-off is usually expressed
as water use efficiency, a measure of the amount of CO 2 fixed per unit of water lost
or per unit of leaf conductance (Pezeshki et al. 1990). When measured per unit of
water lost (AlE), it is generally referred to as the instantaneous transpiration
efficiency (ITE) and expressed in !lmol CO 2 mmol- I H20. When based on stomatal
conductance (A/g), it is usually referred to as intrinsic water use efficiency (IWUE),
also expressed in !lmol CO 2 mmol- I H20. However, when measured per unit of
water lost per unit of vapour pressure deficit (vpd), it is generally referred to as the
water use efficiency (WUE) and expressed in !lmol CO2 /lffiOr l H20 kPa- l •
Transpiration has been most thoroughly investigated in A vicennia marina, at
several locations (Lewis and Naidoo 1970, Leshem and Levison 1972, Steinke 1979,
Attiwill and Clough 1980, Naidoo et al. 1997, 1998), and shows a steady increase
from dawn until it reaches a midday maximum at approximately 1200-1400 hours,
after which there is a steady decrease towards dusk. Under high temperatures, there
may be a mid-morning maximum rather than a midday one (Leshem and Levison
1972, Joshi et al. 1975, Steinke 1979). The afternoon decrease in transpiration does
not appear to be influenced by tidal inundation (Lewis and Naidoo 1970), and it is
assumed that the leaf water potential gradient was so steep that the consequent high
rate of transpiration induced an internal water deficit which resulted in the closure of
the stomata (Steinke 1979). Once the stomata were closed, the continuing high
temperatures prevent stomatal reopening, even when water was freely available
(Steinke 1979).
Naidoo et al. (1992) confirmed a marked midday minimum in leaf water
potential in the South African population of Avicennia marina of -3.5 MPa, while
Attiwill and Clough (1980) found that in a temperate population of this species in
Westernport Bay, Victoria, there was no change in water potential during the day,
even on days of sustained sunlight.
In Avicennia marina, a maximal transpiration rate of 2.1 mmol m- 2 S-I
(equivalent to 37.5 mg H20 m-2 S-I) was recorded in the field around midday (Naidoo
et al. 1992). In hydroponically grown seedlings of Avicennia marina, Youssef and
Saenger (1998a) found transpiration rates of 3.1 mmol m- 2 S-I under aerobic
conditions and rates of 1.8 mmol m-2 S-I under reduced conditions.
Similar diurnal patterns in transpiration rates have been found in all mangroves
investigated to date, including, for example, Aegialitis annulata (Naidoo and von
Willert 1995), Aegiceras comiculatum (Naidoo and von Willert 1995), A vicennia
germinans (Moore et al. 1972, Lugo et al. 1975, Smith et al. 1989), Bruguiera
gymnorhiza (Cheeseman et al. 1991, Naidoo et al. 1998), Conocarpus erectus
(Smith et al. 1989), Lumnitzera racemosa (Kuraishi et al. 1985a), Pelliciera
rhizophorae (Naidoo and von Willert 1999), Rhizophora apiculata (Kuraishi et al.
1985a), R. mangle (Lugo et al. 1975), R. stylosa (Andrews and Muller 1985),
Adapting to the 'Mangrove Environment' 71

Sonneratia ovata (Kuraishi et al. 1985a), Thespesia populnea (Kuraishi et al. 1985a),
and Xylocarpus granatum (Kuraishi et al. 1985a). Some comparative data are given
in Table 3.6.
Table 3.6 Transpiration rates (E in mmol m·2 S·I), water-use efficiency (lTE in Iflllol CO 2 mmol· 1 H 2O
or WUE in IImol CO2 IImol· 1 H 20 kPa"), and stomatal conductance (g in mrnol m· 2 S·I) from
mangroves. Ranges (-) or standard deviations (±) are given where available.

Species E lTE WUE g Ref.


A. annulata (n=6) 6.96±1.60 1.63±O.38 0.43±O.l5 5
A. comiculatum 2.49±O.l5 0.84±O.l1 1
A. comiculatum (n=75) 3.01±1.58 2.84±1.17 1.09±O.67 4
A. comiculatum (n=54) 3.77±1.61 1.61±O.47 0.37±O.l3 5
A. alba 300--400 8
A. alba (n=3) 4.45±O.32 10
A. germinans 3.0--5.1 6
A. germinans 6.l8±O.26 1.63 154 7
A. germinans (n=189) 2.28±O.73 2.70±1.05 160±79 9
A. marina 1.80±0.17 1.01±O.16 I
A. marina 0.50--2.10 78 3
A. marina (n=91) 3.l3±1.69 2.38±1.14 1.14±O.68 4
A. marina (n=66) 4.74±1.99 1.63±O.44 0.4D±D.14 5
A. marina (n=3) 3.70±0.36 10
B. gymnorhiza 1.92±O.11 0.97±O.l3 1
B. gymnorhiza 0.25-0.95 37 3
B. gymnorhiza (n=42) 1.37±O.67 3.56±1.17 1.59±O.81 4
B. exaristata (n=53) 3.95±1.41 1.51±O.40 0.36±O.l3 5
C. erectus (n=185) 2.36±O.81 2.89±1.02 16O±75 9
E. agallocha (n=65) 1.67±O.75 3.16±1.64 1.28±O.71 4
E. ovalis (n=54) 3.64±1.88 1.78±O.55 0.41±O.l7 5
H. tiliaceus 2.70±0.69 0.49±O.29 I
H. tiliaceus (n=13) 3.04±O.77 4. 17±1.04 1.61±O.62 4
La. racenwsa 2.9-6.4 6
La. racenwsa 6.83±O.10 1.22 150 7
La. racemosa (n=185) 2.52±O.77 2.98±1.09 I 82±79 9
L. racemosa (n=3) 2.30±0.18 10
R. apiculata 0-95 8
R. apiculata (n=3) 3.68±O.67 10
R. stylosa 2.04±O.ll 1.19±O.21 1
R. stylosa (n=15) 2.l2±O.58 3.08±1.58 1.10±.49 4
R. mangle 1.8--3.3 125-195 2
R. mangle 6.95±O.35 1.24 200 7
R. mangle (n=192) 1.95±O.73 2.42±1.03 130±66 9
R. mucronata (n=3) 3.55±O.38 10
S. alba (n=5) 3.84±O.90 1.92±O.46 0.51±O.12 5
T. populnea (n=3) 2.60±0.28 10
X. granatum (n=3) 3.64±O.48 10
'Youssef and Saenger 1998a; 2Lin and Sternberg 1992a; 3Naidoo et at 1992; 'unpublished field data,
Ballina; lunpublished field data, Darwin; 6Lovelock and Feller, in press; 7Sobrado 2000; 8Becker et at
1997; 9Snedaker and Arujo 1998; IOKuraishi et al. 1985a.

As the data in Table 3.6 show, transpiration rates are generally low but can
vary with environmental conditions. Compare, for example, the data for Avicennia
marina and Aegiceras comiculatum for Ballina and Darwin. The Darwin data were
collected in April, at the end of the wet season when water was abundant while the
data from Ballina were collected throughout the year; in the presence of abundant
water, transpiration rates increased markedly, accompanied by an equally conspicuous
decrease in ITE and WUE. While water availability undoubtedly influenced these
results, there is also evidence that because of the higher temperatures in Darwin,
72 Mangrove Ecology, Silviculture and Conservation

lower ITE and WUE resulted from the evaporative cooling of the leaf of Aegiceras
comiculatum (Youssef and Saenger 1998b).
Water loss in mangroves is also related to salinity (Scholander et al. 1962,
1965, Smith et al. 1989, Naidoo and von Willert 1999, Takemura et al. 2(00), with
those plants growing in highly saline situations transpiring less than those growing
in less saline conditions. This is partly due to the fact that the water is supplied to
the leaf at a considerable negative hydrostatic pressure potential, and the demand for
water, in terms of the vapour pressure difference at the leaf and air interface, often
can be high. In addition, maintenance of the osmotic potential of the xylem sap
(although considerably higher than the hydrostatic pressure potential) has a high
energy cost because water is taken up against an osmotic gradient. At extreme levels
of salinity, loss of water-use efficiency may occur (Ball and Farquhar 1984a, b, Ball
1988b).
Waterlogging may also depress water uptake and transpiration, although
inconsistent results have been reported. For example, rates of water uptake are lower
in waterlogged mangroves (Avicennia marina, Rhizophora mucronata and Bruguiera
gymnorhiza) than in those grown under well drained conditions (Naidoo 1985) and,
in hydroponic culture, transpiration rates decreased from 3.1 to 1.8 mmol m- 2 S-l
under anaerobic conditions (Youssef and Saenger 1998a). Such depression of
transpiration in hydroponically grown seedlings was most pronounced in Avicennia
marina and in Bruguiera gymnorhiza but insignificant in Aegiceras comiculatum,
Rhizophora stylosa and Hibiscus tiliaceus (Youssef and Saenger 1998a). Naidoo et
al. (1997) re-investigated changes in gas exchange in Avicennia marina under drained
(aerobic) and waterlogged (anaerobic) conditions; in contrast to Youssef and Saenger
(1998a), they found that under aerobic conditions, leaf conductance, transpiration and
internal CO 2 concentrations were generally lower and water-use efficiency (as ITE)
was higher than under anaerobic conditions.
Scholander et al. (1965) discussed some of the physiological costs of these
adaptations and indicated that there is a limit to the amount of water a plant can
effectively take up and transport against the osmotic gradient of its environment;
this limit is reflected in lower transpiration and higher respiration rates. In general,
mangrove roots take up water very slowly, primarily via symplastic pathways
(Moon et al. 1986, Lin and Sternberg 1993), and it appears that the mechanism
involved in salt exclusion from the transpiration stream may also restrict water flow
through the roots (Weiper 1995, Ball 1996). In any case, hydraulic conductances of
roots of Avicennia marina are extremely low, even when grown in freshwater (Field
1984). Thus, Ball (1996) suggested that with the large resistance to water flow, rapid
transpiration rates would produce such low water potentials in the leaves that an
impossibly high concentration of solutes in the cells would be required to maintain
turgor, and the risk of xylem embolism would be greatly enhanced.
Even with low transpiration rates, mangroves are exposed to highly negative
soil water potentials. For example, Naidoo (1989) monitored seasonal water
relations at monthly intervals for a year on Avicennia marina, Rhizophora
mucronata and Bruguiera gymnorhiza in a mangrove swamp in subtropical South
Africa. Soil salinities showed distinct seasonal trends, paralleled by soil water
potentials which ranged from -1.33 to -3.48 MPa. Soil moisture content ranged
from 19 to 40% but did not show any seasonal variation. Minimum leaf water
potentials at midday ranged from -3.28 to -5.97 MPa, being lower in summer
Adapting to the 'Mangrove Environment' 73

primarily as result of high evaporative demand. Maximum leaf water potentials at


dawn varied from -2.07 to --4.48 MPa with no distinct seasonal trends.
Evidence is just emerging that, apart from low transpiration rates, mangroves
deal with highly negative water potentials in a number of ways. In order to maintain
turgor through a range of soil water potentials, mangrove leaves have the ability not
only to increase solute concentrations but also to modify the elasticity of the cell
walls (Rada et al. 1989, Suarez et al. 1998). The elastic properties of cell walls are a
significant component of plant water relations as they affect the change in cell
volume for a given change in pressure and can determine ITE. With a decrease in
relative water content, leaves with more elastic cell walls (e.g. Avicennia germinans;
Suarez et al. 1998) can maintain higher turgor pressures, while those with more
rigid cell walls (e.g. Rhizophora mangle; Rada et al. 1989) can generate greater water
potential differences between leaves and soil. While the value of highly elastic cell
walls has been established in relation to drought resistance, the finding that cell wall
elasticity increased in seedlings of Avicennia germinans in response to salinity
increases (Suarez et al. 1998), suggests that this feature may also be an adaptation to
salt resistance.
From an evolutionary perspective, then, mangroves as a group have acquired a
conservative water-use strategy; low transpiration rates and elastic cell walls avoid
problems associated with very low water potentials, while minimizing the
accumulation of salt around the roots of plants growing in waterlogged soils which
may not be well flushed by tidal water (Passioura et al. 1992, Hollins et al. 2(00).
Given the evolutionary constraints discussed above, it should also be noted that
energy diverted to the uptake of water, or to the production of water-conserving
structures such as hairs, waxy cuticles and scales, is clearly not available for growth
and reproduction. Consequently, a number of ecological strategies have developed in
relation to this energy cost. In some species, the energy expenditure is reduced (1) by
growing only in less saline environments, or (2) by growing in bursts when fresh
water is available and 'marking time' during other periods. Other species have
increased the efficiency of energy production by such means as leaf orientation and a
large photosynthetic surface area. However, even with these strategies the net
productivity is decreased and dwarfing or stunting is likely to occur in highly saline
situations. Dwarfed mangrove systems allocate a larger portion of their energy
supply to respiration and low-loss recycling mechanisms, and a correspondingly
smaller proportion to growth (Lugo et al. 1975, Lin and Sternberg 1992a, b, 1993).

3.3.3 Optimizing Lea/Temperatures


For mangroves with their generally low rates of transpiration, maintammg leaf
temperatures close to air temperature is essential to minimize water loss ani
maximize water-use efficiency. Plants with high transpiration rates can take
advantage of the high irradiances required for maintenance of high photosynthetic
rates with minimal increase in leaf temperatures over air temperature through
evaporative cooling. In contrast, plants (such as mangroves) with highly
conservative water-use must avoid high irradiances if leaf temperatures are to be kept
within physiologically acceptable limits, particularly as evaporative cooling is
costly in terms of water use (Youssef and Saenger 1998b).
If mangroves can avoid high light intensities in the middle of the day, when
heat load is maximal, this would allow the leaves to maintain fairly constant, but
low, assimilation rates throughout the day. This would achieve a greater net carbon
74 Mangrove Ecology, Silviculture and Conservation

gain than if leaves were horizontal and subject to temperature-dependent inhibition of


photosynthesis for extended periods (Ball 1996, Tuffers et al. 1999).
Many mangroves are able to adjust the angle at which leaves are held as a
compromise between requirements for illumination and for maintenance of
favourable leaf temperatures with minimal evaporative cooling (Ball et al. 1988,
Lovelock and Clough 1992, Ball 1996). For example, measurements of leaf
temperatures high in the canopy of Rhizophora stylosa in northern Queensland in
full sunlight in summer showed that leaves at their natural inclination were often
over 5°C cooler than leaves experimentally held horizontal; they also had
correspondingly lower rates of water loss and higher photosynthetic rates than the
horizontal leaves (Clough et al. 1982). Similarly, when exposed canopy leaves of
Rhizophora apiculata were held in a horizontal position, leaf temperatures increased
from 4 to 11°C above ambient air temperatures of approximately 30°C, while
incident light increased from 1,430 to 2,085 ~mol m-2 S-I. In contrast, leaves left in
their natural almost vertical orientation avoided the maximal heat load around noon
when irradiance and air temperatures are greatest. Around noon, these leaves received
only 20% of available sunlight and were approximately 10°C cooler than they would
have been if fully exposed to the sun. Earlier and later in the day, the leaves received
about 500 ~mol m- 2 S-I and leaf temperatures were 30°C, conditions nearly optimal
for photosynthesis (Ball et al. 1988). Such maintenance of leaf temperatures close to
air temperatures by avoiding high irradiance in leaves of Rhizophora stylosa was
critical to maximizing the total integrated carbon gain for a minimum expenditure of
water during a day (Andrews and Muller 1985).
Interspecific differences in the display and properties of foliage reflect the
increasingly conservative water-use characteristics associated with increasing salinity
tolerance (Ball et al. 1988). This is clearly shown by variation in three major
characteristics of leaves that contribute to maintenance of favourable leaf
temperatures with minimal evaporative cooling: leaf angle, leaf size, and heat
capacity per unit leaf area.
Leaf angles in mangroves vary between species and leaf position or age. Thus,
young leaves of A vicennia marina are angled at -45° while mature leaves are held at
_0°, i.e. horizontally. Similarly in Bruguiera gymnorhiza, young leaves are held at
_60° while mature leaves are generally horizontal (Tiiffers et al. 1999). In
Rhizophora spp., Ceriops spp., Bruguiera exaristata, Osbomia octodonta aIXl
Lumnitzera racemosa, fully exposed leaves are held in a near vertical position.
Clough (1998) found that leaf angle decreased from -65° at the top of the canopy to
about 25° at the bottom of the canopy in stands of Bruguiera gymnorhiza aIXl
Rhizophora spp. Ball (1996) found that, in the Rhizophoraceae generally, the greater
the leaf angle, the greater the salinity tolerance of the species.
Leaf sizes in mangroves are variable (Table 3.4). A smaller leaf, however, can
intercept more light with less increase in temperature than a larger leaf. Thus, sun
leaves of mangroves are smaller than shade leaves. In the Rhizophoraceae, the
smallest leaf is the most salt-tolerant and characterizes the water conservative species
(Ball et al. 1988). It follows that leaves of mangrove species from low salinity areas
are much larger than those from species that dominate hypersaline environments.
Heat capacity per unit leaf area is a third leaf property influencing leaf
temperature (Ball 1996). The heat capacity of a leaf increases with dry weight aIXl
water content per unit area. The water content of mangrove leaves (as g plant water
m- 2 leaf area) is given in Table 3.4, indicating a range from 103 in Cynometra iripa
Adapting to the 'Mangrove Environment' 75

to 1,140 in Laguncularia racemosa. Among the Rhizophoraceae, specific leaf weight


and succulence, and thus also heat capacity, increase with salinity (Camilleri and
Ribi 1983), with exposure to sunlight (Ball et al. 1988), and with increase in the
salinity tolerance of the species (Ball 1988a). Accordingly, there is a tendency for
mangrove leaves to have a greater mass per unit area under conditions in which they
are most vulnerable to rapid fluctuations in leaf temperature, i.e. under intense
irradiation or where evaporative cooling is limited (Ball 1996).
As Ball (1996) has pointed out, the characteristics of foliage that contribute to
conservative water use in mangroves, are not without costs to the plant. Increasing
the angle of inclination reduces heat loading but is at the expense of light
harvesting; a large leaf angle can significantly reduce the incident light (and
temperature) around midday (Tiiffers et al. 1999) while marginally extending the
length ofthe photosynthetic day. To compensate for the overall loss of irradiance, a
larger leaf area index may be required to intercept an adequate amount of light.
Similarly, a decrease in leaf size, while enhancing heat transfer rates, requires a
greater investment in supportive and conductive tissue per unit of exposed leaf area
than in larger leaves. Finally, an increase in heat capacity of leaves provides
buffering against rapid changes in temperatures, but occurs at the expense of leaf
carbon which could otherwise be invested in leaf area expansion. Thus, maintaining
favourable leaf temperatures with minimal evaporative cooling is at the expense of
the assimilative capacity of the plant, and this expense increases as water use
becomes more conservative (Ball 1996).

3.4 Root Aeration in Waterlogged Soils


Mangroves occur in waterlogged soils that are low in oxygen (anaerobic) and of a
semi-fluid nature that provides limited physical support. The diffusion resistance to
oxygen is approximately 104 times higher in water than in air, as a result of which
there is a 30-fold decrease in pore water oxygen concentration from that in the
atmosphere. Thus, in waterlogged soils where biological and chemical consumption
of oxygen continues, this high resistance to oxygen diffusion becomes critical in
developing and maintaining anaerobic conditions in the soils. With the onset of
anaerobiosis, chemical transformations occur in the soils, resulting in the
accumulation of phytotoxins (such as sulfides), organic acids, and gases such as CH4
and H 2 S, and in the changed availability of nutrients (such as nitrates). As a
consequence of these changes with waterlogging, the root systems of many
mangroves display morphological, physiological and metabolic adaptations which
aid in overcoming these problems. These are considered in turn.

3.4.1 Morphological Adaptations


Below the surface, most mangroves possess a laterally spreading cable root system
with smaller, vertically descending anchor roots. The latter bear fine nutritive roots.
The root system is shallow, generally less than 2 metres deep. Tap roots have not
been observed. Despite such shallow root systems, the ratio of below- to above-
ground biomass is higher for mangroves than for other vegetation types (Saenger
1982, Snedaker 1995), particularly during early developmental stages. Matsui (1998)
found below-ground (root) biomass (BGB) to be correlated with above-ground
biomass (AGB) in Rhizophora stylosa to give the relationship: BGB = 0.394AGB.
Komiyama et al. (1987) developed the trench method for estimating root biomass
and the results obtained suggest that earlier studies underestimated root biomass.
76 Mangrove Ecology, Silviculture and Conservation

Despite this reservation, the consistently high below- to above-ground biomass ratio
of mangroves compared with other forests (Komiyama et al. 1987, 2000, Saintilan
1997a, b) suggests that mangroves allocate much of their net primary production to
roots. This may be an adaptation to unstable and generally anoxic substrate
conditions, although others have suggested it is the outcome of high soil salinity
(Ball 1988b, Saintilan 1997a, b). Additionally, Passioura et al. (1992) suggested
that shallow roots in mangroves are an adaptation to extract water from near the soil
surface to avoid the potential hazards of salinization of the soil and the difficulties of
coping with waterlogged conditions.
A range of above-ground root types has been reported from mangroves,
including: (1) pneumatophores - roots arising from the cable root system (e.g.
Avicennia. Xylocarpus moluccensis. Heritierafomes and Sonneratia) and extending
upward into the air as small conical projections; (2) knee-roots - modified sections
of the cable root which first grow upward above the substrate and then downward
again (e.g. Bruguiera. fig. 3.10); (3) stilt roots - generally branched roots that arise
from the trunk and grow into the substrate (e.g. Rhizophora and Ceriops. fig. 3.9);
(4) buttress roots - similar to stilt roots in origin but expanding into flattened,
blade-like structures (PelIiciera, Heritiera littoralis and Xylocarpus granatum. fig.
2.2); (5) aerial roots - generally unbranched roots arising from the trunk or lower
branches and descending downward but usually not reaching the substrate (e.g.
Rhizophora. Avicennia and Acanthus. figs. 1.1 and 3.13).

Fig. 3.9 The tangled stilt roots of Rhizophora stylosa show the potential for reducing
wave and tidal energy, thereby allowing sediment accumulation.

While most mangroves have one or more of these types, some species do not
possess a specialised root system (such as Aegialitis. Excoecaria) and their roots lie
near or on the substrate surface. As only relatively small surface areas are available
for the assimilation of oxygen in these species, they tend to be found on less
waterlogged soils (Excoecaria) or on coarser, more aerobic sediments (AegiaIitis).
However, Nypa, the mangrove palm, is an exception. It grows from an underground
Adapting to the 'Mangrove Environment' 77

rhizome and has no specialised aerial root system; yet, it is found in areas of frequent
inundation, generally in waterlogged soils (Tomlinson 1971).

Fig. 3.10 The knee-roots of 8ruguiera gymnorhiza. shown here on Okinawa Island.
Japan. show a classical pattern of negatively geotropic. followed by positively
geotropic. growth phases.

Fig. 3.11 A transverse section of the hypocotyl junction of Rhizophora stylosa showing
a fully developed lenticel. Scale bar represents I mm. (Photo. T. Youssef)

Evidence that these root structures are adaptations providing aeration for
subterranean roots and which physically anchor the plant comes from a variety of
sources. The most apparent is that those mangroves growing at lower tide levels,
and which are, consequently, more frequently inundated, tend to possess the greatest
78 Mangrove Ecology, Silviculture and Conservation

array of above-ground root types. The presence of aerenchymatous tissue and


numerous lenticels (figs. 3.11 and 3.12) in most above-ground and below-ground
roots provides further supporting evidence of their role in root ventilation (Roth
1965, 1992, Curran 1985, McKee and Mendelssohn 1987, Ish-Shalom-Gordon and
Dubinsky 1992, Rovenden and Allaway 1994, Youssef and Saenger 1996, Allaway
et al. 2(01).

Fig. 3.12 Transverse sections through seedling roots (-3 rom in diameter) showing the
extensive development of aerenchyma in (A) Acrostichum speciosum. (8) Bruguiera
gymnorhiza and (C) Avicennia marina. (0) shows the anastomosing network of cells
in the cortex of Avicennia marina at greater magnification. Scale bar represents 0.1
mm. (Photos. T. Youssef)

Curran et al. (1986) were able to demonstrate that the conductance of


pneumatophores in young plants of A vicennia marina was sufficiently large to
resupply the root internal gas space during a normal low tide when the
pneumatophores are exposed to the atmosphere. Curran et al. (1996) measured the
gas space in the roots of this species as ranging between 40-50% by volume. The
pneumatophores also contain chlorophyll and Dromgoole (1988) demonstrated that,
in addition to their primary role in root ventilation, the pneumatophores of the New
Zealand Avicennia marina photosynthesize. In fact, Dromgoole (1988) found that
during the day, these organs are almost autotrophic in terms of carbon balance and,
when corrected for the respiration of non-chlorophyllous tissue, they show
Adapting to the 'Mangrove Environment' 79

photosynthetic rates comparable to those of many other woody plants. Similar


findings have been made on Sonneratia alba and Rhizophora stylosa (Yabuki et al.
1985, Kitaya et al. 2(01).
The actual mechanism of air uptake, through the development of a negative gas
pressure, was investigated by Scholander et aI. (1955) in Rhizophora mangle aOO
Avicennia genninans and, more recently, in Avicennia marina by Allaway et al.
(2001). Scholander et aI. (1955) were able to show that oxygen concentrations inside
the roots declined during high tide, accompanied by a gradual reduction in gas
pressure inside the pneumatophores. During low tide, the low pressure could induce
bulk flow of air into the pneumatophore and allow the rapid re-establishment of
oxygen levels in the roots.

Fig. 3.13 A large specimen of Avicennia germinans on the shores of Lac Aheme,
Benin, a rapidly infilling system, showing prominent development of adventitious
aerial roots. (Photo. F. Blasco)

In A vicennia marina, developing pneumatophores initially show subrisules


followed by lenticels in mature regions (Hovenden and Allaway 1994, Allaway et al.
2(01). When the tide covers the pneumatophores, gas pressure and oxygen
concentrations decrease; pressure by around 1.7 kPa and oxygen by as much as 3
mol m-3• On exposure at low tide, pressure recovers immediately to atmospheric, but
oxygen levels rise more slowly, plateauing below the atmospheric oxygen
concentration. The changes in oxygen concentrations were attributed to simple
diffusion, with the contribution of the influx of air to pressure recovery representing
only a minor fraction of the respiratory oxygen requirement (Allaway et al. 2001).
The cyle of high water with falling oxygen and pressure followed by low water with
rising oxygen and sudden recovery to atmospheric pressure was repeated in the
experimental system over 8 tidal cycles and was altered only by rainfall during the
low tide period; when the pneumatophores are wetted by rain, diffusion via the
lenticels is apparently reduced.
The presence of aerobic conditions within roots was shown by direct
measurement using a polarographic oxygen mini-electrode (Andersen and Kristensen
80 Mangrove Ecology, Silviculture and Conservation

(1988). It is indicated also by leakage of oxygen from roots, which causes


surrounding soils to be more oxidized than would be expected in anaerobic sediments
(Nickerson and Thibodeau 1985, McKee and Mendelssohn 1987, McKee et al. 1988,
McKee 1993, Youssef and Saenger 1996). Such oxidation of the rhizosphere has
major implications for nutrient availability and uptake, and for plant-microbe
interactions under waterlogged conditions.
Mangroves presumably rely on root ventilation only in situations where poor
soil aeration occurs. Gessner (1967) showed experimentally that for Avicennia
germinans growing on coarse coral sand, presumably with reasonable drainage and
aeration, the removal of pneumatophores had no effect on the trees. On the other
hand, in situations where soil aeration is poor, continued covering of the above-
ground roots either by water or by flood-deposited sediments will cause widespread
mortality (Breen and Hill 1969) and a rapid degradation of root tissue (Albright
1976). Snedaker et al. (1981) described the development of anomalous aerial roots in
A vicennia germinans in response to stresses which limit the normal functioning of
pneumatophores - in this case an oil spill which coated the pneumatophores. A
similar response was shown by Avicennia TrUlrina in the Arabian Gulf following the
massive oil spill resulting from the Gulf War (BOer 1993).
The mangrove root system must also be able to respond to changing
sedimentation patterns. Under quiet sedimentary conditions, mud commonly
accumulates at a rate of up to 0.5 cm y-I (Table 4.5 in 4.2.4), but in active areas
such as the Ganges/Mengha/Brahmaputra delta in Bangladesh, localized accumulation
rates of up to 1 my-I have been observed. The complex aerial root systems of
mangroves facilitate this process; Young and Harvey (1996) were able to
demonstrate significant correlation (r=0.72, P<O.OI) between accretion and
pneumatophore densities, ranging from 100-350 pneumatophores m- 2, over a three
monthly interval at Piako River mouth, New Zealand.
Clearly, the mangrove root system must be able to respond in ways that
continue to provide aeration for the roots, as a failure to do so will lead to stress or
mortality (Hutchings and Saenger 1987, Hatton and Couto 1992). In tum, this
suggests that it may be disadvantageous for individual mangroves to produce a high
density of pneumatophores as accretion may be too rapid. In fact, Young and Harvey
(1996) have suggested that mangroves at their study site produce pneumatophores at
a density that balances the benefits of aeration with the drawbacks of sediment
accretion - a positive feedback 'switch' (Wilson and Agnew 1992). Equally,
pneumatophores that continue to grow upwards are an adaptation found in
Avicennia, Xylocarpus moluccensis and Sonneratia. This root growth pattern
appears to maximize the benefits of aeration and accretion (i.e. supply of nutrients).
Other species cope with sediment accumulation by forming extra arches of the stilt
roots (Rhizophora), additional knee roots (Bruguiera), adventitous aerial roots
(Avicennia, fig. 3.13), or the upward secondary thickening of roots or buttresses (X.
granatum and Heritiera spp.).
In exposed situations where sediments are removed by erosion and do not
accumulate, the problem is reversed and plants must remain firmly anchored in order
to survive. Massive development of the various above-ground roots may serve to
reduce water movement around the plant, and an extensive though shallow cable root
system, as in A vicennia and Osbomia for example, may effectively anchor the plant.
Thorn et al. (1975) described the successful colonization of tidal sand flats in the
Adapting to the 'Mangrove Environment' 81

Joseph Bonaparte Gulf, Western Australia, by Avicennia despite rigorous conditions


imposed by the passage of long-period waves.
The effect of water movement and inundation on above-ground root
development of Kandelia candel, was described by Hosokawa et al. (1977). In this
species, the basal part of the stem is buttress-like on plants near creeks with flowing
water, but in comparatively still water typical stilt roots are formed.
The anatomical changes occurring in the aerial and stilt roots of Rhizophora
mangle on penetrating the substrate have been described by Bowman (1921), Gill
and Tomlinson (1971a, 1975, 1977) and Karsted and Parameswaran (1976); the
external colour of the root changes from tan to white as the thin surface layers lose
their chlorophyll, thickened walls are no longer formed, and the ground parenchyma
contains many gas-filled spaces. The aerial root has approximately 5% gas space
before penetration into the substrate compared with about 50% after penetration.
In seedlings of six mangrove species which had not developed pneumatophores
or aerial roots, Youssef and Saenger (1996) showed that significant movement of
oxygen occurs via aerenchyma from aerial parts of the shoots to the absorption roots
despite a constriction of the air flow path at the stem-hypocotyl junction. Air flow
was facilitated by high root porosity, ranging from 45.7% in Avicennia marina, to
Rhizophora stylosa, Bruguiera gymnorhiza and Aegiceras corniculatum (27.9%,
30.0%, 27.4% respectively) and to Hibiscus tiliaceus and Excoecaria agallocha
(14.8% and 17.8% respectively). They suggested that mangrove seedlings appear to
compromise between the need to maintain aerobic metabolism in the apical
meristem of the roots, and rhizosphere detoxification through oxygen leakage. This
compromise is achieved by developing on the root surface an effective barrier to
oxygen leakage and by increasing root porosity which, in turn, increases oxygen
reserves within the root while reducing the volume of respiring tissue.
In some freshwater wetland species, a 'thermo-osmotic' ventilation mechanism
produces a mass flow of gas as a result of pressure differences generated in young
leaves (Grosse and Mevi-Schutz 1987, Armstrong et al. 1988). However, Skelton
and Allaway (1995) found no evidence for mass flow in seedlings of Avicennia
marina with the rates of tracer gas movement consistent with diffusive movement.
Both in seedlings and in adult plants, the development of aerenchyma tissue in
above- and below-ground roots may serve either as an oxygen reservoir, or as a
system which allows the maximum volume of root or rhizome per quantity of
living tissue, thereby achieving an economy in oxygen consumption per unit
volume. The ability of some above-ground roots to photosynthesize and add oxygen
directly into the below-ground roots (Yabuki et al. 1985, Dromgoole 1988) further
enhances the oxygen economy. Nevertheless, the facilitation of gaseous diffusion by
morphological adaptation is not an entirely satisfactory explanation of waterlogging
tolerance in mangroves, and the possible role of physiological and metabolic
adaptations to anaerobic conditions must also be considered (McKee and
Mendelssohn 1987, Youssef 1995).

3.4.2 Physiological Adaptations


Physiological evidence of the aerating function of stilt roots in Rhizophora was
obtained by Canoy (1975), who noted an increase in the number of stilt roots
produced per square metre with increased temperature (and consequently reduced
dissolved oxygen concentrations) in a thermally polluted environment.
82 Mangrove Ecology, Silviculture and Conservation

One immediate effect of waterlogged soils appears to be the induction of a root


oxygen stress which, in turn, leads to decreased permeability of the root membranes,
thereby inducing water stress on the plant. In the short-term, this generally leads to a
reduction in water uptake and transpiration (Naidoo 1985, Youssef and Saenger
1998a).
In the longer-term, auxin accumulates in the stem of the waterlogged plant, arxl
the high levels of this growth hormone result in the formation of adventitious roots.
For example, Snedaker et al. (1981) reported intense aerial root formation in
Avicennia germinans growing in areas subject to oil spills and prolonged
waterlogging. By this strategy it seems that partial restoration of root function
occurs, and water stress in the plant is reduced.
Waterlogging may also lead to decreased cytokinin export from the waterlogged
root system (Itai et al. 1968) and the accumulation of abscissic acid in the leaves
which, in turn, can result in stomatal closure, rapid leaf senescence and shedding, and
a retardation of shoot development and elongation. Growth rates of the shoots may
also be reduced by the altered giberellin balance or the accumulation of ethylene in
the plant (Crawford 1978). All of these responses reduce the water stress induced by
soil anoxia.
Anoxia is generally accompanied by the presence of soluble phytotoxins,
including sulfides, and reduced iron and manganese (Boto 1984, McKee 1993).
Youssef and Saenger (1998a) investigated the effects of root anoxia with and without
the presence of various phytotoxins, singly and in combination, on seedlings of five
species of mangroves. Anoxic treatment per se had little or no effect on the
seedlings, except for those of Bruguiera gymnorhiza and Hibiscus tiliaceus, where a
reduction in carbon assimilation was noted. Reduced iron and reduced manganese
caused some effects in some species; reduced iron lowered carbon assimilation in
A vicennia marina and Bruguiera gymnorhiza, while reduced manganese lowered
carbon assimilation in Aegiceras comiculatum. However, the presence of sulfide (2
mmol) resulted in a complete inhibition of photosynthesis with complete stomatal
closure in all species. Similar effects were noted in all species when a mixture of the
three phytotoxins was added. From these investigations, it is apparent that the
phytotoxins, particularly sulfides, are the major agents determining survival arxl
growth of mangrove seedlings in anoxic soils.
Kimura and Wada (1989) investigated the role of tannins in mangrove tree roots
to determine whether root tannins might alleviate the detrimental effects of the soil
environment. They found that the roots of A vicennia, Rhizophora and Bruguiera
contain a large amount of tannins which not only suppressed the growth of
anaerobic microbes but also combined with ferric ion or H 2 S in the soil solution
resulting in the blackening of the roots due to the formation of a tannin-ferric iron
complex. Based on these phenomena, it was considered that the tannins in mangrove
tree roots do playa significant role in minimizing the phytotoxic effects of reduced
iron and sulfides in the soil solution.
Despite the effective systems for root ventilation in mature mangroves,
waterlogging can adversely affect growth and functioning of mangroves (Naidoo
1983, 1985, Pezeshki et al. 1990, McKee 1993, Youssef and Saenger 1998a). For
example, rates of water uptake tend to be lower in waterlogged mangroves than in
those grown in aerobic conditions (Naidoo 1985, Youssef and Saenger 1998a).
Similarly, mangrove seedlings grown under waterlogged culture conditions showed
increased foliar Na+, and depressed water potentials, stomatal conductance,
Adapting to the 'Mangrove Environment' 83

photosynthesis, oxygen transport through aerenchyma, and growth (Naidoo 1985,


Hovenden et al. 1995). Tolerance to increasing salinity is also reduced by anaerobic
conditions caused by waterlogging, possibly because anoxia interferes both with salt
exclusion and selectivity for K+ over Na+, which can lead to lower shoot
concentrations of K+ and hence to reduction in growth or to metabolic dysfunction
(Ball et al. 1987). High levels of H2S in anaerobic soils can also inhibit nitrogen
uptake and growth (McKee 1993). These effects of anaerobiosis need to be dealt with
metabolically.
McKee (1996) studied the physiological responses of seedlings of Avicennia
germinans, Rhizophora mangle and Laguncularia racerrwsa to hypoxia. Root
hypoxia led to significant decreases in root respiration rates, and in the rates of root
extension in A. marina and L. racerrwsa, but not in R. mangle. Seedlings of the
three species showed little change in net carbon assimilation in reponse to low
oxygen concentrations, suggesting that the balance between carbon acquisitionj ani
carbon loss through root respiration is maintained during low oxygen conditions.

3.4.3 Metabolic Adaptations


Experiments with species other than mangroves during waterlogging have
demonstrated a range of metabolic responses to anaerobic conditions (Crawford
1992). In most plants, products of fermentation (including ethanol, lactic acid ani
malic acid) accumulate to some extent. Ethanol can become toxic to plants in high
concentrations, although toxic levels are rarely attained because of ethanol leakage.
However, when the tissue is re-oxygenated, ethanol is re-metabolized to
acetaldehyde, a strong oxidizing agent that may cause post-anoxia problems. The
carbon loss through ethanol leakage and potential post-anoxia stresses suggest that
pathways other than ethanol production would be more suitable.
Indeed, Crawford and Tyler (1969) examined organic acid accumulation in a
selection of species ranging from those tolerant to those intolerant of waterlogging,
and found that there was an immediate change in organic acid metabolism with the
advent of flooding. This study suggested that under anaerobic conditions malic acid
is more important than ethanol as an end product, that malic, shikimic and quinic
acids accumulate in the roots and rhizomes, and that succinic acid is a common
product of an anaerobic respiration.
The significance of malic acid accumulation is threefold. First, malic acid, like
other organic acids, can accumulate in plant cells in considerable quantities without
injury to the plant, and in this respect differs considerably from ethanol. Second,
malic acid is an alternative product to ethanol in anaerobic respiration in plants ani
it can subsequently be metabolized on the return of aerobic conditions. Third, in
cation absorption, malic acid accumulates in greatest quantity and, thus, preserves
the electrical neutrality of the cell.
The metabolic responses of seedlings of six mangrove species to short- ani
long-term anoxia were investigated by Youssef (1995). After 10 days of anoxic
treatment, there was a significant increase in the amount of ethanol accumulated in
the roots of Aegiceras comiculatum, A vicennia marina, Excoecaria agallocha ani
Rhizophora stylosa (fig. 3.14). On the other hand, neither Bruguiera gymnorhiza nor
Hibiscus tiliaceus showed a significant change in root ethanol concentration. Some
of these differences were due to different rates of leakage to the medium, with
Hibiscus tiliaceus and Excoecaria agallocha losing between 45-75 Ilmol g-l dIy
weight to the external medium. However, it seems that the short-term accumulation
84 Mangrove Ecology, Silviculture and Conservation

of ethanol as the end-product of glycolysis occurs in most mangroves. While the


leakage of ethanol to the external medium might avoid possible toxicity problems,
it seems a costly carbon drain. Interestingly, only Avicennia marina maintained
elevated levels of ethanol under long-term (70 d) anoxic treatment (Youssef 1995).

Ethanol concentration (lJIIlol gDwt -1 )


o 25 50 75 100

H. tiliaceus
E. agallocha 1=======::----
A. comiculatum
B. gymnorhiza
Ethanol accumulation •
R. stylosa
A. marina Ethanolleakage D
Fig. 3.14 Ethanol accumulation and leakage to the external media from roots of
seedlings of six mangrove species subjected to short -term root anoxia. (From Youssef
1995, reprinted with permission)

In terms of organic acid accumulation, under short-term anoxia all species


except Rhizophora stylosa showed increased concentrations of shikimic acid when
compared with the controls but no increases in malic or lactic acids were found
(Youssef 1995). Under long-term anoxic treatment, all species except Rhizophora
stylosa showed a significant increase in malic acid, no species showed any increase
in shikimic acid, and only Rhizophora stylosa showed a significant accumulation of
lactic acid.
Mangroves appear to have a range of metabolic responses to anoxia. In the
short term there is an increase in alcohol dehydrogenase activity (Stewart and Popp
1987) which leads to the initial accumulation of ethanol which, in turn, may be
leaked to the medium or subsequently metabolized once aerobic conditions have been
restored (Youssef 1995). At the same time, there is a reliance on the shikimic acid
pathway. In the long-term, however, there is a change from the ethanol and shikimic
acid pathways to malic acid in the majority of species. As malic acid is also
involved in ionic balance and salt uptake (Werner and Stelzer 1990), this seems an
optimal strategy.

3.5 Surviving Seawater Dispersal


3.5.1 Vivipary and Cryptovivipary
Various types of fruits are found among the mangroves (Table 3.7). In several
genera, the fruit contains seeds which develop precociously; the seed germinates
while still attached to the parent tree. In these species, the embryo develops into a
seedling without any dormant period (Gill and Tomlinson 1969), although a form of
seedling dormancy may be induced by low water content (Sussex 1975). Where the
embryo ruptures the pericarp and grows beyond it, sometimes to considerable
lengths, while still attached to the parent tree (as in all members of the
Rhizophorae), it is referred to as 'vivipary' (Saenger 1982, Tomlinson 1986). In
Adapting to the 'Mangrove Environment' 85

Aegialitis, Avicennia, Aegiceras, Laguncularia, Nypa and Pelliciera, the embryo


develops within the fruit but does not enlarge sufficiently to rupture the pericarp.
This condition is termed 'cryptovivipary'. In the remaining mangrove species, the
seeds, like those of most plants, pass through a resting stage prior to germination,
and do not germinate while still on the parent tree. The resting stage may be very
short (e.g. less than one day in Acanthus) or range up to a month (e.g Cynometra.
Xylocarpus and Heritiera). Interestingly, viviparous and non-viviparous species have
similar buoyancy, seed weight and rates of root and shoot initiation, as well as early
growth and salinity tolerance (Clarke et al. 2(01).
Table 3.7 Reproductive and dispersal units of common mangrove genera. (Compiled from Saenger
1982, Tomlinson 1986, Duke et al. 1998, and Naskar and MandaI 1999)

Genus Reproductive unit Dispersal unit


Acanthus Capsule with 4 flat seeds Capsule
Acrostichum Spore Spore and prothallus
Aegialitis Indehiscent nut Nut
Aegiceras Fleshy capsule, shed with Capsule
calyx attached
Avicennia Fleshy capsule with single Capsule
seed
Bruguiera Fleshy, single-seeded berry, Berry: viviparous
shed with calyx attached hypocotyl
Camptostemon Capsule with 2 to several Capsule
woolly seeds
Ceriops Fleshy berry, usually single- Berry: viviparous
seeded hypocotyl
Conocarpus Cluster of laterally compressed Nutlet
nutlets
Cynometra Wrinkled, one-seeded pod Pod
Excoecaria Trilobed, exploding capsule, Seed
each lobe one-seeded
Heritiera Clusters of woody, keeled Carpels
carpels
Hibiscus Hairy capsules splitting into Seed
5 locules, each with 3 seeds
Kandelia Fleshy, single-seeded berry Berry: viviparous
hypocotyl
Laguncu/aria Indehiscent woody drupe Drupe
Lumnitzera Indehiscent woody drupe with Drupe
thin outer fleshy layer
Nypa Aggregate head of one-seeded Aggregate head
fruits
Osbomia Capsule Capsule
Pelliciera Single-seeded fruit Seed
Pemphis Many-seeded capsule Capsule
Rhizophora Ovoid fleshy berry, usually Berry: viviparous
one-seeded hypocotyI
Scyphiphora Woody capsule, usually with 4 Capsule
seeds
Sonneratia Many-celled, many-seeded Capsule
capsule
Xylocarpus Several-seeded capsule Seed

The absence of seedling dormancy in many mangroves and the rapid loss of
viability if the propagules are dried out appear to be related to the level of abscisic
acid (ABA) in the embryonic tissue (Farrant et al. 1992, Farnsworth and Farrant
1998). In plants generally, ABA co-ordinates both the development of desiccation
tolerance during the onset of seed dormany and the plant's response to flooding.
Viviparous and cryptoviviparous mangroves (Rhizophora. Bruguiera. Ceriops.
Kandelia. Rhizophoraceae; Aegiceras, Myrsinaceae; Aegialitis, Plumbaginaceae; and
86 Mangrove Ecology, Silviculture and Conservation

Nypa. Arecaceae) consistently exhibited lower levels of ABA in embryonic tissues


throughout embryogeny relative to their non-viviparous upland sister taxa
(Cassipourea, Rhizophoraceae; Ardisia, Myrsinaceae; Limonium, Plumbaginaceae;
and Phoenix, Arecaceae). Moreover, while ABA levels were consistently low in
intact mangrove embryonic tissues, mangrove embryos subjected to drying out
exhibited elevated ABA levels, demonstrating that the mangrove embryos had not
forfeited the capacity to produce ABA when required (Farnsworth and Farrant 1998).
It seems, however, that the trade-off for waterlogging tolerance of mangroves is the
absence of seed dormancy coupled with a high seed sensitivity to desiccation.
The frequency with which vivipary and cryptovivipary occurs in mangroves has
prompted the question of whether this is an adaptation to some aspect of the
mangrove environment. The adaptive significance of vivipary has been postulated to
include rapid rooting (Macnae 1968), salt regulation (Joshi 1933, Lotschert and
Liemann 1967), ionic balance (Joshi et al. 1972), development of buoyancy (Gill
1975), prolonged attainment of nutrients from the parent either as a form of
nutritional parasitism (Pannier and Pannier 1975, Bhosale and Shinde 1983, Lin and
Sternberg 1995) or as subsidizing the maternal reserves by hypocotyl photosynthesis
(Smith and Snedaker 2(00), and as a means of minimizing interference with gas
exchange by high tidal levels (Tomlinson and Cox 2000). In the viviparous
seagrasses, Amphibolus and Thalassodendron, vivipary appears to be an adaptation
for rapid root attachment of the plant (Ducker and Knox 1976, Cox and Humphries
1993). However, the occurrence of apparently successful mangroves without
viviparous fruits (such as Osbomia, Sonneratia, Laguncularia, Lumnitzera,
Pelliciera, Xylocarpus and Excoecaria) makes it doubtful whether the possession of
vivipary per se is of any outstanding adaptive advantage.
Tidal buffeting and wavebome objects pose a threat to establishing seedlings,
and it seems likely that the smaller the seedling, or the slower the anchoring
process, the greater the threat (Saenger 1982). Because of this, vivipary in a
mangrove may simply be a means of producing a large seedling with adequate food
reserves, capable of rapid establishment and which is less likely to be damaged by
water movements at this critical stage. Interestingly, the anatomical studies of
Rhizophora propagules show that rapid rooting and erection is facilitated by the
development of eccentric tension wood (reaction) fibres in these propagules
(Tomlinson and Cox 2(00). It is also interesting to note that many of the non-
viviparous genera (such as Xylocarpus, Heritiera, Cynometra and Pelliciera) also
possess large seeds which similarly may be a means of alleviating damage by water
movement.
Elmqvist and Cox (1996) reviewed the phenomenon of vivipary from an
evolutionary perspective, noting its occurrence in only 20 genera in 13 families,
mostly from shallow marine habitats. They suggested that the occurrence of vivipary
in different unrelated angiosperm families indicates independent evolutionary
processes where there are no advantages for plants to disperse their seeds over space
or time. They also suggested that in shallow marine habitats, plants are presented
with rather predictable and uniform environments both in space and time and that
this physical and biotic homogeneity reduces fitness gains from dispersal over time
and space. Hence, taxa with seed dormancy or seed dispersal mechanisms (such as
animal vectors) enjoy no particular advantage. Elmqvist and Cox (1996) concluded
that 'of greater importance is the fact that viviparous taxa in such environments
incur no disadvantage such as they might in highly seasonal or patchy
Adapting to the 'Mangrove Environment' 87

environments'. In turn, the advantage they gain is that large viviparous propagules
establish close to their maternal sources as 'survival outside the parental patch may
often be low'. This finding is consistent with a study on the east coast of Australia
where propagule establishment and survival was monitored for nine years (Saenger
1988, 1996a).
Producing a large propagule with abundant food reserves, while of advantage in
the intertidal environment, also has a cost. As Farnsworth and Ellison (1997a) have
argued, producing numerous seedlings on the parent tree is likely to attract predators
as a result of which pre-dispersal propagule predation is a ubiquitous feature of
mangrove forests world-wide.

3.5.2 Propagule Production


Most mangroves bear mature propagules in the late summer months, February to
March in the southern hemisphere, and July to August in the northern hemisphere
(Tables 3.8 and 3.9). In the tropics and subtropics this generally coincides with, or
is towards the end of, the rainy season (Duke et al. 1984, Jimenez 1988, Fernandes
1999).
In the viviparous species, the time from flower primordium to mature
propagule is considerable, reflecting the energetic costs of the propagules: 2-3.5
years in Rhizophora apiculata (Christensen and Wium-Anderson 1977, Duke et al.
1984, Moriya et al. 1988, Kitamura and Baba 1997); and 0.7-1.5 years in Bruguiera
gymnorhiza, Ceriops tagal, C. australis, Rhizophora stylosa and R. mucronata
(Duke et al. 1984, Moriya et al. 1988, Steinke 1999). From flower buds to mature
propagules has been estimated to take: 12-15 months in Kandelia candel (Nishihira
and Urasaki 1976), Rhizophora mangle (Gill and Tomlinson 1971b) and Aegiceras
comiculatum (Carey and Fraser 1932); and 5 months in Rhizophora apiculata am
Bruguiera gymnorhiza (Moriya et al. 1988). Similarly, from open flower to mature
propagules has been estimated to take: 6 months in Rhizophora mangle (Guppy
1906); and 9-15 months in Rhizophora apiculata, R. mucronta, R. stylosa, Ceriops
spp. and Bruguiera gymnorhiza (Duke et al. 1984). In a detailed phenological study
of the cryptoviviparous Avicennia marina over a large latitudinal range, Duke (1990)
found that the time between flowering and fruiting in this species is variable,
ranging from 2-3 months in tropical sites to 10 months in the southernmost sites.
By way of contrast, in the non-viviparous Conocarpus erectus, the time from mature
flower to mature fruit is 2 months (Tovilla and Da La Lanza 20(0).
In spite of these considerable lags, both flowering and fruiting largely occur in
the summer months, and it would seem that some common environmental parameter
is involved in controlling both processes (Duke et al. 1984). As leaf production in
many species is also seasonal, with the maxima for many species occurring during
summer (Duke et al. 1984, Saenger and Moverley 1985), it seems that fruiting, like
flowering, is timed for the period most favourable for growth (Mulik and Bhosale
1989). Duke (1990) has investigated the onset of the reproductive cycle in A vicennia
marina over a large latitudinal range and concluded that day length and mean
temperature are important determinants in this species (see also 5.2.1). Others have
suggested that in the Caribbean, propagule production is centred on the period
(August to October) when hurricane frequencies are at their highest, and that the
selective advantage of dispersing propagules immediately after storm damage might
have played a role (Roth 1992).
88 Mangrove Ecology, Silviculture and Conservation

Table 3.8 Monthly occurrence of maximal fruiting in the northern hemisphere.

Species Months Location Ref.


J F M A M J J A S 0 N D
A. alba x x India 10
A. bicolor x x x Costa Rica 4
A. germinans x x x x Costa Rica 4
x x x Panama 9
x x Martinique 8
x x x x x Colombia 7
x x x x Brazil 11
x Bermuda 12
x x x West Africa 5
A. marina x x Thailand I
A. officinalis x x India 10
x x Bangladesh 13
B. cylindrica x Thailand 1
B. gymnorhiza x x Bangladesh 13
C. decandra x x x Bangladesh 13
C. erectus x x x x Mexico 6
E. agallocha x x Bangladesh 13
H·fomes x x Bangladesh 13
La. racemosa x x x x Costa Rica 4
x x x Panama 9
x x x x Brazil 11
L littorea x x x x Thailand 1
N. fruticans x x Bangladesh 13
R. apiculata x x x x Thailand 3
x x x Vietnam 14
R. harrisonii x West Africa 5
x x Brazil 11
R. mangle x x x x x x x West Africa 5
x x x x Bermuda 12
x x x Panama 9
x x Martinique 8
x x Brazil 11
R. mucronata x x Thailand 2
R. racemosa x x x x Costa Rica 4
x x x x x West Africa 5
S. hydrophyllacea x Thailand 2
S. apetala x x India 10
x x Bangladesh 13
X. moluccensis x x Bangladesh 13

IWium-Andersen and Christensen 1978; 2Wium-Andersen 1981; 3Christensen and Wium-Andersen


1977; 4Jimenez 1988; 5Saenger and Bellan 1995; &rovilla and Da La Lanza 2000; 7Elster et a1. 1999;
8Imbert and Menard 1997; '1Ubinowitz 1978a; l'i.ahi.ri 1991; IIFemandes 1999; I~llison 1997;
I3Chowdhury 1996; 14Clough et aI. 2000.

Considerable pre-dispersal mortality has been reported for developing seedlings.


Gill and Tomlinson (1971b) showed that for Rhizophora mangle only between 0-
7.2% of flower buds produced mature seedlings, although the number of flowers
produced may be increased markedly by an increase in nutrients (Onuf et al. 1977).
Lugo and Snedaker (1975) followed the development of selected seedlings of R.
mangle while still attached to the parent tree in Florida in the United States, arxl
found a mortality of 9%, 13.4% and 20.9% for the months of January (winter),
April and May (spring) respectively. Similar figures were noted in Rhizophora
apiculata, for which Christensen and Wium-Anderson (1977) reported that only 7%
of flower buds formed flowers, and only 1-3% formed fruits. In Kondelia can,Jelless
than 30% of the flower buds ultimately developed into mature propagules (Nishihira
and Urasaki 1976). For the Australian Sonneratia alba, young fruit developed from
41 of 46 flower buds (Primack et al. 1981).
Adapting to the 'Mangrove Environment' 89

Table 3.9 Monthly occurrence of maximal fruiting in the southern hemisphere.

Species Months Location Ref.


J ASONOJ FMAMJ
A. ilicifolius x x NE Australia 5
A. annulata x x E Australia I
x NE Australia 5
A. comiculatum x x x x x E Australia I
x x x SE Australia 6
x x x NE Australia 5
A. marina x x x E Australia I
x x x SE Australia 6
x NE Australia 5
x x x S. Africa 3
B. cyliruirica x NE Australia 5
B. exaristata x x NE Australia 5
B. gymnarhiza x x x x x E Australia I
x x NE Australia 5
x x x x x x S Africa 3
B. parviflora x x NE Australia 5
B. sexangula x x NE Australia 5
C. schultzii x x x NE Australia 5
C. australis x x x NE Australia 5
C. decandra x NE Australia 5
C. tagal x x x x x x x E Australia I
x x x NE Australia 5
C. iripa x x x NE Australia 5
E. agallocha x x E Australia I
x x NE Australia 5
H. littoralis x x x x NE Australia 5
L littarea x x NE Australia 5
Lu. racemosa x x x E Australia I
x x NE Australia 5
x x x x x S Africa 3
N. fruticans x x NE Australia 5
O. octodonta x x x x x E Australia I
x x NE Australia 5
P. acidula x NE Australia 5
R. apiculata x x x x Bali 4
x x NE Australia 5
R. mucronata x x x x S Africa 3
x x x NE Australia 5
R. stylosa x x x E Australia 1
x x NE Australia 5
S. hydrophyllacea x x NE Australia 5
S. alba x x NE Australia 2
x NE Australia 5
S. casealaris x x NE Australia 2
X. granatum x x x x NE Australia 5
X. moluccensis x x x x x E Australia I
x x x NE Australia 5

lSaenger 1982; 20uke 1988; 3Steinke 1999; 'Kitamura and Baba 1997; 50uke et al. 1984; 6Clarke 1994.

Some of this pre-dispersal mortality may be caused by maternal abortion


(Clarke 1992, Palis 1996) or premature abscission due to such inherent factors as
albinism (Handler and Teas 1983) and other morphogenetic malfunctions (fig. 3.15).
For example, a single recessive gene causes albinism in Rhizophora mangle and
KGJU1elia cGJU1e1 propagules. This albino mutation is in the nuclear genome
(Lowenfield and Klekowsky 1992, Klekowsky et al. 1994a) but adversely affects the
chloroplast ultrastructure, resulting in a marked reduction in development when
compared with sibling wild types (Smith and Snedaker 2000). Rhizophora mangle is
more outcrossed in Puerto Rico, with lower mutation rates for albinism as a
90 Mangrove Ecology, Silviculture and Conservation

consequence (Klekowski et al. 1994a), while Chen (2000) described increased


mutation rates in Kandelia candel in a partially deforested mangrove where
outcrossing had decreased. Maternal abortion of propagules may be a means whereby
mangroves jettison both damaged propagules and predators (Clarke 1992, Farnsworth
and Ellison 1997a).

Fig. 3.15 Green and chlorotic ('albino') propagules on Rhizophora stylosa from Port
Curtis, Queensland.

Undoubtedly, however, most pre-dispersal mortality can be attributed to fungal,


crab, insect and mammal attack on the fruit and emerging propagules (Gill and
Tomlinson 1971b, Lugo and Snedaker 1975, Rabinowitz 1977, Beever et al. 1979,
Robertson et al. 1990, Clarke 1992, Pandit and Choudhury 2001). Many insects
damage propagules while still attached to the tree (see comprehensive review by
Farnsworth and Ellison 1997a); for example, Robertson et aI. (1990) showed that in
north-eastern Australia, between 3.1-92.7% of the seeds or propagules of 12
mangrove tree species had been attacked by insects. Propagules of six species
(Avicennia marina. Bruguiera gymnorhiza. B. parviflora. Heritiera littoralis.
Xylocarpus australasicus and X. granatum) showed consistently high (>40%) levels
of insect damage, with Heritiera littoralis showing around 99% of insect attack. In
the case of Sonneratia caseolaris in India (Pandit and Choudhury 2001), flower and
propagule predation by mammals (rhesus macaques, palm squirrels and rats) caused
74% of buds to be lost and yielded a fruit set of around 6%. In the same study, no
Adapting to the 'Mangrove Environment' 91

predation was noted on the flowers or fruits of Aegiceras comiculatum which showed
a fruit set of 62%.
In addition, a number of crabs are known to climb trees to feed on propagules.
In neotropical mangroves, the crabs Aratus pisonii and Goniopsis cruentata can be
seen to feed on propagules on the trees (Beever et al. 1979, Farnsworth and Ellison
1997a) while in eastern Africa, Sesarma leptosoma has been recorded as climbing
trees to feed on propagules (Vannini et al. 1997). In Asia, tree climbing crabs
include Metopograpsus latifrons, M. oceanicus, Chiromantes bidens, Neoepisesarma
la/ondi and N. mederi (Shokita 2(00).
In their global survey, Farnsworth and Ellison (1997a) reported an overall
propagule attack rate of 23% for all sites and all species, noting that the species-
specific rates varied from 50% in Avicennia marina, to 34% in Rhizophora
mucronata, 28% in Bruguiera gymnorhiza, 27% in Rhizophora mangle, 25% in
Ceriops tagal and Rhizophora apicu/ata, 18% in Rhizophora stylosa, and 10% in
Rhizophora samoensis. They found that propagule predation rates were highly
variable around the world and that no biogeographic trends could be identified. They
concluded that pre-dispersal propagule predation was relatively unimportant in
explaining subsequent seedling establishment.
In more detailed studies with Rhizophora mangle, Farnsworth and Ellison
(1997a) found that much of the crab damage is superficial and probably contributes
less to individual mortality, despite its greater frequency, than scolytids and moths
that burrowed and destroyed the interior of the hypocotyls. Similar findings were
made by Robertson et al. (1990), who showed that Australian mangroves could be
divided into different categories in terms of the effects of propagule attack on
subsequent germination and growth. Interestingly, in Australia, Avicennia marina
propagules most often contain the larvae of the fly Euphranta marina (Tephritidae:
Trypetinae) which tunnels through the cotyledons. Such predated propagules are still
capable of germinating, although the effects of the larvae on subsequent growth are
not known.
In general, Farnsworth and Ellison (1997a) found that growth rates of predated
propagules did not differ significantly from those not predated, but that predated
propagules had premature abscission rates approximately twice those of undamaged
propagules, a figure notably higher than the leaf abscission rate with herbivore
damage (see 4.2.3). Nevertheless, they also concluded that mangrove seedling density
remains to be explained by factors other than pre-dispersal predation.

3.5.3 Propagule Dispersal and Establishment


The dispersal unit ('propagule') of mangroves (Table 3.7) may be a spore or
prothallus (Acrostichum), a single seed (Excoecaria), a one-seeded fruit (Cynometra),
a several-seeded fruit (Sonneratia, Xylocarpus), a multiple fruit (Heritiera), an
aggregated fruit (Nypa), or a precociously developed seedling (Avicennia, Aegiceras,
Rhizophora, Ceriops, Bruguiera).
The ubiquitous characteristic of the propagules of all mangroves trees and
shrubs is their buoyancy (Saenger 1982, Tomlinson 1986). This buoyancy may be
due to several different features of the propagule including the radicle (as in
Rhizophora), the pericarp and cotyledons (Avicennia), the endoderm (Xylocarpus),
the seed testa (Nypa and Sonneratia) or the cotyledon (Pelliciera). In Osbomia
octodonta and Acanthus ilicifolius, the short tomentum of interlocked hairs covering
92 Mangrove Ecology, Silviculture and Conservation

the fruit (Osbomia) or seed (Acanthus) traps air and provides the necessary buoyancy
for water dispersal (Wightman 1989).
Changes in any of these features can alter the buoyancy. For example, Steinke
(1975) showed that propagules of Avicennia marina in South Africa sink after losing
their pericarp, generally within four days. Subsequent investigation of the rate of
pericarp shedding showed that high and low salinities decreased the rate at which they
were shed when compared with the rate in water of intermediate salinity. In
Australian material of A. marina from South Australia, the time required for the
splitting of the pericarp and the separation of the cotyledons also increased with
increasing salinity (Downton (1982). Similarly, using material from south-eastern
Australia, Clarke and Myerscough (1991a) demonstrated that there is considerable
variation both within and between populations, but the pericarp was consistently
shed sooner in dilute than in fully saline seawater. This tendency appears to be
widespread as EI-Shourbagy et al. (1995) also showed, using A. marina from the
United Arab Emirates, that seed-coat shedding, seedling emergence and seedling
survival were unaffected at salinities up to 40 ppt. Above these salinities, however,
seed-coat shedding and seedling emergence were inhibited while seedling survival
declined abruptly. Consequently, propagules in brackish water will disperse less than
those in water of high or low salinity.
High temperatures also· increase the rate of pericarp shedding, consequently
shortening the potential dispersal distance (Steinke 1975). In this respect, Australian
material appeared to differ from South African examples of the same species as no
temperature relationship with pericarp shedding could be discerned in the Australian
material (Downton 1982).
Some natural variability in buoyancy also occurs in the propagules of some
species; in the common neotropical species, some dispersing propagules float while
others sink. For example, McKee (1995b) found that for Avicennia germinans,
about 92% were 'floaters' while 8% sank in seawater (density of 1.025 g cm- 3).
Rhizophora mangle propagules comprised 87% 'floaters' and 13% 'sinkers'. 'Sinkers'
had a density of 1.044 g cm- 3 (for R. mangle) and 1.068 g cm- 3 (for A. germinans)
while 'floaters' were 1.003 and 1.012 g cm- 3 respectively for these two species.
Few data are available on the tidal periodicity of propagule dispersal but Clarke
and Hannon (1971) found that in eastern Australia, release of Aegiceras propagules
coincided with unusually high tides whereas that of Avicennia propagules occurred at
low tides.
A buoyant propagule appears to be an efficient means of widespread, water-
based dispersal. Among the seagrasses, however, only a few genera (Posidonia,
Thalassodendron and Enhalus) have buoyant fruits and, paradoxically, these species
have restricted distributions (Den Hartog 1970). It would appear from this that the
role of buoyancy in effecting widespread dispersal needs experimental evaluation.
In a recent study ofpropagule dispersal, Stieglitz and Ridd (2001) found that in
estuaries with a normal salinity gradient. an axial surface convergence forms in
midstream, generated by a density-driven circulation cell. Such convergences are
often clearly visible in the form of a kilometre-long line of debris floating
midstream. Buoyant mangrove propagules of Rhizophora stylosa, Bruguiera
gymnorhiza, Xylocarpus moluccensis and Heritiera littoralis and other floating
material become trapped in such convergences and move upstream with the
midstream current, estimated in the Normanby River of North Queensland to average
3.2 km d- 1 (Stieglitz and Ridd 2001). On ebb tides, the circulation cell reverses as a
Adapting to the 'Mangrove Environment' 93

result of which entrained propagules move towards the banks, where water velocities
are much smaller than midstream. Over a tidal cycle, a net upstream drift of
propagules occurs. The propagules subsequently accumulate in large numbers in the
hydrodynamic traps upstream from the convergence.
Table 3.10 Distribution of mangroves and associated species from four islands in the Gulf of Guinea"

Speciesb Bioko Principe Sao Tome Annob6n


(Fernando Po)
Avicennia genninans + +c +
Rhizophora harrisonii + + +
Acrostichum aureum + + +
Conocarpus erectus + +
Hibiscus tiliaceus + +
Carapa procera + +
Drepanocarpus lunatus + +
Pandanus candelabrum + +
Rhizophora racemosa +
Rhizophora mongle +
Laguncularia racemosa +
Phoenix reclinata +
ChrysobalanutJ

Area (km2 ) 2,208 126 964 18


Max. Altitude (m) 2,850 948 2,024 655
Dist. to mainland (km) 32 215 235 340
Dist. to landward 32 210 135 180
neighbour (km)

Number of Species 12 5 6 o
Based on data from Hutchinson et aI. 1954/72, Exell1945, 1973, Keay 1953, Adams 1957.
Thespesia populnea has been omitted from this table as it has been widely planted in the towns on
the islands.
This unconfirmed record is based on a description for which no voucher specimen could be located
(Exell 1945).
Includes two species both of which are present on the adjacent mainland (Wouri Estuary,
Cameroon) but not recorded from any of the islands.

Axial convergences are common features of estuaries in North Queensland,


occurring in 80% of the major estuaries in the region (Stieglitz and Ridd 2(01). In
fact, such convergences will form in any estuarine system where, after a short wet
season, salinity gradients are re-established during the ensuing dry season. If, as has
been indicated above, propagule dispersal coincides with the end of the wet season
(Table 3.9), the axial surface convergences will affect propagule dispersal. Clearly,
this mechanism potentially provides an efficient barrier for propagule exchange
across estuaries and results in mangrove popUlations within estuaries forming
floristically and genetically isolated populations with extremely limited gene flow
between adjacent estuaries (Stieglitz and Ridd 2(01).
Once outside an estuary, propagules can potentially disperse over larger
distances. While some anecdotal information exists, no data on wider dispersal is
available. However, the Gulf of Guinea, West Africa, with its relatively linear
coastline and a group of variously distanced islands (fig. 3.16), provides a useful
setting to determine the potential dispersal distance of mangroves (Saenger aOO
Bellan 1995). The data on the distribution of mangrove and associated species on the
offshore islands in the Gulf of Guinea have been collated (Table 3.10) and regressed
against the island area (fig. 3.17) and the distance from the mainland and nearest
landward island (fig. 3.18).
94 Mangrove Ecology, Silviculture and Conservation

Pnncipo {}

146 Ian

SooT..".

2lI01an
1801an

.
AnnoI>on

Fig. 3.16 Map of the Gulf of Guinea, Central West Africa, showing the location and
distances between islands of Bioko, Principe, Sao Tome and Annob6n, and the
adjacent mainland.

As expected. there is a more or less linear relationship with island area,


suggesting that habitat availability is a significant factor. The analysis against
distance suggests that the dispersal ability of mangroves under these equatorial
conditions allows a dispersal distance of 350 Ian from the mainland or 250 Ian from
another island. Given a dispersal longevity in Panama of 35 days for Laguncularia,
ItO days for Avicennia and around 300 days for Rhizophora (Rabinowitz 1978a), arxl
with current speeds in the Gulf of Guinea around 0.3 m S·I, the dispersal distances
obtained from this simplistic analysis seem realistic.
Rabinowitz (l978a) investigated the parameters affecting dispersal of six
Panamanian mangroves, including longevity and vigour, period of floating, period
required for establishment, and the period of obligate dispersal. Two contrasting
dispersal patterns were observed: one for small and another for large propagules. A.
germinans seems to have an absolute requirement for a period of stranding in order to
establish itself. This species is restricted to higher ground where inundation is less
frequent and where it is free of tidal disturbances. The time required for this species
to root is approximately seven days whereas A. marina, whose propagules sink after
approximately four days, becomes firmly rooted in five days (Clarke and Hannon
1970). Laguncularia, whose propagules sink after approximately twenty days, also
Adapting to the 'Mangrove Environment' 95

requires a period of stranding of five days or more in order to become rooted


(Rabinowitz 1978a).

15,----------------------------------,

y =0.OO5x + 2.014 r2 =0.858


'"u
.S!
10
G>

....""0
{12

....
G>

~
Z
5 c

O~----~------~----~------r_----~
o 500 1000 1500 2000 2500

Island Area

Fig. 3.17 Number of species of mangroves in relation to area (km2 ) for four islands
in the Gulf of Guinea. (Data from Saenger and BeHan 1995)

15,------------------------------------,
Y(M) = -0.038x + 13.516 r 2 = 0.965
o Distance from mainland

.; 10
u
G>

....""
{12

o
....G>
~
Z
5
¢ Distance from nearest island

y (I) = -0.053x + 13.091


O+---~ __~~~~~._--~~~~--~~
10 100 1000

Distance (km)

Fig. 3.18 Number of mangrove species on the islands in relation to distance from the
mainland [y(M)I or nearest island neighbour (y(l» in the Gulf of Guinea. (Data from
Saenger and BeHan 1995)

The two genera that have large propagules (RhizopJwra and Pelliciera) tolerate
tidal disturbance better than do either A vicennia and Laguncularia; the propagules of
the former two are capable of taking root in water of various depths because their
96 Mangrove Ecology, Silviculture and Conservation

weight affords resistance to tidal buffeting, and growth continues under water.
Longevity of propagules ranged from thirty-five days in loguncularia to a year or
more in R. mangle (Rabinowitz 1978a).
These findings led Rabinowitz (1978b) to suggest that the seedling populations
of mangroves with smaller propagules tum over annually whereas those with larger
ones are made up of overlapping cohorts. In other words, two reproductive strategies
are involved: mangroves with small propagules pepper the swamp annually with
short-lived seedlings which may become established in gaps that have arisen during
the previous year; those with larger propagules form a persistent seedling bank
which can maintain itself until a gap in the canopy occurs (if shade-intolerant), or
grow in the shade to reach the canopy (if shade-tolerant).
The mechanism of seedling establishment has been widely discussed. Early
studies supported the 'self-planting' theory (LaRue and Muzik 1951) where seedlings
fall from the parent tree and become established in the sediment at the base of the
parent. However, as convincingly shown by Yamashiro (1961) for Kandelia, around
90% of seedlings fall into a horizontal position on release. Thus, 'self-planting' does
not seem a very realistic attribute of mangrove propagules.
Some have suggested that the 'stranding' theory is more generally applicable,
where larger propagules are stranded in deeper water and where only small propagules
can successfully become established in shallow water, i.e. high in the intertidal zone
(Rabinowitz 1975, 1978a, b, c). Others, such as Van Speybroeck (1992) working in
East Africa, argue that both strategies commonly occur, with 'stranding' being the
sole means in areas of regrowth or new colonization, and 'self-planting' being the
most important in mature mangrove communities.
Both the 'self-planting' theory and the 'stranding' theory is based on a restricted
dispersal of mangrove seedlings in the intertidal zone. This raises two questions
concerning (1) the distance over which propagules are generally dispersed from their
point of origin, and (2) whether or not all tidal levels are equally accessible to
variously sized propagules.
Davis (1940) conducted release and recapture experiments to assess the dispersal
ability of Rhizophora mangle propagules, finding that regular dispersal occurs over
several kilometres. Similar release and recapture experiments were conducted on
Halmahera Island, Indonesia by Prawiroatrnodjo (1988) by placing marked
propagules in two mangrove zones, i.e. the Sonneratia zone and the mixed
Rhizophora-Bruguiera zone. After 25 days, around 10% of all propagules placed in
the Sonneratia zone were still within 100 m of the point of release while around
60% of placed propagules in the Rhizophora-Bruguiera zone remained within 10 m
of the point of release. Clearly, tidal dispersal of propagules is highly dependent on
where the propagules are placed; in the dense root network of the Rhizophora-
Bruguiera zone where tidal action is somewhat reduced, few propagules are dispersed
over significant distances, while in the Sonneratia zone, where diurnal tidal activity
is pronounced and where the pneumatophores appear to offer lesser tidal resistance,
the dispersal of propagules is considerably greater.
Working with Kandelia, Yamashiro (1961) demonstrated that around 90% of
seedlings were carried more than 50 m away from the parent tree. Komiyama et al.
(1992) found that the majority of Rhizophora mucronata propagules were stranded
within 300 m of the point of release within a month. Dispersal of Avicennia marina
propagules was limited, with approximately 78% of propagules stranded within 2
km of their starting point, and with the majority well within 50 m (Clarke 1993).
Adapting to the 'Mangrove Environment' 97

With Ceriops tagal in northern Australia, McGuinness (1997a) found that at least
75% of propagules remained within 1 m of the parent tree, and that 91% remained
within 3 m. These dispersal distances suggest that 'self-planting' is likely to be the
exception rather than the rule.
In a nine-year study, Saenger (1982, 1988) showed that most species broadcast
their propagules throughout all tidal levels in the mangrove habitat and that
seedlings become established widely and quickly throughout the intertidal regions.
With time, however, seedling mortality is higher outside the optimal zone of each
species, resulting ultimately in distinct zones which reflect physico-chemical
characteristics (Snedaker, 1982; Youssef and Saenger 1999) that are still being
identified.
The number of mangrove propagules becoming established per adult of the
same species in permanent study areas at Port Curtis, a semi-enclosed bay in central
coastal Queensland, is given in Table 3.11, together with comparative data from a
two-year investigation at Repulse Bay, near Proserpine.
Table 3.11 The annual numbers of propagules per adult of the same species becoming established in
permanent plots during a nine-year study (1974-1983, Port Curtis, Queensland) and a 2-year study
(1980-1982, Proserpine, Queensland), together with mortality rates in the first year after establishment.

Species Port Curtis Proserpine


propagules % mortality propagules % mortality
Aegiceras comiculatum 0.18 14.8 0.17 26.7
Aegialitis annulata 1.50 0
Avicennia marina 1.47 22.1 1.58 38.1
Ceriops tagal 0.13 36.5 0.15 12.5
Excoecaria agallocha 0.21 25.0
Lumnitzera racemosa 1.00 o 0.13 20.0
Rhizophara stylosa 1.64 71.7 0.77 29.2

These rates of propagule establishment are low in view of the apparently high
numbers of propagules borne by most species. However, considerable mortality
occurs prior to dispersal; further mortality occurs during dispersal, including
stranding on unfavourable substances, injury by boring or decomposing marine
organisms, and sinking as a result of the attachment of fouling organisms such as
barnacles and serpulid polychaetes.
Once the propagules are stranded, physical damage by waveborne objects
frequently occurs in addition to their predation which can involve crabs (Smith
198711. b, Smith et al. 1989, Osborne and Smith 1990, McKee 1995a, McGuiness
1997a, b, Dahdouh-Guebas et al. 1997, 1998, Lee 1989a), molluscs (Smith et al.
1989, McKee 1995a), insects (Robertson et al. 1990, Ellison and Farnsworth 1993,
Elster et al. 1999), and a range of other fauna (Siddiqi 1992).
Grapsid crabs (Sesarma and Neosamartium), in particular, can play a
considerable role in the predation of mangrove propagules and seedlings, and may
locally be a threat to the regeneration of mangroves (Smith 1987a, Dahdouh-Guebas
et al. 1998). For example, in Darwin, northern Australia, within 22 days, crabs
consumed 100% of the propagules of Avicennia marina, 63% and 71% of Ceriops
tagal and Bruguiera exaristata respectively, and 19% of Rhizophora stylosa
(McGuinness 1997b). In Belize, McKee (1995a) reported that in 4 days, crabs
consumed 45% of the propagules of Avicennia germinans, and 4% and 13% of
Rhizophora mangle and Laguncularia racemosa respectively. At other localities,
however, their role appears negligible (Siddiqi 1992, Elster et al. 1999).
98 Mangrove Ecology, Silviculture and Conservation

Even where propagule predation by grapsid crabs is relatively high, McKee


(l995a) has demonstrated that there is considerable temporal variation in predation
rates, with high rates at the beginning of propagule fall and relatively low rates near
the end of the dispersal period. Possible reasons for such variable predation rates
include seasonal differences in predator abundance or activity, differences in chemical
composition (and hence attractiveness) of propagules, and differences in the
abundance of propagules and/or alternative food sources.
The rate of seedling establishment and of their subsequent survival depends on a
range of physico-chemical factors over and above the biotic effects of grazing. Thus,
factors such as salinity, pH, Eh, soil hardness, light, soil sulfides, nutrient
availability, and water movement have been identified as important determinants.
For example, Kathiresan et al. (1996) investigated root development (Le. the number
of roots per hypocotyl multiplied by the mean length of the roots in cm) of
Rhizophora mucronata with varying pH and salinity. Root development in this
species was maximal at a pH of 6 and at a salinity of 35 g L- 1 (fig. 3.19).
Komiyama et al. (1996) showed the influence of soil hardness on seedling
establishment and survival, while the effects of light and sulfides are discussed in
section 4.3.3.
100

I~ ~ I
75 T
I:::'
.§'
8
T
;iii ~
50 T jjj ill
~ t... ~
'Cl
il ~~ ~~
-
~
-8c 25 - I:::'
i:'
0
0
..
...
::: III
5 10 15 20 25 30 35 40 45
::: .. ~
Salinity
Fig. 3.19 Root development index (the number ofroots per hypocotyl multiplied by the
mean length of the roots in cm :I: SE) of Rhizophora mucronata after 60 d of
hydroponic culture at various salinities (%0). using Arnon and Hoagland nutrient
medium (Data from Kathiresan et at. 1996)

To determine the relative importance of factors affecting seedling establishment


and survival in neotropical mangroves of Belize, McKee (1995b) examined the
spatial patterns of seedling relative densities in relation to reproductive adults and in
relation to physico-chemical factors, including light and nutrient availability, sulfide
concentrations, redox potentials, soil salinity, soil temperature, bulk density am
relative elevation. She found that the distance from reproductive adults explained 89-
94% of the variation in relative density of Rhizophora mangle seedlings and that the
availability of resources (light and NH4) explained 73-80% of variation in Avicennia
germinans. Immediately after dispersal, 89% of the variation in LaglDlCularia
Adapting to the 'Mangrove Environment' 99

racemosa seedling relative density was attributable to distance from a reproductive


adult but seven months later, 74% of the variation was explained by stresses related
to flooding and salinity.
Among those propagules that become established successfully - that is.
become firmly rooted and possess at least one leaf - mortality rates are variable arxl
site-dependent. At Port Curtis. Queensland, where propagule predation by grapsid
crabs is negligible, mortalities in the first year ranged from 72% in Rhizophora
stylosa to 0% in Lumnitzera racemosa and Aegialitis annulata. At Proserpine.
Queensland. mortality rates during the first year were much more equable (Table
3.11). At these sites. the main factors determining post-establishment mortality
were physical. such as waveborne objects. biological. such as crab damage. arxl
physiological. such as water stress, insufficient light, and high soil salinities. In the
Cienaga Grande de Santa Marta on the Caribbean coast of Colombia, Elster et al.
(1999) reported post-establishment mortality rates of Avicennia germinans of
virtually 100% due to the species-specific predation by the caterpillars of the
butterfly Junonia evarete (Lepidoptera: Nymphalidae). Similarly, high morality rates
were reported in plantations of Rhizophora mucronata in Bali as a result of dense
infestations of the scale insect Aulacaspis marina (Ozaki et al. 2000). These patterns
of mortality between species with differing propagule weights show trends markedly
different from those reported by Rabinowitz (I 978b). who found that mortality rate
was inversely correlated with initial propagule weight.
Thus. it is doubtful that such a simple scheme as 'self-planting' or 'stranding'
consistently operates in mangroves. The ability to utilize sunflecks efficiently. as
can Avicennia marina seedlings (Ball and Critchley 1982). blurs the boundary
between the shade-tolerant and shade-intolerant species with large propagules. arxl
reduces the needs for gaps prior to seedling establishment. Furthermore. newly
arrived seedlings tagged in permanent study areas at Port Curtis. central Queensland.
were able to survive as two- to four-leaved seedlings for up to eight years whether
from small (Lumnitzera). medium (Avicennia. Aegiceras, Aegialitis) or large
(Rhizophora, Ceriops) propagules (Saenger 1982). A persistent seedling bank
appears to be an important survival strategy in mangrove communities, allowing the
broadcasting of propagules throughout the intertidal zone, particularly in years when
propagules are abundant. The broadcasted propagules can persist for some months,
depending on the actual conditions encountered, undergo vigorous growth under
optimal conditions, or gradually deteriorate if the conditions become or remain
unfavourable. This strategy may be termed a 'sow and reap' approach where
propagules are dispersed as widely as locally possible, with outcomes determined by
the specific conditions encountered by each propagule.
Based on the rate of pericarp shedding and, hence sinking of A vicennia marina
propagules, Clarke and Myerscough (1991a) concluded that 'these observations
suggest that most A. marina propagules strand and establish near the parent tree
populations and only a few are dispersed more widely'. Similarly based on
considerations of limited propagule viability and low coastal current velocities,
Lange and Lange (1994) concluded that Avicennia marina is ' ... adapted for restricted
dispersal to ensure that propagules become established in the immediate vicinity of
the parent. It is unlikely that propagules can successfully establish more than 4-5 d
travel from source.'
Genetic analyses of Avicennia marina (Parani et al. 1997b, Duke et al. 1998,
Maguire et al. 2000) and Avicennia germinans (McMillan 1986, Dodd et al. in
100 Mangrove Ecology, Silviculture and Conservation

press) have demonstrated that Avicennia populations are highly genetically


structured, which has led to doubts about the assumed widespread dispersal of the
buoyant propagules of this species. Similarly, morphometric studies of Rhizophora
mangle have indicated that significant differences exist between populations on the
Pacific and Caribbean coasts of Mexico, among populations, and within individual
populations in relation to floral attributes (Dominguez et al 1998). They suggested,
as did Maguire et al. (2000) and Dodd et a1. (in press) for Avicennia marina and A.
germinans respectively, that frequent extinctions and recolonizations by a few
individuals, with some selfing, have led to genetic population differentiation.
Such conclusions are being increasingly supported by molecular genetic studies
in a range of other mangrove species (Lakshmi et al. 1997, 2000, Parani et a1.
1997b, Ge and Sun 1999, 2001). Dispersal of mangrove propagules is not as
effective as 'seawater dispersal' might suggest, and many mangrove populations have
minimal genetic diversity as a result of the limited gene flow between populations.
For example, Huang (1994) first studied the genetic variability between and within
populations of the mangrove Kandelia candel in Taiwan using allozyme analysis. He
concluded that a moderate level of genetic variation existed between the four
mangrove populations studied and there was low within-population diversities.
These findings suggest that local environmental selection and restricted gene flow
between the populations contributed to the limited genetic variability recorded for
this species in Taiwan.
Huang and Chen (1997) extended the study of K. candel to include the Ryukyu
Archipelago. They found that the level of genetic variation in all six populations
studied was lower than previously reported in other plant taxa. This low variation
was largely due to high inbreeding rates, possibly a consequence of founder events in
the recent past. As K. candel was previously shown to be predominantly
outcrossing, and as some genetic differentiation was recorded between populations, it
is likely that the lack of diversity is also due to restricted gene flow between
populations. The results suggest that on a macrogeographic scale, Kandelia
populations consist of isolated and subdivided units with restricted gene flow.
Significant gene flow only becomes apparent in this species on a
microgeographic scale. Thus, in a study of thirteen K. candel popUlations in and
around Hong Kong, Sun et a1. (1998) detected very low genetic diversity despite
high outcrossing rates measured from the study of their mating system. As expected
within such a small geographic scale, these authors found very low genetic
differentiation between sites (with nearly four migrants per generation being
estimated to move across populations). At that scale, homogeneity is likely to be a
result of recent co-ancestry (Sun et a1. 1998).
From the genetic evidence and known dispersal effectiveness, and from the
postulated advantages derived from vivipary, it seems much more likely that
mangroves have a 'sow and reap' strategy, saturating their immediate environment
with propagules at all tidal levels, and forming a seed bank which can persist for a
limited period. It also seems likely that long-distance dispersal is a game of chance,
with only rare success.
4. Physico-chemical Factors and Mangrove Performance
... lhe country was jenile and the vegetation most luxuriant
lhe sago palm and cocoanut being most abundant;
mangroves were also jound
which were 130 jeel in height by sexlant measurement ...
To a distance oj 9 miles from its mouth
the banks were low and densely clothed with high mongroves.
the tide appearing to cover the whole country al times ...

John Sweatman (1846:87)

4.1 Introduction
If the broad ecological prerequisites outlined in the 'Introduction' are fulfilled at any
particular locality, a mangrove community is likely to develop. Such communities,
however, are not necessarily uniform structurally, floristically or functionally when
compared one with another, and even within anyone community, considerable
heterogeneity is apparent. Thus, mangroves can vary from tall, straight-trunked
forests up to 50 m in height, to gnarled, dwarfed communities less than 1 m high.
There can be little doubt that John Sweatmen would have been as amazed by the
dwarfed mangrove communities of the Red Sea. for instance, as he was impressed by
the tall mangroves at the mouth of the Fly River, Papua New Guinea, during the
surveying voyage of HMS Bramble.
Use of the term 'mangrove community' implies that there is some interaction
between the species and one can argue whether the mangroves form a community or
an assemblage (see 5.1). We should note that differences in, and among, mangrove
communities are due to a number of environmental factors, abiotic and biotic, which
act differentially on individual mangrove species. These factors lead to three types of
interactions: (1) those between the physico-chemical environment and the plants, (2)
those among the plants themselves, and (3) those between plants and animals. The
ultimate structure and function of any particular mangrove community is the
outcome of all these interactions, which will be examined in turn. In this chapter,
the interactions between the mangroves and their physico-chemical environment are
examined, while some of the biotic interactions are discussed in chapter 5.
As mangroves are a very specialized group of plants, only those factors, which
either are specific to the mangrove environment, or to which mangroves show an
interesting or unusual response, will be considered here.
A number of physico-chemical factors, arising out of the broad mangrove
environmental prerequisites, have been recognized as primary determinants of
mangrove growth and development. These can operate to modify one or more of the
essential life processes within the mangrove community, and consequently determine
whether or not a species is able to survive and grow at that particular locality.

4.2 Temperature
As mangroves are almost exclusively tropical, temperature has an obvious limiting
role. Because of its critical effect on both photosynthetic and respiratory processes,
temperature regulates a large number of internal energetic processes. Perhaps the
most important of these in relation to mangroves are salt regulation and extrusion,
water uptake, growth, and root respiration. How do mangroves deal with high and
102 Mangrove Ecology, Silviculture and Conservation

low temperature extremes and are there interspecific differences in the responses of
mangroves? These questions will be discussed in turn.

4.2.1 Mangroves and Low Temperatures


Frosts can severely affect plant growth, reproduction, and community processes
(Inouye 2(00), and mangroves appear to be particularly sensitive in this regard. The
limited tolerance of mangroves to low temperatures, and its effect on geographical
distribution, has been discussed already; Avicennia marina appears to be the most
tolerant of low temperatures, extending outside tropical and subtropical latitudes in
Australia, New Zealand and southern Africa. Chapman and Ronaldson (1958) and
Farrell (1973) considered Avicennia marina to be limited by the occurrence of killing
frosts, i.e. around -3°C. Nevertheless, Lange and Lange (1994) have shown that in
New Zealand, frosts are not the dominant factor restricting the present-day southern
limits of Avicennia marina; they suggest that the present mangrove distribution is
not in equilibrium with climatic conditions and that planted mangroves have
survived and reproduced further south. During the Holocene, mangroves occurred as
far south as Poverty Bay (Mildenhall and Brown 1987, Mildenhall 1994) but have
contracted northwards during subsequent cooler phases. At present, however,
mangroves are not limited in their distribution by climatic factors such as frosts, but
by an inability to traverse the large distances between suitable habitats (Lange and
Lange 1994).
Kangas and Lugo (1990) have similarly suggested that on the Atlantic and Gulf
of Mexico coasts of Florida, mangrove distributional limits are broadly controlled by
frost stress. In Florida, the frost line migrates up and down the peninsula over time
and appears to correspond with the transition zone of mangroves and salt marsh
(Snedaker 1995). Interestingly, Kangas and Lugo (1990) found that competition with
mangroves determines the southern limit of salt marsh ecosystems. On temperate
coasts, salt marshes are competitively superior due to adaptations for frost tolerance,
while on tropical coasts without frost, mangroves are competitively superior as they
are able to allocate more resources into structures that allow them to outcompete salt
marshes where they co-exist in south Florida (Kangas and Lugo 1990). This
competitive interaction is much less significant in those regions (such as India and
northern Australia) where tropical salt marshes co-exist with mangroves.
McMillan (l975a) and Markley et al. (1982) showed that both Avicennia
germinans and A. marina, collected from a range of localities and subjected to frost
under identical conditions, have populations selectively adapted to a latitudinal range
of habitats, including those with recurrent low winter temperatures. However, there
are limits; Sherrod and McMillan (1985) and Sherrod et al. (1986) reported that
despite the fact that viable propagules of red mangroves, Rhizophora mangle,
regularly arrive on Texas beaches, they are not naturalized on the Texas coast.
Propagules from Florida that were experimentally established on the southern Texas
coast at South Padre Island and at the mouth of the Rio Grande in February and April
1983 survived until the record freeze of December 1983. Propagules of unknown
origin that had apparently been artificially established on South Padre Island prior to
1983 had achieved heights ~2.5 m, and had reproduced successfully, prior to their
destruction in the subfreezing conditions in December 1983. Clearly, in the Gulf of
Mexico, the frequency, duration and/or severity of cold winter temperatures is a
prime factor governing the distribution and abundance of mangrove species.
Physico-chemical Factors and Mangrove Performance 103

Leaf scorch seems to be the predominant symptom of frosts (Chapman and


Ronaldson 1958), often followed by a reduction in the leaf area index due to leaf
mortality (Lugo and Patterson-Zucca 1977).
Smillie (1984) investigated the cold tolerances of Australian mangroves;
susceptibility to cold injury was assessed by the decrease in the rate of induced rise of
chlorophyll fluorescence in darlc-adapted leaves kept at O°C, a technique initially
developed for crop plants (Smillie and Hetherington 1983). Cold tolerance was
measured in 27 species of mangroves. Certain species such as Bruguiera exaristata
and Ceriops decanJra were very intolerant, accounting for their confinement to the
tropics and the warmer subtropics. Overall, a wide range of cold tolerances was
found, but, within genera, the cold tolerances of species were correlated with their
latitudinal distribution. In other words, the further south the species occurred, the
greater its cold tolerance. For Avicennia marina. the most poleward extending
mangrove, there was considerable cold adaptation in the southern populations
compared with the more northerly ones, confirming the glasshouse studies of
Markley et al. (1982) and McMillan (l975a).

4.2.2 Responses to High Temperatures


At the other extreme, the ecological response of mangroves to high air or water
temperatures is not well known. For example, mangroves growing in the discharge
areas of coastal power stations show little or no visible effects (Thorhaug et al.
1973, Saenger 1988). On the other hand, Canoy (1975) showed that Rhizophora
mangle in Puerto Rico developed more stilt roots per unit area where it was
subjected to a 5°C temperature increase from a cooling water discharge point, and
that in temperature-stressed areas this species formed more, but significantly smaller,
leaves (Lugo and Snedaker 1974, Canoy 1975). McMillan (1971) reported that
young seedlings of A vicennia germinans were killed by water temperatures of 39-
40°C in the Gulf of Mexico, although established seedlings and trees were not
damaged.
Temperature optima for photosynthesis in Florida mangroves was subject to
some seasonal variation, but for all species were below 35°C with little or no
photosynthesis occurring at 40°C (Moore et al. 1972, 1973). For Australian
mangroves, both assimilation rate and stomatal conductance are maximal at leaf
temperatures ranging from 25-30°C, and decline sharply with increases above 35°C
(Clough et al. 1982, Andrews et al. 1984, Andrews and Muller 1985, Ball et al.
1988).
Smillie (1984) investigated the heat tolerances of Australian mangroves by
measuring the decline in induced chlorophyll fluorescence following application of a
heat stress to the leaf tissue. Heat tolerance was determined by the decrease in
chlorophyll fluorescence after heating in water to 49°C for ten minutes in 20
mangrove species from tropical areas. All species showed a very high degree of heat
tolerance compared with other plants tested by the same technique; mangroves appear
to be at the extreme high end of the heat tolerance range for non-arid tropical plants.
The most heat-sensitive species were Acrostichum speciosum. Acanthus ilicifolius
and Rhizophora stylosa. Both Acrostichum and Acanthus grow in sunflecked shade
and are subject to short periods only of solar heating. Rhizophora stylosa. on the
other hand, most commonly grows in full sunlight, a situation difficult to reconcile
with its apparent heat sensitivity. There was no evidence for any latitudinal
104 Mangrove Ecology, Silviculture and Conservation

differentiation of heat tolerance in Avicennia marina and Aegiceras comiculatum, the


only two species tested (Smillie 1984). Not surprisingly, Aegiceras comiculatum
has shown an adaptability to varying temperature regimes by evaporative cooling
(Youssef and Saenger 1998b).

4.2.3 Inter- and Intra-specific Differences


Three thermal groups of plants have been identified in the Australian vegetation,
based both on species distribution and on the threshold temperature at which shoot
growth is initiated (Specht 1981a. c). This approach has global application. Thus, in
the tropical-subtropical group of plants, shoot growth is initiated when the mean air
temperature rises above 25°C; the warm-temperate group shows shoot growth
between 15 and 25°C; and the cool-temperate group shows shoot growth when the
mean air temperature rises above lOoC. This grouping is similar to groupings used
elsewhere, such as in China, where Li and Lee (1997) classed mangroves into (1)
cold-resistent eurytopic species such as Kandelia candel, Avicennia marina and
Aegiceras comiculatum; (2) cold-intolerant (thermophilic) eurytopic species such as
Rhizophora stylosa, Bruguiera sexangula, B. gymnorhiza, Excoecaria agallocha and
Acrostichum aureum; and (3) thermophilic stenotopic species such as Rhizophora
mucronata, R. apiculata, Lumnitzera littorea. Nypafruticans and Pemphis acidula.
From species distributions alone, mangroves belong predominantly to the
tropical-subtropical group, although some species extend considerably beyond the
sUbtropics (Tables 2.2 and 2.3). However, based on leaf-growth data (as very few
shoot-elongation datasets are available for mangroves), a better resolution of thermal
groupings can be obtained.
Saenger and Moverley (1985) and Hutchings and Saenger (1987) presented data
on the monthly production of new leaves for nine species of mangroves from
Gladstone (24° S) and sixteen species from Proserpine (20° 30' S), eastern Australia,
in relation to air temperature. These data permitted a tentative allocation of
mangroves into Specht's (1981a) three thermal groupings. Of the mangroves studied,
the majority were classified as warm-temperate, with leaf production ceasing below
16-18°C and maximal leaf production just under 30°C. Only one of the studied
species, Xylocarpus granatum, was classified as tropical-subtropical, with leaf
production ceasing below 26°C and maximal leaf production above about 30°C.
Clearly, however, such mangroves as Camptostemon and Scyphiphora, which only
occur further north than the study areas, should probably also be classified as
belonging to the tropical-subtropical group.
Only Avicennia marina was classified as cool-temperate by Hutchings and
Saenger (1987), with leaf production ceasing below l2°C and maximal leaf
production at 20°C. These temperature optima are reflected in its broad geographical
range. In South Africa, however, Avicennia marina appears to respond differently to
temperature. Steinke and Naidoo (1991) investigated respiration and net
photosynthesis of discs excised from cotyledons of A vicennia marina at temperatures
of 17, 21 and 25°C in a Gilson respirometer. Rates of respiration and net
photosynthesis were significantly different at the three temperatures, with the lowest
rates recorded in each case at l7°C and the highest at 25°C. While seedlings at 25°C
grew well, very little growth was recorded at l7°C. Similarly, at saturating light
intensities, maximal CO2 exchange in Avicennia marina occurred at a temperature of
31°C in South African plants (Naidoo et al. 1998), suggesting that there may be
Physico-chemical Factors and Mangrove Performance 105

clinal physiological adaptation between populations of this species throughout its


geographic range, a finding consistent with that described by McMillan (1975a) from
seedling growth studies.
The only other species which can also be classified as cool-temperate is
Kandelia candel which occupies the northern limit of global mangrove distribution
and shows differential cold tolerance throughout its range (Maxwell 1995).
In Rhizophora stylosa leaf formation ceased below 16°C. Interestingly, working
with R. mangle, Miller (1975) demonstrated, by measuring leaf resistance, that the
stomata of this species are only fully open above 18°C, thereby restricting
transpiration and photosynthetic gas exchange at low temperatures. Thus, it appears
that both R. stylosa and R. mangle fit well into the warm-temperate group.

4.3 Light and Photosynthesis


The physiological and ecological role of light is one of the critical aspects of
mangrove growth and survival. Photosynthesis is necessary for production of food
by plants, and many plants have adaptations or physiological responses that
optimize this process. However, high light intensities are usually associated with
higher temperatures, higher risk of photo-oxidation, high levels of ultra-violet
(particularly UVh) radiation, and increased water losses (see 3.3.2). Consequently,
mangroves need to strike a balance between photosynthetic advantage and harmful
effects of other physical conditions associated with intense radiation.

4.3.1 Photosynthetic Efficiency of Mangroves


Mangroves show light response curves similar to other plants (fig. 4.1),
characterized by an initial, steep, linear response below a photosynthetic photon flux
(PPF) of about 300 Ilmol m-2 S-I, followed by a broad convexity transition until
saturation is reached at higher irradiance levels. However, because of their generally
low stomatal conductance and intercellular CO2concentrations, the light intensity at
which mangrove p~otosynthesis becomes saturated is relatively low (Attiwill and
Clough 1980, 1998, Ball 1988a, Cheeseman et al. 1991, Clough 1992, Tiiffers et
al. 1999).
In their study of mangrove photosynthesis, Attiwill and Clough (1980)
examined the relationship between light and photosynthesis in branches of Avicennia
marina; for branches both within and at the top of the canopy, they found that
photosynthesis became light saturated at a total short-wave radiation between 200-
400 W m- 2 (approximately equivalent to 400-800 PPF Ilmol m-2 S-I). The
photosynthetic efficiency of this species (0.0135 Ilmol C021lmol-1 m-2 S-1 PPF) was
low compared with non-mangrove plants. Furthermore, this quantum efficiency (QE)
was obtained only at low light intensities; at full midday light intensities
(approximately 1,000 W m-2) the quantum efficiency was reduced to approximately
one-tenth of this level. While such a decline in efficiency is shared with many
species, it appears to be of greater magnitude than in many non-mangrove species,
suggesting that although the photosynthetic mechanism of Avicennia is inefficient,
it is reasonably well adapted to shade conditions (Attiwill and Clough 1980).
lO6 Mangrove Ecology, Silviculture and Conservation

15

a a
a a
00 a
a
IO

O~---r---..---~---r--~
o 500 1000 1500 2000 2500
Photosynthetic Photon Flux

Fig. 4. I Photosynthetic light response curve for leaves of Aegieeras eomiculatum.


Units for net photosynthesis are in "mol m·l s'\ and PPF in "mol m- 2 sol. The fitted
curve represents the equation y=xI(24.942+O.I4Ox). r 2=0.47. p<O.OOI. (Data from
Youssef and Saenger 1998b)

By way of contrast, the photosynthetic responses of seedlings of A vicennia


marina suggest that they are best adapted to growing in exposed conditions and
appear to have a low capacity to acclimate to low light intensities. However, the
seedlings are able to use sunflecks very efficiently, and consequently are able to
survive in the understorey environment (Ball and Critchley 1982).
Table 4.1 Quantum efficiency of mangroves derived from field measurements. (Data collected by the
author)
Species Quantum Efficiency Location.
("mol of COz "mol-I PPF m-z S·I) number of replicates
Mean (:tsd) Range
A. speciosum 0.005±O.002 0.002-0.007 Ballina. n=9
A. annulata 0.006±0.002 0.004-0.009 Darwin. n=6
A. eomieulatum O.OO9±O.OO6 0.002-0.022 Ballina. n=75
A. eomieulatum 0.008±O.007 0_001-0.045 Darwin. n=54
A. marina 0.017±O.012 0.002-0.048 Ballina. n=92
A. mariTUJ 0.0IO±O.007 0.002-0.032 Darwin. n=66
B. exaristata 0.01 I±O.O 10 0.002-0.040 Darwin. n=53
B. gymnarhiza 0.0 17±O.0 19 0.002-0.084 Ballina. n=42
E. agal/oeha 0.018±O.023 0.002-0.166 Ballina. n=73
E.ovalis 0.0 II ±O.OO9 0.002-0.034 Darwin. n=54
H. tiliaceus O.OIO±O.OO4 0.003-0.015 Ballina, n= 13
R. stylosa 0.018±O.016 0.004-0.054 Ballina. n= 15
S. alba 0.014±O.008 0.004-0.024 Darwin. n=5

Data on quantum efficiency of selected species are presented in Table 4.1, and
indicate that other mangroves are similar in terms of their photosynthetic efficiencies
to Avicennia marina. What seems clear is that, at increasingly lower latitudes, light
is present above saturation levels for most mangroves for more and more of the time
and, apart from the problems of raising leaf temperatures (3.3.2), excessive irradiance
may predispose mangroves to photoinhibition. Thus, Attiwill and Clough (1980)
showed that at Westernport Bay in Victoria a decrease in photosynthesis of
Physico-chemical Factors and Mangrove Performance 107

Avicennia marina occurred on days of sustained high radiation levels. ascribing this
to a photochemical inhibition of the photosynthetic mechanism.
Different light and shade requirements have been noted in mangroves. and
geographic variation among adults. seedlings and saplings is apparent (Hutchings and
Saenger 1987). In their detailed study of Rhizoplwra mangle. Farnsworth and Ellison
(1996) showed that some species have properties of both 'light-demanding' and
'shade-tolerant' species. defying designation. In addition. the mode and magnitude of
the adaptability of R. mangle in relation to light intensity changes with age.
Nevertheless. two broad groups of mangroves seem to emerge (Table 4.2): (1)
those which are somewhat shade-tolerant both as seedlings and as adults and (2) those
which appear to be shade-intolerant. Some species may be shade-intolerant in the
seedling stage but shade-tolerant as a tree or vice versa. The shade-tolerance of
mangroves is further discussed in 4.3.2.
Table 4.2 Presumed shade-tolerance characteristics of mature mangroves based on field and laboratory
observations.

Species Shade-tolerant Shade-intolerant


Acanthus ilicifolius II 5,9
Acrostichum speciosum I 5,9
Aegialits annulata 9
Aegiceras comiculatum 4,8
Avicvennia germinans 13
Avicennia marina 4, 7 5,8,10
Bruguiera spp 5,11
Bruguiera parvijlora II
Camptostemon schultzjj 11
Ceriops australis 11
Ceriops decandra 11
Ceriops tagal 8 5,9,11
Exceocaria agallocha 6, II
Laguncularia racemosa 11, 12, \3, 14
Lumnitzera spp. 11
Osbomia octodonta 6
Pelliciera rhizophorae 11
Rhizophara mangle 3,14 2, II
Rhizophora racemosa 11
Rhizophora spp. 5 9,11
Scyphiphora hydrophyllacea 5
Sonneratia alba 9, II
Xylocarpus spp. 6, II

lMedina et aI. 1990; 2Snedaker 1995; lparnsworth and Ellison 1996. 4Clarke and Hannon 1971;
5Macnae 1966; 6Hutchings and Saenger 1987; 1Attiwill and Oou¥.h 1980; l7hom et at. 1975; 9Macnae
1968; lOSall and Critchley 1982; l1Smith 1992; 12McKee 1995c; I.Rabinowitz 1978; 14Ba1l1980.

Leaves developing in high light intensity show a higher degree of xeromorphy


than those protected from it (Wylie 1949. Saenger 1982. Givnish 1988. Farnsworth
and Ellison 1996). Consequently. it is possible that some of the xeromorphic
features discussed earlier (3.3.1) may be responses to high light intensities rather
than (or in addition to) water shortages. Isobilateral leaf anatomy is generally
regarded as a xeromorphic character. but when it is combined with a mechanism for
orientating the leaf towards the sun (as occurs for example in Lumnitzera.
Laguncularia and Pemphis). it seems reasonable to assume that a light response also
may be involved. Other leaf characteristics associated with high light intensities
include such xeromorphic features as a high ratio of volume to surface area and a
well-developed. highly differentiated. often isobilateral palisade mesophyll. These
features are present in many mangrove species and. in conjunction with the leaf
108 Mangrove Ecology, Silviculture and Conservation

pattern of arrangements on the shoots (Tomlinson and Wheat 1979), may constitute
an adaptation optimizing exposure to light under varying conditions of light and
shade (Farnsworth and Ellison 1996).
As leaves developing in intense light show a greater degree of xeromorphy, one
can distinguish between 'sun' and 'shade' leaves. Several mangrove species showed a
marked morphological differentiation between sun and shades leaves, particularly the
palaeotropical Avicennia, Aegiceras, Bruguiera; Ceriops, Lumnitzera and Rhizophora
(Saenger 1982, Lovelock et al. 1992) and the neotropical Avicennia germinans,
Rhizophora mangle and Laguncularia racemosa (Lugo et al. 1975, Farnsworth and
Ellison 1996).
Shade leaves of Ceriops fagal, when compared with sun leaves, are larger,
thicker, have a higher volume-to-surface ratio, possess fewer stomata per unit leaf
area on the lower leaf surface, and possess a proportionately thicker tannin-filled
hypodermis on both upper and lower surfaces. In addition, shade leaves have a
proportionately thinner upper palisade mesophyll. lower epidermis and lower cuticle
(Saenger 1982). In Rhizophora mangle, Farnsworth and Ellison (1996) found that
total foliar thickness was higher in sun leaves, with thickening of the hypodermis
and, to a lesser extent the cuticle, contributing to leaf thickness. Shade leaves were
generally heavier, larger and longer than sun leaves, and were displayed more
horizontally. Physiologically, shade leaves of trees and seedlings (though not of
saplings) of R. mangle had higher quantum yields (Table 4.3), smaller dmX
respiration rates and maximal assimilation at lower light intensities than sun leaves.
Lugo et al. (1975) compared sun and shade leaves of three neotropical mangrove
species, showing that net daytime photosynthetic rates in Rhizophora mangle and
Avicennia germinans were about twice as high in sun leaves as in shade leaves. At
night, the shade leaves had respiration rates that were four times as high as those of
sun leaves, a finding that does not appear to be supported by the studies of
Farnsworth and Ellison (1996). R. mangle and A. germinans behaved differently in
terms of transpiration; Rhizophora sun leaves had a higher transpiration rate than the
shade leaves while in A vicennia, and also in Laguncularia racemosa; the sun leaves
had lower transpiration rates when compared with shade leaves.
Table 4.3 Photosynthetic efficiency of 'sun' and 'shade' leaves of Rhizophora mangle. (Measurements
conducted on Wee Wee Cay, Belize; n;;: 50 for all means; data from Farnsworth and Ellison 1996).

Leaf type Quantum Efficiency


(fJmol of CO2 fJmol-1 PPF m·2 S·I)
'Sun' tree leaves 0.016
'Shade' tree leaves 0.020
'Sun' saplings 0.027
'Shade' saplings 0.010
'Sun' seedlings 0.017
'Shade' seedlings 0.071

Seedling leaves of A vicennia marina also can be subdivided into sun and shade
leaves on a morphological basis; the shade leaves contained more chlorophyll on
both a leaf-area and fresh-weight basis, were rich in chlorophyll b relative to a, and
had a lower specific weight and greater leaf area than sun leaves (Ball and Critchley
1982). In terms of gas exchange and photosynthetic characteristics, however,
distinction between sun and shade leaves could not be made; both leaf populations
were typical of sun leaves. In southern Australia, A. marina sun leaves have a rate of
photosynthesis approximately 4.5 times that of shade leaves (AttiwiII and Clough
Physico-chemical Factors and Mangrove Performance 109

1980), although at high light intensities both types of leaves can potentially reach
the same maximum rate of photosynthesis per unit of leaf surface.
As discussed above, the photosynthetic rates of mangroves saturate at relatively
low light levels despite their occurrence in tropical environments with high intensity
sunlight. Clearly, the low photosynthetic rates of mangroves result in light
requirements for maximal photosynthesis being considerably less than amounts of
light available on cloudless days. Field-grown mangroves become light-saturated at
25-50% full sunlight (Ball and Critchley 1982) and low photosynthetic rates under
saturating irradiance inevitably result in an excess excitation energy. A
superabundance of PPF becomes potentially hazardous, particularly for photosystem
II (PSII), which may be photoinhibited when more light is absorbed than can be used
in photosynthetic photochemistry. Photoinhibition is a light-dependent loss in
photosynthetic functioning of PSII, which is manifested in whole leaves as a decline
in quantum efficiency under low light conditions. Furthermore, salinity-induced
limitations in rates of photosynthesis may cause leaves of plants in highly saline
environments to be even more subject to photoinhibition than those with higher
photosynthetic capacities in less saline habitats (Attiwill and Clough 1980,
Bjorkman et al. 1988, Cheeseman 1994, Ball 1996, Sobrado I 999a, Sobrado and
Ball 1999).
Using chlorophyll fluorometry, Tiiffers et al. (1999) showed that under fully
exposed conditions, there was a significant mid-day depression in the intrinsic PSII
quantum yield in dark-adapted leaves of both A vicennia marina and Bruguiera
gymnorhiza. While complete recovery occurred in the afternoon, the degree of the
mid-day depression was strongly correlated with the direct solar radiation received by
the leaves which, in tum, was related to their leaf angle. While both species
possessed photoprotective mechanisms, as evidenced by the complete aftemoon
recovery of photosynthesis, avoidance of excess radiation by steep leaf angles was
important for both species but particularly for B. gymnorhiza during months with
high solar altitude and high radiation loads (i.e. summer).
Mangroves appear to have the ability to protect PSII by dissipating excess
radiation energy as heat (Bjorkman et al. 1988), via the epoxidation of xanthophylls,
particularly zeaxanthin (Lovelock and Clough 1992, Gilmore and Bjorkman 1994),
and through the conversion of the superoxide anion 02· to phenolics and peroxidases
(Cheeseman et at. 1997). However, despite these photoprotective mechanisms,
adjustment of leaf angles to reduce incident PPF (and temperature) remains a major
adaptive device of mangroves to their environment (Ball 1996, Clough 1998, Tiiffers
et al. 1999).
In the tropics, mangroves are also subject to elevated levels of ultraviolet
radiation when compared to plants of subtropical and temperate regions. Ultraviolet
radiation, particularly UV-b, inhibits plant growth by disrupting protein synthesis,
and depressing photosynthesis via stomatal closure. UV -absorbing phenolic
compounds that protect against the damaging effects of UV -b have been found in the
leaf epidermis of mangroves (Lovelock et al. 1992). The sun leaves of Bruguiera
parvijlora, B. gymnorhiza, Rhizophora stylosa and Xylocarpus granatum have
greater contents of phenolic compounds than shade leaves, and more saline sites have
plants with greater levels in their leaves than less saline sites. Sun leaves were
generally more succulent than shade leaves and these authors concluded that the
higher levels of soluble phenolic compounds (probably flavonols) and succulence in
110 Mangrove Ecology, Silviculture and Conservation

sun leaves compared with shade leaves of the investigated species jointly provide
effective protection against UV radiation.
Finally, the photosynthetic carbon metabolism of two Indian species of
mangroves was investigated by Joshi et al. (1974, 1975b) who concluded that they
were of the aspartate type. They suggested that this type of metabolism may be due
to the inhibition of the enzyme malic dehydrogenase in the presence of high sodium
chloride concentrations. To test this hypothesis, plants were labelled with radioactive
carbon and the early products were identified; aspartate and alanine both had become
heavily radioactive and, consequently, Joshi et al. (1974, 1975b) concluded that the
C 4 carbon fixation pathway operated in these species.
Subsequent studies, however, have shown that this biochemical adaptation of
photosynthetic carbon fixation does not occur in mangroves, and evidence, drawn
from leaf anatomy, the 13C/12C carbon isotope ratio, and gas exchange properties
(Cowan 1978, Clough et al. 1982, Ball 1986. Naidoo and von Willert 1999),
indicates that mangroves possess the more common C 3 photosynthetic carbon
metabolism. The ol3C values for mangroves consistently fall within the range of
-20 to -30%0 (Rico-Gray and Sternberg 1991, Ball 1996), indicating that they are C 3
plants.
Overall, the photosynthetic apparatus of mangroves does not appear to differ
greatly from that of other plants, although the light intensity at which mangrove
photosynthesis becomes saturated is relatively low. As a result, those features
involved in minimizing the adverse effects on photosynthesis of high leaf
temperatures, high light intensities, and high levels of UV are particularly well
developed (see 3.3.3). In combination, these features suggest that mangroves as a
group are well adapted to a low light environment. Yet, paradoxically, as discussed
in the next section, this is not the case.

4.3.2 Canopy Gaps: Gap-enhancement of Reproduction and Growth


A further aspect of light that we need to consider is the role of canopy gaps in
mangrove regeneration and stand composition. Canopy gaps in mangroves may
result from natural events such as storms and lightning strikes (Johns 1986, Smith
1992, Sherman et al. 2000, Duke 2001), from biotic agents such as fungi, termites
or wood-boring beetles (Pegg et al. 1980, Tho 1982, Feller and Mathis 1997, Feller
and McKee 1999, Wier et al. 2000), or human activity (Putz and Chan 1986, Ewel
et al. 1998c, Allen et al. 2001). However formed. gaps show differences in the
physico-chemical environment when compared with adjacent intact forests, including
increased light, nutrients, and soil temperatures, and decreased humidity (Ewel et al.
1998c, Feller and McKee 1999, Sherman et al. 2(00). In other respects, they may be
similar to areas under an intact canopy; for example, Feller and McKee (1999) found
no differences between gaps and intact canopies for sulfide concentrations, interstitial
salinity and redox potential while Clarke and Kerrigan (2000) found no differences in
total phosphorus (TP), concluding that soil resources and conditions were more
strongly controlled by the range of sediment types than by the modifying effects of
natural gap creation within them. Some biological processes may also differ in gaps
compared with intact canopies; these include predation of propagules by crabs
(Osborne and Smith 1990).
Koch (1997) investigated the development of seedlings (25-85 cm) to the
sapling (>85 cm) stage in Rhizophora mangle in southern Florida, concluding that
canopy gaps are important for this transitional development. Growth of R. mangle
Physico-chemical Factors and Mangrove Perfonnance 111

seedlings in gaps was two to five times greater than in adjacent closed canopies with
around 80% PAR extinction. While other factors (such as labile inorganic P, TP,
TN and redox potential) also influenced stem elongation and leaf production, Koch
(1997) found that light availability (around light saturation levels of >500 ~mol m· 2
S·I) was the primary detenninant under low salt stress. Based on these results, and on
the findings of some earlier studies, Koch (1997:436) suggests that 'a majority of
tropical forest species including mangroves are characterized as gap-dependent'.
Some data seem to support this suggestion. For example, Smith (l987c)
showed that with 80-85% reduction in PPF, seedling survival and growth of
Avicennia marina, Ceriops tagal and Rhizophora stylosa were significantly lowered
in north-eastern Australia. Similarly, Clarke and Allaway (1993) concluded that full
sunlight was essential for the growth of A. marina seedlings to the sapling stage in
southern Australia. In Thailand, Tarnai and Iampa (1988) showed that survival and
growth of Rhizophora and Bruguiera was considerably enhanced in gaps. Increasing
light availability by canopy removal increased growth rates of suppressed
Rhizophora mangle seedlings in Belize (ElIison and Farnsworth 1993) and small gap
(12-72 m 2) fonnation by tree fall or wood-boring beetles was reported to enhance
survival of seedlings of Rhizophora mangle. Avicennia germinans and Laguncularia
racemosa (Feller and McKee 1999).
On the other hand, Farnsworth and Ellison (19%) have demonstrated the
extremely wide phenotypic plasticity of Rhizophora mangle in relation to light,
indicating that this species is capable of growing (albeit slowly) under closed
canopies. Similar findings were made by Kathiresan and Ramesh (1990), who
examined seedling establishment of Rhizophora species in the open and under
canopies in the Pichavaram forest in Tamil Nadu, India. They found that
significantly more seedlings (6.1 seedlings m· 2 as against 1.5 seedlings m· 2) became
established and survived in 100 m 2 plots under intact canopies than beyond the
canopy. While the protection offered by proproots under the canopy could not be
ruled out as a contributing factor, they concluded that seedling establishment under
the canopy occurs widely and should be used for enrichment planting. Turner et al.
(1995) investigated the architecture and allometry of mangrove saplings (1.5-3 m) at
Sungei Merbok, Malaysia, both in open and shaded conditions. They found
considerable allometric plasticity in mangroves growing under canopies in
comparison to those in the open. For a given height of saplings, those growing in
the shade had narrower crowns, fewer orders of branching, and fewer leaves at the
tops of the crown compared with those species or conspecifics growing in the open.
This relative absence of leaves at the base of the sapling, very pronounced in the
shaded popUlations, results in monolayered crowns which maximize light capture
under low irradiance.
On Kosrae, Micronesia, Ewel et al. (l998c) examined seedling densities in
natural and artificial gaps (mean size 158 m1), both in fringe and basin mangrove
forests. Seedlings of Rhizophora apiculata and Sonneratia alba were confined to gaps
in the fringe forest while other species occurred at similar densities in gaps and under
the canopy. In basin forests, however, no significant differences in seedling densities
were observed between sites in gaps and under canopies for any species. Additionally,
there were significantly more seedlings of all species, and of Rhizophora apiculata in
particular, under the canopy in the basin forest (0.84 seedlings m o2 ) than in the fringe
forest (0.13 seedlings mo2 ).
112 Mangrove Ecology, Silviculture and Conservation

Shennan et al. (2000) investigated regeneration in small gaps in mangroves of


the Dominican Republic. They found that seedling densities under intact canopies
and in canopy gaps were highly variable but, when compared across all plots and
gaps, the seedling density in canopy gaps was not different compared from that under
the intact canopy for Rhizophora mangle, Avicennia germinans and Laguncularia
racemosa. However, when they examined seedling densities in the different
vegetation zones, a different pattern emerged. Density of R. mangle seedlings was
significantly greater under the canopy than in canopy gaps in the Rhizophora forest.
In the Rhizophora-Laguncuklria forest, A. genninans and L. racemosa seedling
densities were greater in the understorey than in gaps, while R. mangle seedling
density was similar in gaps and under the intact canopy. In the Laguncularia forest,
seedling densities were not different between the closed canopy and gaps for any of
the three species. In tenns of sapling densities, Shennan et al. (2000) found that
those of R. mangle were significantly greater in canopy gaps than under the intact
canopy in all vegetation zones and that the density of R. mangle saplings was
significantly greater than that of the other two species both in gaps and under forest
canopy, suggesting that gaps favoured R. mangle.
In a detailed study of the influence of gaps on population structure and species
composition of mangrove stands in north-eastern Australia, Clarke and Kerrigan
(2000) examined the physico-chemical, structural and floristic differences between
canopy gaps and adjacent forests with intact canopies. They found that propagule
abundance was similar between all sites, and sapling (>1 m high) abundance, while
highly variable, was often higher in gaps than in the intact forest. Nevertheless,
saplings were common in the understorey of forests with continuous canopies,
indicating that recruitment to the sapling stage is possible in the absence of gaps.
No significant differences in the abundance of seedlings «1 m high) of the
three most common species (Rhizophora stylosa. Bruguiera gymnorhiza and Ceriops
tagal) were detected between canopy gaps and intact canopy plots. The floristic
composition of their plots varied more among study areas and sites within areas than
between gaps and surrounding forests, and the hypothesis that floristic composition
of canopy gaps is distinct from the surrounding forest was rejected. Similarly, they
were unable to detect any significant canopy effects on the species richness, although
they found a slight indication that Avicennia marina was more abundant in gaps than
under intact canopies, a feature they ascribed to the resprouting ability of this species
after canopy breakage.
As Clarke and Kerrigan (2000) concluded, there is little evidence of gap-
dependency in mangrove forests, although gap-enhancement of recruitment and
growth of all species appears to be widespread, particularly in small «100 m 2) gaps.
Gap-phase regeneration of mangroves appears to follow a simple 'direct replacement'
model in which a population of a particular species is replaced by members of the
same species in forest gaps. This contrasts markedly with tropical rainforests where
canopy gap specialists are common (Whitmore 1989). Such specialists are generally
characterized by donnant and persistent seeds with protracted viability, and high
initial growth rates following gennination (Stewart 1995). It would appear that the
very constraints of the mangrove environment which have led to large, nondonnant
propagules with limited viability (Saenger 1982) in mangroves as a group (see
3.5.1), have thereby prevented the development of a 'gap specialist strategy' in that
environment.
Physico-chemical Factors and Mangrove Performance 113

4.3.3 The Mangrove Understorey and the Role of Sulfides


If gap specialists have not evolved in mangrove systems, it might also be pertinent
to ask whether understorey specialists have evolved. As discussed earlier, seedlings
and saplings of the canopy species can regenerate under intact canopies, but are other
species present as a distinct understorey community?
In a provocative paper entitled 'Mangroves: where's the understory?', Janzen
(1985) suggested that mangroves are unique among tropical forests in lacking
reproducing herbs, shrubs and vines, with only the mangrove fern, Acrostichum
approximating to a mangrove understorey herb. Janzen (1985) then offers three
hypotheses: (1) plants with low light resources cannot accumulate enough reserves
to meet the metabolic demands of the drain of the machinery and morphology for
salt tolerance; (2) the herb, shrub and vine life forms are intrinsically incapable of
growing in saline soils; and (3) the characteristically small seeds of herbs, shrubs
and vines do not produce a sufficiently robust seedling to withstand the abiotic
conditions and/or grazing conditions of a mangrove swamp.
Before dealing with these hypotheses, it is instructive to examine whether
herbs, shrubs and vines are invariably absent from the mangrove understorey. Corlett
(1986) has suggested that, although most mangrove forests do lack an understorey
and vines, in Singapore and southern Malaysia the shrub Brownlowia tersa forms a
dense, if patchy, understorey, while the large woody vine Finlaysonia obovata is
found rooted among and scrambling over Rhizophora species in areas inundated by
normal high tides. In the mangroves of north-eastem and northern Australia, patchy
understories are commonly present under intact canopies and include species of non-
canopy mangroves such as Acanthus ilicifolius, Acrostichum speciosum, Aegialitis
annulata, Clerodendrum inerme and Aegiceras corniculatum. and salt-tolerant herbs
such as Suaeda arbusculoides, Batis argillicola and Tecticornia cinerea. Vines are also
frequent in the moister mangroves of north-eastern Australia (see listing of species
in Hutchings and Saenger 1987).
Lugo (1986) drew attention to the existing information on known mangrove
understorey plants and reiterated Chapman's (1975) findings that the records of
understories and vines are from high rainfall locations, the landward portion of
mangroves, and from low salinity soils.
More significantly, Lugo (1986) points out that stressors other than shade and
salinity exist in mangroves, and he generalized Janzen's (1985) hypothesis (1) as
follows: understorey plants grow in those mangrove ecosystems where combinations
of nutrients, light energy, soil oxygen, and freshwater meet the metabolic costs of
all environmental stressors converging on the site. Snedaker and Lahmann (1988)
have 00ded an evolutionary slant, suggesting that the very high metabolic cost of
existence in the mangrove environment has precluded the evolution of plants with
salinity tolerance, waterlogging tolerance, and shade tolerance. Thus, they suggested
that the absence of an understorey is associated with the combined effects of shade,
salinity, and the persistence of anoxia, particularly with high concentrations of
sulfides in mangrove sediments.
In stagnant, flooded soils, roots of many mangroves develop a very thin,
slightly oxidized zone that can effectively isolate the actively growing root area from
the highly concentrated phytotoxins by oxidative detoxification (Thibodeau and
Nickerson 1986, Youssef and Saenger 1996). This process of radial oxygen loss via
fine rootlets is particularly significant for seedlings (of either canopy mangroves or
understorey species) which do not have the well-developed array of proproots and
114 Mangrove Ecology, Silviculture and Conservation

pneumatophores of their adult counterparts. Radial oxygen loss relies on the internal
ventilation system which, in tum, depends on photosynthesis. As seedlings of
mangroves are very sensitive to soluble sulfide concentrations (Youssef and Saenger
1998a), any reduction in photosynthesis (e.g. shading) will reduce the protective
oxidized zone around the roots of the seedling, with soluble sulfides damaging the
root cell membranes and interfering with the roofs ability to discriminate against
sodium, chloride and other toxic ions.
Understorey plant survival in mangroves depends on specific environmental
conditions, including light availability and dissolved sulfide concentrations in the
sediments (Lugo 1986). Where light is available, the process of radial oxygen loss
from the roots of seedlings will detoxify the soluble sulfides; where light is limited,
soluble sulfides become a potent phytotoxin. As light is nearly always limiting
under intact canopies, the presence of understorey plants is indicative of low sulfide
concentrations. Localized areas of low sulfide concentrations are the result either of
coarser sediments which generally do not allow sulfides to accumulate, groundwater
or surface flows which continuously remove or oxidize soluble sulfides, or of
burrowing activities of invertebrates that assist in the oxidation of soluble sulfides.

4.4 Wind and other Atmospheric Phenomena


Mangroves, like other plant communities, are affected by many meteorological
events and conditions. Wind effects are most consistent but unusual events such as
tropical storms, hail-storms and thunderstorms cause major damage despite their
infrequency. Each of these phenomena is reviewed below.
Wind affects mangroves in many ways: coastal water drift and tidal currents are
modified by wind direction and speed; wave action is accentuated, especially at high
tides, by stormy conditions; both waves and water movement affect sediment
transport; wind plays a major part in causing evaporation and in increasing salinity,
and it can cause physical damage to canopies and desiccate foliage under more
extreme conditions. On the positive side, it may facilitate pollination and the
dispersal of propagules in a number of wind-pollinated species such as Rhizophora
and Excoecaria. There are, however, three aspects of wind that impinge directly on
the physiological performance of mangroves: its evaporative capacity, its effect on
sealevel, and its role in regulating evapotranspiration from leaves.

4.4.1 Ambient Winds


Most of the standard wind data do not relate directly to the mangrove environment
because wind recording is usually done close to the ground and at some distance from
the coast. The sea surface causes less mechanical and thermal obstruction to air flow
than land and wind speeds are therefore greater over water. As winds flow on to the
shore, frictional drag of the land surface reduces wind speeds, but this is progressive
and the mangrove zone tends to experience the speeds characteristic of winds over the
sea. Recording stations are often on the landward side of the sharp decrease in wind
speeds from sea to land.
Despite this shortcoming in wind data, it seems clear that climatic factors such
as humidity, wind velocity and higher solar irradiance, together with the degree of
plant cover, have a significant influence on evaporative losses from the mangrove
environment. These evaporative losses, together with the frequency of inundation,
largely determine soil salinity. The most severe conditions characterize the arid and
semi-arid climates in the subtropics and tropical margins where highly seasonal or
Physico-chemical Factors and Mangrove Perfonnance 115

low rainfall combines with high all-year-round evaporation. For example, in the Red
Sea, the Arabian Gulf, the Horn of Africa, and in Mauritania, where extremely arid
conditions exist, only dwarfed and species-poor mangroves occur (Saenger and Bellan
1995, Dodd et at. 1999).

How Avicennia marina got its name!


The evergreen vegetation fringing the desert landscapes of the Red Sea and the
ArabianlPersian Gulf have long aroused curiosity; already in the fourth century se,
Theophrastus (the pupil of Plato and Aristotle) described the mangroves of the Red
Sea in his Peri phyton historia (Enquiry into Plants). At around the same time, Admiral
Nearchus, the commander of the fleet of Alexander the Great, described the
mangroves of Tylos, present day Bahrain, while conducting a military reconnaissance
between the Indus delta and the Euphrates. Later in the eighteenth century, the Danish
botanist Pehr Forsskal, a student of Linnaeus, commenced his extended exploratory
travels to the Middle East in 1761. In his Flora Aegyptiaco-arabica (published in 1775,
twelve years after his death from malaria near Sanaa, Yemen), he first described
Avicennia marina, the most widely distributed of all mangrove species, from the Red
Sea. He called it Sceura marina to latinize. and perpetuate, the Arabic name of this
species - schura, characterizing it as '... frequens in Insulis ad littoribus maris rubri ...
Folia pabulum praebent Cumelis, asirus, ovibus narrarunt'. Unbeknown to Forsskal, his
mentor had described a plant from India in his 1753 Species Plantarum as Avicennia
officinalis - after the famous Persian philosopher-scientist of Islam, A vicenna or Ibn
Sina, author of a Book of Healing, which wa.~ the medical authority in Europe for
several centuries. Ultimately, Sceura marina became Avicennia marina, thus
combining the generic name of Linnaeus, the ma.~ter, with the specific name of
Forsskal, the loyal student! Additionally, but quite fortuitously, the Middle Eastern
connection was also maintained in the new name. Less than 20 years later, in 1794,
Jean Louis Poiret described the second most widespread mangrove, Rhizophora
mucronata, from Madagascar.

The inland margin of mangroves is particularly prone to high evaporative


losses and drying out of the substrate. An edge effect is often noticeable where
mangroves abut salt flats. The evaporative build-up of soil salinity results in
mangrove dieback and gradual expansion of the salt flats. Similarly, where breaks
occur in the canopy, especially in the mangroves towards the landward margin,
evaporation may lead to increased soil salinities which, in turn, may prevent the
regeneration of mangroves (SpenceJey 1976). In the more humid tropics, on the
other hand, rain wetting of leaf surfaces, cloud cover. and high humidities reduce
evaporative losses, and the tendency towards salt flat fonnation is not so great.
Wind affects evapotranspiration from mangroves by the same mechanism as in
other plants. Because mangroves are at the land-sea interface, however. they tend to
be more consistently exposed to windy conditions, and wind probably assumes a
greater importance in relation to evapotranspiration in mangroves than in other plant
communities. As transpiration occurs, there is a tendency for a moist layer of air to
form next to the leaf surface. This layer, termed the boundary layer, is variable in
thickness, but in those mangroves with epidermal hairs or scales it is thicker than
around those leaves with untextured surfaces. Wind conditions also affect the
thickness of this layer, with the greatest thickness in still air. The boundary layer
decreases the diffusion gradient between the leaf and the atmosphere, and
transpiration consequently decreases. On the other hand, air movement carries away
this layer of humid air, replacing it with drier air, thereby causing an increase in
transpiration. The more rapid the air movement. the faster the moist air will be
carried away and the higher will be the rate of transpiration. If the wind is strong,
stomata may close, possibly as a result of excessive water loss. and transpiration is
116 Mangrove Ecology. Silviculture and Conservation

subsequently reduced. Temperature also affects the opening and closure of stomata
(Miller 1975. Steinke 1979. Naidoo et al. 1998).
The inference to be drawn from the control of evapotranspiration by wind is
that a plant within the general mangrove canopy will experience different growing
conditions from one of the same species growing as an isolated individual at the
front or the back of the stand. Indirect evidence of this was noted when
Phytophthora-induced dieback became prevalent in the Port Curtis area on the central
Queensland coastline. Pegg and Foresberg (1981) showed that Phytophthora killed
its host (Avicennia marina) only when the host was also under some other sort of
stress. especially water stress. The first trees to die were those growing as isolated
specimens on mudbanks well away from dense mangrove stands; these were followed
by the very tall specimens which emerged beyond the general level of the canopy;
and it took approximately another year before the A vicennia within the mangrove
canopy showed any signs of dieback.
Seasonal wind shifts can also affect mean sealevels. Persistent onshore or
offshore winds can raise or lower the effective mean sealevel to a considerable degree.
Storm surges accompanying cyclones are a dramatic illustration of this (although in
this case there are pressure effects also). Wind patterns during normal conditions
operate on a smaller scale. but nevertheless have physiological consequences for the
mangrove community.
Seasonal changes of mean sealevel of up to I m occur in the southeastern part
of the Gulf of Carpentaria, northern Australia, owing to seasonally changing wind
patterns (south to southeast in winter and mainly north to northeast in summer).
Similar but less pronounced sealevel changes were detected from yearly tide
recordings at Port Curtis on the central Queensland coastline. where winter winds
from the south to southwest are offshore and summer-autumn winds from the north
to northeast are onshore. In Bangladesh. persistent onshore winds during the
monsoon period (July to August) causes mean sealevels to rise in the Gulf of Bengal
by between 0.3-0.7 m (Siddiqi 2(01). In Terminos Lagoon. Mexico. mean sealevels
rise by around 0.3 m during periods of persistent onshore winds in October and
November each year (Fuentes-Yaco et al. 200 1).
Even changes of 0.3 m on a relatively flat coastline represent a significant
increase in depth and frequency of tidal inundation as well as the area subject to tidal
inundation. However. the season during which such sealevel changes occur seems to
be ecologically important. For example. in the Gulf of Carpentaria as in the Bay of
Bengal. mean sealevel is raised during the summer months when the river discharges
are at their maximum (Saenger and Hopkins 1975). Consequently. flooding is
frequent and freshwater inundation of the mangroves and salt flats aids in the
leaching of salt from these communities. At Port Curtis. on the east coast of
Australia. however, sealevel is raised during the dry autumn season and. rather than
remove salt. may in fact contribute salt to those communities at or near high water
spring levels. particularly when winter evaporation is high.
Seasonal changes in mean sealevels are important for the distribution and
survival of mangrove species and. in terms of projected sealevel rises due to global
warming. this is an area worthy of detailed investigation (see 9.3.5).

4.4.2 Extreme Winds and Mangrove Vegetation


Extreme wind events present a different set of problems for mangroves. All coastal
areas situated on the western side of the major oceans. with a latitude greater than
Physico-chemical Factors and Mangrove Performance 117

about 5° N or S, are frequently subjected to extreme winds in the form of tropical


storms variously described as hurricanes, typhoons or cyclones; regions particularly
affected include the Caribbean, the Bay of Bengal, Vietnam and the Philippines.
Occurring at the land-sea interface, mangroves generally bear the full brunt of
such tropical storms; they can severely alter species composition and forest structure
as a result of sustained wind velocities of >200 km h-'. Extensive damage may be
sustained by the mangroves in the path of a tropical storm (Roth 1997); generally
this damage is immediate in the case of the vegetation, although some effects of
storm damage may not become apparent until after the storm has passed. Thus,
Smith et al. (1994) and Davis (1995) reported that mangrove tree mortality
continued for several months after hurricane 'Andrew' struck Florida and that the
resultant erosion caused ongoing damage to propagules and seedlings. Changes in
species composition may result from the differential loss of species during the storm
(McCoy et al. 1996) and may persist because of changed recruitment patterns
(Baldwin et al. 1995).

Fig_ 4.2 Damage to the mangroves at the mouth of the Wildman River, northern
Australia, seen here in 1981, 7 years after cyclone Tracy'.

Comparing the effects of extreme wind events in many parts of the world has
often been hampered by the unique criteria for measuring damage. Thus, following
tropical cyclone 'Tracy' (December 1974) in northern Australia (fig. 4.2), Stocker
(1976) assessed the damage to mangroves around Darwin, classifying wind damage
into four types: (I) windthrow, where the tree is felled; (2) crown damage, where the
leaves and twigs are removed and/or branches are tom off; (3) bole damage. where the
trunk is broken, severely fractured or leaning; and (4) standing dead. Because all these
damage types also can be caused by wave action (which generally accompanies high
winds), no distinction was made between wind and water damage (Steinke and Ward
1989).
Hurricane 'Joan' struck Isla del Venado, on the Nicaraguan Caribbean coast, in
October 1988 and selectively damaged and disproportionately killed the larger trees
(Roth 1992). Around 36% of the trees, representing 68% of the basal area of the pre-
118 Mangrove Ecology, Silviculture and Conservation

existing stand, were killed. Trees showing the highest rates of recovery one and a
half years after the storm were those of short stature and small diameter. Of the
surviving trees, 42.4% of trees 'well refoliated', and another 21.2% 'poorly
refoliated'. In addition, seedling densities in the three study areas ranged from 0.68-
1.82 seedlings m·2 with a stem elongation rate of around 2.7 mm d· 1 (Roth 1992), a
value remarkably similar to the stem elongation rate (3.1 mm d-I in open sites)
reported for Rhizophora mangle in Florida mangroves recovering from hurricane
damage (Koch 1997).
Hurricane 'Hugo' struck the Caribbean in September 1989 with sustained winds
of 230 km h-I and gusts up to 296 km h-I (Imbert et al. 1996). It passed directly over
the Bay of Grand Cul-de-sac Marin on the Island of Guadeloupe, with its 3,000 ha of
mangroves and, thus, has provided an opportunity to assess the damage to
neotropical mangroves. Typically, these mangroves are zoned, with an outer
Rhizophora coastal fringe, followed by a mixed dwarf zone, a Rhizophora-dominated
tall forest zone, and a landward A vicennia-Laguncularia zone.
Imbert et al. (1996) used a system of damage classification similar to that of
Stocker (1976) with 'minor' (defoliation and loss of twigs), 'major' (loss of main
branches, broken boles or leaning) and 'lethal' damage (windthrown or standing dead).
Overall, the mangrove forests were heavily affected by hurricane 'Hugo' but
significant differences occurred among the four vegetation zones. The tall mixed
forest was most affected, with a 78% decrease in average stem density and a 71 %
decrease in mean basal area, followed by the coastal fringe (59 and 68%), the mixed
dwarf (26 and 23%) and the Avicennia-Laguncularia (3 and 3%) zones (Table 4.4).
Stem mortality rates increased with diameter at breast height (DBH),
particularly in the tall mixed zone, while minor damage correspondingly decreased.
Over 50% of Rhizophora trees suffered lethal damage in the tall mixed zone arxl
around 20% of Avicennia trees showed lethal damage. No Rhizophora survived
broken boles while 40% of A vicennia trees survived stem breakage below breast
height.
Table 4.4 Mangrove stem density and basal area of trees with a DBH >3.8 cm before and after the
passage of hurricane 'Hugo' across Guadeloupe (data from Imbert et al. 1996).

Mangrove Stem density (ha· l ) Basal area (m 2 ha- I )


Zone Before After Decrease Before After Decrease
(%) (%)
Coastal fringe 5,334 2.200 59.0 19 6 68.0
Mixed dwarf 2.583 1.900 26.0 II 8 23.0
Tall mixed 1.983 433 78.2 25 7 71.0
Landward zone 2.633 2.550 3.0 24 24 3.0

Imbert et al. (1996) concluded that the extreme susceptibility of Rhizophora


mangle. notably due to its inability to coppice or resprout from broken boles arxl
main branches, was responsible for the catastrophic effect of hurricane 'Hugo' on the
mangroves of Guadeloupe. Such a species effect, which may prevail over other
structural parameters at the stand level in mangroves, induces a relative uniformity
of damage over larger areas. This apparent extreme susceptibility of Rhizophora
raises the question of whether the frequent recurrence of tropical storms will
ultimately favour the dominance of species more resistant to storm effects. At
present, it seems that the high reproductive output and success of Rhizophora
mangle may offset its susceptibility.
Physico-chemical Factors and Mangrove Performance 119

Given these accounts of the damage caused by extreme winds, what features of
mangroves can be recognized to minimize damage from such events, and what
strategies are there to aid recovery from damage?
Windthrow is the severest form of damage and, in Australia, Stocker (1976)
found several mangroves to be particularly susceptible, including Camptostemon
schultzii, Ceriops tagal, Rhizophora stylosa, Bruguiera parvijlora, and Excoecarin
agallocha. Other species such as Xylocarpus moluccensis, Aegiceras comiculatum,
Aegialitis annulata, and Lumnitzera racemosa showed little or no windthrow, and
they rapidly developed new crowns. It seems likely that windthrow-susceptible trees
are those with weakly developed cable root systems, or whose root systems are
weakened by erosion or bank-slumping, or by some biological agency such as
infestation by isopods or wood-boring molluscs. For most species, windthrow
results in death, although for Sonneratia and Avicennia epicormic shoots will
usually develop if some root connection remains.

Fig. 4.3 Mature stands of Sonneratia alba on the cyclone-prone Kimberley coast of
north-western Australia. Despite a high cyclone frequency, these stands reach
maturity due to the physical shelter provided by the deeply indented coast with narrow
bays and inlets with generally steep shorelines.

In the neotropical mangroves, both Roth (1992) and Imbert et al. (1996)
showed that the largest trees, predominantly Rhizophora mangle, appeared to be
most susceptible to windthrow and that shorter communities were less damaged.
However, in a more recent but milder hurricane ('Rosa') which struck the mangroves
of the Pacific coast of Mexico, Rhizophora mangle was the least affected with 65%
of trees found in a well-vegetated condition three years after the storm (Kovacs et al.
2(01). Interestingly, in this region, lAguncularia racemosa was the dominant tree
and it sustained the greatest damage with only 34% of trees remaining in a well-
vegetated condition three years later. Using a logistic regression model, Kovacs et al.
(2001) found that as tree diameter (DBH) increases, the likelihood of lethal damage
increases regardless of species.
120 Mangrove Ecology, Silviculture and Conservation

Crown damage is the most common type of damage, with the plant being
completely defoliated in extreme cases. Leaves of most mangroves are leathery arxl
strengthened by various sclerenchymatous cells, and in strong winds leaf-bearing
twigs appear to be shed rather than individual leaves. Recovery from twig or leaf
damage is usually rapid; A vicennia, Excoecaria and Sonneratia have abundant reserve
buds in the stem. In Rhizophora. buds are present in the stems of saplings but
become restricted to thin terminal branches as the tree matures (Gill and Tomlinson
1969). Conditions severe enough to remove or kill all branches possessing viable
reserve buds will kill species of the Rhizophoraceae.
Susceptibility to bole damage varies considerably among species. The
anomalous wood structure of Avicennia, with its non-concentric, non-annual growth
rings of alternating bands of xylem and phloem (Baker 1915, Gill 1971), gives the
wood unusual qualities: (1) it is extremely strong for its weight; (2) it is extremely
difficult to split radially yet it is easy to do so tangentially (hence it was used to
make shields by Australian Aborigines - Dick 1915); and (3) the unusual ring
structure ensures that, if any part of the trunk is damaged, sufficient intact
conductive tissue remains to supply the crown and epicormic shoots. As a
consequence of this distribution of xylem and phloem tissue, Avicennia cannot be
killed by ringbarking, an apparently useful adaptation in minimizing damage from
waterborne objects.
Only limited studies have been conducted on the secondary wood anatomy of
other mangrove species (Panshin 1932, Marco 1935, Venkatiswarlu and Rao 1964,
van Vliet 1976, 1979, Tomlinson 1986). In Ceriops. thick-walled bast fibres form a
mechanical tissue cylinder giving strength and rigidity to the stem (Rao and Sharma
1968). In Rhizophora. abundant sclereids occur in non-functional phloem tissue
(Karsted and Parameswaran 1976, van Vliet 1976) and stone cells and fibres occur
throughout the plant. The wood of Bruguiera has been described as extremely strong
(Banerji 1958) as has that of Heritiera. Rhizophora apiculata and Lumnitzera littorea
(Panshin 1932).
In the case of a broken bole, a few species are able to regrow from the stumps.
A vicennia. Sonneratia. Xylocarpus. Excoecaria, Lumnitzera. Laguncularia arxl
Conocarpus coppice readily if the bole is broken some distance above ground level.
When the tree is dead but remains standing, a number of causative factors may
be involved, including changes in the substrates, fatal root or bole damage caused by
wind sway, or stress following the near-total loss ofleaves. Steinke and Ward (1989)
reported on the flood effects of cyclones 'Domoina' and 'Imboa' in South Africa.
suggesting that these may be commensurate with wind effects.
Recovery patterns following storm damage are diverse. Reserve buds in many
species and the ability to coppice are an obvious means to quickly recover from
damage. As long as some root connections remain, several species have the ability
to resprout from uprooted trees, producing a straight row of trees of similar diameter
which mayor may not remain attached to the initial (now horizontal) trunk. This
feature has been observed in Laguncularia. Avicennia and Sonneratia. Finally, based
on the extremely high densities of seedlings, particularly of Rhizophora mangle,
Roth (1992, 1997) suggested that in the Caribbean, propagule production is centred
on the period (August to October) when hurricane frequencies are at their highest,
and that the dispersing propagules immediately after storm damage might have
played a selective role.
Physico-chemical Factors and Mangrove Performance 121

Fig. 4.4 Gaps in the mangroves caused by hurricane 'Gilbert' in Celestun, Mexico,
appear initially to be invaded by the fern, Acrostichum aureum.

4.4.3 Hail
Despite the fact that mangoves occur in tropical and subtropical climatic regions,
they are nevertheless subjected to hailstorms, albeit infrequently. As hailstorms
result from cumulonimbus or other convective clouds with strong updrafts, they are
generally accompanied by strong winds and thunderstorms.
The unpredictability of hailstorms means that their effects are rarely
documented. Exceptions are cases where other studies have led to the establishment
of reference sites, e.g. Clarke (1992) documented the effects of hail on predispersal
propagule mortality in Avicennia marina in south-eastern Australia. Similarly,
Houston (1999) was able to document the effects of a severe hailstorm in Port
Curtis, central Queensland, from study sites established to monitor the effects of
reclamation due to port development. All mangrove species in the area showed
evidence of hail damage, including the dominant species, Rhizophora stylosa,
Ceriops tagal and A vicennia marina. The effects of hail included defoliation,
punctured leaves, bruising to bark, divots removed from bark, and branch and plant
death. Differences between species were noted, with Ceriops tagal showing higher
mortality rates than the other two common species. Structural changes noted include
reductions in stem densities, basal area, and canopy cover, and changes in the relative
abundance of species. Houston (1999) reported that recovery was observed in some
stands, but others had not recovered to pre-hail levels of canopy cover two years after
the hailstorm.
Clearly, this form of natural disturbance can exert considerable localized
influence on the structure and composition of mangrove communities.

4.4.4 Lightning
In the tropics and subtropics, thunderstorms are frequent. occurring on average on
20-60 days per year. Lightning strikes occasionally fell trees, although the result is
more often patches of dead standing trees in a 20-30 m diameter circle. For example,
in Papua New Guinea, stand-level dieback has frequently been observed in mangrove
122 Mangrove Ecology, Silviculture and Conservation

forests (Paijmans and Rollet 1977, Arentz 1988) with groups of largely even-aged
trees dying simultaneously due to lightning strikes. Paijmans and Rollet (1977)
found that lightning-created gaps visible in 1957 aerial photographs were no longer
visible in 1972, indicating a maximum closure time of less than 15 years.
In mangroves of the Los Haitizes National Park, Dominican Republic,
Sherman et a1. (2000) found lightning-created gaps to range in size from 300-1,600
m 2, averaging 724 m 2 • They were able to identify between 17-54 gaps km- 2,
comprising approximately 2% of the total forest area. Persistence of canopy gaps
was estimated from a series of aerial photographs spanning 1959-1996. Most gaps
present in one photograph closed during the time interval until the next, leading to
gap closure estimates from 8-16 years.
Sherman et a1. (2000) also found that Rhizophora mangle was favoured by gaps
in that peat mat subsidence in gaps resulted in increased levels of standing water.
They concluded that the present disturbance regime would gradually lead to an
increase in the distribution and abundance of this species, thus bringing about of
marked change in the species composition of the stands.
Smith et a1. (1994) reported on the interaction between lightning strikes arxl
recovery after hurricane 'Andrew' in Florida. During overflights, small circular
patches of living mangroves were observed within large expanses of dead trees.
These patches of surviving saplings represented the colonizers of lightning-created
gaps which had survived the hurricane because of their smaller size. It was concluded
that these sapling-sized survivors may provide an important source for recolonization
of the destroyed forests.

4.5 Sedimentary Conditions and Processes


Mangroves grow in highly variable sediments which they, in turn, modify.
Nevertheless, the substrate characteristics are extremely important in terms of direct
influence on mangrove growth and productivity. It must be noted, however, that the
type of soil and its physico-chemical state are the result of the interactions between
such factors as topography, climate, hydrodynamic processes, tidal range, and long-
term sealevel changes. In this sense, mangrove sediments have a unique history at
anyone site which, on an accurate reading, can provide significant insights.

4.5.1 Mangrove Soil Development


Mangrove communities develop best in sheltered depositional environments where,
in the absence of drastic resculpturing of the coastline, there is a steady accretion of
sediments. This accretion of sediments is facilitated by mangrove vegetation,
especially by those species with a complex matrix of roots such as Avicennia
(Young and Harvey 1996) and Rhizophora (Furukawa and Wolanski 1996), and by
the flocculant effect of salinity which causes fine clay particles, brought down by
rivers, to aggregate and be deposited when reaching the brackish waters of estuaries.
Furukawa and Wolanski (1996) showed that the sediment particles carried in
suspension into tropical mangrove forests during tidal inundation are cohesive,
mainly clay and fine silt, and form large flocs, with a mean particle size of around
100 ~m (Wolanski et a1. 1997. 1998). These are maintained in suspension by the
high turbulence as a result of complex flows around the vegetation until settling
occurs at high slack tide. The settled sediment is not re-entrained at ebb tide because
the high vegetation density inhibits currents capable of eroding the sediments. In
Physico-chemical Factors and Mangrove Performance 123

this regard, mangroves are not merely passive colonizers of mudbanks, but actively
capture mud to continuously replenish their own environments.
Accretion rates measured from mangrove areas (Table 4.5) show that vertical
accretion is variable but commonly approaches 0.5 cm y.l. As a result of this
mangrove-facilitated accretion, there is a gradual elevation of the sediment surface in
relation to sealevel and with it a gradual change in soil water characteristics. Such
gradual and directed changes are often interrupted, however, by periods of erosion
occasioned by storms or flooding which can rapidly reverse the biologically mediated
depositional phase (Bird 1971, 1972, Spenceley 1982). The likelihood of such
disruptive change depends on the geography of the coast and on its
geomorphological history. In active areas such as deltas, where allochthonous
sediment inputs can be large, mangroves can modify the rate at which sedimentary
processes take place but generally do not change the pattern of landform evolution
(Woodroffe 1992).
Two major types of intertidal landforms can be recognized: those which contain
a veneer of transported or trapped sediment over a consolidated parent material, ml
those which are the result of sedimentary accretion, producing prograding shorelines
(Thorn 1982, 1984). The former type may be important regionally, for example,
where sediments accumulate over fossil coral platforms or where comparatively
narrow terrigenous fringes occur along sunken river valleys. The latter type is the
more common and includes many fringing substrates and deltas.
Table 4.5 Accretion rates from mangrove areas using various measurement techniques.

Location Method Accretion Rate Ref.


(cm y.l)
Auckland. NZ Marker horizon 0.17 I
Hauraki Plains. NZ Grids of stakes 1.68·3.84 5
Hauraki Plains. NZ Air·photo analysis 0.64 5
Tonga Radiocarbon dating 0.01 2
S Australia Grids of stakes 0.1-1.6 3
NE Australia Grids of stakes -1.1-0.46 4
NE Australia Grids of stakes 0.5-1.0 14
NE Australia Sediment mass balance 0.1 6
N Australia Radiocarbon. dating 0.6 7
E Malaysia Slabs of perspex 0.64-1.46 8
Bangladesh Radiocarbon dating 0.7 9
Mexico 210Pb and J31Cs isotopes 0.10-0.44 10
Florida 210Pb and 137Cs isotopes 0.14-0.17 10
Florida Marker horizon 0.44-0.72 II
Florida Elevation table 0.25-0.37 11
Florida J37Cs isotopes 0.58-1.33 15
Florida Radiocarbon dating 0.08-0.13 15
Grand Cayman Is. Radiocarbon dating 0.07-0.14 16
Bermuda Radiocarbon dating 0.08-0.11 17
Colombia Radiocarbon dating 0.4 12
E Venezuela Radiocarbon dating 0.13 13
IChapman and Ronaldson 1958; ~lIison 1989; JBird 1971; 4Spenceley 1977. 1982; 5Young and Harvey
1996; 6Furukawa et a\. 1997; 7Woodroffe 1990; KSaad et al. 1999; 9Stanley and Hait 2000; JOLynch et
al. 1989; IlCalJoon and Lynch 1997; 12Jaramillo and Bayona 2000; iJRull et al. 1999; 14Bird and Barson
1977; 15Parkinson et al. 1994; 16Woodroffe 1981; 17ElIison 1993.

Stability of these landforms is strongly influenced by differing


geomorphological origins. Accretion or erosion may be a continuing, seasonal or
periodic process in depositional substrates (Stanley and Hait 2000). Modification of
124 Mangrove Ecology, Silviculture and Conservation

the landforms of more consolidated shores may be intermittent and arise from
catastrophic events such as severe storms.
Thom (1967, 1975) and Thom et al. (1975) studied the ecology of mangroves
in terms of the response of the plants to habitat change induced primarily by
geomorphic processes. Given the climatic-tidal environment and a pool of mangrove
species, each of which possesses a certain physiological response to habitat
conditions, they considered the history of the land surface and contemporary
geomorphic processes jointly to determine the nature of the soil surface on which
mangroves grow (Thom 1982, 1984). Such attributes of the substrate as moisture
content, texture, salinity, redox potential and chemical composition are, to a large
extent, functions of past and present geomorphic processes. The mangroves reflect
each of these geomorphic situations by responding to the environmental gradients of
elevation, drainage, stability, soil characteristics, and nutrient input which each of
these situations produces. According to the physiological response of species to
moisture and/or salinity stress, for example, there will be more or less plant growth
in a particular habitat. Thus, landform properties and geomorphic processes find
expression in the variation in growth, morphology and metabolism of mangroves
along environmental gradients. It should be noted, however, that geomorphological
processes influencing mangrove development in different parts of the world may
differ not only because sedimentary characteristics differ in relation to supply of
sediments, but also because the history of sealevel change may have been different
(Woodroffe 1983, 1992, Fujimoto et al. 1996).
Thus, although mangrove development is bound historically to the geomorphic
processes of a region. it is an expression of the resultant properties of the soils that
occur there. From an ecological viewpoint. a study of the soil relationships of the
mangroves will provide more direct information on mangrove growth performance
than will historical (geomorphological) analysis. This is not to deny the importance
of geomorphological studies, for these place mangrove soil characteristics into a
broader, more causally related context. For example, Spenceley (1983) showed that
there are differences in elemental concentrations between open accreting shores and
estuarine coastlines, and that the temporal and spatial behaviour of the elements also
differ. The nature of the soil in a mangrove community is largely determined by a
range of geological, geomorphological and hydrodynamic processes. Some of these
processes, such as sealevel change, erosion or large inputs of allochthonous
sediments, may affect the mangroves directly. More often, however, they change
certain characteristics of the sediment which, in turn, renders it more (or less)
suitable for mangrove growth and development.
Where freshly deposited sediments have accumulated, the nature of the soil is
primarily determined by pedogenic processes. These processes include the nature of
tidal movement, climate, and hydrology. Additional pedogenic processes commence
with the establishment of vegetation on freshly deposited sediments and gradually
lead to the 'ripening' of the soil.
The mineralogical composition of coastal sediments is largely determined by
the mineralogy of the catchment areas where the sediments originate. In humid
tropical areas, the dominant constituents include kaolinite, iron oxides and quartz.
while in more arid areas, sediments are less kaolinitic, contain less iron oxides, and
are often dominated by illites and minerals of the smectite group. In seasonally arid
areas, montmorillonite may be the most abundant clay mineral, followed by illite.
Physico-chemical Factors and Mangrove Performance 125

The organic matter content of freshly deposited coastal sediments of fine clays
in the tropics is of the order of 1 to 2% on an oven-dry basis. Such sediments
generally contain only small amounts of primary pyrite (i.e pyrite deposited with the
sediments), and secondary and tertiary pyrite formation may occur within mangrove
and other plant debris subsequent to sedimentation (Bush and Sullivan 1999). As a
rule, clayey sediments of the coastal fringe in the tropics tend not to contain free
carbonates unless the catchment areas contain abundant limestones.
Soil 'ripening' starts with the establishment of vegetation on freshly deposited
muds and it can be subdivided into two phases: ripening in the reduced muds, am
ripening which takes place with aeration and oxidation of the sediments.

Fig. 4.5 Pyrite is typically found as framboidal complexes in mangrove soils. The
indivual pyrite crystals within each framboid are uniformly sized, ranging up to 211m.
Scale bar 10 11m. (Photo. L.A. Sullivan)

In the reduced phase, there is accumulation of secondary organic matter, mainly


from roots and leaf litter. Under the prevailing anaerobic conditions, decomposition
and mineralization of this organic material is slow and depends to a large extent on
the rate of the siltation process. In slowly accreting parts of the coastal fringe, time
for accumulation in the superficial layers is long and the organic content is high. In
rapidly accreting areas, the organic content in the superficial layers tends to be
considerably lower. In addition, the amount of undecomposed or partially
decomposed organic material in the soil in a given area will be influenced by the
type of vegetation in that area. For example, soils under Rhizophora have the
highest contents of fibrous organic matter (Hesse 1961a, b, Lacerda et al. 1995).
This may lead to peat formation (Odum et al. 1982, Fujimoto 2000), with estimates
ranging from 11.6-79.5 kg C m· 2 (Fujimoto et at. 1999) or 3-40 cm 100 y.1 (UNEP
1994).
As a result of bacterial respiration, soil oxygen becomes limited. Further
microbial decomposition and mineralization of the organic matter under reducing
126 Mangrove Ecology. Silviculture and Conservation

conditions is facilitated by facultative anaerobes. which do not require oxygen but


which do need an alternative oxidant source (such as the nitrate. ferric or sulfate ion)
for respiration (Boto 1984). As illustrated in fig. 4.6. when all the oxygen is
consumed. then Mn4+ and N03- are converted to Mn 2+ and N2 (gaseous) respectively.
In tum. when all the manganese and nitrate ions are completely exhausted. then Fe3+
is reduced to Fe2+ until the soil eventually reaches a highly anaerobic condition. At
an Eh of -150 mY. Sol is redured to S2- while under extreme reducing conditions
(where the Eh is approximately -230 mY). CO2 in the soil will be reduced to CH4
(marsh gas). All of these redox changes can obviously have profound effects on plant
growth (Boto 1984); however. the redox changes associated with Fe 3+ and SOl are
probably of greatest significance.

500

Disappearance of 02 0x
Disappearance of N03' )00 it
!!
Appearance of Mn 2+ 0-
::s
Appearance of Fe2..
100
"(I
I:l.
~
~
0-
::s
'"
0
;-
Disappearance of 504
2· - 100 :!.
E
,.....
S
Appearance of CH 4 <:
........

-)00

Fig. 4.6 Chemical changes in soils at different levels of oxidation-reduction potential


(Eh, corrected to pH 7). (Redrawn from Youssef 1995).

Under reducing conditions. ferric oxides present in the sediment and sulfates
brought in by tidal waters act as the principal oxidant sources for the anaerobes. As a
result. ferric oxides are reduced to ferric hydroxides and sulfates are reduced to
sulfides. The energy for these processes is supplied by the decomposition of organic
material in the soil. In tum. the ferric hydroxides react with the sulfides to form
ferrous sulfides. predominantly pyrites (Bush and Sullivan 1999. Aragon et al. 1999.
Lin et al. 2000). The degree to which secondary pyrites accumulate in the tidal
sediments is determined by the amounts of organic matter present. the level of ferric
oxides. and the concentration of sulfates brought in by tidal waters. If pyrite levels
exceed -0.1 %. the potential exists for the formation of acid-sulfate soils on
oxidation of the soils during the second phase of soil ripening.
The second phase of soil ripening begins with the aeration of mangrove muds
and this may occur simultaneously with the above processes. especially at the soil
surface. It becomes more significant when tidal flushing diminishes due to further
sediment accretion or to the construction of coastal embankments which reduce tidal
Physico-chemical Factors and Mangrove Performance 127

inundation to the areas behind them. This ripening process can be followed visually
by the change of colour of the soil matrix from neutral or bluish-grey to olive-grey
or greyish-brown with the development of rusty mottling (Lin et al. 2(00). Both
physical and chemical changes are involved in this phase of soil ripening.
The physical aspect of ripening is dominated by the dehydration of soil
colloids. The rate of water extraction from the soil depends on direct evaporation
from the soil surface, on the transpiration rates of the plants, and on internal
drainage of the soil. Transpiration. however, is generally the dominant process.
Dehydration causes shrinkage of the soil and a change in its consistency from semi-
fluid to firm and ultimately to hard. Soil structure develops and, in soils rich in fine
clays and organic particulates (colloids), cracks are formed which accelerate the
internal drainage and aeration of the subsoil. The most common condition in rapidly
accreting mangrove land is that the superficial layers are physically ripe while the
underlying ones are progressively less ripe. Soil shrinkage leads to subsidence which
can cause a deterioration of the hydrology of the land, especially when it is uneven,
causing areas of standing water.
Chemical ripening of the soil commences with the oxidation of pyrites. Where
the organic content is high, the oxidation of this material enhances subsidence ani
the formation of cracks, and assists in the oxidation of iron sulfides or pyrites (Lin
et al. 2(00). Such oxidation of pyrites is complicated, involving a number of
intermediate steps, but it may result in the formation of jarosite, KFe3(S04MOH)6'
which can be hydrolyzed to form sulfuric acid in the absence of substantial quantities
of free carbonates (Lin et al. 2(00). In soils with an excess of alkalinity (mainly
CaC03) over oxidizable sulfur compounds, the sulfuric acid formed is inactivated at a
near neutral pH level and CaS04 (gypsum) is formed. Under these conditions, a non-
acid soil is formed.
When pyrites are present in excess, acidification occurs after the alkalinity is
exhausted. Such acidification leads to an increase in H+, the liberation of A1 3+, and a
decrease in the adsorbed bases on the soil colloids, and results in soils with low pH,
low in exchangeable bases, and high in exchangeable and free aluminium. Termed
actual acid-sulfate soils, they are generally characterized by the presence of straw-
yellow mottles of jarosite; in peaty soils, however, jarosite rarely occurs.
A complication occurs in the ripening process when ripened, non-acid soils are
again inundated or saturated with seawater. In such circumstances, calcium saturated
clays become saturated with Mg and Na, which, in turn, may lead to severe
structural degradation of the soil, and possibly a lowering of the surface. As a result
of lowering of the soil surface, ponding of seawater may occur, and in seasonally
arid climates, a build-up in soil salinities to several times that of seawater may
result. Such high salt levels, in combination with a poorly developed (or degraded)
soil structure, may make these soils unsuitable for mangrove species and result in
unvegetated salt flats.
Understanding mangrove-soil relationships is further complicated by the ability
of most mangrove species to grow on a variety of substrates, and to often alter the
substrate through (1) peat formation, (2) modifying the pattern of sedimentation, ani
(3) altering the microbial decomposition of organic matter in the sediment (Locerda
et al. 1995). Mangrove trees are found on a wide variety of substrates including
muds, silts, peat, sand, and even rock and coral shingle, provided there are sufficient
crevices for root attachment. Extensive mangrove communities appear to be best
developed, however, only on muds and fine-grained sand (Butler et al. 1977a,
128 Mangrove Ecology, Silviculture and Conservation

Galloway 1982); these muds are often highly saline and gypseous, with soft loose
surfaces showing neither seasonal cracking nor change in texture with depth. These
physical characteristics are important in terms of the drainage and aeration of the
soils.

4.5.2 Aeration and Drainage Properties of Mangrove Soils


Because mangrove soils are mostly derived from alluvial materials deposited in a
sedimentary environment, very fine sand or silt particles predominate. When
combined with regular tidal inundation, drainage and aeration are generally limited,
and the soils are generally waterlogged and highly reduced, with Eh ranging from
+100 to -250 mY.
Soil aeration is important in mangrove environments in supplying oxygen for
root respiration. Aeration is directly related to soil drainage and is therefore highly
variable. It depends upon elevation, steepness of the topography, and the physical
characteristics of the substrate, particularly texture and structure. Both aeration and
drainage are relatively good in coarsely textured sediments. However, clays can
accumulate in sandy sediments where these are frequently inundated and even a low
concentration of clay (Le. -15%) can greatly reduce aeration in coarsely textured
sediments. Such clay accumulation results from the flocculant action of seawater on
the one hand, and the physical reworking of sediments by tides and waves on the
other. In consequence, drainage in the lower to middle regions of the intertidal zone
may be reduced, except in areas of turbulence owing to wave action or where there
are high current velocities.
Clays and sands are not the only soil types on which mangroves will grow; in
terms of aeration and drainage, however, sandy and clayey soils usually can be
considered as the two extremes in which mangroves will grow. The properties of
these two kinds of soils are compared in Table 4.6.
Table 4.6 Comparison of various characteristics in unconsolidated sand and clay sedimenL~.

Parameter Sand Clay


Infiltration High Low
Porosity Low High
% 35 45
Permeability High Low
Seepage velocity High Low
Water-holding capacity Low High
% 20-35 35-50
Salt retention Low High
Leaching Rapid Slow
Capillarity Low High

In generally finely textured sediments, however, subsurface horizons of greater


permeability, or structural features such as animal burrows, can sometimes provide
good drainage and replenishment of oxygen through subsurface drainage (Ridd 1996,
Ridd and Sam 1996). On the other hand, finely textured sedimentary soils can
prevent drainage from higher elevations (Aucan and Ridd 2000). Drainage problems
at such sites may last for extended periods and impose stresses on the plants growing
there, including significant increases in soil soluble sulfide and ammonium levels.
Low rates of water infiltration into many mangrove soils can compound the soil
conditions described above. Clarke and Hannon (1967) measured the infiltration rates
into mangrove and saltmarsh soils in Sydney and, with the exception of rates
Physico-chemical Factors and Mangrove Performance 129

measured near crab holes, the infiltration rates were low despite the sandy texture of
these soils.
At Hinchinbrook Island, however, Wolanski and Gardiner (1981) observed that
rainwater percolated directly into the mud with no surface run-off; in other words, the
Hinchinbrook muds have extremely high infiltration rates. In this system, structural
features such as crab holes were the dominant water transport routes. The
effectiveness of crabs holes in facilitating groundwater flows was demonstrated by
Ridd (1996) using dye injections. He estimated that in a I km 2 area of north-eastem
Queensland mangroves, 1<Y-105 m 3 groundwater move through crab holes in each
tidal cycle.
Occasionally, mangroves occur in weathered limestone landscapes as, for
example, in Yucatan, Mexico and in Sulawesi, Indonesia. Because of the high
porosity of the underlying limestone, high infiltration rates and high groundwater
flows are common, and the groundwater outflows (fig. 7.9) are sometimes
spectacular. While such edaphic conditions appear highly favourable for mangroves,
the strong tendency for CaC03 to reduce the availability of essential micronutrients
such as iron, copper and zinc, partially offsets any advantage.
When tidal inundation is superimposed on the drainage and aeration properties
of mangrove soils, they are characterized by a high water content (Hesse 1961 b,
Clarke and Hannon 1967, Naidoo 1980), low oxygen content, often high levels of
salinity, and high levels of soluble sulfides. In addition, these soils are often semi-
fluid and poorly consolidated.
Although the root adaptations (described in 3.4.1) facilitate survival in the
generally adverse soil environment of mangroves, species vary in their tolerance to
soil conditions and in the extent to which they directly influence soil microbial
processes. Thus, Avicennia, Rhizophora and Bruguiera grow on soils with different
characteristics in sulfide content, acidity, organic matter content, redox potential,
nutrients, cation exchange capacity, exchangeable bases and acidity, and clay content
(Hesse 1961 b, Naidoo 1980, Spenceley 1983, Nickerson and Thibodaeu 1985,
Thibodeau and Nickerson 1986, McKee 1993, Lacerda et al. 1995), and consistent
patterns are not always apparent. For example, Nickerson and Thibodeau (1985) and
Thibodeau and Nickerson (1986) have suggested from observations that differential
distribution of Avicennia germinans and Rhizophora mangle may be due to the
latter's preference for areas of low soluble sulfide concentrations. Other experimental
(McKee 1993) and field studies (Carlson et al. 1981), however, have reported the
opposite results.
Experimental evidence is scanty, but it does appear from field observations of
distribution and growth that mangroves differ in their sensitivity to poorly drained
and poorly aerated or anaerobic soils. Field data on the soil water content in which
various species grew were collected from Proserpine, north-eastern Queensland, from
four sites over one-and-a-half years (Hutchings and Saenger 1987). Three groups,
based on the water content of the soil on which they grow, were recognized.
Osbomia octodonta and Bruguiera parvijlora grow on soils that have low water
contents, either because of good drainage (B. parviflora) or because of their location
on the landward margins of the mangroves where tidal inundation is infrequent (0.
octodonta). The second group contains those species growing on soils with an
intermediate water content and includes Bruguiera gymnorhiza, B. exaristata,
Clerodendron inerme and the mangrove fern Acrostichum. The third group,
containing eleven species, grows in soils that have high water contents, either
130 Mangrove Ecology, Silviculture and Conservation

because of frequent tidal inundation (such as R. stylosa and A. marina), or because


high freshwater run-on occurs (such as Heritiera littoralis and Cynometra iripa).
Although this grouping is tentative, it provides some indication of the soil water
regime under which the various species grow.

Fig. 4.7 Well-developed stands of Rhizophora stylosa on waterlogged and


unconsolidated sediments at Princess Charlotte Bay. north Queensland.

The height of A vicennia marina appears to depend on drainage properties of the


soil, with the tallest trees growing on well-drained banks close to streams (Chapman
and Ronaldson 1958, Macnae 1966). Macnae (1966) maintained that Ceriops tagal in
Australia is found only on well-drained soils, and he suggested that its virtual
absence from areas of high rainfall may be as much due to drainage irregularities as
to rainfall. Measurements of soil water content within Ceriops tagal stands along the
Queensland coastline (Saenger and Robson 1977) do not support his suggestion, but
it is possible that considerable geographic variation occurs.
From field observations, Macnae (1966) cites two examples of contrasting
responses to soil drainage: first, Rhizophora stylosa grows on well-drained soils in
Malaysia (Ding Hou 1958), whereas in Australia it grows on a range of substrates,
but with the tallest trees occurring on soft, waterlogged muds (fig. 4.7); and second,
in Australia, Lumnitzera racemosa is recorded from well-drained sandy soils on the
landward fringe while in southern Africa, Macnae (1966:96) records it 'as a true
mangrove extending down to almost high water neaps'.
Apart from their draingage and aeration characteristics, there are other
difficulties posed by mangrove soils. As these soils are generally also high in sulfur
content and as some of this may be bound with iron to form pyrites, the pH of the
soils tends to be between 6-7. Such pH levels will influence nutrient availability on
the one hand, while biogenic sulfides, particularly pyrite and hydrated iron-
monosulfides, may provide a geochemical trap for heavy metals (Harbison 1986,
Saenger et al. 1991). However, the oxidation of the sulfides when the sediment is
dried, can lead to the remobilization of metallic species with the development of
actual acid sulfate soil conditions (Clark et al. 1997).
Physico-chemical Factors and Mangrove Perfonnance 131

4.5.3 Nutrients in Mangrove Soils


Mangroves require a range of inorganic salts, many in small quantities, but others,
such as the nutrients N and P, in larger amounts. The mean concentrations of total
Nand P in a range of mangrove leaves are shown in Table 4.7. These data indicate
that considerable quantities are required of both N and P, with approximately a
tenfold greater requirement for N than for P. The annual minimal requirements for
mangroves in north-eastem Australia have been estimated to be around 200-300 kg
N ha'! y-! and 20-30 kg P ha-! y-! (Clough et at. 1983, Boto 1991, Robertson and
Phillips 1995). What are the sources of nutrients to allow mangroves to remain
productive?
Table 4.7 Nutrient concentrations (as total nitrogen and phosphorus) in mangrove leaves (% of dry
weight) for selected species and localities.

Species C TN TP Location Ref.


A. aureum 0.89-1.48 0.11-0.14 Mexico 6
1.3 0.18 Sri Lanka 12
2.24 0.18 Thailand 15
A. germinans 50.9 l.l 0.2 Guiana I
0.03-0.04 Guadeloupe 2
2.21-3.10 0.10-0.14 Guadeloupe 14
2.0-2.1 Florida 10
A. marina 45.4-50.8 0.32-1.25 Australia 3
2.11 0.12 Australia 4
43.9-48.0 1.69-2.57 0.28-0.31 Australia II
A. of/icinalis 44.3 0.81 0.06 India 5
A. schaueriana 1.9-2.3 Brazil 9
Bruguiera spp. 1.17 0.07 Thailand 15
C tagal 44.7 0.63 Kenya 8
La. racemosa 42.9 0.8 0.2 Guiana I
0.48-0.57 0.06-0.10 Mexico 6
0.03 Guadeloupe 2
1.0 Florida 10
2.3-2.5 Brazil 9
R. apiculata 48.0 1.64 0.02 Malaysia 7
45.7 0.49 0.04 Australia 3
44.0 0.68 0.06 India 5
1.95 0.09 Thailand 15
R. mangle 47.2 0.6 0.1 Guiana I
0.55-2.07 0.06-0.09 Mexico 6
0.01-0.02 Guadeloupe 2
0.51 0.01 Belize 16
1.3-1.5 Florida 10
4.5 Brazil 9
46.3 0.40 0.15 Venzuela 13
R. mucronata 46.0 0.65 Kenya 8
45.2 0.62 0.12 India 5
0.77 0.08 Thailand 15
R. stylosa 45.7 0.34 0.05 Australia 3
X. granatum 1.59 0.11 Thailand 15
X. moluccensis 1.08 0.05 Thailand 15
IBetoulle et al. 2001; 2Imbert and Portecop 1986; 3Bunt 1982; 4CIough and Attiwill 1975; 5Wafar et al.
1997; 6Medinaetal. 1995; 7Gong and Ong. 1990; 8Slim et al. 1996; 9Lacerda et al. I 986b; iOMcKee
and Faulkner 2000; IIDay et al. 1999; 12Balasubramaniam et al. 1992; BJayasekera 1991; 14Saur et al.
1999; 15 Aksomkoae and Khemnark 1984; 16Feller et al. 1999.

With the exception of ammonia, soluble fonns of inorganic Nand P are low to
undetectable in most mangrove sediments (Boto 1991, Alongi 1998) and total soil N
and P (Table 4.8) are orders of magnitude greater, indicating that the vast majority of
132 Mangrove Ecology, Silviculture and Conservation

soil N and P is contained in organic forms. What little inorganic P is present is


bound within hydrated iron and aluminium oxyhydroxides, severely limiting its
availability to plants;
In a pioneering study, the response of an entire mangrove community to
nutrient input and cycling was studied by means of a mathematical model (Lugo et
aI. 1976) based on a mangrove forest in southern Florida. In this model, the
incoming radiation interacts with nutrients and mangrove plants to form organic
matter through the process of photosynthesis. Some of this gross production is
respired by the forest, some is stored as a net increase in forest biomass, and some is
deposited on the forest floor as detritus. The detritus may be exported from the forest
to the estuary by tidal action. Some of it is grazed in situ by mangrove consumers,
and some decomposes, or accumulates as peat. Decomposition may occur under the
influence of oxygen-saturated waters of incoming tides, or of atmospheric oxygen
when the forest floor is exposed to the air. Decomposition of detritus within the
mangrove system represents a source of nutrients for photosynthesis. Other nutrient
sources are from terrestrial drainage, tidal waters, rainfall, and sediment storage. Of
these, terrestrial drainage is the most significant. Some nutrients are not used and are
lost from the system; the rest are taken up by the plants, thus completing the loop
in the model. This model was validated using field data for the various forcing
functions, state variables and flows within the system, including nutrient input,
nutrient uptake and nutrients exported in the detritus.
The model was highly sensitive to nutrient availability, and when nutrient
input to the system was set at zero, mangrove biomass decreased steadily. The
decrease in mangrove biomass, gross photosynthesis and nutrient storage with the
three levels of nutrient input (high, medium and low) indicated the dependency of
mangrove communities on nutrient inputs derived from the land. Decomposition
within the forest and inputs from seawater did not seem to be enough to maintain
the observed rates of metabolism because of losses through detrital export.
Table 4.8 Total nitrogen and phosphorus concentrations from selected mangrove soils.

Locality TN TP P20 5 Ref.


(% D Wt.) (% D Wt.) (ppm)
Brazil 0.16-0.35 5
Colombia 0.37-1.25 0.09-0.15 I
Guiana 0.06-0.08 2
Florida Bay, USA 0.20-0.40 II
Senegal 0.06-0.45 9
Sierra Leone 0.08-0.16 10
Sri Lanka 0.14-0.26 4
Indonesia 0.22-0.24 1.10-1.75 3
0.35-0.47 8
Vietnam 0.14-0.27 0.06-0.11 6
Thailand 0.10-0.91 7
0.08-0.10 13
lriomote Is., Japan 0.05-0.09 0.02-0.04 14
NE Australia 0.10-0.40 0.02-0.05 12

'Cardona and Botero 1998; 2Pabre et aI. 1999; JSukardjo and Yamada 1992; 4Balasubramaniam et
aI. 1992; lLacerda et aI. 1995; 6Hong 2000; 7Chan¥.prai 1984; 8Sukardjo et al. 1984; 9Debenay et
al. 1997; 'DJfesse 1961, 1963; "Rosenfeld 1979; , Boto and Wellington 1984; 'JKristensen et al.
1988; '4Kuraishi et al. 1985a.

Lugo et al. (1976) found that the contribution of nutrients from land drainage
was ten- and twenty-times that from seawater during the dry and wet seasons
Physico-chemical Factors and Mangrove Performance 133

respectively. They concluded that gross photosynthesis appears to be sensitive to


terrestrial nutrient input and that the development of mangrove biomass is dependent
on the quantity of nutrients and the efficiency of nutrient intake. In addition, they
concluded that during succession, mangroves exert significant control over the
amount of nutrients in adjacent water but, if terrestrial nutrient input is reduced, they
do not have the capacity to maintain themselves at the same level of production.
This is due to a loss of nutrients through export from the mangrove community,
which suggests that there must be selective pressure for mechanisms of recycling.
This modelling was based on mangroves from a carbonate setting which tends
to retain phosphates (Silva and Mozeto 1997, Koch and Snedaker 1997). So, how
applicable is this model to other situations? Are there sources other than terrestrial
input of these essential nutrients available to mangroves in other parts of the world?
Another source of nutrient input, at least for nitrogen, is from nitrogen-fixing
bacteria and cyanobacteria associated with mangrove mud and with above-ground root
systems (Zuberer and Silver 1975, 1978, Kimball and Teas 1975, Potts 1979, van
der Valk and Attiwill 1984, Sheridan 1991, 1992, Pelegri and Twilley 1998, Saur et
al. 1998). In Caribbean mangroves, nitrogen-fixation may be a critical factor in
mangrove establishment during early stages of stand development when trapped
detritus is minimal and nitrogen is limiting. On the other hand, in Guadeloupe,
French West Indies, Sheridan (1991) found extremely high rates of nitrogen fixation
by epicaulous cyanobacteria (principally Nostoc punctifonna and Stigonema sp.)
growing on mature trunks of Rhizophora mangle and Avicennia germinans, ranging
from 174-2,012 nmol C 2H 4 mg Chla· 1 h-I. These rates were calculated to yield
approximately 42.3 g N m· 2 y-I. While such nitrogen yields are potentially
substantial, these microepiphytes were constrained in their nitrogen fixation by their
consistently high demand for humidity and their sensitivity to salinity; in these
mature forests, nitrogen fixation was inhibited at moisture contents below 75% and
salinities above 1,500 ppm, suggesting that lower and pulsed yields of nitrogen
would be most common in low salinity mangroves.
In other parts of the world, however, the importance of nitrogen fixation to
mangrove communities is more variable; preliminary studies of northern Queensland
mangroves have not revealed any potential areas of high nitrogen fixing activity
(Boto 1982, Boto and Robertson 1990, Alongi et al. 1992). Van der Valk and
Attiwill (1984) found low rates of nitrogen fixation associated with the root zone
sediments of Avicennia marina in Westernport Bay, Victoria, concluding that the
processes of ammonification and nitrification may be more important for
maintaining nitrogen levels in Australian mangrove soils. Similarly in South
Africa, Mann and Steinke (1989) found low rates of nitrogen fixation by blue-green
algae (cyanobacteria) associated with Avicennia marina pneumatophores except under
optimal conditions of temperature. Moreover, inorganic nitrogen concentrations at 1-
5 mg L- I significantly depressed nitrogen fixation in all habitats. As these
concentration levels are commonly found in estuarine mangrove systems, it seems
unlikely that nitrogen fixation is crucial in early stages of mangrove stand
development in these areas. In India, nitrogen fixation in root associations of various
developmental stages of tropical mangroves in the Ganges river estuary was found to
be patchy, with high rates of nitrogenase activity (64-130 nmol C 2H4 g-I dry root h-
I) in some pioneer mangrove stands and low in mature stands (Sengupta and
Chaudhuri 1991).
134 Mangrove Ecology, Silviculture and Conservation

What other sources of nutrients are there? Work on Hawaiian mangroves by


Walsh (1967) demonstrated that there is considerable removal of nitrate aOO
phosphate from the tidal water entering the mangrove community. Lugo et al.
(1976) and Davis et al. (2001) confIrmed these observations for mangrove
communities in Florida. However, while removal of nutrients from incoming tidal
waters may be substantial, concentrations of nitrate and phosphate in tidal waters is
generally low and it does not constitute an ample supply.
Perhaps the colonies of seabirds and bats that roost in mangroves constitute a
more signifIcant, if localized, supply. Onuf et al. (1977) studied the effects of
nutrient enrichment of Rhizophora mangle islands in Florida by nesting pelicans aOO
egrets. Two islands were compared: one with a breeding colony of pelicans aOO
egrets (high nutrient), and one without (low-nutrient). Their data indicate higher
growth rates, greater additions of leaves, reproductive parts and new lateral branches,
and larger increases to existing stems at the high-nutrient site. Growth in the
fertilized stand began earlier in the year and had a second peak of growth not shared
by mangroves at the low-nutrient site. These responses to nutrient enrichment in the
fIeld appear to verify the more theoretical predictions regarding the effect of nutrients
on a mangrove community made by Lugo et al. (1976).
Experimental fertilization, conducted over a year in north-eastern Australia,
showed a signifIcant growth response to soil ammonium enrichment, suggesting
that nitrogen limitation was common to all sites (Boto 1983, Boto and Wellington
1983). On the other hand, phosphate limitation was indicated only in those areas
where phosphates were less than 5 JIg P g.l dry soil. Preliminary data from six Cape
York river systems indicate that all sites had low to extremely low phosphate
concentrations (0.1-5 JIg P g.l dry soil), whereas ammonium levels were similar to
those at Hinchinbrook Island (Boto 1983). Nevertheless, in more detailed studies at
Hinchinbrook Island, Boto (1983) reported that mangrove standing crop biomass was
signifIcantly correlated with (1) extractable phosphorus, (2) soil salinity (negative
correlation) and (3) the redox potential or reducing conditions of the soil. Mean
ranges of phosphate and ammonium concentrations were 5-20 JIg P g-l dry soil aOO
4.6-7.3 JIg N g-l dry soil. These fIndings strongly suggest that phosphorus status is
an important influence on mangrove primary productivity, at least in northern
Australia. Similar experimental fertilization trials in the Caribbean involving
mature, dwarfed Rhizophora mangle stands showed that primary production in these
was also severely P-Iimited (Feller 1995, 1996, Feller et al. 1999).
In general, however, it seems that the bulk of the nutrients come from land
drainage, but that this is augmented by some seawater sources, nitrogen fIxation, aOO
localized inputs from bird or bat roosts. In some instances, ground water may also
be a source of nutrients (Kitheka et al. 1999). More importantly, it seems clear that
mangroves, and particularly mature systems, have very effective mechanisms to
retain and recycle nutrients (Boto 1982, Twilley et al. 1986, Alongi et al. 1992,
Feller et al. 1999). From fertilization experiments, it also seems clear that as the
availability of a limiting nutrient increases, the recycling mechanisms become less
efficient (Feller et al. 1999).
These mechanisms in mangroves commence with the resorption of N and P
prior to leaf fall. While resorption efficiency is inversely linked with nutrient
availability, resorption rates in mangroves range between 30-70% for P and 20-60%
for N (Twilley et al. 1986, Alongi et al 1992, Slim et al. 1996, Wang and Lin
1999, Feller et al. 1999, McKee and Faulkner 2000).
Physico-chemical Factors and Mangrove Perfonnance 135

According to Alongi et al. (1992), other processes leading to this tight cycling
of nutrients in mature mangroves include: (1) burial and processing of detritus by
crabs; (2) efficient nutrient uptake; (3) tidal fluctuations and topography; (4)
sediment types; and (5) climatic disturbances.
Leaf processing in mangroves by crabs enhances rates of nutrient cycling as
crabs break down litter at >75 times the rate of microbial decay alone, thereby
facilitating high rates of productivity by sedimentary bacteria (Robertson and Daniel
1989a). These sedimentary bacteria may also contribute N via nitrogen fixation.
Efficient nutrient uptake by bacteria, algae and mangroves ensures low
dissolved nutrient concentrations in mangrove creeks and in porewater (Alongi et al.
1992). This removal of dissolved nutrients is further enhanced by changes in soil pH
and Eh as a result of tidal fluctuations (Boto 1984, Alongi 1998). Thus, with
impeded drainage and little or no internal water movement, the interstitial water
bathing the roots could quickly become exhausted of major plant nutrients. Low
infiltration and the generally mediocre cation exchange capacity of the soils mean
that little replacement of such nutrients occurs around the roots. Furthennore, in the
immediate vicinity of the roots, the development of anaerobic conditions can lead to
pH changes which, in turn, can change the availability of nutrients. For example,
phosphorus becomes unavailable at low pH values.
Soils in infrequently flooded areas near the landward limits of mangrove
vegetation undergo periodic changes in redox potential (Boto 1982, Clark et al.
1998, Shennan et al. 1998), changing from saturated, anaerobic conditions to a
partially aerobic state. Under such aerobic conditions, phosphorus is readily adsorbed
onto clay and complexes with calcium, magnesium and iron, limiting its
availability (Silva and Mozeto 1997, Silva et al. 1998). In contrast, the availability
of phosphorus is much higher in anaerobic sediments where bacteria and H 2S reduce
ferric iron to ferrous iron. This process increases the availability of dissolved P in
soil water because ferrous iron is much less effective at adsorbing P than ferric iron.
In addition, the presence of abundant organic matter in anaerobic soils may
contribute to an accumulation of dissolved P by inhibiting its adsorption onto clay
surfaces. In aerobic soils, the ratio of adsorbed to dissolved P is 1,000: I whereas in
reduced soils, this ratio is around 5: I. These differences between aerobic and
anaerobic sediments in the availability of P may explain why mangrove growth was
enhanced by the addition of phosphate at an infrequently flooded site whereas
mangroves failed to respond to addition of phosphate at a frequently flooded site
(Boto and Wellington 1983, 1984).
The availability of N is also affected by flooding regimes. At high tidal levels,
concentrations of nitrate and ammonia are very low as they are readily leached from
oxidized mineral soils. At low tidal elevations, ammonia, which is the major fonn
of combined inorganic nitrogen under reducing conditions, can reach relatively high
concentrations because it is adsorbed onto organic paricles. However, anaerobic
conditions inhibit the uptake of nitrogen in wetland plants. This may explain why,
despite relatively high concentrations of ammonia, mangroves responded to the
addition of nitrogen when growing in waterlogged and anaerobic soils at frequently
flooded sites (Boto and Wellington 1983, 1984).
For the two major plant nutrients, nitrogen and phosphorus, microbiological
processes are the main detenninants of their release in a form available for plant
growth. In the mangrove environment, nitrogen becomes available through
microbial fixation of atmospheric N2 and through the biological decomposition of
136 Mangrove Ecology, Silviculture and Conservation

organic matter in the soil. Nitrogen bound up in proteins is converted to ammonia


by numerous proteolytic bacteria and fungi; this process is termed 'ammonification'.
Ammonia can serve directly as a source of nitrogen but, more importantly, it
provides energy for nitrite bacteria which, in the presence of oxygen, oxidize the
ammonia to nitrite. As a rule, the nitrite is further oxidized to nitrate by another
group of nitrifying bacteria; the whole process is termed 'nitrification'. Losses of N
through 'nitrification' from mangroves are usually small and generally only increase
in the presence of excess N due to pollution (Rivera-Monroy et al. 1995, Rivera-
Monroy and Twilley 1996, Corredor et al. 1999).
With the seasonal discharge of fresh water, a short-term increase in phosphate
occurs in estuaries (Pailles et al. 1993). This phosphate does not remain in the water
column but becomes adsorbed onto aerobic sediments and suspended particulate
matter containing iron oxyhydroxides as insoluble ferric phosphate (FeOOHP04).
Under reducing conditions resulting from an oxygen deficiency in the overlying
water or through bacterial activity, ferric phosphate is reduced to ferrous phosphate
and may subsequently be released from the sediments to the porewater. Thus, the
sediments from mangrove areas are able to remove dissolved phosphates from the
overlying water and bind them. In the reduced mangrove sediments, they would then
be available for uptake by mangrove rootlets, and the generally clayey nature of the
soils would prevent leaching and loss of these soluble phosphates to the overlying
waters. As phosphate retention is more efficient in fine sediments than in coarse
ones, fine sediments generally possess a greater store of available phosphates; this
may partly explain the better development of mangrove communities on fine
sediments such as alluvial muds in deltaic environments. Thus, porewater P
concentrations can be strongly regulated by Fe chemistry and the redox state of the
sediments which, in tum, are largely determined by the activities of sulfate-reducing
bacteria (Sherman et al. 1998).
Nutrient recycling is also influenced by litter quality which, in tum, affects
rates of microbial decomposition and macrofaunal consumption. Low N
concentrations in senescent leaf (as in Rhizophora when compared with Avicennia or
Laguncularia) reduce the rate of microbial decomposition and macrofaunaI
consumption (Sherman et al. 1998). In their study of mangroves in Guiana, Betoulle
et al. (2001) found that litter fall varied from 8.8 and 8.7 t ha-' y-' at the pioneer and
senescent stages respectively to 12.5 and 12.6 t ha- I y-' for young and mature stages
where development conditions are optimal. Nitrogen and carbon inputs were
estimated at 1.3 x 10-2 and 6.4 t ha-' y-' respectively. Litter appeared rich in
phosphorus, corresponding with high concentrations in the sediments of around 600-
800 Ilg g-' (Fabre et al. 1998).
In their comparison of Rhizophora mangle and Avicennia schaueriana soils in
Brazil, Lacerda et al. (1995) found that the organic carbon and nitrogen contents were
consistently higher in Avicennia than in Rhizophora soils. The contribution of
sugars and amino acids to the total organic carbon pool was constant with depth in
Rhizophora soils whereas in the Avicennia soils it decreased with depth. These
findings indicated that under Rhizophora stands there is an accumulation of organic
matter whereas Avicennia soils undergo continuous breakdown of organic matter.
This reduction in microbial organic matter breakdown under Rhizophora appeared to
be linked to the generally higher acidity produced by Rhizophora litter and to its
more refractory nature. As a result, recycling of N was much more rapid in
Avicennia soils (Lacerda et al. 1995).
Physico-chemical Factors and Mangrove Performance 137

Clearly, nutrient retention and recycling is a significant process affecting


nutrient availability, which detennines whether a mangrove system functions as a
nutrient sink or source to adjacent systems.
Table 4.9 Cation content of mangrove leaves (% of dry weight) for sclected species and localities.

Species Ca2+ Mg2+ K+ Na+ Location Ref.


A. aureum 0.27 0.28 1.5 0.49 Sri lanka II
0.14-0.17 0.19-0.22 2.98-3.47 0.11-0.12 Mexico 7
0.06 0.18 1.06 1.60 Thailand 13
A. germinans 0.5 0.9 0.5 1.6 Guiana 1
1.24 0.36 2.10 0.46 Brazil 2
0.3-0.7 1.0-1.3 0.5-1.2 1.8-2.6 Guadeloupe 3
A. marinLl 0.38 0.94 1.09 2.82 Australia 4
A. schaueriana 0.13 0.76 1.58 2.29 Brazil 2
0.3-0.8 1.2-1.4 0.3-1.2 2.6-4.0 Brazil 5
2.8-4.0 Brazil 10
B. gymnorhiza 1.77 0.53 0.34 1.98 Indonesia 6
H. tiliaceus 1.12 0.40 1.53 0.34 Indonesia 6
La. racemosa 1.7 0.7 0.5 1.5 Guiana 1
0.81-1.98 0.22-0.35 0.47-0.56 0.53-1.07 Mexico 7
2.7 0.4 0.4 1.8 Guadeloupe 3
0.24 0.71 1.35 1.20 Brazil 2
1.6-1.8 Brazil 10
0.8-1.6 0.4-0.5 0.4-0.6 1.8-2.1 Brazil 5
Lu. racemosa 1.46 0.14 0.78 5.40 Thailand 13
R. apiculata 0.44 0.77 0.52 1.29 Malaysia 8
2.10 0.44 0.84 1.29 Indonesia 6
0.61 0.28 1.14 4.00 Thailand 13
R.mangie 0.70-1.29 0.23-0.53 0.81-1.36 0.58-0.87 Mexico 7
1.22 0.47 0.84 0.98 Panama 9
1.3-1.9 0.6-0.8 0.3-0.4 1.4-1.7 Guadeloupe 3
0.34 0.85 1.38 1.45 Brazil 2
0.9-1.3 0.6-0.7 0.4-0.8 1.9-2.3 Brazil 5
1.38 0.41 1.78 1.96 Venezuela 12
R. mucronata 0.42 0.13 0.53 4.40 Thailand 13
S.alba 1.02 0.45 2.09 1.44 Indonesia 6
x. granatum 3.57 0.26 0.69 1.56 Indonesia 6
0.86 0.15 1.11 5.90 Thailand 13
X. moluccensis 0.74 0.32 1.06 3.10 Thailand 13
IBetoulle et at. 2001; 2Lacerda et at. 1986a; 11mbert and Portec~ 1986; 4Clough and Attiwill 1975;
5Lacerda et al. 1985; 6Komiyama et a1. 1988; 7Medina et a1. 1995; ong and Ong. 1990; 9Golley et al.
1975; l'Lacerda et at. 1986b; IIBalasubramaniam et at. 1992; 11Jayasekera 1991; 11Aksomkoae and
Khemnark 1984.

Other nutrients are often divided into two classes: those of which the
concentrations vary according to the environment (Na+ and K+), and which are
involved with osmotic regulation, and those (Ca2+ and Mi+) which show a certain
stability and participate in cell formation (Ball 1988a). The leaf concentrations of
these important cations in various mangroves (Table 4.9) clearly indicate that the
order of cation concentration in mangrove leaves generally is Na~Mg2~K+=Ca2+.
One exception is Acrostichum aureum which appears to have an exceptional
selective ability for K+ uptake (Balasubramaniam et al. 1992, Medina et al. 1995).

4.6 Salinity of the Soil Water


Salinity of the interstitial soil water has long been recognized as an important factor
regulating growth, height, survival, distribution and zonation of mangroves
(Bowman 1917, Cintr6n et al. 1978, Semeniuk 1983, Ball and Pidsley 1995,
Saintilan 1997a, b). The physiological importance of salinity has been investigated
138 Mangrove Ecology, Silviculture and Conservation

using culture experiments. For example, in culture, Avicennia and Aegiceras showed
maximal growth at 25% seawater (Burchett et al. 1984, 1989, Ball 1988b). By way
of contrast, in the field the response to salinity is more variable, and mangroves
have been found at salinity levels that exceed those suggested by laboratory
experimentation (Table 4.10). For example, Macnae (1968) showed that Avicennia
marina and Lumnitzera racerrwsa can tolerate salinities of up to 90%0 in the soil.
Rhizophora mangle is probably limited to soil salinities below 65%0 (Cintron et al.
1978, Teas 1979), Avicennia genninans becomes dwarfed and gnarled in Florida
when soil salinities approach 60-80%0, and Laguncularia racel1wsa occurs at
maximal salinities of around 80%0 (Jimenez 1984); on the other hand, culture
experiments indicate salinity limits for Rhizophora mangle, Avicennia genninans
and Laguncularia racerrwsa to be 130%0, 100%0 and 80%0 respectively (McKee
1995b).
Table 4.10 Maximal soil salinities at which field-grown mangroves have been observed.

Species Salinity (%0) Location Reference


Acanthus ilieifolius 65 N Australia 3
Aegialitis annulata 114 E Australia 4
Aegieeras eomieulatum 148 E Australia 4
Avieennia germinans 100 Costa Rica I
Avieennia integra 63 N Australia 3
A vieennia bieolor 90 Costa Rica I
Avieennia marina 90 E Australia 2
300 E Australia 4
Bruguiera gymnorhiza 85 E Australia 4
Ceriops tagal 60 E Australia 2
300 E Australia 4
Laguneularia raeemosa 80 Costa Rica I
Lumnitzera raeemosa 90 South Africa 2
110 E Australia 4
Pelliciera rhizophorae 37 Costa Rica I
Rhizophora harrisonii 4S Costa Rica I
Rhizophora mangle 60 Costa Rica I
Rhizophora stylosa 55 E Australia 2
148 E Australia 4

IJimenez 1984; 2Macnae 1966; 3Wells 1982; 4Hutchings and Saenger 1987
Soil salinity is regulated by a number of factors, including tidal inundation,
soil type, topography, depth of impervious subsoils, amount and seasonality of
rainfall, freshwater discharge of rivers, run-off from adjacent terrestrial areas, and
evaporation. In tidally inundated situations, however, evaporative losses and the
frequency of flooding are the major factors determining soil salinity. Other climatic
factors, such as humidity, wind velocity and high solar radiation, together with the
extent of plant cover, exert a significant influence on evaporative losses from the
mangrove community.
Particularly where the clay content is high, soils have a high resistance to
internal salt and water movement (Hollins et al. 2000). As a result, tidal inundation,
rainfall and evaporation principally affect the soil surface, although with time an
equilibrium with the soil at considerable depths will be reached. As an
approximation, however, an initial understanding of the regulation of soil salinities
can be made considering only the surface processes. At any particular point in the
intertidal gradient the soil salinity can be directly related to: (I) salinity of the tidal
Physico-chemical Factors and Mangrove Performance 139

water; (2) time interval between inundations; (3) rainfall; (4) evaporation rate; (5)
retention properties of soil; and (6) run-on minus run-off.
All of these, except time interval between inundations, can be relatively
constant at a particular locality. Thus, at any particular locality, the soil salinity
along the intertidal gradient is determined more or less by the time interval between
inundations. At a locality with low rainfall and high evaporation, a soil salinity
maximum may be broad and located high in the intertidal gradient where inundation
is infrequent. On the other hand, where rainfall greatly exceeds evaporation as, for
example, on Kosrae Island, Micronesia (Ewel et al. 1998a), no soil salinity build-up
occurs, and the soil salinity will simply show an approximately linear decrease in
the landward direction.

Fig. 4.8 Satellite view of the coast of French Guiana. showing the Sinnamary estuary
with its zoned mangrove communities. The patchy clouds indicate the freshwater
swamp forest~ dominated by Pterocarpus ojficinalis with swampy grasslands further
inland. (Part of SPOT KJ 690/339 dated 02109/97. © cnes 1997 - distribution Spot
Image. Reprinted with permission.)

Thus, two extremes of intertidal environments with respect to interstitial soil


salinity can be recognized. First, areas of high rainfall that are permanently wet (fig.
4.8), where the soil salinity decreases progressively from the sea towards the land,
and where the vegetation ranges from mangroves to inland vegetation without
discontinuity. Such areas include Papua New Guinea (Taylor 1959), Malaysia (Chan
et aJ. 1993), Kosrae Island (Ewe! et al. 1998a), Nigeria, Cameroon and Gabon (Boye
140 Mangrove Ecology, Silviculture and Conservation

et al. 1975, Saenger and Bellan 1995) as well as in parts of Central America
(Jimenez 1984, Fromard et al. 1998). The second extreme type consists of areas of
low rainfall where dry and wet seasons alternate, where evaporation rates are high,
and where salt flats and/or saltmarshes appear between the mangrove community and
the inland vegetation (fig. 4.9). Such areas are well represented in Australia (Fosberg
1961, Saenger et al. 1977, Eisol and Saenger 1983), in Senegal, Gambia and Guinea
(Saenger and Bellan 1995), and in the Caribbean (Cintr6n et al. 1978)
Evaporation rates across saItflats have been measured at Cocoa Creek near
Townsville in north-eastern Australia and found to vary from about 4 x 10-3 m dol
when the surface is wet to less than 2 x 10- 3 m dol when the surface is dry (Hollins
and Ridd, 1997). Most of the evaporation occurs during the day time due to the
extreme temperatures (>50°C) of the surface sediments. Hollins and Ridd (1997)
found that in the 28 days of a spring tide cycle, the saItflats were covered for four
consecutive days and remained damp for a further four. For the other 20 days, they
remained dry. With a total monthly evaporation of about 7 x 10- 2 m of water and no
tidal flushing, elevated salinities both in the surface and groundwater of the saltflat
inhibit colonization by macroflora.

Fig. 4.9 In those areas where the rainfall is highly seasonal and usually less than 1.500
mm per annum. extensive saltflats are formed on the landward margins of mangroves.
as shown here near Gladstone, Queensland.

The tolerance of mangroves to various levels of soil salinity is poorly known,


and until more laboratory and field data are available conclusions must remain
tentative (Snedaker 1982, Jimenez 1984, Hutchings and Saenger 1987, Ball 1988b,
Duke 1992, Smith 1992, Ball and Pidsley 1995). Ranking mangrove species for the
various salinity parameters shows no consistent sequence of species either in the
field or in culture. The foIl owing points, however, are worth noting. Avicennia
marina appears to grow over the largest salinity range, whereas Cynometra iripa and
Heritiera littoralis seem to have the narrowest ranges and do not grow where salinity
is high. Rhizophora stylosa, Aegialitis annulata, Bruguiera gymnorhiza and Ceriops
tagal grow at salinities up to three to four times the concentration of seawater.
Physico-chemical Factors and Mangrove Performance 141

Aegiceras comiculatum will tolerate extremes of salinity both at high and low
concentrations, although its overall range is less than that of A vicennia marina.
Factors associated with salt tolerance in mangrove species include the carbon
cost of water uptake, and the water use efficiency. The carbon cost of water uptake
increases with increasing salinity and is greater in the more salt tolerant species (Ball
1988b), manifested in the field by an increase in root biomass along gradients of
increasing soil salinity (Soto 1988, Saintilan 1997a, b). Water use becomes
increasingly conservative with increasing salinity and with increase in the salt
tolerance of a species (Ball and Farquhar 1984a, b), generally manifested in the field
by decreasing light interception along gradients of increasing salinity. Thus, salt
tolerant species tend to grow more slowly than less tolerant species even under
optimal salinities for growth (Ball 1988a). In southern Florida, Rhizophora mangle
decreases its salt stress by using surface water as its sole water source (Lin and
Sternberg 1994). During the wet season, the fine root biomass increases in response
to decreased salinity of the surface water, directly facilitating the uptake of surficial
low-salinity water.
Interestingly, where data are available for congeneric species pairs, generally
one species is found at high-salinity levels and the other at lower levels: high - A.
marina, B. gymnorhiza, C. tagal, L. racemosa, R. stylosa, S. alba, X. moluccensis;
low - A. integra, B. sexangula, C. decandra, L. littorea, R. apiculata, S. caseolaris,
X. granatum. Similarly, Smith (1988) showed that Ceriops australis and C. tagal
are often segregated along gradients of interstitial soil salinity, with C. australis
occurring in higher salinity regions. In the Fly River delta in Papua New Guinea,
Pernetta (1993) found a similar occurrence of high and low salinity pairs amongst
related species.
Ball and Pidsley (1995) examined the effects of soil salinity on the growth of
two closely related species, Sonneratia alba and S. lanceolata in northern Australia in
relation to their differential distributions along naturally seasonal salinity gradients
(fig. 4.10). They showed that there were interspecific differences in salt tolerance
which were founded on the inherent growth characteristics of the two species. In fact,
these species showed a neat trade-off between growth and salt tolerance with S.
lanceolata growing in salinities of up to 50% that of seawater while S. alba can
grow in 100% seawater. Both species, however, showed optimal growth in culture at
5% seawater. Nevertheless, at optimal salinity the growth of S. alba, measured as
biomass, height and leaf area, is less than half that of the less salt-tolerant S.
lanceolata, indicating that S. lanceolata will be the successful competitor even at a
salinity that is optimal for both species. On the other hand, S. lanceolata could not
grow or compete at high salinities. In this species pair, as in the others listed earlier,
a species can apparently opt for salt tolerance, or for rapid growth and competitive
ability under low salinity conditions, not both. Growth analysis showed that change
in net assimilation rate rather than a change in their leaf area ratios accounted for
most of the differences in growth between species, and for changes in growth by a
species with increase in salinity from 0 to 100% seawater (Ball and Pidsley 1995).
Without pre-empting the discussion of plant growth strategies in section 5. 1.4,
it is worth noting that this comparative study of two species of Sonneratia provides
elegant physiological support for the ecological classification of plants into
competitors, stress tolerators and disturbance-tolerators, the three primary growth
strategies of plants (Grime 1977, 1979).
142 Mangrove Ecology, Silviculture and Conservation

A different approach was used by Bunt et al. (l982a) to investigate the soil
salinity effects on growth. They examined upriver distribution patterns of mangroves
and related these to the upriver salinity gradients in order to determine the extent to
which salinity explains the observed pattern of distribution (see also 6.2.2). Five
rivers were examined in north-eastern Queensland and, although some variation was
found among river systems, nine species showed significant correlation with upriver
salinity gradients. Of these nine species, four showed positive correlations, that is,
they occurred at the high-salinity areas and were absent at lower salinities. These
were Rhizophora stylosa, R. apiculata, Sonneratia alba and Ceriops taga!. Five
species (Heritiera littoralis, Excoecaria agallocha, Acrostichum sp., Aegiceras
comiculatum and Rhizophora mucronata) showed negative correlations in the above
ranking, indicating their presence in the upstream, low-salinity areas. Three further
species (Avicennia marina, Bruguiera gymnorhiza and Xylocarpus granatum) showed
no significant correlations, suggesting that they can grow over almost the complete
salinity range from freshwater to seawater, and that their upriver distributions are
determined by factors other than salinity.

Fig. 4.10 Sonneratia lanceolata in its typical upriver habitat, fringing creeks that drain
freshwater from seasonally flooded floodplains around the Wildman River, Northern
Territory.

Based on these patterns, Bunt et al. (1982a) ranked the species in order of
decreasing tolerance of or adaptation to seawater: Rhizophora stylosa, R. apiculata,
Sonneratia alba, Ceriops tagal > Aegiceras comiculatum, Bruguiera parvijlora >
Excoecaria agallocha, R. mucronata, Acrostichum sp., Heritiera littoralis, Nypa
jruticans > Barringtonia sp.. Bruguiera sexangula, Sonneratia caseolaris, Hibiscus
tiliaceus.
Although this ranking shows some similarities to that of Wells (1982). it must
be remembered that salinity of the tidal water is only one of the variables
determining soil salinities, and that the salinities immediately adjacent to the roots
of mangroves - be they maxima, minima, means or ranges - will ultimately
determine growth and success of mangroves in particular situations. As Bunt et al.
Physico-chemical Factors and Mangrove Performance 143

(1982a) have emphasized, tidal water salinity is only one of the factors affecting the
distributions of even those species with significant correlations. In the case of those
species not showing significant correlations, factors other than salinity are probably
more important. Similar studies of the upriver distributions of mangroves have been
described from other regions (see 6.2.2).
In parts of the Caribbean (fig. 4.11), Rhizophora mangle occurs both in a dwarf
form (scrub forests rarely exceeding 1.5 m, litter fall of 1.3 t ha- I y-I; Teas 1979) or
as a tree, ranging in height up to 30 m (fringe or basin mangrove forests, litter fall
up to 16.3 t ha- I y-I; Lahmann 1988) and salinity has long been suspected to be the
primary causal agent (Lugo and Snedaker 1974, Lin and Sternberg 1992a). The dwarf
forests also have smaller leaves, a lower leaf area index (LAI of 3.0 as against 5.7),
and a more open canopy than did the basin forests (Araujo et at. 1997).
In the scrub forests, R. mangle had significantly lower CO 2 assimilation rates,
stomatal conductance to water vapour, and intracellular CO 2 concentrations, but
higher water use efficiency (ITE) and leaf oI3C values than those in the fringe
forests, indicating higher long-term water use efficiency (Lin and Sternberg 1992a).
During the wet season, these scrub mangroves rely heavily on rain-derived
freshwater, as indicated by the lower oD and OIBO values of their stem water and by
the increase in fine root biomass (Lin and Sternberg 1992a, 1994). During the dry
season, however, there is a shift of water sources in scrub mangroves and they
utilise the same water source as the fringe mangroves, reflected in their similar oD
and OIBO values of stem water. Water use efficiency (ITE) was significantly higher
for scrub mangroves both in the short-term and, as indicated by the higher leaf oI3C
values, in the longer-term. Higher water use efficiency results from stomatal
limitations on photosynthesis which, in turn, reduces the potential for growth.
However, Lin and Sternberg (1992a) tentatively suggested that the strong seasonal
fluctuations in salinity and water availability are the main cause of dwarfing.

Fig. 4. 11 Dwarf scrubs of Rhizophora mangle in the more elevated parts of the saline
Florida Everglades.
144 Mangrove Ecology, Silviculture and Conservation

When grown hydroponically in a greenhouse under twelve different growth


conditions, combining a range of salinities, nutrient levels and sulfide concentrations
(Lin and Sternberg 1992b), the two growth forms showed similar physiological and
growth responses to these treatments, suggesting the growth forms are
environmentally induced rather than genetically fixed.
High salinity, low nutrient levels, and high sulfide concentration all
significantly decreased CO 2 assimilation, stomatal conductance, and plant growth,
but only salinity significantly decreased intercellular CO 2 concentration and leaf
isotope discrimination, suggesting that the lower isotope discrimination, or higher
ITE, observed for scrub mangroves in the field, is caused solely by high salinity
during the dry season. Lin and Sternberg (1992b) thus concluded that hypersalinity
seems to be the overriding stressful environmental condition common to all scrub R.
mangle stands in southern Florida.
Similar changes in the physiognomy of other mangroves have also been found.
Thus, the variation in the physiognomy of the mangroves Avicennia germinans,
Rhizophora mangle and Laguncularia racemosa, as related to salinity in two different
geomorphic habitats (mudflats and interdistributary basins), was investigated by
L6pez-Portillo and Ezcurra (1989b). The highest species diversity and the maximum
height and diameter for the three species is at low salinities and in interdistributary
basins. However, the cover of A. germinans is higher on the mudflat and under high
salinities. The range of response to environmental change is wider in A. germinans.
There was a negative association between the cover of A. germinans and that of the
other two species, and the slope of the regression line suggests a substitution of one
unit of cover of A. germinans by one unit of cover of any of the other two species
in the interdistributary basins as salinity decreases.
Interaction of soil texture with salinity tolerance of mangroves was suggested
by experimental studies on the effects of hypersalinity (McMillan 1975b). Two
neotropical mangroves, A vicennia germinans and Laguncularia racemosa, were
experimentally subjected to hypersaline conditions while growing in a range of soils
with differing clay contents. Seedlings of various ages up to three-and-a-half years
were subjected to salinities up to five times that of seawater for forty-eight hours,
and their responses noted. In soils with a high sand content, whether coarse- or fine-
grained, the plants failed to survive this treatment. In soils with a clay content of 7-
10%, the hypersaline exposure was tolerated. Avicennia seedlings tested over a broad
range of salinity survived 48 h exposure to 60%0 in sand and water culture but failed
to survive at higher salinities.
It was suggested by McMillan (1975b) that, although the actual mechanisms
underlying interaction of soil texture and salinity tolerance are not understood,
depression of the pH in all the experimental soils indicated the involvement of
cation exchange. It was suggested that in clay soils the exchange of Na+ and H+ ions
may reduce the salinity of the interstitial water immediately around the roots, and
that the adsorption of Na+ ions would cause the clay particles to deflocculate. In turn,
this would reduce the contact of roots and hypersaline water, thereby facilitating the
uptake of water, and simultaneously reduce the uptake of salt; wilting and salt
extrusion by the experimental plants suggested that this took place.
As a result of studies of potassium depleted soils in northern Australia (Keene
and Melville 1999), an alternative explanation of McMillan's (1975b) findings is
now possible. It appears that K+ availability in the soil is a primary factor in
determining the growth performance of Avicennia marina. Oxidation processes in
Physico-chemical Factors and Mangrove Performance 145

mangrove soils, particularly the oxidation of iron sulfides and the deposition of
jarosite, lead to the mobilization and subsequent depletion of K+. In turn, the reduced
availability of K+ was associated with reduced mangrove growth. As Avicennia
marina requires uptake of soil K+ into its leaves to maintain photosynthetic
efficiency and, as this uptake is inhibited by high salinity (Rains and Epstein 1967,
Ball et al. 1987), it appears that one result of high salinity is a salinity-induced
potassium deficiency. By the addition of clay to his cultures, McMillan (1975b) may
have simply reversed a salinity-induced potassium deficiency rather than identified a
soil texture interaction with salinity tolerance.
Some mangroves have the ability to maintain leaf K+ by selective ion uptake
even in the presence of high concentrations of Na+, thereby minimizing the effects of
a salinity-induced potassium deficiency. Thus, Medina et al. (1995) showed that
Acrostichum aureum has the ability to maintain high K+/Na+ ratios even at high
salinities, ranging from 14.1 to 18.1 at salinities of 81 - 159 mmol kg" and 55-86
mmol kg" respectively. In contrast, Rhizophora mangle and Laguncularia racemosa
have K+/Na+ ratios of 0.6-1.4 and 1.9-3.3 respectively.

Fig. 4.12 Saline grasslands adjacent to stunted mangroves of A vicennia germinans in


Senegal. Note the large Adansonia digitata in the background, indicating dry climatic
conditions. (Photo. F. Blasco)

There can be no doubt that the availability and proximity of freshwater affects
mangrove development. For example, more luxuriant and species-rich mangroves
occur in high rainfall areas (Tomlinson 1986, Hutchings and Saenger 1987). Bunt et
al. (1982b) examined species assemblages within 56 coastal systems between
Rockhampton and Cape York, north-eastern Queensland, and related these to
prevailing hydrological conditions. Their data showed that species distribution is
strongly influenced by the extent of freshwater influence either from rainfall or from
'wet' rivers, that is, rivers that flow reliably for most of the year. Obversely, Diop et
al. (1997) have shown significant changes in the mangroves of the Saloum River
estuary, Senegal (figs. 2.5 and 4.12), as a result of a decrease in total rainfall during
the 1980s and a shortening of the rainy season. The waters of the estuary have
146 Mangrove Ecology, Silviculture and Conservation

acidified and become hypersaline, the mangroves have declined and bare saltflats
('tannes') and saline grasslands have increased in surface area.
Besides flow reliability, catchment characteristics will also influence the
hydrological regime; thus, two catchments with similar total rainfall may, through
their characteristics of geology, topography, soils and vegetation cover, effectively
absorb and utilize different proportions of that rainfall, with the proportion
remaining after evaporation becoming part of the surface run-off from that
catchment. For example, the continental islands examined by Bunt et al. (1982b) and
shown to be depauperate, have small catchments, largely of rock, with steep slopes,
skeletal soils and sparse vegetative cover. Their run-off coefficients, primarily the
ratio of run-off to rainfall, are high because little of the rain is able to infiltrate the
soil where it can be retained and utilized. Those catchments, on the other hand, with
extensive swamps, overflow basins, dense vegetation or sandy landscapes, have low
run-off coefficients and allow considerable retention of water with more regulated and
sustained release to their drainage river systems. However, some catchments without
these characteristics may also have low run-off coefficients because of low rainfall
and high evaporation or transpiration over the catchment.
Run-off coefficients for all drainage basins along the coast of Queensland were
examined in relation to their mean annual rainfall (Hutchings and Saenger 1987). A
number of discrete groups of drainage basins were recognized which could be broadly
arranged in some order on two hydrological gradients: increasing freshwater inputs
and increasing catchment retention of flows. A third dimension should be added to
this arrangement, namely the reliability of the freshwater input. These groupings
show some similarity to the detailed classification based on floristic data (Bunt et al.
1982b) and suggest that catchment characteristics and, to a lesser extent, flow
reliability are about equal in importance to average rainfall of the catchment area in
shaping the floristic composition of the mangroves in the various coastal regions.
It should be noted that north-eastern Australia does not have any drainage basins
with both high and reliable rainfall and high catchment retention. This probably
reflects the general aridity and low relief of the continent. It results in the absence of
such river basins as, for example, the Fly River or Purari River deltas in Papua New
Guinea, with extensive mangrove development and vast Nypa forests in the reduced-
salinity reaches (Womersley 1975, Percival and Womersley 1975, Robertson et al.
1991).
5. Biotic Interactions and Mangrove Performance

I do not recall any scene more expressive of hideous horror


than a mangrove morass swarming with ferocious mosquitoes,
filthy-looking saurians,
and slimy snakes of various hues,
whose lighlest sting is as fatal as a dose of prussic acid,
while the deadly miasmo which fills the air
is quite palpable.
It is fortUlUlte that such spots are infrequent,
and more so thot they are difficult of approach,
for they are usually located amid a dense mass of green shrubbery,
which shows light and life above and gloam and death below.
J.M. Murphy (1899:355)

5.1 Plant·Plant Interactions


By way of introduction, it seems instructive to consider whether mangroves form an
assemblage or a community; in other words to what extent do the individual species
interact with each other? An assemblage can be considered to be a more or less
random aggregation of species that are ecologically independent but that co-exist in
the same space. While such associations may exist in nature as subsets of taxa
within particular habitats, their practical recognition is difficult. In contrast, a
community is a local set of functionally interacting species of plants, micro-
organisms and animals, living together at some locality.
That is not to say that discrete communities exist in mangroves, as elsewhere,
with areas of relatively uniform and constant vegetation, and with pronounced
changes in species composition between them (Bastow and Chiarucci, 2(00). Thus,
whilst we can speak of 'communities' for convenience, the variation that exists at
the 'community' level can be seen as only a larger-scale manifestation of variation at
the micro-habitat level.
The constituent plants of the mangrove community interact with one another,
often in specific or defined ways; many of these interactions are subtle, and most are
poorly studied and little understood. There is, however, a gradually increasing
awareness that plant-plant interactions within the mangrove community are
important because the distribution and success of the mangroves cannot be
adequately explained solely in terms of their interaction with the physico-chemical
environment.
Several categories of plant-plant (or plant-microbe) interactions are recognized
as important in determining the structure and/or function of the mangrove
community. They include: parasitic, antagonistic, mutualistic, and competitive
interactions, which are discussed below.

5.1.1 Parasitism
Parasitic relationships are those in which the parasite obtains food from its host,
which mayor may not suffer harm as a result. Many such relationships occur in the
mangrove community.
Mistletoes (family Loranthaceae) are parasitic plants which, although capable of
photosynthesis, tap into the host's vascular system to obtain water and mineral
nutrients (fig. 5.1). They are relatively benign parasites and rarely kill the host
148 Mangrove Ecology, Silviculture and Conservation

plant. Nevertheless, they deprive the host of desalinated water and nutrients, both of
which may be scarce resources for mangroves. As well they cause growth
modification and shading of the affected branches. A small number of species of
mistletoes occur on mangroves, and some are only found on specific mangrove hosts
(Hutchings and Saenger 1987).

Fig. 5.1 Amyema congener is one of the mistletoes that commonly occurs on rainforest
trees but which occasionally occurs on several mangrove species. including
Excoecaria agallocha a~ shown here in eastern Australia.

The neotropical mistletoe Phthirusa maritima commonly occurs on Conocarpus


erectus and the strand plant Coccoloba uvifera. Goldstein et al. (1989) and Orozco et
al. (1990) studied the gas exchange, water relations and carbon balance of this
mistletoe and its hosts on the Venezuelan coast. Under similar light and humidity
conditions, mistletoes had higher transpiration rates, lower leaf water potentials, aOO
lower water use efficiencies than their hosts. Potassium content was much higher in
mistletoes than in host leaves, but mineral nutrient content in the xylem sap of
mistletoes was relatively low. The resistance of the liquid pathway from the soil to
the leaf surface of mistletoes was larger than the total liquid flow resistance of host
plants. The costs of infestations to the hosts included the induction of higher
transpiration rates, reduced CO 2 assimilation rates, and lower water-use efficiency
(Orozco et al. 1990).
Biotic Interactions and Mangrove Performance 149

Being relatively benign and uncommon, it seems that mistletoes exert little
influence on the mangrove community as a whole, although individual host plants
may suffer considerable stress. Similarly benign, are the members of the
Orobranchaceae, a family of root parasites that tap into the host's vascular system.
Cistanche tubulosa is particularly common on Avicennia marina and salt marsh
plants around the Arabian Gulf but the specific effects on their hosts are unknown.
Unlike the parasitic flowering plants discussed above, parasitic fungi can have
devastating effects. Many parasitic fungi occur in the mangroves from the canopy to
the root. Usually some equilibrium is established with the host, but sometimes that
eqUilibrium is disturbed and considerable mortality results. For example, on the
Central Queensland coastline, a previously undescribed species, Phytophthora
(Halophytophthora) operculata, caused considerable mortality in Avicennia marina
(Pegg et al. 1980, Pegg and Alcorn 1982). This fungus is normally a leaf litter
decomposer and, as such, it occurs throughout Australian mangrove communities.
However, it does have the capacity to become pathogenic, attacking the roots of its
mangrove host whose susceptibility has been heightened by natural or anthropogenic
stress (fig. 5.2). As a result of fungal invasion, the roots function inefficiently, or
even cease to function, and severe water stress is induced. leading ultimately to death
of the mangrove.

Fig. 5.2 Dieback of the mangrove AviceMia marina at Gladstone. Queensland. The
fungus Phytophthora operculata was found to be the pathogen responsible.

So far, wherever outbreaks of Phytophthora and high mortality have been


recorded, only one host, Avicennia marina, has been involved, and it appears that its
susceptibility to this parasite is considerably higher than that of other mangrove
species (Pegg and Alcorn 1982, Weste et al. 1982). The resultant selective mortality
has led to a dramatic change in the species composition at particular localities. For
example, in Port Curtis, where almost pure stands of A vicennia once occurred, the
mangrove community contains virtually no mature A vicennia marina at present, aOO.
Rhizophora stylosa is rapidly filling the gap. Hence, the change from an A vicennia
150 Mangrove Ecology, Silviculture and Conservation

marina community to one dominated by Rhizophora stylosa has occurred as a result


of a fungus over a relatively short period, less than four years.
Wier et a1. (2000) have identified an imperfect fungus, Cytospora rhizophorae,
that causes dieback and mortality in Rhizophora mangle in Puerto Rico. This fungus
is a facultative pathogen and requires a wound to infect the host; wounds may be of
biotic (macroinvertebrate herbivores) or physical (storm or debris damage) origin. In
their survey, Wier et a1. (2000) found that 261 of 767 (33%) Rhizophora mangle
seedlings had wounds capable of leading to infection. Such a high incidence of
wounding in combination with salinity stress appears to predispose the southern
Puerto Rican mangroves to this particular fungal infection and contributed
significantly to localized dieback.

5.1.2 Antagonism (Ammensalism)


Antagonistic relationships are those in which the growth of a particular plant is
inhibited or interfered with through the creation of adverse conditions by another
plant, generally through the production and secretion of toxic or inhibitory
substances. Numerous examples have been reported in which vegetational
composition and species distribution have been attributed to the action of chemical
inhibitors of both shoot and root origin.
Clarke and Hannon (1971) investigated the potential importance of inhibitory
compounds in the mangrove communities around Sydney, Australia. Examination of
the physico-chemical environment of the Sydney mangrove and saltmarsh species
indicated the importance of salinity and waterlogging in determining distributional
patterns of species (Clarke and Hannon 1969, 1970). It also suggested that additional
factors such as inhibitory substances must be operative in maintaining the sharply
defined vegetation zones. For example, species that overlap in their tolerance of
salinity and waterlogging form mixed stands unless interaction between them
favours only one species. In New South Wales, Sarcocornia quinquejlora rarely
grows beneath mangroves, although it is not limited by salinity or by waterlogging,
nor by light where the canopy is open. Similarly, seedlings of Casuarina glouca
have not been observed beneath mature trees in the Casuarina zone.
Clarke and Hannon (1971) used leachate and macerate extracts of both bark and
leaves from A vicennia and cladodes and litter from Casuarina to detect any growth-
inhibiting substances that might be derived from rain falling through the tree
canopy. Leaf detritus from the soil surface of the Avicennia and Casuarina zones was
also tested for the accumulation of any toxins. No inhibition of Sarcocomia
seedlings or of mature plants of any of the associated species (Suaeda, Triglochin and
Sporobolus) by mangrove extracts was found. Juncus plants were healthy only in
the control treatment (tap water), and yellowing of the shoots was common where
leachates and macerates were included. Survival and growth of seedlings of both C.
glauca and J. maritimus were inhibited by the presence of a layer of Casuarina litter
on the soil surface, but no significant differences in growth were found using
Casuarina extracts. These investigators concluded that in the Sydney communities it
appears highly unlikely that phytotoxic exudates influence the establishment and
maintenance of zonation patterns, but that Casuarina litter creates physical
difficulties for germination and seedling growth.
In contrast to these temperate mangrove communities, it seems likely that in
the more tropical and species-rich mangroves, toxic and inhibitory exudates are of
greater significance. For example, phytotoxic exudates from leaf litter of some·
Biotic Interactions and Mangrove Performance 151

mangroves, including Lumnitzera racemosa, Ceriops decandra and Rhizophora


apiculata, have been reported to inhibit the growth of roots and shoots of
Rhizophora apiculata and R. mucronata seedlings (Kathiresan and Thangam 1989,
Kathiresan et al. 1993). Similarly, phytotoxic substances are contained in Nypa
jruticans leaves which inhibit both germination and seedling growth of Pennisetum
polystachion, Euphorbia heterophylla, Phaseolus lathyroides and Centrasema
pubescens (Wongkaew and Techapinyawat 1996).
Phytochemical screening of mangroves has been inadequate but a range of
exotic compounds has nevertheless been reported from mangroves (see also 5.2.3).
Catechol-type tannins are abundant in the bark, wood and leaves of many mangroves
(Brunnich and Smith 1911, Hogg and Gillan 1984, Achmadi et at. 1994). Brugine
has been recorded from stem and bark extracts of Bruguiera sexangula, B. exaristata
and B. cyliruirica (Loder and Russell 1969, Kato 1975) and a triterpenoidal saponin
has been recorded from the roots of Acanthus ilicifolius (Minocha and Tiwari 1981).
Pentacyclic triterpenoids were isolated in considerable quantities from leaves of
. Avicennia officinalis, Acanthus ilicifolius, Bruguiera gymnorhiza, Ceriops decandra
and Rhizophora mucronata (Ghosh et al. 1985). The fish-poisoning properties of the
bark and stems of Ba"ingtonia, Thespesia and Derris, three common associates of
mangroves (fig. 5.3), are well known and exploited by Australian Aborigines
(Everist 1974). Leaves of Avicennia germinans contain iridoid glucosides (Fauvel et
al. 1995). A novel complex of taraxeryl compounds has been described from
Rhizophora apiculata (Kokpol et at 1990), and Cleroderuirum inerme contains a
range of flavonoids, terpenoids, the caffeic glycoside ester, verbascoside (Fauvel et
al. 1989), and a novel4a-methylsterol (Akihisa et al. 1990).

Fig. 5.3 De"is trifoliata is a common climber in the mangroves of northern Australia,
throughout southeast Asia, and extending as far as East Africa. In Australia, coastal
Aboriginal people used the crushed leaves as a fish poison.

All of these compounds are physiologically active, capable of regulating or


inhibiting growth; at present, however, no information is available to indicate that
they do so under field conditions.
152 Mangrove Ecology, Silviculture and Conservation

5.1.3 Mutualism
Mutualism is the association of individuals of different species such that their ability
to survive and reproduce is greater when together that when apart. Numerous
mutualistic associations have been documented for plants in general, but the few that
are known from mangroves mostly involve interactions of plants and micro-
organisms (Hutchings and Saenger 1987, Holguin et al. 2(01). Several types of
mutualistic interactions between mangroves and micro-organisms have been
identified, although details of their frequency of occwrence are not available. The
known and potential mutualistic interactions are discussed below.
The first, and probably the most widespread interaction, occurs in the
rhizosphere, that region immediately surrounding the fine roots which is
characterized by enhanced microbial activity (Smith and Delaune 1984). Although
generally not intimately connected with root cells. the fungi and bacteria modify the
micro-environment around each root through their metabolic activity, releasing
nutrients and altering the pH of the soil. In turn, this microbial flora probably
depends on the leakage of organic material from the roots, which is used as a source
of energy (Holguin et a1. 2(01). A number of soil fungi are characteristically
associated with mangrove roots and form part of this rhizosphere flora (Hyde and Lee
1995).
One micro-organism that forms nitrogen-fixing root nodules with higher plants
is the actinobacterial genus Frankia. A number of unrelated genera of flowering
plants including Casuarina form nodules with Frankia (Bond 1956, Fleming et a1.
1988). Plants with nodules grow much better in nitrogen-deficient media than those
without nodules (Bond 1963, Kumar and Gurumurthi 1999). For Casuarina giauca, a
common inhabitant of the landward margins of mangroves in eastern Australia where
they abut freshwater swamps (and in other areas such as Florida and Bangladesh
where it has been introduced), these nodules may be of ecological significance.
The bacterial genera Rhizobium and Bradyrhizobium form nitrogen-fixing
nodules almost exclusively on the roots of the angiosperm families Fabaceae and
Caesalpiniaceae (Allen and Allen 1981). Nodulated legumes grow more vigorously
in nitrogen-deficient soils than do non-nodulated ones. In view of the low nitrogen
status of most mangrove soils (Boto 1983) and adjacent freshwater swamps, root
nodules may be important to the legumes commonly found in or around mangroves,
including the mangroves Cynometra iripa and C. ramiflora, the climbing associates
Derris and Dalbergia, and the 'almost mangrove' Pterocarpus officinalis (Saur et al.
1998). To date, however, no information on root nodules in mangroves is available.
Another potential type of interaction is more intimate and involves fungi which
form a direct association with roots of higher plants. These fungal associations,
which are known since the Middle Eocene (Lepage et al. 1997), are tenned
mycorrhizae. In some cases, the fungi are unicellular and live within the individual
root cells ('endomycorrhizae'), but in many cases the fungi cover the root tips in a
thick mat and penetrate the intercellular spaces of the cortex ('ectomycorrhizae').
Although no mycorrhizae have been reported specifically from mangroves, they are
frequently found in forest and swamp soils that are rich in organic matter. In view of
their habitat diversity, it would seem likely that some mangrove mycorrhizae 00
occur. Like the rhizosphere flora, mycorrhizae facilitate the movement of
phosphorus, potassium and calcium into the roots, and the movement of metabolites
from the roots to the fungus.
Biotic Interactions and Mangrove Perfonnance 153

Mutualistic interactions involving bacterial leaf nodules are common in over


400 species of the angiosperm families Rubiaceae and Myrsinaceae (Lersten am
Homer 1976). The bacteria are maintained as a colony in the closed shoot tips of the
plant and enter developing leaves through the stomatal pores, ultimately forming
chambers along the leaf margin (Miller et al. 1983). These bacterial leaf nodules
have been shown to be capable of nitrogen fixation (Van Hove 1976) and they may
be involved in synthesis of cytokinin (Miller et al. 1983). Three mangrove species,
Aegiceras comiculatum, A. floridum and Scyphiphora hydrophyllocea, belong
respectively to the Myrsinaceae and Rubiaceae in which bacterial leaf nodule
formation is common. These species would be worthy of detailed study in this
respect.
Other interactions with micro-organisms and mangroves have recently been
comprehensively reviewed (Holguin et al. 2(01). Many of these microbial
interactions involve nutrient transformation and recycling (Alongi et al. 1992) am
they can be considered to provide benefits to both parties. The dissolved organic
carbon present in interstitial water as a result of the degradation of plant material arxI
from root exudation, is consumed by bacterial populations which result in this form
of carbon (and the associated nutrients) being recycled rather than exported to adjacent
ecosystems. Thus, in return for mangrove carbon, bacteria fix N2 (see 4.5.3),
solubilize phosphates where soils are not always completely anoxic (VIIZQuez et al.
2(00), mineralize organic sulfur, and produce soluble iron, all of which, in turn,
become available to the mangrove plants.
Due to the increased awareness of the significance of these microbial
interactions, attempts have been made to inoculate seedlings of A vicennia germinans
with the cyanobacteria, Microcoleus chthonoplastes, to facilitate N 2-fixation and to
improve seedling growth performance (Toledo et al. 1995). Inoculated seedlings
showed significantly increased total N contents (Bashan et al. 1998). Similar
attempts have been made on Avicennia germinans seedlings with the terrestrial
halotolerant growth-promoting bacteria, Azospirillum halopraeferens and A.
brasilense (Puente et al. 1999). Thirteen species of phosphate-solubilizing bacteria
were isolated from the rhizosphere microbial community of Avicennia germinans
and Laguncularia racemosa (VIIZQuez et al. 2(00). These bacteria appear to produce
organic acids which solubilize tribasic calcium phosphate in the rhizosphere.
When the N2-fixing bacteria, Phyllobacterium sp. was combined with the
phosphate solubilizing bacterium, Bacillus licheniformis (both isolated from the
mangrove rhizospbere), nitrogen incorporation into mangrove leaves virtually
doubled (Holguin et al. 2(01).
Although the information available on mutualistic interactions in the mangrove
flora is limited, there is sufficient to suggest that the study of mutualism, am
particularly with microbes, may be a productive line of investigation. The fact that
most of the interactions involve the availability of nitrogen, which is generally in
short supply in mangroves around the world, suggests that considerable ecological
significance may be attached to a greater understanding of the relationships between
mangroves and micro-organisms.
Some of the epiphytes occurring in mangroves may also have a mutualistic
association but, whereas the benefit to the epiphyte is readily discernible, the
benefit, if any, to the mangrove partner remains questionable. For example, the
Bromeliaceae contains approximately 50 genera and 2,600 species of which >50%
are epiphytic (fig. 5.4). They occur in the Americas from 37°N to 44°S. They are
154 Mangrove Ecology, Silviculture and Conservation

bird-pollinated (humming birds) and their earliest pollen is known from the late
Eocene (-40 my BP) in Panama. The distribution of this family is entirely confined
to the Americas, except for a single species (Pitcaimjafeliciana) in West Africa.

Fig. 5.4 Pelliciera rhizophorae with it~ complement of bromeliad epiphytes on the
Pacific coast of Costa Rica.

In epiphytic bromeliad species, water is absorbed through the leaves which are
covered by penneable scales which fonn an impervious seal when dried out. They
also possess C 4 carbon fixation, a further adaptation towards water-use efficiency
(Benzing, 1980). Bromeliads are characteristic epiphytes of mangroves in tropical
and SUbtropical regions of Central and South America. Common mangrove
epiphytes include Aechmea bracteata. Tillandsia dasyliriifolia, T. paucifolia and T.
streptophylla (Olmstead and G6mez-Juarez 1996). No known species are exclusive to
mangroves and most bromeliads extend over large altitudinal ranges. While the
bromeliads clearly benefit from this relationship, it is difficult to identify any benefit
to the mangroves. However, as there is one record of the humming bird Amazilia
tzacatle pollinating Pelliciera rhizophorae (von Prahl 1986), perhaps the pollinators
of the bromeliads also playa role in pollinating the mangroves.

5.1.4 Competition
Competition between plants has been defined as the tendency of neighbouring plants
to utilize the same quantum of light, ion of mineral nutrient, molecule of water, or
Biotic Interactions and Mangrove Performance 155

volume of space (Grime 1973). According to this definition, competition refers


exclusively to the capture of resources and is only one of the mechanisms whereby a
plant may inhibit the growth of a neighbour by adversely modifying its
environment. In this sense, competition is strongly contrasted with antagonism, two
interactions which are often lumped together in the more traditional usage of the
term 'competition'.
The competitive ability of a plant is a function of the area, activity, aJXl
distribution in space and time, of the plant surfaces through which resources are
absorbed. Therefore, it depends upon a combination of plant characteristics including
storage organs, height, lateral spread, phenology, growth rate, response to stress, aJXl
response to damage (Grime 1979). Several of these characteristics have been
discussed already under the heading of adaptation - the selective change of a
particular set of characteristics in a way suited to a particular environment. Stated in
another way, plants will tend to disperse as widely as possible. This may take them
into habitats where their physiological optima are exceeded. If they encounter other
individuals better suited to the prevailing environment, differences in growth
potential, either above or below ground, will result in the suppression of the less-
suited individual.
Within the mangrove environment, most plant species are relatively widely
dispersed. However, large differences in the environmental conditions also occur,
particularly in relation to water, salt, nutrients and light, and it seems clear that the
sharp boundaries between areas dominated by different species are often the direct
result of competition.
Even within communities, species composition may be determined, or at least
influenced, by competitive interactions between component species (Ball 1980). For
example, in a detailed study of the mangroves of the open shoreline at Princess
Charlotte Bay, Queensland (Elsol and Saenger 1983), it was found that the
distributions of Ceriops tagal and Avicennia marina overlapped to a large extent.
Their relative importance values indicate that two broad bands are recognizable: first,
on the landward side, the importance values of the two species vary inversely with
the other, and second, further seawards, both species vary in direct proportion to each
other. This suggests that on the landward side of the transect, the environmental
conditions are favourable for both species and they compete with each other. Further
seawards, the conditions are no longer so favourable, and both species together
decline in importance. The similarity of these two species, in terms of their salinity
and waterlogging tolerances and their leafing and dominance characteristics (fig. 5.7),
supports the notion of such a competitive interaction.
A similar competitive interaction between A vicennia germinans and Batis
maritima has been described from Mecoacan Lagoon, Mexico (L6pez-Portillo aJXl
Ezcurra 1989a). At this site, Batis is competitively excluded by Avicennia from
mudflats which it is capable of colonizing. Avicennia is capable of displacing Balis
in two ways, not mutually exclusive: root competition for space or by
overshadowing the Batis shrubs. Where A vicennia is dwarfed or where it has been
clear-cut, Batis grows abundantly (L6pez-Portillo and Ezcurra 1989a). In southern
Florida, Kangas and Lugo (1990) also found direct competition between mangroves
and salt marshes (see 4.2.1). They suggested that on frost-free coasts, mangroves are
competitively superior as they are able to allocate more resources into structures that
allow them to outcompete salt marshes where they co-exist.
156 Mangrove Ecology, Silviculture and Conservation

In their discussion of tree mortality in mangrove forests, Jimenez et al. (1985)


reviewed what they termed normal and massive mortality in mangroves. They
suggest that normal mortality is due to initial high rates of propagule establishment
soon after space has become available. This is followed by a period of competitive
thinning which results in the domination of a stand by fewer but larger trees. During
this period, intense intra- and inter-specifc competition occurs, not only for space
and light, but also for nutrients and water. A period of maturity follows in which
competition for space is reduced, and mortality is mostly due to the death of
suppressed individuals or of later recruitment. Most of the available energy is used
for stand maintenance and less is available to cope with significant changes in the
external environment. The stage of senescence, with few old and large trees, with
wide gaps and a lack of regeneration, is rarely reached in mangrove forests. It is often
by-passed by major environmental perturbations such as cyclones, frost, erosion, or
hydrological or depositional events (Lugo and Patterson-Zucca 1977, Wells am
Coleman 1981, Blasco et al. 1996, Duke 2001), or by major biotic events such as
fungal or insect outbreaks (Pegg et al. 1980, Piyakarnchana 1981, Wier et al. 2000)
which cause massive mortality of an already stressed or senescent system (fig. 5.5).
These findings led Jimenez et al. (1985) to postulate a generalized model of
mangrove stand evolution which relates the stem density of a particular stand to the
age of that stand. The progressive changes in stem densities are linked by density-
dependent mortality as a result of intra- and inter-specific competition, on the one
hand, with habitat changes on the other.
In crowded plant populations in temperate and cold regions, 'self-thinning'
occurs over time (Ham 1984), largely via intra- and inter-specific competition for
scarce resources, resulting in declining stand densities (White 1981, Westoby 1984).
Under natural dynamics, a model called 'the self-thinning rule' predicts that stand
biomass and stand density are related as:

log (AGB) = log C - 0.5 log (D)

where AGB is above-ground dry biomass (g), D is stand stocking density (stems m-2)
and C is a parameter. For forests of different species, log C has been found to fall in
the range of 3.5-4.4.
From their studies in French Guiana, Fromard et al. (1998) demonstrated that
mangrove forests can be described by this model because of their structural
simplicity, a feature in which mangrove forests are unique amongst tropical forests.
Mangrove forests have few species and are often even-aged. They found the
following relationship after omitting two stands which were affected by clear-felling
and sediment burial:

log (AGB) = 3.85 - 0.53 log (D) (r=O.93, n=7)

with log C, at 3.85, falling midway in the observed range for forests. Thus, Fromard
et al. (1998) concluded that the studied stands could be considered as different stages
of the same ecosystem and that the density level reached by each stage corresponds to
an equilibrium state from a competitive (trophic and structural) point of view.
Furthermore, they showed that their stand densities and estimated ages (fig. 5.6)
agreed well with, and quantified, the general evolutionary model proposed by
Jimenez et al. (1985) to account for density-dependent mortality.
Biotic Interactions and Mangrove Performance 157

Fig. 5.5 Wherever rapid accretion of sediment occurs periodically. a~ in the delta of
the Orinoco River, Venezuela, where sediments from the Amazon River are
deposited, extensive mangrove mortality can result. Similar 'cemetery' mangroves also
occur in the Guiana~. (Photo. F. Blasco)

Competition can also be viewed in the context of major adaptive strategies


which have evolved in plants (Grime 1973, 1979), and it is important to relate these
strategies to the processes which determine the structure and species composition of
vegetation. In other words, what specific characteristics of a plant give it an
advantage (a competitive edge) over another plant which is occupying the same space
and/or utilizing the same resources?
Two categories of external factors limit the amount of living and dead plant
material in any habitat. The first is stress, which includes unfavourable temperatures
or elevated salinities, and other factors which restrict photosynthetic production,
such as shortages of water, light or nutrients. The second category is disturbance,
which includes those factors involved with the destruction of plant tissue, such as
the activities of herbivores, pathogens, and humans, and such phenomena as wind,
frost, fire and erosion.
Grime (1973, 1979) and Grime and Campbell (1991) contend that three
strategies have evolved among established plants which relate to stress am
disturbance. These are the 'competitors', which exploit conditions of low stress am
low disturbance, the 'stress-tolerators' (high stress, low disturbance), and the
'disturbance-tolerators' or 'opportunists' (low stress, high disturbance). These three
strategies are extremes of evolutionary specialization, and many plants have adopted
various combinations which adapt them to habitats with intermediate intensities of
stress and disturbance.
A triangular ordination technique was developed (Grime 1977) which provides a
basis for classifying plants and vegetation types. Species are classified with respect
to (1) potential maximum rate of dry-matter production and (2) a morphology index,
reflecting the maximum size obtained by the plant under favourable conditions. This
approach assumes that the three primary strategies correspond to the three
158 Mangrove Ecology, Silviculture and Conservation

permutations of primary production rates and morphology, that is, rapidly growing
and large ('competitors'), rapidly growing and small (,disturbance-tolerators'), aOO
slow growing and small ('stress-tolerators').

,,/ I I

"
1#

~~
~~
4,500
00 it;' ~
,......
.....
3,500
~
e
.~ 2,500
'"c
~
"0 1,500
~
(/)

500
0

0 10 20 30 40 50 60 70 80 90
Estimated Age (y)

Fig. 5.6. Changes in stem density as a mangrove stand matures in the absence of
catastrophic environmental change. The underlying model is that of Jimenez et al.
(1985) with the data of Fromard et al. (1998) superimposed. The arrow indicates the
trend where the model differs significantly from the data.

As much of the required data for this type of analysis is not available for
assessing mangroves in this context, Saenger (l985a) used some equivalent data
from permanent study sites at Proserpine (22°S), Queensland. which have been
adapted as follows: the maximum monthly rate of leaf production was used as a
measure of potential maximum dry-matter production; and a dominance index
consisting of the product of the maximum height and mean density of each species
in the area was used. The results of this strategic ordination for Australian
mangroves (Saenger 1985a, Hutchings and Saenger 1987) are shown in figure 5.7.
Clearly, the spread is a relative one in that the scales have been suited to mangroves
and, as such, cannot be compared with similar ordinations of other plant
communities. Nevertheless, the ordination gives a relative indication of the tendency
towards the three strategies adopted by the various species of mangroves at
Proserpine.
The ordination suggests that, even on a relative scale, none of the mangroves
has adopted the strategies of extreme 'competitors' or 'disturbance-tolerators', but
there is a general distribution of the species towards the 'stress-tolerator' strategy.
The numerically most abundant and widespread species in Australasia. A vicennia
marina, Rhizophora stylosa, Aegiceras comiculatum and Ceriops tagal, appear to
have a combined 'competitor/stress-tolerator' strategy which would enable them to
persist during unfavourable periods on the one hand, and to exploit favourable
periods reasonably efficiently on the other. Avicennia is probably the most-studied
member of this group; its ability to grow in a wide range of habitats appears to be
Biotic Interactions and Mangrove Performance 159

due to its capacity to reduce its growth rate and adjust its growth habit in response to
increasing stress. It is also worth noting, in relation to the previous comments
concerning the competitive interactions between Avicennia and Ceriops, that these
two species appear to be almost identical in their adopted strategy. Together with
their similar tolerances to certain environmental conditions, this suggests that they
are indeed competing with each other where they co-occur. Acanthus ilicifolius am
Aegiceras comiculatum is another possible competitive pair that should be
investigated where their distributions overlap.
COMPETITION

\.
104 \

A. ilicifolius E. agallochu -
'\
"I-
~. comiculatum

. .
_ B. exnristata
- H. littoralis
_ B. gymnorhiza
- -
X.granatum
_c. mpa -B. parviflora O. octodontu

STRESS 0.5 . 2.5 5 10 25 50 100 DISTURBANCE


Maximum leafing rate (lvs/l000/day)

Fig. 5.7 Mangrove ordination according to the method of Grime (1977). (Data from
Saenger J985b).

Based on growth experiments under contrasting conditions of resource (light and


nutrients) availability, McKee (l995c) concluded that Rhizophora mangle, Avicennia
marina and IAguncuiaria racemosa were also generally 'stress-tolerators'. They
showed, however, variable degrees of flexibility in terms of their response when
stresses were reduced, with some species tending towards either the 'disturbance-
tolerator', or 'opportunists' strategy. McKee (1995c) found that when resource
availability was low, the mangrove seedlings repartitioned biomass to enhance root
growth or leaf area, maximizing their potential to capture nutrients and light -
typical responses of 'stress-tolerators'. Under such conditions, few interspecific
differences were apparent, suggesting that Avicennia germinans and IAguncularia
racemosa would be unlikely to outperform Rhizophora mangle. Under conditions of
high resource availability, however, major interspecific differences became apparent:
L racemosa and A. germinans maximized their potential for carbon gain and growth,
and minimized allocation to roots, while R. mangle maintained a similar allocation
as under low resource availability. Thus, under high resource availability, L.
racemosa and A. germinans outperformed R. mangle; interestingly, L racemosa
outperformed A. germinans as well, exceeding the relative growth rate, leaf
production and branch growth of A. germinans by about two-to-one. These growth
and partitioning patterns suggest that L racemosa would be at a significant
160 Mangrove Ecology, Silviculture and Conservation

advantage against both A. germinans and R. mangle where light and nutrient levels
are high (but note opposing conclusions concerning A. germinans and L racemosa
by Lovelock and Feller, in press). Rhizophora mangle. in contrast, maintains its
conservative growth and partitioning patterns, at least to the seedling stage. The
growth characteristics of A. germinans appear to fall somewhere between that of R.
mangle and L racemosa.
In the ordination of Australian mangroves (fig. 5.7), it may seem surprising to
find species such as Excoecaria agallocha, Xylocarpus spp., Lumnitzera racemosa
and Osbomia octodonta tending towards the 'opportunist' strategy. As apparently
slow-growing members of the landward fringes, they do not give the impression of
being opportunistic species. However, all of these species are deciduous or near-
deciduous (Saenger and Moverley 1985), producing a new canopy of leaves over a
very short time at the most favourable period of the year. The number of leaves in
their canopy varies greatly from year to year, and presumably reflects the degree to
which conditions are favourable during their leafing period. In this sense, these
species respond very rapidly to favourable or unfavourable conditions during their
leafing period and, at least during this restricted time, can be viewed as
·opportunists'. The inherent growth flexibility of lAguncularia racemosa (McKee
1995c) may well be a phylogenetic characteristic shared by other species of the
Combretaceae, such as Lumnitzera racemosa and Conocarpus erectus.
To what extent this approach can be generalized, is questionable, because of a
lack of data. Certainly in West Africa. Paradis (1979) has described lAguncularia
racemosa as an 'opportunist' which characterizes secondary successional stages in
mangroves. As a shade-intolerant species with abundant and widely dispersed seeds,
this species is able to rapidly colonize any gaps formed in the mangroves due to
human disturbance. This feature of lAguncularia helps to explain the various
accounts of its distribution in relation to the other species. For example, Rosevear
(1947), Lebigre (1983) and Din (1991) describe lAguncularia as a pioneer species,
colonizing newly deposited mudflats in the very frontal zone. Others describe it as
occurring at the landward boundary of the mangroves where soils are better drained
and of lower salinity. lAguncularia seems to be able to colonize virtually any
unoccupied area in sheltered intertidal zones, becoming outcompeted or excluded by
other species with denser canopies. Part of this wide ecological ability may be
attributable to its flexible growth and reproductive pattern (McKee 1995c) and to the
facultative peg-root development (Jenik 1970) that occurs in lAguncularia when
growing on newly deposited muds.

5.2 Plant-Animal Interactions


Plant-animal interactions in mangroves are extremely diverse; there are intimate
links between certain animals and mangrove pollination, the mutualistic
cohabitation of mangroves with such animals as ants and butterflies, and the trophic
dependence of certain faunal groups (Hutchings and Saenger 1987. Saenger 1994b,
Mastaller 1997, Hogarth 1999). The plant-animal interactions discussed in this
chapter are the ones that are of widespread significance. or which directly or
indirectly affect the physical environment in which mangroves grow.

5.2.1 Flowering and Pollination


Flower primordia develop on young plants of most mangroves when little more than
three or four years old (and even at 18 months under greenhouse conditions). The
Biotic Interactions and Mangrove Perfonnance 161

initiation of flowering appears to be independent of size. but the actual factors


involved are largely unknown. Most species of mangroves begin flowering in spring
(Tables 5.1 and 5.2) and continue through the early summer months which. in the
tropics. often coincides with the dry hot summer prior to the rainy season (Saenger
1982. Duke et al. 1984. Tomlinson 1986. Mulik and Bhosale 1989. Steinke 1988.
1999. Saenger and BeHan 1995. Fernandes 1999).
Table 5.1 Monthly occurrence of maximal flowering in the northern hemisphere.

Species Months Location Ref.


J F M A M J J A S 0 N D
A. marina x x Thailand I
A. bieolor x x x x Costa Rica 4
A. germinans x x x x Costa Rica 4
x Panama 6
x x x x x x x Martinique 8
x x x x Brazil 9
x x x West Africa 5
x Bermuda lO
A. officinalis x x Bangladesh 11
B. cylindriea x x x Thailand 1
B. gymnorhiza x x Bangladesh 11
C. erectus x x West Africa 5
x x Mexico 7
C. deeandra x x Bangladesh 11
C. taga/ x Thailand I
E. aga/locha x x Bangladesh 11
L littorea x Thailand I
H·fomes x x Bangladesh 11
La. raeemosa x x x x Costa Rica 4
x x x Brazil 9
x x West Africa 5
N. frutieans x x Bangladesh 11
x x Malaysia 12
R. apieulata x x x Thailand 3
x x Vietnam I3
R. harrisonii x x Brazil 9
R. mangle x x Martinique 8
x x Brazil 9
x Bermuda 10
R. mucronata x Thailand 2
R. racemosa x x x x Costa Rica 4
x x x West Africa 5
S. apetala x x Bangladesh 11
S. hydrophyllaeea x Thailand 2
x. nwluceensis x x Bangladesh 11
IWium-Andersen and Christensen 1978; 2Wium-Andersen 1981; 3Christensen and Wium-Andersen
1977; 4Jimenez 1988; SSaenger and Bellan 1995; 'Lefebvre and Poulin 1997; 7Tovilla and Da La Lanza
2000; 81mbert and Mmard 1997; 9pernandes 1999; IGellison 1997; IIChowdhury 1996; l2pong 1992;
13Clough et a1. 2000.

Duke (1990) investigated patterns of reproductive phenology with changing


latitude in Avicennia marina, using litter fall collections from 25 sites in Australia.
Papua New Guinea and New Zealand. The combined data revealed that flowering
shifted from November-December (late spring) in northern tropical sites. to
May-June (late autumn) in southern temperate sites. Duke (1990) found that
initiation of the reproductive cycle occurred when daylength exceeded 12 h. The
subsequent development rates to flowering and fruit maturation were determined by
mean daily air temperature. In combination. daylength and mean daily temperature
explained up to 92% of variance in the total cycle duration and timing.
162 Mangrove Ecology, Silviculture and Conservation

Table 5.2 Monthly occurrence of maximal flowering in the southern hemisphere.

Species Months Location Ref.


J A S 0 N D J F M A M J
A. ilicifolius x x x x NE Australia 5
A. annulata x x x x E Australia I
x NE Australia 5
A. comiculatum x x x x x x x x E Australia I
x x x SE Australia 4
x x x x NE Australia 5
A. marina x x x x E Australia I
x x x SE Australia 4
x NE Australia 5
x x x x S Africa 3
B. cylindrica x NE Australia 5
B. exaristata x x NE Australia 5
B. gymnarhiza x x x x x x E Australia I
x x x x x NE Australia 5
x x x x S Africa 3
B. parviflora x x x NE Australia 5
B. sexangula x x x x x NE Australia 5
C. schultz;; x x NE Australia 5
C. australis x NE Australia 5
C. decandra x x x x NE Australia 5
C. tagaJ x x x x E Australia I
x x x NE Australia 5
x x x x S Africa 3
C. iripa x x x x x x NE Australia 5
E.agallocha x x x x x E Australia I
x NE Australia 5
H. littoralis x x NE Australia 5
L littorea x NE Australia 5
Lu. racemosa x x x x E Australia I
x NE Australia 5
x x x x S Africa 3
N. fruticans x NE Australia 5
O. octodonta x x x x E Australia I
x x x NE Australia 5
P. acidula x x NE Australia 5
R. apiculata x x NE Australia 5
R. mucronata x x x x S Africa 3
x x x NE Australia 5
R. stylosa x x x x x x x x E Australia 1
x x x NE Australia 5
S. hydrophyllacea x NE Australia 5
S. alba x x NE Australia 2
x NE Australia 5
S. caseolaris x x NE Australia 2
x x x x NE Australia 5
X. granatum x x NE Australia 5
X. moluccensis x E Australia 1
x x NE Australia 5

ISaenger 1982; 2Duke 1988; 3Steinke 1999; ·Clarke 1994. Clarke and Myerscough 1991b; 5Duke et al.
1984.

Little is known about the breeding mechanism in mangroves. Most mangrove


flowers are hermaphroditic but are thought to be mainly outcrossing (Primack ani
Tomlinson 1980, Tomlinson 1986). In some species, the maturation of the style
occurs before that of the stamens (protogyny; e.g. Sonneratia) and the style is far
exserted above the stamens (herkogamy; e.g. Sonneratia) (Tomlinson 1986, Aluri
1990, Pandit and Choudhury 2(01). Other mangroves have protandry (e.g. Avicennia
marina; Clarke and Myerscough 1991b), have separate male and female flowers (e.g.
Excoecaria) or distylous flowers (e.g. Pemphis), which presumably enhance
Biotic Interactions and Mangrove Performance 163

outcrossing (Lewis and Rao 1971). Primack and Tomlinson (1980) divided the 54
species of mangrove for which they had information into hermaphrodites (85%),
monoecious (9%) and dioecious (6%).
Genetic studies have indicated that Rhizophora mangle (Lowenfield and
Klekowski 1992, Klekowski et al. 1994a, b) and Aegiceras comiculatum (Ge and
Sun 1999) are predominantly selfing species while Kandelia candel (Sun et al. 1998)
and A vicennia marina (Maguire et al. 2(00) are primarily outcrossing species,
although the outcrossing rate in Kandelia candel (fig. 5.8) decreases with decreasing
plant density and increasing plant age (Chen 2(00).
Outcrossing species rely on a pollinating mechanism more than species which
self-pollinate. Cross-pollination is almost entirely by animals, and the range of
pollinators is diverse and generalized, so that no plant is highly dependent on a
specific pollinator (Tomlinson 1986, Pandit and Choudhury 2001). The presumed
pollinators of mangroves are given in Table 5.3. With few exceptions, the
pollination mechanism in most mangroves is merely inferred from flower and pollen
morphology and from observations of potential pollen vectors visiting the flowers.
In that sense, much of our knowledge of mangrove pollination is anecdotal rather
than systematic. Nevertheless, pollination in most mangroves appears to occur
through the agency of wind, insects or birds, and most species possess small, non-
sticky pollen grains which are distinctive for most species (Muller and Caratini
1977).
Two detailed investigations of pollination and breeding systems in mangroves
are available from the east coast of India: Aluri (1990) investigated the floral biology
of Acanthus ilicifolius (fig. 5.8), Avicennia officinalis. Aegiceras comiculatum and
Lumnitzera racemosa in the Godavary Delta, while Pandit and Choudhury (2001)
investigated pollination and reproductive success in Sonneratia caseolaris (fig. 5.8)
and Aegiceras comiculatum in the Bhitarkanika Wildlife Sanctuary (200 30' N). A.
ilicifolius and S. caseolaris were outcrossers, requiring a pollen vector, and A.
officinalis was also found to be an outcrosser although capable of self-pollination.
L racemosa and A. comiculatum were both found to be autogamous. Microsatellite
analysis of Avicennia marina from twelve populations (Maguire et al. 2000) showed
a similar pattern to A. officinalis; it was largely outcrossing but variable levels of
self-pollination were evident in the different populations. Taken together, the floral
biology and the genetic data seem to indicate that the majority of mangroves are
largely outcrossers and thus rely on pollen vectors. Aegiceras comiculatum was
found to be autogamous (Aluri 1990, Ge and Sun 1999, Pandit and Choudhury
2001), but fruit set was found to be pollinator-dependent in Hong Kong (Ge and Sun
1999) and unaffected by pollinator-exclusion in India (Pandit and Choudhury 2001).
It should also be noted that not all flower visitors are pollinators. Inouye
(1980) classified flower visitors on the basis of their behaviour, as related to
methods of pollen or nectar harvest, into (1) pollinators, (2) thieves, that obtain the
reward without damaging the flower, but do not pollinate it because of a mismatch
of morphologies, and (3) robbers, that obtain the reward by damaging floral tissues
without resulting in pollination. In their detailed studies, Pandit and Choudhury
(2001) found that for S. caseolaris, flower visitors included 17 species of
Lepidoptera, 7 spp. of Hymenoptera, 3 spp. of Diptera, 5 spp. of birds and 3 spp. of
mammals. All species of Lepidoptera, Hymenoptera and birds were pollinators, and
the Diptera were thieves. The mammals were robbers, predating the flowers and
fruits, accounting for 74% of fallen buds. For A. comiculatum, flower visitors
164 Mangrove Ecology, Silviculture and Conservation

included 16 spp. of Lepidoptera, 9 spp. of Hymenoptera, 2 spp. of Diptera, 1 sp. of


Coleoptera and 3 spp. of birds. All the visitor categories were pollinators except for
the Coleopterans, which were nectar thieves. No robbers (predators) were noted on
the flowers or fruits of this species.

Fig. 5.8 Flowers of (A) Kandelia caruJel with prominent white sepals and fibrous
ee)
petals; (B) blue flowers of Acanthus ilkifo/ius; and red flowers of Sonneratia
caseolaris with its exserted style.

Aoral or extrafloral nectaries occur in several species, providing rewards for


animal pollinators. For example, in Sonneratia caseolaris, nectar containing about
20% sucrose was available for pollinators; more nectar was available in the
mornings for diurnal visitors than in the evening for nocturnal ones (Pandit aOO
Choudhury 2001). In Bruguiera haines;;, Noske (1993) reported small amounts of
nectar with around 20% sucrose. He also reported the occurrence of nectarivorous
mites, Hattena panopla (Acarina: Ameroseidae) on both the flowers (B. hainesii, B.
gymnorhiza and B. sexangula in Malaysia and B. e.mristata in northern Australia)
and the birds (Brown-throated Sunbird, Anthreptes malacensis in Malaysia, and the
Red-headed Honeyeater, Myzomela erythrocephala in northern Australia). The mites
apparently use the birds as a means of transport from one flower to the next.
In South Africa, Bruguiera gymnorhiza is pollinated by insects and sunbirds,
and the petals of this species are peculiarly adapted to this method of pollen dispersal
(Davey 1975). A similar mechanism has also been noted in Bruguiera hainesii
(Noske 1993). Bruguiera petals possess a heavily cutinized epidermal region which,
on the application of gentle pressure, causes the petal lobes to spring apart, thereby
releasing the stamens together with a puff of pollen. While birds are capable of
triggering this pollen release mechanism, it seems that insect visitors are too small
to do so (Kondo et al. 1987). Other large-flowered species of Bruguiera (such as B.
sexangula, fig. 5.8) are bird-pollinated whereas small-flowered species (such as B.
parviflora and B. cyUndrica) are pollinated by butterflies (Tomlinson et al. 1979).
Other species that appear to have specific adaptations for specialized pollinators
includes Acanthus ilicifolius (fig. 5.8), an outcrosser whose flower structure requires
a strong pollinator to separate the four stamens in order to reach the nectar at the
base of the ovary; during field observations only the yellow-breasted sunbird
Nectarina jugularis was noted to probe the flowers (Primack et al. 1981).
Interestingly, this sunbird is confined to north-eastern Australia but, as Acanthus
ilicifolius has a wider northern Australian and extra-Australian distribution, other as
yet unidentified pollinators must be involved as well.
Biotic Interactions and Mangrove Performance 165

Fig. 5.9 Bruguiera sexangula is an Indo-West Pacific mangrove which was


successfully introduced to Oahu, Hawaii. in 1922. This Hawaiian specimen, shown
flowering (A), regularly sets fruit although its pollen vector here can only be guessed
at. Inflorescence on a mature plant of Nypa fruticans in Panama (B). Note the flies
attracted to the scent and nectary. (Photo. N. Duke)

Flower colour also appears to have a role to play; thus, the white-flowered
Lumnitzera racemosa is insect-pollinated while the red-flowered L. littorea is
pollinated largely by honeyeaters, particularly Meliphaga gracilis (Tomlinson et al.
1978). The distributional range of L littorea extends beyond that of this honeyeater
and other pollinators are yet to be identified.
The time of flower opening also appears to be important. Sonneratia spp. are
reported to release copious amounts of dry or slightly sticky pollen at dusk when the
flower opens (Muller 1969). According to Tomlinson (1986), the flower only
remains open for 12 h during the night at which time pollen transfer occurs via
nocturnal bats and moths. However, in a detailed study of Sonneratia caseolaris in
India, Pandit and Choudhury (2001) found that the flowers remained open for 56 h
and that they had both diurnal and nocturnal visitors, whose separate exclusion did
not affect fruit set differentially.
Pollination by wind would appear to be an effective mechanism as mangroves
occur in windy coastal environments. However, few species have apparently made
use of this form of pollination. Excoecaria, which is dioecious, bears flowers in
catkins and possesses a two-celled pollen grain (Venkateswarlu and Rao 1975), can
be presumed to be wind-pollinated. When flowering, however, numerous insects also
visit the catkins. Rhizophora also appears to be wind-pollinated, although Primack
and Tomlinson (1978) noted that glands on the inner surface at the base of the
stipules produce a sugary secretion which is attractive to birds.
166 Mangrove Ecology. Silviculture and Conservation

Table 5.3 Presumed pollination agents reported for various species of mangroves.
Species Pollinator R~on Ref.
A. Uicifolius Sunbirds E ustralia. India 1.2
A volubilis Bees India 3
A annulata Ants N Australia 4
A. rotundifolia Bees India 3
A. comiculotum Bees E Australia 5
Insects. birds India 6
Bees, butterflies Hong Kong 23
A. marina Bees E Australia, India 5. 7
A. officinalis Flies India 2
A. germinans Bees Caribbean 8
Insects Colombia 9
B. exaristata Birds, honeyeaters Australia 10,4
B. cylindrica Butterflies E Australia 10
B. gymnorhiza Insects, sunbirds S Africa II
Birds E Australia. India 10
Honeyeaters Ryukyu Islands 12,3
B. hainesii Sunbirds Malaysia 21
B. parviflora Butterflies N Australia. India 4,3
B. sexangulo Birds Australia 10.4,3
C. schultzii Wind, insects N Australia 4
C. australis Moths E Australia 10
C. decandra Insects India 3
C.tagal Moths E Australia 10
C. erectus Wind, insects Mexico 13
C. iripa Bees E Australia 5
E. agal/oeM Wind India 14
Bees E Australia. India 15.3
Insects N Australia 4
H·fomes Bees India 3
H. tiliaceus Bees India 3
K. candel Insects E Asia, India 15,3
Bees, butterflies Hong Kong 22
L racemosa Insect., E Australia 16
Bees, Moths India 3
L littorea Honeyeaters E Australia 16
La. racemosa Insects Colombia 9
N. fruticans Insects, flies SEAsia 17.18
Wind India 3
O. octodonta Insects E Australia 15
P. rhizophorae Humming bird Colombia 19
P. acidulo Bees N Australia 4
R. apiculota Wind E Australia 10
R. stylosa Wind E Australia 10
R. mucronata Wind E Australia 10
s. hydrophyllocea Insects, bees E Australia. India 15.3
S. alba Bats. hawk-moths SEAsia 19
Bats W Australia 20
Moths E Australia I
S. caseoloris Bats, moths SE Asia, India 19,6
Moths E Australia I
T.populnea Bees, birds India. N Australia 3,4
x. granatum Bees India 3
X moluccensis Bees E Australia 15
lPrimack et aI. 1981; 2Aluri 1990; 3Na.,kar and Mandai 1999; 4Wightman 1989; sBlake and Roff 1972;
6Pandit and Choudhu~ 2001; 7Chanda 1977~Percival 1974; "Elster et al. 1999a; l'7omlinson et al.
1979; llDavey 1975; 2Kondo et aI. 1987; l~rovilla & Da La Lanza 2000; 14Venkateswarlu & Rao
1975; l~omlinson 1986; l~omlinson et al. 1978; 17Uhl 1972; l&a~ig 1973; 19yan Prahl 1986; 19paegri
& van der Pij11971; 20Semeniuk et aI. 1978; 21Noske 1993; 22Sun et al. 1998; 23Ge and Sun 1999.

5.2.2 Bioturbation of Sediments


Probably one of the best examples of direct interaction between the flora and fauna
consists of the reworking of sediments among the mangroves by crabs. mud-lobsters
Biotic Interactions and Mangrove Performance 167

and callianassid shrimps (Macnae 1966, Havanond 1987). Mud-lobsters (Thalassina


anomala) build large tunnelling burrows in tropical and subtropical mangroves.
These burrows are generally recognized by the mound of fresh mud up to 1 m high
around their entrances. The burrows are U-shaped and extend up to 1.5-2 m below
the surface. Mounds are continuously increased on the top and sides until they
connect with each other. Their entrances generally are blocked by a mud plug during
the day, but at night they are opened when the mud-lobster emerges to feed on
surface muds (Pillai 1985).

Fig. 5.9 Like mud-lobsters, callianassid shrimps also rework the sediment~ in and
around mangroves as shown here on the sandflats in front of mangroves (mostly
Rhizophora stylosa) in the Kimberley region of Western Australia.

Burrow densities up to 442 ha'i have been observed (Chai and Lai 1984). These
burrowing activities have various effects, but the enormous amounts of sediment
these animals bring to the surface help to mix the soils and to change their surface
characteristics. Often, the soil brought to the surface is anoxic and rich in organic
matter and sulfides (FeS) (Andriesse et at. 1973). This fresh mud oxidizes on the
surface and often forms localized patches of highly acidic muds (acid-sulfate soils)
characterized by the yellow mottling of jarosite (see 4.5.1). Natural mangrove
regeneration is inhibited by the low pH (Kathiresan et at. 1996). Gradually,
however, as the mud mounds age, the sulfur content decreases as a result of leaching,
and these slightly raised areas then become suitable for mangrove colonization. In
Rhizophora or Bruguiera forests, such elevations are initially colonized by the
mangrove ferns Acrostichum speciosum and A. aureum in small discrete patches,
although other species such as Derris trifoliata. Flagellaria indica and Finlaysonia
maritima have also been recorded·(Havanond 1987).
The mud-lobster provides an example of a species which can markedly alter the
mangrove environment. The burrows allow drainage of and interchange between
surface water and subsoil water; the mud is turned over, with subsurface muds placed
on the surface where they can be oxidized, leading to their acidification. Once the
sulfides are oxidized to sulfates, they can be leached from the mounds, allowing the
168 Mangrove Ecology, Silviculture and Conservation

mounds to be colonized by the mangrove fern which is not able to grow at the lower
level of the surrounding mud surface. This topographical change can also alter
drainage patterns, immersion time in seawater, or access to freshwater run-on which.
in tum. can lead to localized changes in species abundance and distribution.
Other burrowing organisms have similar effects, although generally on a
smaller scale (Warren and Underwood 1986). The burrows of fiddler crabs. mud-
skippers and even the mud crab (Scylla serrata) allow drainage, mixing and a degree
of aeration of subsurface waters in the mangroves, and in this way enhance the
growth of mangroves. The significance of these effects was demonstrated by
removing crabs from 15 m x 15 m enclosures over a period of 12 months in
Queensland mangroves (Smith et al. 1991). The removal of crabs resulted in a
significant increase in soil sulfides and ammonium. and a reduction in cumulative
growth and reproductive output of the trees.

5.2.3 Grazing and Trampling (Leaf Herbivory)


Another important example of plant-animal interaction in the mangroves is that of
grazing and trampling. The importance of grazing by large herbivores in the
mangrove ecosystem is not well documented, but probably has the effect of
maintaining the mangrove community at a lower level of plant biomass than would
occur in its absence. In this sense, grazing and trampling are not unlike other
regularly continuing disturbances. Where grazing is species-specific, changes in
species composition may be profound.
Mangrove foliage contains significant quantities of minerals, vitamins. amino
acids. proteins, fat, and crude fibre, and is thus a nutritious food source for
herbivores (Kehar and Negi 1953, Tanaka et al. 1994). Consequently, it is not
surprising that mangrove foliage is grazed by cattle, sheep, goats and camels (see
7.2.6).
Probably of far greater significance, however. is the widespread grazing of
leaves by insects (Johnstone 1981, Murphy 1990, Farnsworth and Ellison 1991.
1993) and crabs, which can exert effects at the individual and community levels.
Crab and insect herbivory of propagules has been previously discussed in 3.5.3.
Leaf herbivory by crabs can account for considerable leaf losses am
reprocessing in mangroves. Thus, Beever et al. (1979) found grazing by an arboreal
grapsid crab (Aratus pisonii) on Rhizophora mangle leaves to range from 0.4-7.1 %
of total leaf area. In south-east Queensland, Camilleri (1989) found that individual
Sesarma erythrodactyla, one of the most abundant species of crabs in the forest.
processed approximately half a leaf from any of the three species of mangroves
(Avicennia marina, Bruguiera gymnorhiza and Rhizophora stylosa) in 24 h under
laboratory conditions. Of this material, 20% was lost from the mandibles due to
'sloppy feeding', 68% was egested as faeces, and only 12% was converted into crab
biomass. In Thailand. up to 82% of the diet of sesarmid crabs was reported to
consist of mangrove material (Poovachiranon and Tantichodok 1991) while
Emmerson and McGwynne (1992) indicate that Sesarma meinertii consumed
Avicennia marina leaves at a rate of 0.78 g m-2 d-·, accounting for around 44% of the
leaf fall in Natal mangroves.
Many of the crabs are selective rather than general feeders on mangrove foliage.
Camilleri (1989) found that feeding preference in S. erythrodactyla was ranked A.
marina> Bruguiera gymnorhiza > Rhizophora stylosa. probably because leaves of
A. marina have a lower tannin and a higher nitrogen content. Micheli et at (1991)
Biotic Interactions and Mangrove Perfonnance 169

found that in East Africa. SesamuJ meinertii generally prefers leaves of Bruguiera
over Avicennia, although Steinke et aI. (1993) suggested that the age of the litter
was more important than its source, with crabs choosing yellow leaves of either
Bruguiera and Avicennia over green leaves of either species. SesamuJ messa and S.
smithii both prefer leaves that are decaying rather than freshly fallen (Micheli 1993a)
and Kwok and Lee (1995) showed that Chiromenthes bidens and Parasesanna plicata
do best when fed brown rather than yellow leaves. Smith (1998) found from
controlled feeding experiments in Darwin with PerisesamuJ darwinensis and P.
semperi that the crabs preferred green and brown leaves over yellow leaves of most
spercies, although brown leaves of Sonneratia alba was the most preferred type.
By way of contrast, the tropical sesarmid Neosamartium meinerti prefers freshly
fallen leaves (Dahdouh-Guebas et al. 1997); fresh leaves of Ceriops tagal contain
flavologlycans and consumption rates of fresh leaves correlated with flavologlycan
concentrations (Neilson et aI. 1986). With seawater leaching, flavologlycan
concentrations were reduced but when flavologlycans were experimentally restored on
older leaves, Neosamartium resumed grazing (Neilson et al. 1986).
Table 5.4 Weekly leaf consumption by Helice leachi of green, yeJlow and brown leaves of 8ruguiera
gymnorhiza together with their nutritional composition. (Data from Shokita 2(00)
Parameter Green Yellow Brown
Feeding rate (g leaf g-I crab) 0.27 0.18 0.31
Water (%) 71.4 70_1 65.1
Energy (kcal 100 g-I dry wt.) 317 328 339
Protein (g 100 g-I dry wt_) 7.3 2.0 2.9
Fat (g 100 g-I dry wt_) 2.5 4.7 5.7
Carbohydrate (g 100 g-I dry wt.) 78.0 79_6 79_7
Ash (g 100 g-I dry wt.) 12.2 13.7 11.7

Other factors responsible for feeding preferences are still unclear. Micheli
(1993a, b) found that preferences were not affected by tannins, water content, %
organics, C:N ratio, or leaf toughness. Many crabs store the leaves in their burrows
(Robertson 1986) but the benefit, if any, of this activity is still unknown. Neilson
and Richards (1989) found that leaves of Ceriops tagal were greatly altered by
degradation processes, with pectates rapidly degrading while polysaccharides resisted
degradation. As a result, all component sugars were still present in 8-week-old leaves
although the acetone-water soluble material had been completely removed. Shokita
(2000) examined the feeding preference of Helice leachi when green, yellow and
brown leaves of Bruguiera gymnorhiza were simultaneously made available. He also
examined the composition of the leaves (Table 5.4) concluding that brown leaves
were preferred by this species because of their elevated levels of fat and
carbohydrates, and their total energy content.
The arboreal insect fauna of mangroves appears to be characterized by an
abundance of Hymenoptera. Diptera and Psocoptera. differing from tropical
rainforests and swamp forests by the abundance of Psocoptera and the scarcity of
Collembola (Abe 1988); insect densities in Indonesian mangroves ranged from 13.1-
48.1 m- 2 and similar densities are likely in other tropical mangrove areas (Murphy
1990).
Exceptionally, major defoliation by insects has been reported from Hong Kong
(Avicennia marina; Lee 1991, Anderson and Lee 1995), Colombia (Avicennia
genninans; Elster et al. 1999a), Ecuador (Rhizophora mangle; Gara et a!. 1990),
Sumatra (Excoecaria agallocha; Whitten and Damanik 1986), Bangladesh (Sonneratia
170 Mangrove Ecology, Silviculture and Conservation

aptelala; Saenger and Siddiqi 1993), and Thailand (Avicennia alba; Piyakarnchana,
1981). Such defoliation events, while rare, may lead to mortality, particularly in
seedlings and saplings, and to changes in species composition.
Heald (1969) estimated a mean grazing effect on Florida mangrove leaves of
5.1% of the total leaf area, ranging from 0-18% on a leaf area basis. Johnstone
(1981) has suggested that approximately one-fifth of all mangrove leaf material at
his study site in Papua New Guinea is diverted to herbivorous rather than detrital
food chains, an estimate similar to that reported from Florida Bay (26% of total leaf
area; Onuf et al 1977), Braiil (16% of total leaf area; Lacerda et al. 1986b) arrl
Guadeloupe (4.5% of total leaf area; Saur et a1. 1999). In Australia, Robertson arrl
Duke (l987a) estimated that only 2.1 % of the annual mangrove leaf production was
grazed while Lee (1991) showed that Kantlelia cantlel in Hong Kong loses, on
average, 10.3% of its leaf area, with >90% of the leaves suffering <30% loss. Thus,
it seems clear that although considerable geographic and interspecific variation exist,
leaf herbivory in mangroves is considerable and is likely to exert marked effects at
the individual and community levels.
Based on recent investigations, mangrove herbivory was 3-6 times higher when
measured over the entire lifespan of leaves than when based on discrete studies
(Burrows 2001). Using long-term observations of tagged shoots of Avicennia marina
and Rhizophora stylosa, Burrows (2001) found that Rhizophora stylosa lost only 2-
3% of its leaves in discrete samples, but lost 7-11% of its leaves in long-term
samples. Avicennia marina had 5-6% leaf loss in discrete samples but 28-36% loss
in long-term samples. For both species, most herbivory occurred whilst the leaves
were young, and loss of these leaves explained most of the differences between the
discrete and long-term results. For R. stylosa, once past the juvenile phase, leaves
had a very high chance of reaching senescence. In contrast, for A. marina, there was
also significant damage and leaf loss on mature leaves by a different suite of
herbivores.
In Rhizophora mangle stands on cays in Belize, wood-boring moths arrl
beetles, particularly Elaphidion mimeticum (Coleoptera: Cerambycidae), excavate
tunnels in mangrove twigs which, in turn, provide habitat for more than 70 species
of ants, spiders, mites, moths, cockroaches, termites and scorpions (Feller arrl
Mathis 1997, FeUer and McKee 1999). When loss ofleaves due to herbivory (3-8%)
is combined with the subsequent loss of leaves due to wood-boring (3-8%), a total
leaf loss of around 6-16% was recorded (Feller and Mathis 1997).
Over 100 insect species were found on small mangrove islands in Florida Bay
(Wilson and Simberloff 1969, Simberloff and Wilson 1969, 1970). The insect
population re-established quickly after fumigation of these islands with methyl
bromide, and reached pre-fumigation levels within one year. After that, total species
numbers remained more or less constant, although there was considerable species
turnover from year to year.
The size and diversity of the insect populations suggest that insect grazing is a
significant factor in the structure and function of mangrove communities. Onuf et al.
(1977) tested this experimentally by comparing two mangrove islands in Florida:
one island (high-nutrient area) had breeding colonies of pelicans and egrets, and the
other (low-nutrient area) did not. The effect of the input of nutrients by birds on
plant growth and reproduction has been noted already (see 4.5.3). More striking,
however, was the significant stimulation of herbivory by insects in response to
nutrient enrichment. Larvae of five lepidopteran species that fed on leaves and/or
Biotic Interactions and Mangrove Performance 171

buds were either more abundant in the high-nutrient area, or were present in that area
only. Similarly, the mangrove borer (Poecilips rhizophorae) that infested propagules
before they dropped from the parent tree, was more abundant in the high-nutrient
area. This resulted in a fourfold greater loss to herbivores, and more than offset the
increased leaf production owing to high nutrient input.
The observed difference in grazing in the two areas disappeared when the birds
seasonally migrated from their nesting area at the high-nutrient island. This
relationship between birds, nutrient enrichment, and insect damage, illustrates the
complex interactions that occur in mangroves, as in other vegetation types.
As Onuf et al. (1977) were working in overwash forests consisting only of
Rhizophora mangle. further investigation with multi specific mangroves was clearly
needed; such a study was provided by Lacerda et al. (1986b) who were able to
examine herbivory in mangroves of Sepetiba Bay (control site) and Guanabara Bay
(nutrient-enriched site) in Brazil. They found no consistent differences between the
sites to indicate any increase in herbivorous intensity with pollution load. They did,
however, find significant differences in herbivory between species; Avicennia
schaueriana showed significantly lower values for leaf areas consumed and for the
number of leaves attacked at both sites compared with Rhizophora mangle am
Laguncularia racemosa, which had similar values.
When leaf composition between sites and species were examined (Lacerda et al.
1986b), only Na and CI varied between sites in some species. All other parameters,
including total N content, showed no significant differences between sites. This
finding clearly differs from that of Onuf et al. (1977), who found that total N content
of the leaves was increased at the high-nutrient site. Given the role of denitrification
in removing excess N from mangrove systems (Nedwell 1975, Rivera-Monroy et al.
1995, Rivera-Monroy and Twilley 1996), a direct relationship rarely exists between
leaf N content and organic pollution loads.
Similarly, Johnstone (1981) found that the quantity of leaf material consumed
could not be correlated with the following parameters: species richness, distance of
the plant from the upper tidal limit, plant density, season, energy value of the
leaves, the presence of 'protective' ants, and the chloride and nitrogen content of
leaves. Feller (1995, 1996) also found no general increase in herbivory with N-, P-,
or NPK-fertilization of Rhizophora mangle in Belize. However, two specific
instances of herbivory did increase with P-enrichment: Ecdytolopha. an endophytic
insect that feeds in apical buds, and Mannara, a stem borer, both increased
significantly (Feller 1995).
Stowe (1995) investigated intracrown distribution of herbivore damage in
riverine Laguncularia racemosa in Costa Rica. finding that exposed foliage incurred
significantly more damage than tidally submerged foliage (mean % total leaf area
damaged 9.5±1.l % and 1.9±O.5% respectively) and that a small but significant
change in herbivory guilds resulted. Clearly, tidal submersion of foliage affects the
interaction between L. racemosa and its herbivores by concentrating herbivory in the
upper portions of the canopy and by differentially affecting the members of the
herbivory community.
In an attempt to determine what, if any, role is played by various leaf
characteristics on insect grazing, Lacerda et al. (1986b) correlated mean percentages
of total leaf area consumed in each species with the mean leaf concentration of water,
CI, Na, N, phenols, ash, soluble carbohydrates, and crude fibre. Chloride, Na am
total N concentrations were not significantly correlated with herbivore intensity. The
172 Mangrove Ecology, Silviculture and Conservation

contents of water, ash and crude fibre were significantly negatively correlated with
herbivore intensity. Soluble carbohydrates were strongly positively correlated with
herbivore intensity; herbivory was also higber in plants with high phenol content,
but the correlation was not significant.
Leaf toughness, measured as the ratio of protein to fibre, appears to reduce
palatability and digestibility (Choong et al. 1992) while Kathiresan (1992) suggested
that tannin content is related to herbivory level, with low tannin species such as
Avicennia suffering greater herbivore damage than high tannin species such as
Rhizophora, a finding apparently opposed to that ofLacerda et al. (1986b).
These data seem to indicate that the patterns of insect grazing in mangroves
may differ from those in terrestrial communities. For example, the negative
correlations with total N and water contents of the leaves in mangroves differs from
what generally occurs in most terrestrial communities (Coley 1983); however, a
recent review of temperate vegetation suggests that while N fertilization has a
beneficial effect on the individual herbivore, at the population level it is often
insignificant or even negative, except for sucking insects such as aphids (Kyto et al.
1996).
Lee (1991) showed that Kandelia candel was able to regulate resource allocation
by differential abscission of damaged leaves and that the life span reduction was
directly proportional to the leaf area loss; leaf life span reduction occurred at an
approximate rate of one day per percent of leaf area removed. It seems that, like the
premature abscission of predated propagules, abscission of damaged leaves may be a
means of defence or reduction of wastage in response to herbivore attack. However,
as <5% of leaves suffered >40% area loss under natural grazing pressure, and natural
leaf longevity was 11.5 months, insect herbivory probably plays a minor role in the
regulation of leaf dynamics of this species. Furthermore, as leaf litter accounted for
53.5% of the total litter fall in Hong Kong (Lee 1989b), and annual net above-
ground primary production of K candel averaged between 19.5-24.4 t dry wt. ha- I y-I
(Lee 1990), insect herbivory probably removes only 2.8-3.5% of the annual net
above-ground primary production of this mangrove community (Lee 1991), a figure
remarkably similar to the 2.1 % of annual mangrove leaf production removed in
north-eastern Australia (Robertson and Duke 1987a).
Burrows (2001) similarly found that for Avicennia marina and Rhizophora
stylosa, loss of leaf material through premature abscission of damaged leaves was as
great as that actually consumed by insects, indicating a role for herbivory in
promoting aseasonalleaf fall. Herbivore damage on early-falling leaves was much
greater than that on leaves that had reached senescence. In R. stylosa and A. marina,
4-5% and 19-29% of leaves, respectively, were either consumed or prematurely
abscissed due to insect damage.
The energy expenditure for each mangrove leaf varies with each species and
depends largely on the size (or dry weight) of the leaf (Table 5.5). The calorific value
of dried leaf material from Australian mangroves is not particularly variable, ranging
from 17.43 kJ g-I in Sonneratia caseolaris to 20.92 kJ g-I in Osbomia octodonta,
with its high oil content. Because of the variable sizes (and weights) of mangrove
leaves, the total energy expediture for each leaf (or leaflet in the case of Cynometra
iripa and Xylocarpus moluccensis) varies from around I kJ in Osbomia and
Cynometra, to around 30 kJ in Heritiera and Rhizophora mucronata, with other
species having intermediate values (Table 5.5). To put this into perspective, I g
Biotic Interactions and Mangrove Perfonnance 173

sucrose = 16.5 kJ. Clearly, each leaf (or leaflet) lost to a herbivore represents a
considerable and measurable loss of energy, as well as foregone leaf producti vity.
Table 5.5 Total calorific value of typical mangrove leaves (or leatlets) from north-eastern Australia
(unpublished data collected by the author)

Species Calorific Leaf Dry Total


value per Weight Energy
Dry Weight (g) per Leaf
(kJ g.l) (kJ)
A. ilicifolius 18.57 0.40 7.43
A. marina
E Australia 17.44 0.39 6.80
NE Australia 17.59 0.37 6.51
B. gymnorhiza 16.72 0.61 10.20
C. iripa 19.25 0.04 0.77
H. littoralis 20.02 1.71 34.23
O. octodonta 20.92 0.06 1.26
R. apiculata 17.84 1.17 20.87
R. mucronata 18.12 1.75 31.71
R. stylosa 17.75 0.70 12.43
S. alba 17.54 0.35 6.14
S. caseolaris 17.43 0.49 8.54
X. moluccensis 17.98 0.37 6.65

Whether any mangroves have developed an effective, specialized defence


mechanism against grazing is uncertain. Lacerda et al. (1986b) and McKee (l995a)
have suggested correlations between grazing, and chemical constituents of leaves and
propagules. Many mangrove leaves are extremely tough, and contain high, but
variable, amounts of tannins (Choong et al. 1992, Kathiresan 1992, Basak et al.
1998), features that may reduce their palatability and discourage grazing. Schoener
(1988), on the other hand, has suggested that pubescence in Conocarpus erectus is a
deterrent against herbivory, based on a study of this species on islands in the
Bahamas. This mangrove occurs as a glabrous ('green') or densely pubescent ('silver')
form on these islands and the silver-leafed form was significantly less damaged by
herbivores. Additionally, however, some of the islands have arboreal lizards of the
genus Ano/us, which prey on herbivorous insects, particularly lepidopteran larvae.
By comparing grazing on green and silver leaves on both 'lizard' and 'non-lizard'
islands (Table 5.6), an interesting interaction between insect herbivory, leaf
pubescence and lizards was observed. Clearly, lizards reduce grazing of 'green' and
'silver' leaves, but in the absence of lizards, grazing of 'green' leaves increases
considerably more than that of 'silver' leaves.
Table 5.6 Mean leaf area (%) lost to insect herbivory on islands in the Bahamas in relation to leaf
pubescence and the occurrence of Anolus lizards. (Data from Schoener 1988)

Island type Leaf pubescence


Green Silver
'lizard' 7.1 5.3
'non-lizard' 12.2 8.4
Significance p=O.OO2 p=O.OO6

For plants in general, herbivory poses a dilemma in terms of allocating


resources into further growth or into herbivory defence. Two types of defensive
strategy have been recognized, immobile and mobile defences (Coley 1988). Mobile
defences involve the production of toxic, inhibitory or repellant compounds that
reduce herbivory, and the optimal defensive investment is inversely proportional to
174 Mangrove Ecology, Silviculture and Conservation

the inherent growth rate of a plant (Basey and Jenkins 1993). Immobile defences are
those of a structural nature, including thoms, hairs, scales, lignin, tannins, thick
bark and high wood density, which hinder or discourage herbivory (Coley 1988,
Basey and Jenkins 1993). The optimal defensive investment for immobile defences
is low for plants with high growth rates, and high for plants with low growth rates,
with no intermediate level of optimal defence (Basey and Jenkins 1993, Loehle
1996). Thus, Loehle (1996) demonstrated two distinct groupings in temperate trees:
a short-lived, fast-growing group with low investment in immobile defences such as
tannins and wood density, and a longer-lived, slower-growing group with a high
investment in immobile defences.
Coley (1988) has also shown that in resource (light) limited lowland rainforests
of Panama, species with long-lived leaves had significantly higher concentrations of
immobile defences such as tannins and lignins. Similar trends have been shown in
other resource-limited habitats (e.g. nutrient-poor sandstone flora; Wright and
Cannon 2(01).
Given the generally low growth rates of mangroves and their variable but
relatively long leaf lifespans (Table 3.1), a considerable investment in mobile
defences or a high investment in immobile defences could be expected if mangroves
conform to the general patterns of plant responses. However, as a group, mangroves
are not remarkable in terms of immobile defences. Some mangroves have spines
(Acanthus), hairs (Conocarpus, Avicennia) and scales (Heritiera, Camptostemon),
and many have abundant lignin and tannins, but their investment in wood density or
bark thickness does not appear exceptional as evidenced by the high incidence of
wood-boring in mangroves (Baksha t 996, Feller and Mathis 1997, Feller and McKee
1999). It seems, therefore, that mangroves have either invested in mobile defences,
or have traded-off a level of herbivory for the potential gains in nutrient input or
turnover associated with it.
Indeed, apart from the tannins and phenolics, many exotic chemicals are known
from mangroves which may deter herbivory (also see 5.1.2). Thus, leaves of
Acanthus ilicifolius contain saponins (Minocha and Tiwari 1981) which may render
them unpalatable. Clerodendrum inerme contains neo-clerodane diterpenes which are
responsible for growth inhibition and anti-feeding activity in insects, a steroidal
glycoside (Atta-Ur-Rehman et al. 1997), and three iridoid biglycosides which have
pesticidal properties (<;alis et at. 1994a, b). Several species of Melaleuca occurring
near swampy margins of mangroves contain oils (Jones and Harvey 1936), of which
nerolidol has been demonstrated to have anti-feeding properties effective against
larvae of the gypsy moth, Lymantia dispar (Doskotch et al. 1980). Modified
terpenes, limonoids. have been identified from various mangroves (Mulholland and
Taylor 1992, Alvi et al. 1994) and shown to have a marked insect antifeeding
activity (Champagne et al. 1992). Leaf extracts from Rhizophora apiculata have a
mosquito larvicidal activity, although the specific chemicals involved have not been
identified to date (Thangam and Kathiresan 1994). The mangrove E:xcoecaria
agallocha contains a milky latex which possibly discourages grazing, either by its
taste or by its toxic properties.
From the discussion above, it seems clear that further study into the resource
allocation by mangroves into herbivore defences, particularly mobile ones, is needed.
For example, McKee (t 995c) found that high or low resource availability (light and
nutrients) had little effect on concentrations of total phenolics, gallotannins and
condensed tannins in Rhizophora mangle and LaguncuLaria racemosa seedlings. Only
Biotic Interactions and Mangrove Performance 175

high light caused an increase in total phenolics in Avicennia germinans which, in


any case, had low levels of all of these chemicals. She also found that relative
growth rates of the three species were not inversely correlated with the amounts of
secondary chemicals in the seedlings. Perhaps compounds other than phenols and
tannins comprise the chemical arsenal (mobile defences) of the mangroves and, as
McKee (1995c) suggested, the major role of tannins and phenolics may be linked to
protecting the photosynthetic system against photo-oxidation rather than to defend
against herbivory. Perhaps, too, the high, multifaceted stress levels of the 'mangrove
envrionment' fundamentally alters the usual plant growth-defence responses of
mangroves, in ways that only further studies are likely to reveal.
Large nesting or roosting aggregations in mangroves (fig. 5.10) such as those
of birds and bats undoubtedly have some effects similar to insect herbivory by the
physical damage such aggregations cause to the mangrove foliage. In addition, their
faeces may enhance the local nutrient status, thereby simultaneously affecting the
general level of herbivory. A tradeoff may be involved where large aggregations of
nesting birds or bats occur. Nesting birds may use twigs pruned from the top of the
mangrove canopy to build massive nests (for example, white ibis, Threskiomis
molucca in northern Australia, or the black-headed ibis, T. melanocephalus in
Indonesia), and they commonly peck at and damage the young growing tips within
reach while attending their young. In return, their droppings provide nutrient inputs
- even though the threat of herbivory may be increased.

Fig. 5.10 Brown pelicans and darters roost in Avicennwgerminans in the Sian Ka'an
Biosphere Reserve, Quintana Roo, Mexico. Whatever nutrient enrichment this may
cause, it does not offset the physical damage caused to young branch tips.

Except for the early studies of Onuf et a1. (1977), who were primarily interested
in the relationship between nutrient enhancement and degree of herbivory, the
broader effects of large nesting and roosting aggregations have not been examined in
detail.
176 Mangrove Ecology, Silviculture and Conservation

5.2.4 Other Mutualistic Interactions


One case of mutualism between mangrove epiphytes and ants illustrates a close
dependency between the partners, reflecting a long co-evolutionary relationship
(Brouat and McKey 2(00). These plants are termed ant-fed plants or 'myrmecophytes'
because in all instances, ants live within caulinary domatia (hollow and usually
swollen stems), which appear to be specialized for harbouring ants and absorbing
nutrients from the ants' piles of debris (Thompson 1981, Huxley and Cutter 1991).
In the tropics around the world, myrmecophytes occur in at least 141 plant genera
from 47 families, including several mangrove epiphytes (fig. 5.11). The number of
ant species nesting exclusively inside myrmecophytes is unknown but
representatives occur in 29 genera from five subfamilies (Davidson and McKey
1993).
The ant-plant relationship has been established in several mangrove epiphytes,
including Dischidia (Janzen 1974, Rintz 1980), Hoya (Merrill 1945), Hydnophytum
(Huxley 1978, Rickson 1979) and Myrmecodia (Janzen 1974, Huxley 1978). It has
been suggested that the relationship evolved in response to a nutrient-poor (low
nitrogen) habitat, and this has been demonstrated experimentally in Hydnophytum
formicarium (Janzen 1974, Rickson 1979) and Myrmecodia tuberosa (Huxley 1978).
Myrmecodia beccarii is common in the more northerly mangroves of Australia
and a plant may weigh up to several kilograms. Their swollen bulb-like stems are
honeycombed with tunnel-like galleries in which an ant species, Iridomymex
cordatus, and the larvae of a butterfly, Hypochrysops apollo, live in an apparently
symbiotic relationship.

Fig. 5.11 Myrmecodia beccarii. a common epiphyte of the mangroves of south-east


Asia and northern Australia. is commonly inhabitated by ants of the genera
lridomymex or Pheidole.

In Hydnophytum formicarium, two types of cavities occur in the swollen base:


rough-walled, water-absorbing chambers, and smooth-walled non-absorbing cavities.
The ant Iridomyrmex cordatus places its debris in the rough-walled cavities instead of
tossing it out of the nest (Janzen 1974). Rickson (1979) showed that when
Biotic Interactions and Mangrove Performance 177

radioactively-labelled Drosophila larvae were carried into the plants by the ants,
radioactive compounds moved into the plant and were translocated up the stem.
Huxley (1978) showed that when ants were given food tagged with radioactive
tracers, the tracers were preferentially deposited in the rough-walled cavities of
Myrmecodia, most probably through defaecation. She also found that survival of
Myrmecodia seedlings in the field was significantly higher when the plants were
associated with l. cordatus than when the plants either had no ants or were associated
with ant species other than I. cordatus.
Thompson (1981) argued that the phenomenon of ant-fed plants is associated
with the following ecological conditions: nutrient-poor environments, epiphytic
habit generally on open-canopy trees, and low moisture availability. These
conditions are typical of those prevailing in the tops of mangrove canopies, arxl
possibly explain the disproportionately large number of ant-fed species among
mangrove epiphytes.
Other ant-plant interactions include the relationship between various mangroves
and leaf-weaver ants. In mangroves of the old world, leaf-weaver ants are common,
weaving leaves into complex nests within the mangrove canopy. The green leaf-
weaver ant, Oecophylla smaragdina (Formicidae: Formicinae), is common from
India, throughout most of tropical Asia, to the Solomon Islands and northern
Australia (Lokkers 1986). The African weaver ant Oecophylla longinoda occurs
throughout the forested areas of tropical Africa. Although neither species is confined
to mangroves, they are abundant in mangrove forests, with territories approximately
200-800 m 2 comprising 12-44 mangrove trees (Holldobler 1983). The nests are
constructed directly from leaves still attached to the host plant, bound together with
silken threads spun by the last instar larvae, which the workers hold between their
mandibles and move back and forth like spinning shuttles (Wilson and Holldobler
1980, Holldobler 1983).
Multiple brood-nests are maintained within the territories and a system of
peripheral guard-nests defines the territorial boundaries (Holldobler 1983). Such nest-
decentralization optimises patrolling and cropping within the territory while
economizing the defence of the territory.
Both O. smaragdina and O. longinoda react aggressively both to con specific
aliens and to many other ant species, calling up reinforcements, if needed, by an
elaborate communication system (Holldobler 1983, Holldobler and Wilson 1990).
Thus, these ants may provide some protection to the mangroves by preying on
herbivorous insects within their territories. In his study of leaf herbivory in Papua
New Guinea, Johnstone (1981) however, found no reduction in insect leaf herbivory
in mangroves with or without leaf-weaver ants. On the negative side, for mangroves
of the family Rhizophoraceae which have strictly terminal growing points, these ant
nests effectively inhibit the further development of the affected shoot, and thus can
impair the full development of the tree. Clearly, some tradeoffs are involved; for
possible protection against herbivorous insects, some structural damage is
acceptable.
In a striking case of convergent evolution, an unrelated ant, Azteca trigona
(Formicidae: Dolichorderinae) establishes and defends non-overlapping territories in
neotropical mangroves (Collins et al. 1977, Adams 1994). Based on his studies at
Coco Solo on the Atlantic coast of Panama, Adams (1994) found that each colony
occupies several interconnected trees, constructing nests out of fibrous material in
each tree, defending them against other ants (conspecific or otherwise) and other
178 Mangrove Ecology, Silviculture and Conservation

insects. Tomlinson (1986) suggested that the association between Azteca arxl
Pelliciera, in particular, may be mutualistic as the ants obtain some nutrients from
the pair of extrafloral nectaries at the base of each leaf blade, as well as from the
insects they attack. Like their old world counterparts, when their territories are
threatened, this species uses pheromonal and tacticle communications to call up
reinforcements to attack, immobilize and kill intruders.
Interestingly, ants of the genera Azteca and Oecophylla both tend scale insects,
prey on living insects, scavenge opportunistically, and have mUltiple nests
throughout their territories which are defended vigorously against any intruder
(Holldobler and Wilson 1978, 1990, Adams 1994). The effect of such exclusive
foraging territories being maintained by these ants could be expected to be a benefit
to the mangroves through some reduction in insect herbivory. The only study to
examine this aspect (Johnstone 1981), however, found no evidence to support such a
protective effect.
Ants often tend scale insects, consuming their exudates as an important source
of nutrition and, in tum, protecting them against predators and parasites. However,
Ozaki et al. (2000) report an interesting example where ants predate scale insects arxl
thereby protect mangrove plants. In a planted monoculture of Rhizophora mucronata
on Bali, the scale insect Aulacaspis marina (Homoptera: Diaspididae) occurred at
high densities (>200 mature females leafl), causing leaf discoloration and premature
leaf fall, and ultimately causing the death of the saplings. In contrast to its high
population density in the plantations, however, this scale insect was scarce in
natural mangrove forests adjacent to the plantations. Manipulative experiments
established that in the natural mangrove forests, the ants Monomorium floricola
(Myrmicinae) and Paratrechina sp. (Formicinae) were feeding on the scale insects to
the extent that only very small populations remained. In the plantations, on the
other hand, the survival of Aulacaspis marina was six times higher than in the
natural forests due to a lower level of ant predation caused by a lack of ant nesting
sites due to the young age of the saplings.
Another interesting mutualistic interaction is that of grazing by ocypodid arxl
grapsid crabs of macroalgae and bark from pneumatophores (Wada and Wowor 1988,
1990). These investigators found significant grazing of macroalgae and bark from the
pneumatophores of Sonneratia alba by seven species of ocypodid crabs in Indonesia.
The proportion of pneumatophore-foraging individuals among active crabs was
highest in Macrophthalmus quadratus. The timetable of pneumatophore foraging by
individuals of M. quadratus exhibited no clear tendency in relation to low tide. When
the bark of pneumatophores was scraped off together with growing macroalgae, M.
quadratus foraged less frequently on it, compared with intact pneumatophores. When
comparing M. quadratus foraging on pneumatophores with those foraging on the
substratum, the former had bark of pneumatophores and macroalgae in their
stomachs more frequently and in larger quantities than the latter. These findings
suggest that bark and macroalgae are important food items of M. quadratus in
pneumatophore foraging.
Wada and Wowor (1988, 1990) concluded that this activity is beneficial to the
mangrove trees as wen as to the grazers. As pneumatophores are capable of
photosynthesizing via the chlorophyll layer beneath the cuticle, and as the oxygen
produced by this process is transported to the root system (Yabuki et al. 1985,
Dromgoole 1988, Kitaya et al. 200 I), keeping the photosynthetic surface of
Biotic Interactions and Mangrove Performance 179

pneumatophores free of macroalalgal growth as well as bark, enhances the amount of


photosynthetic oxygen produced and available to the root system.
Koch and Wolff (1996) have described a similar example from Costa Rica
where the mangrove snail, Thais kiosquifonnis. removes barnacles from the
mangrove root systems, fulfilling a mutually-beneficial 'cleaning' function.
An interesting case of aquatic facultative mutualism between mangrove trees
and invertebrates was reported by Ellison et a!. (1996) from Belize. The proproots of
Rhizophora mangle support a sponge fouling community, and it appears that
sharing of nitrogen and carbon resources between the trees and the fouling sponges
results in increased growth and production of both partners.

5.2.5 Trophic Pathways


Mangrove forests in coastal deltas, estuaries and embayments provide an invaluable
habitat resource for numerous organisms, including those supporting important
sustainable fisheries. This biotic interaction between mangroves and nearshore
fisheries has often been invoked as grounds for mangrove conservation and
sustainable use.
Early studies in the Caribbean suggested that mangroves provide the primary
carbon source fuelling detrital-based foodwebs, and therefore, that they represent the
major route of energy flow in such systems (Heald 1971, Odum 1971, Odum and
Heald 1972, 1975a, b, Twilley 1995). In this model of organic matter production by
mangroves, physical forcing functions such as solar and tidal energy drove the
dynamics of organic matter production and export. The biotic processes within the
system such as herbivory or other forms of in situ processing of the organic material
were ignored.
This model of organic matter production and fate, termed the 'Florida' model by
Lee (1999), provided the theoretical framework for mangrove trophic structure for
nearly two decades until it was challenged by mangrove litter dynamics in north-
eastern Australia, and by the use of stable isotope analysis over traditional gut
content analysis in determining real pathways of energy flow in mangrove
ecosystems.
Boto and Bunt (1981) estimated litter export from a north-eastern Australian
mangrove system from mass balance calculations based on the difference between the
rate of litter production and the amount remaining on the forest floor after high tide.
Any missing biomass was, according to the 'Florida' model, assumed to have been
exported by tidal flow. Later investigations on the same system demonstrated that
biotic processes within the system accounted for a significant fraction of the missing
biomass (Boto 1982, Boto et a!. 1989, Robertson et al. 1991). Some of the missing
biomass had been reprocessed by sesarmid crabs, which Robertson (1986) had shown
could remove around 28% of the litter produced in mixed Rhizophora forests.
Similar findings were subsequently reported from mangroves in south-east Asia (Lee
1989a), south-eastern Africa (Micheli et a!. 1991, Emmerson and McGwynne 1992,
Steinke et al. 1993), and in other Australian mangroves (Micheli 1993a, b). Recent
studies in Ecuador also demonstrated that crabs consume considerable quantities of
mangrove leaf litter even in high tidal range neotropica1 systems (Twilley et a!.
1997).
Amphipods also playa significant role in litter reprocessing. Large populations
(up to 7,000 m- 2) of the amphipod Parhyale hawaiensis occur in the mangroves of
north-eastern Australia (Poovachiranon et al. 1986). Using measurements of faecal
180 Mangrove Ecology, Silviculture and Conservation

production as an index of feeding rate, these authors showed that this amphipod
could consume large quantities of decomposing Rhizophora stylosa leaves, with
maximal faecal production in the order of 1,700 mg dry weight g-I amphipod d-t,
depending on favourable salinity and food source conditions. Feeding rates were not
significantly different over the salinity range 15-350/00, although significant decreases
were noted at further extremes, i.e. 10 and 40%0.
The quality of organic matter is also enhanced by biotic reprocessing within
mangroves, particularly by sesarmine crabs (Nielson and Richards 1989, Lee 1997).
Sesarmine crab faecal pellets contain finely fragmented mangrove material, low in
deterrent chemicals such as tannins, and readily colonized by microbes, comprising
the basis of a coprophagous food chain both in the mangrove benthos and among
such pelagic consumers as copepods (Lee 1999).
As discussed in 5.2.3, insect herbivory is widespread in mangroves, although
accounting for only around 2-3.5% of the annual net above-ground primary
production (Lee 1991). Nevertheless, when taken together with the activities of crabs
and amphipods, it seems clear that a considerable amount of mangrove primary
production is diverted into internal, herbivorous pathways rather than the external,
detrital pathways suggested by the Florida model. Thus, biotic reprocessing within
mangroves may be as important, if not more important, than the export of leaf
material from mangrove systems, at least on a regional basis (Lee 1999). When
internal reprocessing is taken into account, as for example in studies of organic
export from the Fly River delta in Papua New Guinea (Table 5.7), organic export
remains sizeable but considerably less than the total production.
Table 5.7 Areas covered by major forest types in the Ay River delta. Papua New Guinea. and mean
estimates of forest productivity. together with daily production and export. taking internal reprocessing
into account. (Data from Robertson et al. 1991)
Forest type Area Productivity Daily Daily
(ha) kg C ha- I d" Production Export
tCd- ' tCd- 1
Rhizophora- 31,500 26.7 841 394
8ruguiera
Nypa 38,400 27.1 1041 208
Avicennia-
Sonneratia 17,500 19.0 333 76
Total 87,400 2,214 678

According to the Florida model (Odum and Heald 1975a, b), the principal
energy pathway was considered to be from mangrove leaf detritus to bacteria arxl
fungi and then to detritivores which, in tum, were consumed by carnivores. This
pathway was established by gut content analysis which had revealed that more than
20% of material contained in the digestive tracts of all herbivores and omnivores in
the estuarine mangrove system of Florida contained mangrove detritus. However,
simple ingestion of detritus does not necessarily imply any direct assimilation
(energy transfer) ofthat material.
Initially, stable carbon isotope ratios (o\3C) were used to study assimilation,
although later work has additionally used stable nitrogen isotope ratios (OI5N). As
mangroves are depleted in both of these isotopes (because of heterotrophic activity),
the ratios of both stable isotopes become more negative in mangroves and in
organisms feeding on, and assimilating, mangrove organic matter. Thus, Rodelli et
al. (1984) recorded mangrove detritus in the guts of offshore marine consumers in
Biotic Interactions and Mangrove Perfonnance 181

Malaysia. although they could find no isotopic evidence for any mangrove
assimilation.
Similarly. France (1998). using both isotopes. concluded that the fiddler crab
Uca vocator from a mangrove-fringed. land-locked lagoon in Puerto Rico
differentially assimilated material from ingested sediments. Benthic microalgae were
preferentially assimilated over mangrove detritus. indicating that the presence of
mangrove detritus in the gut does not constitute evidence of an energy pathway.
Similarly. penaeid prawns have been assumed to benefit from mangrove detritus.
However. Stoner and Zimmennan (1988) showed that Penaeus notialis, P. subtilis
and P. brasiliensis in a mangrove-fringed lagoon in Puerto Rico contained less than
25% detrital material in their diets. the bulk of which consisted of capitellid
polychaetes and amphiphods. In a detailed study using three isotope ratios (o 13C.
0'5N and 034S). Loneragan et al. (1997) studied the amounts of carbon from
mangroves. seagrasses and macroalgaelseston incorporated into penaeid prawns in
northern Australia. particularly those species commonly associated with mangrove
habitats (e.g. Penaeus merguiensis). Although the carbon source in prawns depended
a little on their location within the estuary. it was concluded that the contribution of
mangroves to the prawn food web was insignificant.
Clearly. it seems that in situ degradation and recycling within mangroves is
sizeable while export of mangrove material. either as macro-detritus (leaves arxi
wood) or as micro-detritus (particulate organic matter) is highly variable (Robertson
and Daniel 1989a. Lee 1995). depending on the hydrodynamics of the estuaries arxi
the mangrove-nearshore environment (Wolanski et al. 1992. Stieglitz and Ridd
2001). Within estuaries. the axial convergence will tend to move macro-detritus
upstream (Stieglitz and Ridd 2(01). while the stable coastal boundary layer water
will reduce the extent of any outwelling from mangroves to offshore areas (Wolanski
et al. 1992).
Nevertheless. most studies conducted in macrotidal mangrove communities
suggest that mangroves do serve as net exporters of organic carbon and nutrients. but
the amounts involved are highly variable in space and time (Lee 1995. Dittmar arxi
Lara 2(01). However. there are also reports of mangroves acting as net importers of
organic carbon and nutrients. storing large proportions of litter production for in situ
decomposition. mainly as a result of a restricted inundation regime (Twilley et al.
1986. Flores-Verdugo et al. 1987. Lee 1990. Simpson et al. 1997. Alongi et al.
1998. Ayukai et al 1998).
In a recent study of a mangrove system in Braganlta. northern Brazil. Dittmar
and Lara (2001) investigated the driving forces behind the coastal outwelling of
nutrients and organic matter from mangroves. They concluded that the tidal range arxi
porewater concentrations were the major driving forces behind coastal outwelling.
They suggested. on the basis of equilibrated fluxes. that outwelling probably occurs
only from mangroves where the nutrient concentration in porewater exceeds the
demands of the benthic community and trees. a phenomenon caused by sediment
accumulation rates or by high N-fixation rates. Furthennore. outwelling only occurs
in macrotidal regions where porewater can flow in considerable amounts to tidal
creeks and the ocean (Dittmar and Lara 200 1).
The fate of this outwelling of organic matter and nutrients to mangrove creeks
and nearshore areas is largely unknown (Alongi 1998). While mangrove creeks arxi
nearshore areas are known to act as nursery areas for juvenile fish (Thayer et al.
1987. Robertson and Duke 1987b. 199030 b. Laegdsgaard and Johnson 1995.
182 Mangrove Ecology, Silviculture and Conservation

Kimani et al. 1996, Halliday and Young 1996, Ley et al. 1999), spiny lobsters
(Acosta and Butler 1997), and prawns (Staples 1980a, b, Vance et al. 1990,
Primavera 1998, de Graaf and Xuan 1998), this function is largely dependent on the
physical habitat conditions, including such features as the highly complex structure
of the habitat offering shelter and an array of osmotic refuges. The direct trophic
subsidy to this nursery function of mangrove creeks and nearshore areas from
mangrove detritus is minimal at best (Loneragan et al. 1997). Nevertheless, a
considerable body of data has been collated to suggest that prawn and fish catches can
be correlated with the available areas of intertidal vegetation, particularly mangroves.
Thus, Pauly and Ingles (1999) have reviewed the prawn catches from 38 areas around
the world and have found a highly significant relationship as follows:

Log1oMSY =0.874 + 0.48410g wintertidal vegetation - 0.021latitude


~ = 0.53 p<O.O I

where MSY is the mean shrimp (prawn) yield in t x 103, intertidal vegetation is in
km2 x lW, and latitude in ON or oS.
Similar relationships have been found in relation to finfish catches (Blaher et
al. 1989, Chong et al. 1990). However, despite these statistically significant
relationships, the assumption of direct subsidy via mangrove detritus, the central
tenet of the Florida model, remains unconvincing while the role of the structured
habitat provided by mangroves appears to be increasingly significant.
6. Mangrove Structure and Classification

The mangrove spreads widely.


its roots looping. braru:hing. grappling. entwining.
revelling in the sludge;
from its braru:hes it drops aerial roots
into the static convulsions below.
And over all is a dense roof of glossy leafage
which effectively excludes sunlight
from the stinking. impenetrable chaos beneath.
One can do nothing with mangroves but avoid them.
Edgar Beale (1977:172)

6.1 Classification of Mangrove Communities


The physico-chemical and biotic interactions discussed in the previous chapters
produce a range of mangrove communities, which differ in their function and
structure. These functional and structural attributes can be used to classify mangrove
communities, a process which, through its data reduction, can provide an overview
of the types of mangrove communities and how dominant interactions shape them.
Any of the various attributes (physico-chemical, functional, structural or
floristic) of mangrove communities may be useful in classifying them, and the
appropriate selection of attributes depends upon the purpose of the classification.
The classificatory schemes discussed here, however, appear to be of some universal
value in comparing mangrove communities over a range of scales.

6.1.1 Phytosociological Classification


The consistent co-occurrence (and mutual absences) of particular pairs of plant
species, together with their dominance (cover abundance), were used by the Ziirich-
Montepellier (sometimes referred to as the Braun-Blanquet) School of European
ecologists to establish a system of discrete plant communities (Braun-Blanquet
1964). Each of the communities (now termed syntaxonomical units) recognized were
given names, governed by rules contained in the 'International Code of
Phytosociological Nomenclature' (3rd Edition) (Weber et at. 2(00). The scientific
names of the dominant species are used and under Articles II of that Code, the
syntaxa have standardized endings, depending on their rank (Table 6.1).
Table 6.1: Codified tenninations for variously ranked syntaxa according to the International Code of
Phytosociological Nomenclature.

Rank Tennination
Association -etum
Alliance -ion
Order -etalia
Class -etea
Subassociation -etosum
Suballiance -enion
Suborder -enalia
Subclass -enea

While such community labels are very convenient and likely to facilitate a
degree of standardization, there can be no doubt that discrete, consistent and
persistent plant communities are a fiction (Bastow and Chiarucci 2(00).
184 Mangrove Ecology, Silviculture and Conservation

In his salt marsh studies, Chapman (1960) used a phytosociological


classification; consequently, it is not surprising that he applied this approach to
mangroves (Chapman 1970). He identified and named 8 Alliances, 15 Orders and 40
Associations, a number that almost matched the number of mangrove species (55)
that he recognized. With only minor modifications, Chapman (1975, 1976, 1977)
used this classificatory scheme and attempted to link it to successional sequences.
From time to time, this approach has been used by others in salt marsh aOO
mangrove studies, particularly when comparing reasonably distinct forest
community associations at broad regional scales (Bridgewater 1975, 1985, 1989,
Suzuki and Mochida 1982, Miyawaki et al. 1983, Suzuki and Saenger 1996).
Although the phytosociological approach has never been widely used outside
Europe, and sometimes bemused American ecologists (Snedaker and Brown 1982), it
nevertheless survives. Thus, as recently as 2001, Bouzulle et al. used this approach
to describe syntaxonomical units colonizing abandoned salt pans in western France.

6.1.2 Classification Using Structural Attributes


Specht (1970) developed a structural classification of evergreen plant communities
that uses those properties which reflect the amount of photosynthetic tissue
(contributing to energy input) and the biomass of respiring aerial plant tissue
(involved in energy output). The properties used are (1) the height and life form of
the tallest stratum (which provides an estimate of the biomass) and (2) the 'foliage
projective cover' (FPC) of the tallest stratum. The FPC is the areal proportion of
photosynthetic tissue vertically above the landscape. Ideally, it should be measured
using some crosswire device to determine the presence or absence of foliage
vertically above a large number of randomly selected points in the community. More
often it is estimated using photographic techniques.
Using these two properties, the identification of structural formations can be
achieved. These formations (such as forest or woodland) can be defined further by
including the name ofthe dominant genus or species (such as Avicennia woodland).
The general absence of well-developed understorey and shrub strata in mangrove
communities (Janzen 1985) and the marked tendency towards dominance by one
species of canopy tree mean that it is rarely necessary to seek further precision.
Using this classification, the range of mangrove structural formations that have been
encountered are shown in Table 6.2. The most common of these are closed
communities of variable height (Blasco and Aizpuru 1997). Only rarely d>
mangroves exceed 30 metres in height. Open canopies are associated with high
salinity sites, often at or near high-water spring levels, where rainfall or run-off are
low or moderately seasonal. Open canopies also may occur where persistent
waterlogging is a feature of the environment and, in some other instances, in dwarfed
communities whose structure results from seasonally high salinities (Lin aOO
Sternberg 1992b).
The assumption is made in this classification that the communities are mature,
that is, fully reflecting the constraining effects of water balance, soil fertility,
temperature and light. In practice, this assumption is not difficult to meet as
successional response by FPC is rapid (Specht and Morgan 1981), minimizing any
error arising from this parameter; the age/size structure of the population in relation
to neighbouring sites permits a reasonable assessment of the developmental
(successional) phase of the ecosystem. This scheme can be used validly also where
Mangrove Structure and Classification 185

the community is not mature due to disturbance if that disturbance is regular. In this
instance, regular disturbance can be viewed as an integral part of the environment.
In terms of understanding the ecological biogeography of mangrove plant
communities, this classification is extremely useful. The selected parameters (FPC
and height) relate to community growth and, consequently, allow some functional
interpretations of structural variation to be made (Specht and Specht 1999). An
assessment of mangrove plant communities in these terms conveys clues about the
environment and will enable some approximate predictions to be made about the
direction of any change in the community following environmental manipulations.
Table 6.2 Common structural formations of mangrove communities, classified according to the scheme
of Specht (1970).

Life form Height Foliage projective cover of tallest stratum


of tallest of tallest Dense Mid dense Sparse
stratum* stratum 70-100% 30-70% 10-30%
Trees >30m Tall-closed forest
Trees 10-30m Closed forest
Trees 5-lOm Low closed forest Low open forest Low woodland
Shrubs 2-8m Closed scrub Open scrub Tall shrubland

* A tree is defined as a woody plant >5 m taIl, usually with a single stem; a shrub is defined as a woody
plant <8 m tall, usually with many stems near the base.

This structural classification of mangroves combines easily with aerial


photograph interpretation and various other forms of remote sensing (Terchunian et
at 1986, Ibrahim and Hashim 1990, Ramsey and Jensen 1996, Blasco et al. 1998,
Ramirez-Garcia et al. 1998). For example, Ibrahim and Hashim (1990) identified
species groups in a mangrove forest from aerial photographs, delineating their
distribution and quantifying their coverage. Aerial photographs at a scale of
1:40,000, taken in 1983, were used to delineate mangrove forest from Kemaman to
Kuantan, on the east coast of Malaysia. Results indicated that three mangrove forest
types, namely the Rhizophora type, the AvicennialSonneratia type, and the mixed-
mangrove type, could be delineated with 90% accuracy. A total mangrove area of
2,214 ha was mapped, of which 2% is of the Avicennia/Sonneratia type, 24% the
Rhizophora type, and 74% the mixed-mangrove type. AvicenniaiSonneratia was
found to have the highest stand density with 13,348 trees ha· 1, followed by
Rhizophora with 6,697, and mixed-mangrove forest 1,997 trees ha- 1•
Similarly, mapping with satellite imagery can be incorporated into this
scheme, although maps based on satellite imagery are generally less accurate than
those derived from aertial photographs (Manson et al. 200 1). Nevertheless,
following field sampling near Marco Island, Florida, in October 1988 to collect
information on mangrove type, maximum canopy height, and percent canopy
closure, correlations were investigated with selected vegetation indices, derived from
SPOT multispectral (XS) data obtained at the same time (Jensen et at. 1991). The
Normalized Difference Vegetation Index (NOVI) information was the most highly
correlated index with percent canopy closure. Thus, Jensen et at. (1991) concluded
that % canopy closure information could be used as a surrogate for mangrove
density.
Other forms of remote sensing data, often in combination with geographic
information systems (GIS), have been used to describe major mangrove formations.
For example, the mangroves of the Turks and Caicos Islands were mapped using
186 Mangrove Ecology, Silviculture and Conservation

multispectral airborne imagery (Green et al. 1998), which aIlowed the mapping of
zonation patterns and the dominant species in each zone. Ramsey et al. (1998)
combined Landsat TM, colour infra-red photography, and ERS-l synthetic aperture
radar to map coastal wetlands in Florida. Pasqualini et al. (1999) combined SPOT
data with SIR-C radar data to monitor mangroves during the establishment of a large
aquaculture facility. Ramirez-Garcia et al. (1998) combined Landsat TM imagery
with historical air-photos to detect longer term changes in the mangroves around the
mouth of the Santiago River, Mexico. Long and Skewes (1996) combined Landsat
TM imagery with codified rules for a GIS to distinguish mangroves from other
coastal vegetation in the south-eastern Gulf of Carpentaria.

~ inlerl1dai Mudnal LOll « 7m) open scrub


Medium (7-1 Om) closed fore 1
D SaltDal herbland 'lannes'

Tall (-10m) closed fore 1


D.,
Rlccpond

IIIIIrn 1.011 ( 7m) cio ed !.crub


D
... uprnlldal Upland

Fig. 6.1 Littoral vegetation map of the delta of the Soumba River, Guinea, (prepared
from an analysis of SPOT image 30-330 taken on 19.4.1986), showing the
distribution of tall, medium and low mangrove communities. (From Saenger and
Bellan 1995)
Mangrove Structure and Classification 187

Because height (H) and FPC are readily ascertainable from air-photographs and
remotely sensed imagery, the use of this classificatory scheme is both informative
and convenient at a range of scales.

6.1.3 Classification Using Geomorphological Settings


Comparisons of 34 major river systems by Wright et al. (1974) in terms of
particular sets of physico-chemical variables (such as river discharge, wave energy
regimes, river-mouth morphology, and delta-plain landform suites) revealed that
deltas tend to cluster together into relatively few categories. Further generalization
(Coleman and Wright 1975, Wright 1978) resulted in the classification of a number
of general delta types, with each reflecting a particular combination of processes and
physico-chemical controls.
Thorn (1982, 1984) used this classification as the basis for a broader
classification of coastal settings in which mangroves grow. He has described five
types of terrigenous sedimentary coasts, and three carbonate (coral coast) settings,
where sediment accumulation is either from in situ growth of coral reefs or from
deposition of carbonate particulates. Woodroffe (1987, 1992) has further elaborated
this classification scheme and added a further carbonate setting.
As Galloway (1982) has shown, this scheme has considerable practical value
when used on a regional scale, as it is based on both structural and dynamic
characteristics of a particular section of coastline. For these reasons, this scheme
also may have considerable value when applied on a more local scale. The six broad
types of settings are described below.
River-dominated Alluvial Plains: This setting is characteristic of coasts with a
low tidal range and where the discharge of fresh water and sediment leads to rapid
deposition of terrigenous sands, silts and clays to form deltas. Examples of this
setting include the delta of the Ganges-Bramaputra-Mengha Rivers of India and
Bangladesh, the delta of the Orinoco River in Venezuela (fig. 5.5), the delta of the
Atrato River in Colombia, and the deltas of the Fly and Purari Rivers in Papua New
Guinea.
These deltas build seawards over flat offshore slopes composed of fine-grained
sediments. Such slopes help dampen wave energy and any tendency for longshore
drift. The delta consists of multiple branching distributaries forming elongate,
finger-like protrusions, resulting in a highly crenulate coastline with shallow bays
and lagoons between and adjacent to the distributaries.
The active distributary region is predominantly an area of high freshwater
discharge, so that salt-tolerant plants are not common. However, where abandoned
distributaries occur into which saline waters penetrate seasonally or more frequently,
salt-tolerant vegetation will develop. The area around these distributaries is also
relevant to this setting as longshore drift of muds and reworking of sands and shells
by waves influence plant establishment and regeneration, a striking phenomenon on
chenier plains. Thus, parts of the alluvial plain may contain an array of habitats
where mangroves can become established or be maintained. Such plains are highly
dynamic, and subject to rapid fluctuations in rates of subsidence, deposition, or
freshwater discharge. They are consequently characterized by a high degree of
physico-chemical diversity and rapid habitat change.
Tide-dominated Tidal Plains: This setting occurs on coasts where high tidal
ranges and associated strong bidirectional tidal currents predominate. These currents
are responsible for the dispersion of sediments brought to the coast by rivers, and in
188 Mangrove Ecology, Silviculture and Conservation

the offshore zone they form elongate sand bodies. Wave power is often low because
of frictional damping over broad intertidal shoals. The main river channels are
typically funnel-shaped and are fed by numerous tidal creeks; these creeks are often
separated by extensive tidal flats. Extensive low-gradient intertidal zones are
available for mangrove colonization. Examples of this setting include the Ord River,
Fitzroy River and King Sound in Western Australia (Semeniuk 1985a, b), and the
Klang Delta of western Malaysia.
Wave-dominated Barriers and Lagoons: This setting is characterized by much
higher wave energy than previous types, and by relatively low amounts of river
discharge. Offshore barrier islands, barrier spits and bay barriers are typical of this
setting. Small finger-like deltas prograde into these water bodies without significant
opposition from marine forces. Considerable tidal modification may occur within
the barrier system. Where the barriers project from the coast or link islands to the
mainland, sheltered water in their lee provides sites for extensive mangroves if a
sediment supply is available (fig. 6.2). Salt-tolerant plants occur around the margins
of the lagoon in a variety of habitats. Examples include the lagoonal barrier
coastlines of Mexico, EI Salvador and Brazil, the coastal lagoons of Cote d'Ivoire,
Ghana and Benin (Saenger and BeHan 1995), and the western shores of the Gulf of
Carpentaria, in the Northern Territory.

Fig. 6.2 Typical mangrove communities of the coastal lagoons of Benin. In the
background can be seen coconut palms on tbe strandline separating the lagoon from
the Gulf of Guinea. Clumps of Rhizophora racemosa line the shorelines of the lagoon.
(Photo. F. Blasco)

Composite Alluvial Plains and Barriers: This setting represents a combination


of high wave energy and high river discharge. Sand carried to the sea by the river is
rapidly redistributed by waves along shore to form extensive sand sheets. Much of
the sand deposited on the inner continental shelf during lower sealevels is reworked
landward during periods of rising or stable sealevels.
Mangrove Structure and Classification 189

Fig. 6.3 The Rio Sierpa River delta on the Pacific coast of Costa Rica. is a typical
composite of aIlivial plans and barriers. Surf-swept beaches protect an extensive
alluvial plain covered with dense mangrove forests.

The result is a coastal plain dominated by sand beach ridges and narrow
discontinuous lagoons with an alluvial plain to landward. Examples include the
Grijalva Delta in Mexico (Thorn 1967) and the Niger Delta in Nigeria (John ani
Lawson 1990). Salt-tolerant plants such as mangroves are concentrated along
abandoned distributaries and in areas near river mouths and adjacent lagoons. Where
the tidal range is large and the climate dry, there is a spread of saline habitats to
interdistributary areas which are periodically inundated by high spring tides.
Drowned Bedrock Coasts: Many large coastal embayments in the tropics ani
sUbtropics have been inundated by post-glacial rise in sealevel and are known as rias
(Semeniuk 1985a, Woodroffe 1992). This setting can be described as a drowned river
valley complex where estuarine muds gradually infill the area (Roy et al. 1980).
Neither marine nor river deposition has been sufficient to infill what is an open
estuarine system. However, the heads of the valleys may contain relatively small
river deltas which are little modified by waves, and often maintained by self-scouring
(Bunt and Wolanski 1980). At the mouth of the drowned valley bordering the open
sea, a tidal delta may occur, composed of marine mud and sand reworked landward
during rising sealevels.
Examples include the Kimberley Coast of Western Australia, Hinchinbrook
Channel (fig. 6.4) in Queensland, and many estuaries in eastern Australia.
Mangroves flourish in the fine sediments at the heads of the drowned tributary
valleys, and in lagoons behind bay barriers near the mouth of the estuary.
Carbonate Coasts: Mangroves may grow on the sparse terrestrial sediments
which have accumulated on or behind fringing reefs, or they may occur on coral
sediments (sand or limstone surfaces) on modern or fossil platform reefs where
calcareous sediment production dominates (fig. 6.5).
Thorn (1975) has described the general response of mangroves on carbonate
coasts to varying substrate and energy conditions, while Stoddart (1980) has
attempted to relate the mangrove occurrence on cays to the evolutionary stage of the
190 Mangrove Ecology, Silviculture and Conservation

reefal substrate. Buckley (1983) and Woodroffe (1987) have reviewed the
biogeography of mangroves in the northern Great Barrier Reef and the Pacific
respectively, while Saenger (1984) provided descriptions of the mangrove
communities occurring on the coral islands of the northern Great Barrier Reef.

Fig. 6.4 Hinchinbrook Channel is a good example of a drowned bedrock coastal


landscape. providing habitat for extensive and productive mangrove communities.

Fig. 6.5 Dwarfed A vicennia marina growing on a coral fringing reef at AI Tafiah. a
little south of Jeddah, on the Red Sea coast of Saudi Arabia.

6.1.4 Classification Using Physiographic and Structural Attributes


As mangrove communities exhibit a tremendous range of form, a convenient system
of classification can be based on the geomorphic and hydrological processes that
Mangrove Structure and Classification 191

induce that fonn. From their work in Florida. Lugo and Snedaker (1974) identified
six major community types based largely on their physiographic setting. They found
the correlation between community physiography and community structure to be
high. and led them to recognize. on structure alone. one community type whose
physiography was extremely variable and poorly understood. Each of the six
community types has its own characteristic set of environmental variables. such as
soil type and depth. soil salinity range. and tidal flushing rates. In addition. each
community type has characteristic ranges of primary production. litter
decomposition. and carbon export. along with differences in nutrient recycling rates
and community components. The types are described below.
Overwash mangrove forests: These occur on the smaller low islands and finger-
like projections of large land masses in shalIow bays and estuaries (fig. 6.6). Their
positions and alignments obstruct tidal flow. and thus they are overwashed frequently
by tides and much of the organic matter is washed away. In Florida. all local
mangrove species may be present but Rhizophora mangle usually dominates.
Maximum height is about 7 m.
Fringe mangrove forests: These fonn thin fringes along protected shorelines and
islands. being best developed along shorelines whose elevations are higher than
mean high tide. This community type generally shows characteristic zonation. The
low velocities of the incoming and retreating tides and the dense. well-developed stilt
root systems entrap all but the smalIest organic debris. Because of the relatively
open exposure along shorelines. the fringe forest is occasionally affected by strong
winds. causing breakage and resulting in the accumulation of relatively large
amounts of debris among the stilt roots.

• • .u.'
- , . • •• I ~

Fig. 6.6 Overwash mangrove forest on one of the many small island~ in the Horida
Keys.

Riverine mangrove forests: These include the tall (up to 20 m) floodplain


forests along flowing waters such as tidal rivers and creeks (fig. 6.7). Although a
shaIIow benn often exists along such creeks. the entire forest is usually flushed by
daily tides. This forest type is often fronted by a fringe forest occupying the slope on
192 Mangrove Ecology, Silviculture and Conservation

the creek side of the berm. During the wet season, water levels rise and salinity
drops because of upland terrestrial drainage. Low flow velocities over the surface
preclude scouring and redistribution of ground litter.

Fig. 6.7 Riverine mangrove forests near Cape Upstart. north-eastern Queensland.

Basin mangrove forests: These occur in inland areas along dminage depressions
channelling terrestrial run-off towards the coast. Close to the coast, they are
influenced by daily tides and, in Florida, are dominated by R. mangle. Moving
inland. the tidal influence lessens and the dominance is increasingly shared with
Avicennia germinans and Laguncularia racemosa. Trees may reach 15 metres in
height.
Hummockforests: These are similar to the basin type except that they occur on
ground that is slightly elevated (about 5-10 cm) relative to surrounding areas, often
by underlying peat deposits.
Scrub or dwaif forests: In Florida, this community type is limited to the flat
coastal fringe of southern Florida and the Keys (fig. 4.11). In the wider Caribbean,
similar scrub forests also occur in Belize (Feller 1996) and Mexico (Olmstead and
Juarez 1996). Individual plants rarely exceed 1.5 metres in height, except where they
grow over the depressions fiUed with mangrove peat, and many trees (shrubs) are
forty or more years old. Nutrients appear to be limiting (FeUer 1996) although
highly calcareous substrates and high salinities during the dry season also may play
a role (Lin and Sternberg 1992b). All three main neotropical species of mangroves
may occur in this situation.
While this classifactory scheme was based on Floridian mangrove
communities, Lugo and Snedaker (1974) report comparable forest types in similar
environments in Mexico, Puerto Rico, Costa Rica, Panama and Ecuador. The
apparent success of this scheme strengthens the idea that physiographic control, via
surface hydrology and tidal dynamics, is important in the distribution of mangrove
species and structural units.
Woodroffe (1992) first related this functional classification of mangroves to
their geomorphological settings (fig. 6.8), emphasizing that the two most important
Mangrove Structure and Classification 193

physical processes are riverine unidirectional flows and tidal bidirectional flows.
More recently, Ewel et al. (1998b) have proposed a simplified hybrid classification,
which distinguishes three extremes, based on dominant physical processes: river-
dominated mangroves as riverine mangroves, tide-dominated mangroves as fringe
mangroves, and interior mangroves as basin mangroves. Intermediate kinds of
forests, including the six New World types identified by Lugo and Snedaker (1974),
can be viewed as subgroups within the three extremes.

BASIN MANGROVES
INTERIOR

bidirectional
flux

RIVER DOMINATED TIDE DOMINATED


Increasing
RIVERINE MANGROVES salinity of FRINGE MANGROVES
floodwaters
Fig. 6.8 Classification of the three functional types of mangrove forests, as proposed
by Ewel et al. (I 998b), in combination with their dominant physical processes as
outlined by Woodroffe (1992).

The three extremes are easily described (Ewel et al. 1998b). Fringe mangroves
receive the brunt of the tides, which are often full-strength seawater. Stilt roots,
buttresses and pneumatophores are common among trees in this part of a forest.
Riverine mangroves are flooded by river water as well as by tides, so that salinity is
moderate. Trees in this zone are likely to be among the most productive in a forest.
Basin mangroves generally cover large areas behind fringe and riverine mangroves,
and only occasionally do tides inundate an entire basin forest. Soil salinity may be
very high at higher elevations where evapotranspiration causes salts to accumulate.
In small forests that are frequently flooded, or where rainfall is high or ground water
flow is substantial, basin forests can be of moderate or even low salinity
This hybrid scheme (fig. 6.8) highlights the substantial differences in
hydrology, nutrient cycling and productivity between these three types of forests
which can be expected to assist in formulating generalizations that may be
particularly helpful in establishing appropriate management policies (Ewel et al.
1998b).
194 Mangrove Ecology, Silviculture and Conservation

6.2 Zonation of Mangroves


There are several situations where the mangroves are zoned, and these provide a
particularly good opportunity to investigate and perhaps answer many ecological
questions, for these zoned communities can be treated as the outcome of a natural
experiment (Pielou 1977). Especially where uniformly sloping environmental
gradients are involved, the zonation pattern can be interpreted and used to study some
of the plant-environment interactions of the constituent species.
Two types of zonation are discussed below: the parallel zonation along open
shorelines, and the longitudinal zonation along rivers. As pointed out by E1s01 and
Saenger (1983), both types are superimposed in the lower reaches of rivers and their
deltas, and may produce diverse floristic assemblages with highly complex patterns
that cannot be interpreted even with sophisticated techniques of pattern analysis
(Bunt and Williams 1980, 1981, Bunt et al. 1985, Sinclair 1985, Williams et al.
1991, Bunt 1996, 1999, Ellison et al. 2(00). Consequently, both types of zonation
are discussed separately.

6.2.1 Shoreline Zonation


The presence of rather predictable, often monospecific zones of mangroves parallel to
shorelines has been described from many parts of the world (figs. 6.9 and 6.10),
including Asia (Watson 1928, van Steenis 1957, Macnae 1968, Mall et al. 1982,
Amarasinghe and Balasubrananiam 1992a, Satyanarayana et al. in press), Australia
(Macnae 1966, Saenger et at. 1977, Bunt and Williams 1981, Elsol and Saenger
1983, Johnstone 1983, Smith 1992), East Africa (Walter and Steiner 1936, Macnae
1968, Gallin et al. 1989, Beeckman et al. 1990, Ruwa 1993), West Africa (Kunkel
1966, Villiers 1973, Saenger and Bellan 1995), and the Americas (Rodriguez 1987,
L6pez-Portillo and Ezcurra I989a, McKee I995b, Santos et al. 1997, Vegas
VilarrUbia 2(00).
In the more species-rich Indo-West Pacific mangroves, there is some regional
consistency in the sequence of zones (Macnae 1966, 1967, Saenger et al. 1977). The
underlying causes as to why mangroves so frequently appear in zones are, however,
far from clear (Snedaker 1982, Rodriguez 1987, Smith 1992, Ball and Sobrado
1998). In his critical review of zonation, Snedaker (1982) examined suggested causes
and found that they fell into the following general categories: plant succession,
geomorphology, physiological ecology, and population dynamics. To these must be
added the role of chance events at least on the local scale (Bunt and Stieglitz 1999).
Zonation as the spatial expression of plant succession was the earliest view,
going back to Curtiss (1888). This view interprets zoned mangrove communities as
a sequence of seral communities from seawards to landwards, ultimately progressing
through to terrestrial vegetation. It hinges on the apparent ability of Rhizophora to
build and colonize new land (primary succession) by trapping sediments among its
root system into which fall the viviparous propagules that colonize the newly won
land (Davis 1940, Richards 1964). Further build-up of the substrate allows other
mangrove species to invade and eventually replace the Rhizophora (secondary
succession), until build-up exceeds the level of tidal inundation. At this point,
terrestrial species that are less salt-tolerant invade and replace (out-compete) the
mangroves.
The view that plant succession is the basis for mangrove zonation is logical
and appealing (Snedaker 1982). With more detailed study of succession in
ecosystems, however, it is becoming increasingly clear that mangrove zonation does
Mangrove Structure and Classification 195

not conform to the general characteristics of secondary succession (Johnstone 1983,


Tomlinson 1986, Smith 1992).

Landward Seaward
Station
Fig. 6.9 Seaward to landward changes in the importance values (sum of relative
density, frequency and dominance) of mangrove and associated species in the littoral
vegetation of Gabon. (Data from Vi11iers 1973)

As mangroves can trap sediment and thus build land, mangrove zonation has
been interpreted as a response to geomorphic change. Snedaker (1982) and Woodroffe
(1983, 1992) summarized the evidence for geomorphic control over vegetational
patterns and species assemblages, particularly landform patterns and vegetation.
Thorn (1967, 1975), Thorn et al. (1975) and Woodroffe (1987) investigated
vegetation in Mexico, Australia and the Pacific. They were able to relate species
assemblages, distributions, and overall spatial organization to the depositional and
erosional histories and to subsidence. compaction, freshwater discharge, and sealevel
history. Undoubtedly, mangrove development and zonation are historically bound to
the geomorphic process of a region through the particular soils and soil conditions
that these processes have produced. For example, Saenger and Bellan (1995) have
suggested that in contrast to zonation in the Indo-West Pacific, mangrove zonation
in West Africa may not reflect intertidal gradients of waterlogging or salinity but,
rather, may be more closely related to the varying soil texture preferences of the
different species. Clearly, what is needed now is more detailed information on
mangrove growth. development, and zonation from physiological and ecological
studies of the soil-mangrove relationship in each of the geomorphic settings that
have been identified.
196 Mangrove Ecology, Silviculture and Conservation

Zonation, as a physiological response to tidally maintained gradients, has


received considerable attention since the classical work of Watson (1928). The
interactions between surface hydrology and salinity on the one hand, and mangrove
zonation on the other, have been reviewed (Hutchings and Saenger 1987, Smith
1992, Ball 1996, Ellison et al. 2(00).

Fig. 6.10 Zonation of the mangroves of the open shoreline between the East and West
Alligator Rivers, Northern Territory.

From a series of detailed physiological studies, Ball (1988a, b, 1996) proposed


a model to explain how interspecific differences in salt tolerance might contribute to
the segregation of species along a salinity gradient. Species intolerant of high
salinity, such as Bruguiera gymnorhiza. operate with relatively high transpiration
rates and low water use efficiencies. These species can maintain larger leaves with
greater projected leaf areas than more salt tolerant species. In low salinity
environments, stands of B. gymnorhiza have dense canopies that permit little
penetration of direct sunlight to the forest floor. In contrast, highly salt-tolerant
species, such as Ceriops australis. are very conservative in their use of water aM
maintain small leaves with a low proportion of projected leaf area. Thus, slowly
growing species typically form stands with porous canopies, which under low
salinity regimes could not exclude the more rapidly growing and densely canopied
species characteristic of low salinity environments. Thus. despite growing
maximally under relatively low salinity conditions, the very attributes that enable C.
australis to tolerate highly saline conditions may reduce its competitive ability under
Mangrove Structure and Classification 197

the low salinity conditions where it grows optimally. Individual C. australis


occasionally occur in low salinity environments, but the species is limited largely to
highly saline habitats where it is a superior competitor or where competition with
other species may be absent. Thus, species from an available pool could become
distributed differentially along a salinity gradient because of differences in tolerance
limits and because of the ways in which physiological attributes associated with
differences in salt tolerance might affect competitive interactions for resources along
the gradient.
As Ball (1996) emphasizes, however, while physiological studies of species
responses to static salinity gradients contribute to our understanding of salt
tolerance, under natural field conditions salinities fluctuate in time and space, with
the time scale and magnitude of fluctuations dependent on the climate and
hydrological characteristics of the catchment area. These fluctuations in salinity are
an important determinant of the distribution and relative abundance of species along
spatial salinity gradients. For example, Ball and Pidsley (1988, 1995) showed that
seasonal fluctuation in salinity contributes to survival of both Sonneratia alba and
S. lanceolata in many environments (see 4.6). Similarly, Lin and Sternberg (1992a,
b) found that dwarfing in Rhizophora mangle in Florida is associated as much with
salinity fluctuations as with the actual levels of salinity. When tested experimentally
under constant and fluctuating salinity regimes in which the mean salinities were
identical (Lin and Sternberg 1993), seedling growth in constant salinity regimes
exceeded growth in a fluctuating regime at all salinity levels, confirming the
findings of Camilleri and Ribi (1983) concerning the greater response of this species
to salinity fluctuations than to actual levels of salinity.
Ball and Sobrado (1998) have discussed the various interactions between
salinity and humidity, high and low irradiance, and nutrients, and have suggested that
these various interactions are important in the performance of different species along
salinity gradients. In other words, it is not the differential ability to tolerate salinity
stress alone that will determine where a species will occur along a salinity gradient.
Rather, how salinity affects other physiological functions of a species will determine
the relative performance of that species. For example, interspecific differences in the
capacity of leaves of mangroves to maintain relatively high photosyntheic rates and
to provide protection from excessive irradiance, with spatial and temporal variation
in environmental factors that affect photosytnhesis, will influence the relative
performance of species along salinity gradients (Ball and Sobrado 1998).
Working in northern Brazil, Santos et al. (1997) proposed a salinity-based
zonation model in which salinity was largely determined by topographic level (and
thus, frequency of tidal inundation) and upland run-off (surface or subsurface). They
found that this model provided a good explanation of the various mangrove and salt
marsh distributional patterns in their study areas.
A range of studies have investigated environmental factors other than salinity;
in some studies good correlations were found while in others, no clear patterns could
be discerned. For example, working on the mangroves of Gazi Bay, Kenya, Matthijs
et al. (1999) investigated the relationship between soil redox state, sulfide
concentration, salinity, and mangrove zonation. Of the major species present, only
the distribution of only one, Rhizophora mucronata, could be explained by the
measured soil variables; R. mucronata did not occur in the less-reduced (more
oxidized) zone with high salinity (Matthijs et al. 1999). Similarly in eastern
Australia, Youssef and Saenger (1999) investigated the segregation of seven species
198 Mangrove Ecology, Silviculture and Conservation

of mangroves across intertidal gradients of topography, interstitial chlorinity, pH,


Eh, canopy cover, and interstitial concentrations of sulfides, reduced iron aIXl
manganese. They concluded that the zonation pattern of mangrove vegetation at their
study site could not be explained on the basis of anyone of the studied parameters,
suggesting instead that the segregation of the species across the site was best
explained by the cumulative interaction of three environmental gradients, namely
interstitial sulfide concentration, canopy cover, and height above the water table.
In a series of investigations, Ukpong (1989, 1992, 1994, 1995, 1997, 2(00)
investigated intertidal species distributions of the mangroves of Nigeria in relation to
various soil parameters, including pH, soil structure and nutrient gradients.
Although some correlations were found (such as the close relationship between
Avicennia germinans and Nypafruticans and high calcium sediments), most species
displayed wide ecological amplitudes with largely overlapping ranges.
Sherman et al. (1998) investigated the distribution of mangroves in the
Dominican Republic in relation to various soil chemical parameters including
alkalinity, TN, TP and DOC. They found that Laguru:ularia racemosa was
significantly correlated with TP and DOC concentrations, and suggested that a plant-
soil-microbial feedback system involving sulfate-reducing bacteria contributes to the
spatial patterning of vegetation and soil variables.
In a detailed analysis of data from II blocks, each with 3 transects and 20
quadrats in the Sundarbans of Bangladesh, Ellison et al. (2000) failed to detect
species zonation at any scale, concluding that this is a reflection of the underlying
biology of the system rather than an artefact of long-term human disturbance. Such
disturbance of the Sundarbans is well-documented (see recent review by Blasco et al.
2001) and the fact that this disturbance has been driven by the economic importance
of individual mangrove species has resulted in a situation where 'the relationships
between abundance and distribution of woody species in the Sundarbans ... do not
properly reflect the response of individual species or populations to local ecological
conditions' (Blasco et al. 2001 :252). Moreover, as Ellison et al. (2000) conclude, in
a hydrologically and sedimentologically complex environment such as the
Sundarbans, there may be a lack of concordance between differing edaphic
characteristics (salinity, field capacity, cation exchange capacity, grain size
distribution) across the intertidal zone. Differential responses of individual species to
independently varying edaphic factors result in high variability which is difficult to
explain. Similar findings were made by Bunt and Stieglitz (1999) and Youssef aIXl
Saenger (1999) from studies in complex settings in north-eastern and eastern
Australia respectively.
In 1982, Snedaker pointed out that, although good correlations may exist
between salinity, tidal inundation and mangrove zonation, the physiological
response of mangroves to many of these features is so incompletely known for most
species that it is premature to conclude that such correlations imply causality.
Despite a few excellent examples of physiological studies which have demonstrated
the significance of salinity (and salinity fluctuations) in species segregation aIXl
growth (Ball 1988a, b, Ball and Pidsley 1988, 1995, Lin and Sternberg 1992b,
1993), we are little closer to identifying causal connections.
Experimental and field studies have shown that species of salt-tolerant plants
near Sydney, Australia, do have definable tolerances and optima under specific
conditions, and that these can be used to explain landward and seaward boundaries for
each species (Clarke and Hannon 1971). Similar boundary conditions have been
Mangrove Structure and Classification 199

identified at other Australian localities, namely Port Curtis (Hutchings and Saenger
1987) and Mobbs Bay, New South Wales (Youssef and Saenger 1999), while similar
field optima for mangrove diversity, especially in relation to salinity and
waterlogging have been established (Ball 1998).
The metabolic basis of responses of species and communities to salinity, for
example, has been examined, and rests on a decrease in transpiration rates with
increasing salinity, an increase in respiration with increasing salinity, and a
maximization of photosynthesis at particular salinity levels in each mangrove
species (Hicks and Bums 1975, Ball and Pidsley 1995, Ball and Sobrado 1998,
Naidoo and Willert 1999).
It thus seems likely that if a salinity gradient is present then a gradational
sequence of species can result. Although a relationship between soil salinity and
mangrove metabolism and zonation is generally accepted, the extreme variability in
soil salinity makes the concept difficult to apply to specific field conditions.
Snedaker (1982) argued that short-term measurements do not necessarily reveal the
long-term mean to which mangroves must adapt; however, he indicated this to be an
error in technique, not in concept.
The view has also been put forward that zonation is a consequence of
differential dispersal and survival of propagules. In a series of studies based on field
work in Panama, Rabinowitz (1975, 1978a, b, c) found that (I) mangrove genera
which dominate lower tidal levels have large propagules whereas those that dominate
at high elevations further inland have smaller propagules requiring a period of
stranding prior to becoming established, (2) seedling mortality rates were inversely
correlated with propagule weight, and (3) seedlings did not exhibit better growth
under the canopies of their respective adults.
Based on these findings, Rabinowitz (1978c) concluded that zonation was
probably the result of differential tidal sorting and dispersion according to propagule
size and the frequency of tidal inundation of potential sites. She postulated further
that, following establishment, the competitive interaction between seedlings and
adults dominates subsequent survival.
These findings differ from those of others studying neotropical mangroves. For
example, Ball (1980:233) found that the propagules of Rhizophora mangle (large)
and those of Laguncularia racemosa (small) were 'widely dispersed and are readily
established everywhere, as shown by the presence of seedlings of both species
throughout the intertidal forests.' McKee (1995a, b, c) examined the spatial patterns
of seedling survival and growth in neotropical mangroves (see 3.5.3) and found that
the distance from reproductive adults explained much of the variation in the
performance of the major species. From transplanting experiments in Belize, Ellison
and Farnsworth (1993) similarly showed differential mortality of Avicennia
germinans and Rhizophora mangle seedlings associated with different tidal levels.
They suggested that zonation is maintained by differential seedling survivorship and
growth.
In the species-rich mangroves of northeastern Australia, the three underlying
conditions identified by Rabinowitz (l978c) for tidal sorting are not present. For
example, on the basis of long-term data from Port Curtis, Queensland, a region with
a 4 m tidal range, Saenger (1982) showed that the largest propagules (those of
Rhizophora stylosa) were the most widely dispersed throughout the intertidal zone,
and also had the highest first-year mortality rate. In addition, propagule survival in
nearly all species was enhanced when they occurred close to (within 2 m of) their
200 Mangrove Ecology, Silviculture and Conservation

respective adults (Saenger 1982). Similarly, Clarke et al. (2001) found that large,
viviparous propagules had no advantage over small or non-viviparous propagules in
terms of such early growth traits as lateral root and shoot initiation. They concluded
that tidal sorting was not important for zonation, as propagule buoyancy,
orientation, lateral root initiation, shoot initiation, and early shoot extension differed
among the 14 mangrove species investigated, but none correlated with adult
shoreline or upriver zonation.
It seems likely from the discussion above that simple tidal sorting cannot be
invoked as a universal mechanism and that in at least some mangrove communities
it has a minor role, if any at all (see section 3.5.3). One example of this minor role
was investigated by Patterson et al. (1997) in coastal Louisiana where only one
species of mangrove occurs (Avicennia genninans) and where this mangrove zone is
sharply demarcated from the low tidal level saltmarsh comprising Spartina
altemiflora. It was found that while propagules of A vicennia could become
established in the Spartina zone, they rarely did so because of tidal removal from this
zone due to their buoyancy. Thus, tidal action limited retention and settlement of
Avicennia propagules in the Spartina zone, and a combination of predator damage
and frequent flooding led to rapid decay of any propagules that were stranded there.
The notion that predator damage may affect zonation patterns by altering the
population dynamics has been suggested by Smith (1987a, b), based on the apparent
preference of grazing mangrove crabs for propagules of particular species of
mangroves (see 3.5.3). Species of Avicennia experience high rates of consumption
by predators compared to other mangroves and, at some sites, in apparent patterns
consistent with the spatial distribution of adult conspecifics (Smith 1987a, b, Smith
et al. 1989). For example, Smith (1987a, b) found that predation on Avicennia
marina propagules eliminated this species from the mid-tide levels although it could
survive and grow if predators were excluded; this could result in a bimodal
distribution of this species across the intertidal zone, a feature not uncommon in
northern Australia.
Smith (1987a) also found a negative correlation between predation and canopy
dominance in four out of the five species studied. On that basis, he modified the
general seed predator model, which suggests a relationship between predation and
distance from conspecifics in tropical forests, into the mangrove 'dominance-
predation' model which hypothesizes that significantly higher losses of propagules
will occur in mangrove forests where conspecifics are rare or absent than in forests
where conspecifics are dominant. While this model would not explain what caused
zonation in the first place, it would provide insights into how the species
composition of the various zones might be maintained.
Smith et al. (1989) tested this model in three widely differing locations
(Australia, Malaysia and Central America) and found that the 'dominance-predation'
hypothesis was supported by the results for Avicennia at all locations, and for
Rhizophora only in the old world mangroves.
Other studies (both in old and new world mangroves) have found that, although
crab predation is significant and varies between species. no relationship existed
between the amount of predation on the propagules of a species and the abundance of
conspecifics in the canopy at their study sites (McKee 1995a, McGuinness 1997b).
While regional differences in the seed predator guild might explain some of these
contradictory observations, it emphasizes the pitfalls inherent in extrapolating the
results from a few carefully chosen sites too widely.
Mangrove Structure and Classification 201

In his review of zonation, Snedaker (1982) concluded that those advocating


geomorphology and environmental physiology appeared to be most relevant in
furthering an understanding of zonation and plant succession in the intertidal zone.
He argued that there is a temporal tendency for each species to assume competitive
dominance in its preferred zone. Whether this occupation is guided by physical forces
or results from interspecific competition, the species which can maximize its
photosynthetic output with greatest metabolic efficiency dominates in competition
with other species. The concept of a zone or environmental preference implies that
each mangrove species does have a preferred optimum and a limit of tolerance related
to the metabolic cost of existence along an environmental gradient. Variations in
that gradient may either last long enough to result in competitive exclusion or
domination by a previously subordinate competitor, or fluctuate around a long-term
mean which enhances the likelihood of survival of the existing dominant and, thus,
of the zone.

6.2.2 Longitudinal Upriver Zonation


Although sharing some features with the parallel zonation of shorelines, upriver
zonation has been recognized as a distinct phenomenon since the descriptive accounts
of Myers (1935) of the riverine vegetation of South America. Myers defined upriver
zonation as the definite sequence of plant communities along the course of a stream.
This sequence was not determined by edaphic factors of the area through which the
river flows, but by factors dependent on the stream itself (for example, its width in a
given place, or the distance from the sea). As a result, essentially similar sequences
recurred in all the streams of a region where modification by humans have not
obscured them.
Hutchings and Saenger (1987) provided some Australian examples of
longitudinal upriver species distributions (fig. 6.11), including along the East
Alligator, Watson and Calliope Rivers. These river systems, despite the great
geographical and geological differences between them, show mangrove sequences
which have some similarities. For example, both Avicennia marina and Excoecaria
agallocha or E. ovalis have wide upriver distributions, whereas Ceriops tagal and
Lumnitzera racemosa have limited, downriver distributions. On the other hand,
Aegiceras comiculatum and Xylocarpus moluccensis show very different upriver
distributional patterns in the three river systems (Hegerl et al. 1979, Bunt et al.
1982a, Eisol and Saenger 1983).
Additional data from other river systems in northern Australia. namely the
Adelaide River, Northern Territory (Ball 1998), the South Alligator River, Northern
Territory and Murray River, NE Queensland (Duke 1992), the Daintree River,
northeastern Queensland (Duke et al. 1998), and the Normanby River, NE
Queensland (Bunt and Stieglitz 1999), show somewhat similar longitudinal species
sequences.
Longitudinal upriver species sequences have also been described from rivers on
the Atlantic and Pacific coasts of Panama (Duke et al 1998). Kuraishi et al. (1985b)
described the following sequence along the Shiira River on Iriomote Island, Japan: 0
km, Sonneratia alba; 0-0.4 km, mixed forest of Rhizophora stylosa and 8ruguiera
gymnorhiza; 0.4-1.5 km, pure forest of Bruguiera gymnorhiza; 1.5-2.0 km, a mixed
forest of 8ruguiera gymnorhiza and non-mangrove plants. From their data, salinity
was not the causal factor as it was always less than seawater due to high, non-
seasonal rainfall. Kuraishi et al. (1985b) suggested that soil differences were
202 Mangrove Ecology, Silviculture and Conservation

involved including higher upstream ratio of di- to mono-valent cations, less sand and
more clay upstream, more humus upstream, lower phosphate adsorption upstream,
higher upstream CEC, and more highly reduced soils upstream.

L racemosa
A. ItWrina
, R. stylosa
A. comiculatum
E. agal/ocha
, C. tagal
r---
'--- A. (mnulara
B. gymnorhiza
C. iripa
C. schultzii
H. tiliaceus
R. apicuuua
X.granatum
T. populnea
A. ilicifolius
B. sexangula
Melaleuca >pp.
X. moluccem'is
A. speciosum

o 10 20 30 40 50 60 70

Distance from the mouth (km)

Fig. 6.11 Mangrove species distributions in relation to distance from the mouth of the
Watson River, north-west Queensland. (Redrawn from Hutchings and Saenger 1987)

Initially, Myers (1935) tentatively identified three factors which influence the
upriver zonation, including (l) width of the river, (2) character of the water and (3)
distance from the sea. According to him, the width of the river was of importance
because it determined whether or not the waterway acted as a light gap, exposing the
vegetation on the bank to full sunlight. The character of the water largely depends on
catchment characteristics, and the distance from the sea is important because salinity
is a function of that distance and of the size of the river.
Two additional factors can be suggested which appear to correlate with upriver
distance: upstream gradients of decreasing salinity fluctuation (or range), and
upstream gradients of increasing turbulent flow, including the phenomenon of axial
convergence as recently described by Stieglitz and Ridd (2001).
Clearly, the salinity range to which a mangrove species is exposed may be just
as important as the mean levels of salinity, and may influence upriver distributions
of individual species (Duke 1992, Pidsley and Ball 1995). Leaf thickness in
Rhizophora mangle. for example, is related more to salinity fluctuations than to
absolute levels of salinity (Camilleri and Ribi 1983).
Few direct effects of turbulent flow on mangroves have been reported. The
significance of axial convergences in estuaries for upriver dispersal of propagules and
the subsequent concentration in hydrodynamic traps has only recently been identified
(Stieglitz and Ridd 2001). Similarly, turbulent flow characteristics can affect the
Mangrove Structure and Classification 203

meanders in a river, which, in turn, may affect species absence or presence. Erosion
of concave banks and accretion on convex lobes are largely associated with the
intermittent seasonal flow of floodwaters. The species compositions of actively
accreting convex lobes are generally strikingly different from nearby, non-accreting
river banks (Elsol and Saenger 1983).
Bunt et al. (l982a) related longitudinal upriver distributions of mangroves to
upriver distance and salinity gradients (see fuller discussion in section 4.6) in rivers
of north-eastern Australia. They found that Rhizophora stylosa. R. apiculata,
Sonneratia alba and Ceriops tagal were mainly from downstream, high-salinity areas,
whereas Heritiera littoralis. Excoecaria agallocha, Acrostichum sp., Aegiceras
comiculatum and Rhizophora mucronata occurred principally in upstream, low-
salinity areas. Three other species, including Avicennia marina, showed no
correlations with site and were found over almost the entire salinity range of the
river systems. Correlations with distance from river mouth were always better than
those with salinity, and in the Lockhart River, where virtually no salinity gradient
was found, they concluded that certain of the mangroves were responding to some
aspect of distance other than salinity. Furthermore, their study indicated that distance
does not simply act as an integrated measure of salinity, or at least not universally.
Duke (1992) clearly showed that upriver and downriver distribution limits in
Sonneratia alba, S. lanceolata and Avicennia marina were correlated with dry and wet
season salinity levels. In a detailed study of the Adelaide River in northern Australia,
Ball (1998) investigated the pattern of mangrove species richness in relation to the
combination of soil water content and soil water salinity. She found that species
richness was minimal in areas experiencing prolonged exposure to extremes of either
freshwater or hypersaline conditions (regardless of whether those conditions are
products of the river salinity regime and/or pronounced seasonal cycles of
waterlogging and drying), and maximal in areas where moderate salinities and high
soil water contents prevail in the late dry season.
Three ecological hypotheses were provided by Ball (1998) to explain why
species richness was greatest in environments of moderate salinity. First,
interspecific differences in salt tolerance may limit the pool of species potentially
able to colonize sites along a salinity gradient. Second, physiological attributes
associated with increasing salt tolerance (such as less dense canopies) may affect <»-
existence of species (such as understorey species) along a salinity gradient. Third,
seasonal variation in salinity may affect co-existence of species sharing at least part
of the same range of salt tolerance. Ball (1998) has suggested that the pairing of
deciduous and evergreen species might be one form of <»-existence facilitated by
salinity fluctuations (see also 3.2.3); thus, the drought deciduous Excoecaria ovalis
occurs as scattered trees overtopping a shorter, denser canopy of Ceriops australis.
Excoecaria bears its foliage during the wet season, when heavy rains and low salinity
floodwaters leach salts from the sediments. Vigorous growth of E. ovalis during this
period may adversely affect growth of C. australis. However, the more salt-tolerant
C. australis would be able to take advantage of the much longer growing season, as
E. ovalis loses its foliage when soil salinity increases during the onset of the dry
season. Neither one species nor the other has a constant advantage in terms of
monopolizing resources.
Clearly, the phenomenon of longitudinal upriver zonation is still poorly
understood. Gradients of absolute salinity or of degree of salinity fluctuation are
undoubtedly involved (Bunt et al. 1982a, Duke 1992, Ball 1998), but other factors
204 Mangrove Ecology, Silviculture and Conservation

such as the width of the stream and its geomorphological characteristics may also
have an effect. As has been discussed already under parallel shoreline zonation, if a
salinity gradient is present, then a gradational species sequence can result, although
in river systems the sequence may be secondarily modified by river width and
sedimentary characteristics. The fact that Bunt et al. (1982a) found better correlations
with distance than with salinity may be referable to shortcomings of measuring
salinity. As previously discussed, short-term measurements in a river system are
unlikely to reveal long-term means or ranges to which mangroves must adapt. As in
the case of soil salinities in shoreline zonation, this can be viewed as an error in
technique rather than concept (Snedaker 1982). Consequently, it seems appropriate
that, until detailed long-term salinity studies can be related to upriver mangrove
distributions, the view that salinity gradients are important should not be discarded,
despite some of its presently known imperfections.
In contrast to longitudinal upriver zonation, Bunt et al. (1985), Bunt (1996,
1999) and Bunt and Stieglitz (1999) and have studied changes in parallel zonation
patterns with distance upriver in northern Australia. They found considerable
diversity in zonal patterns, partly the result of tloristic differences between and along
rivers, but also arising from variability in the centres of distribution of species
across the intertidal surface. Specifically, they found that there are marked
distributional differences in individual species in relation to topographic height
between long and short rivers. As a result, sequencing of species across the intertidal
zone with increasing distance upriver is highly variable and unpredictable both
within and between tropical river systems.

6.2.3 Similarities and Differences in Shoreline and Upriver Zonation


Hutchings and Saenger (1987) examined boundary conditions for various mangroves
and associated ecosystems in Australia and they found that there was general
agreement in the boundary conditions for Port Curtis and Repulse Bay, Queensland.
Boundary conditions investigated included percent submergence, number of tides per
year, soil water content, and soil salinities. All boundary conditions coincided, with
the exception of the number of tides per year for the salttlat and saltmarsh
boundaries. This is significant in that the three boundaries involved are clearly at
different tidal levels at the two localities, yet, ecologically, there is little to separate
them. For example, critical percent submergence time is 1 and 10 at Repulse Bay
and Port Curtis respectively, and both soil salinity and soil water levels are similar
at the two sites. The same conditions in terms of soil salinity and waterlogging
obviously can be attained at different tidal levels at the two localities. This suggests
that boundaries between mangroves, salt marshes and salttlats may be determined by
levels of soil salinity and waterlogging which are largely determined by elevation,
but are not necessarily directly correlated with it. Similar observations of non-
concordance of edaphic factors have been noted across the intertidal zone elsewhere
(Youssef and Saenger 1999, Ellison et al. 2(00). Interestingly, Saintilan and Wilton
(2001) concluded that a shift in the boundary between mangroves and salt marshes
on the east Australian coast was due to an increase in the delivery of freshwater and
nutrients to the intertidal environments in response to decadal increases in rainfall in
combination with catchment urbanization.
Individual mangrove and saltmarsh plants can be broadly grouped by their
tolerance with respect to salinity and waterlogging. This has been illustrated in fig.
1.6 and the results may be interpreted as follows. On the open shoreline where
Mangrove Structure and Classification 205

salinities range from medium to high, the plant sequence is likely to follow the
waterlogging gradient, that is, high at the seaward margin and low at the landward
margin. On the other hand, in the upriver situation where waterlogging ranges from
medium to high, the plant sequence is likely to follow the salinity gradient, that is,
high salinities at the mouth and low salinities in the upper reaches. Thus, as a broad
generalization, shoreline zonation may be viewed as a response to waterlogging-
salinity gradients (Santos et al. 1997) while upriver zonation is influenced by
salinity-waterlogging gradients on a spatial and temporal scale (Ball 1998).
The vegetation sequences on open shorelines differ from those along the length
of an estuary. Salt flats, which may be abundant on open shorelines, are absent
upstream because very high salinity levels do not occur there. In addition, the tidal
influence decreases upstream with the result that the tidal zone becomes narrower and
the terrestrial fringing vegetation approaches the river bank. Those plants adapted to
medium-low salinities become increasingly common and are able to out-compete
most of the species tolerant of high salinities. This, in turn, would result in the loss
of the most seaward mangrove zones somewhere in the lower reaches of the river at
the same time that the driest landward zones (salt flats and saltmarshes) are lost. The
middle and landward mangrove zones, because of their medium tolerance of both
salinity and waterlogging, would extend farthest upriver. Gradually, the mangrove
zone would become dominated by those species able to optimize growth in medium
to low salinity conditions (such as Aegiceras comiculatum, Heritiera littoralis, H.
fomes, Hibiscus tiliaceus, Cynometra spp., Sonneratia spp., X. granatum and
Acrostichum spp. in the old world and Pelliciera rhizophorae, Acrostichum aureum,
Hibiscus tiliaceus, Tabebuia palustris and Mora oleifera in the new world).
Irregular shorelines disrupt the idealized pattern and may introduce various site-
specific anomalies. Nevertheless, the general patterns described can be recognized
with sufficient frequency to suggest that gradients of salinity and waterlogging are
the major interconnecting features of longitudinal upriver and shoreline zonation.
7. The Value of Mangroves
The channel by which we went to and returnedfrom Olinda,
was bordered on each side by mangroves.
which sprang like a miniature forest out of the greasy mud-banks.
The bright green colour of these bashes
always reminded me of the rank grass in a churchyard;
both are nourished by putrid exhalations;
the one speaks of death past,
and lhe other too often of death to come.

Charles Darwin (1836:432)

7.1 Introduction
Despite the above comments from Charles Darwin, homesick after crUlsmg the
world's oceans for five years, and 'having met with a want of politeness' at the hands
of Brazilians, mangroves are not harbingers of death - although such a negative view
was the prevalent one around that period. Only a few years earlier, the acclaimed
naturalist Alexander von Humboldt had described the South American mangroves in
similar terms: 'the mangle trees produce miasmas ... all settlers in the tropics are well
familiar with the noxious perspirations of those plants. They attribute the unhealthy
air to the root-stocks of the mangle trees ... we assume that chemical gases are
generated which defy all chemical investigations:

Mangroves. Vespucci and Venezuela

Between May 1499 and June 1500. the Florentine. Amerigo Vespucci. sailing under
the Spanish flag along the east coast of South America. noted in the Gulf of
Maracaibo near the mouth of the Orinoco River, the use of mangroves by coastal
people. He was particularly interested by entire villages built over the water,
supported by stilts. apparently of mangrove origin, where the people lived on the
estuarine products available. The area reminded him of Venice and, consequently, he
named it Venezuela ('Little Venice'). The enthusia.~m generated by the voyage in 1500
of the Portuguese explorer, Pedro Alvares Cabral, to the more southerly parts of South
America, heightened Vespucci's desire to return to the same region. Not supported by
Spain. presumably because of the papal division of the world into Spanish and
Portuguese hemispheres by Pope Alexander VI on 7 June, 1494, this Italian navigator.
like Columbus, now sailed under the Portuguese flag, reaching a spacious bay
(Guanabara Bay) with a large river on I January, 1502. Not surprisingly, he promptly
named the river. Rio de Janeiro!

Similar views of mangroves are contained in the various quotes commencing


some chapters of this book. Such perceptions prevailed in one form or another into
the late 20th century and imbued the mangroves with at least a nuisance value if not
a downright dangerous and sinister mystique (Saenger 1985b). With increasing
knowledge of mangroves, they came to be viewed as interesting scientific curios
before a more utilitarian view became prevalent. In fact, increasingly today,
mangroves are being protected and actively managed because of their perceived values
in providing products and services.
So, what are the values of natural mangrove wetlands and which are the
products and services that might be obtained from them? How many of these can be
obtained from managed mangrove communities? The values of a mangrove system
have many facets (Saenger 1985b) but four broad categories can be identified, i.e.
economic, usefulness, intrinsic, and symbolic values.
208 Mangrove Ecology and Conservation

Economic values are readily recognized in the direct and indirect products
obtainable from mangroves. Direct products consist of various wood products and
related materials such as tan-bark and fodder while indirect products include fish,
crustaceans, shellfish and honey (Saenger et at. 1983). These economic values can be
easily quantified as, for example, the annual value of a mangrove fishery or timber
yield.
Usefulness values ('instrumental values', 'free services' or 'ecological functions')
include the provision of habitat, shoreline protection, chemical buffering, water
quality maintenance, recreational and education opportunities, and reservoirs of
genetic materials. These values are difficult to quantify in monetary terms although a
number of approaches have been used with varied success (Lugo and Brinson 1978,
Thurairaja 1994, Gilbert and Janssen 1998, Perrings 2(00).
The acceptance of intrinsic values, i.e. that organisms, communities and
ecosystems have an inherent right to exist independent of human interest in them, is
becoming more widespread, forming the basis of much of the rationale of the
conservation and animal welfare movements and underpinning our efforts at heritage
and biodiversity conservation. Because these values cannot be quantified (and,
perhaps, should not be in any case - see review by O'Neill 1997), they are often
overlooked.
Symbolic values, derived from religious, totemic or mythical beliefs, are
probably attached to many mangrove areas by indigenous people. Thus, the sacred
groves in India, the sacred sites of the Australian Aborigines, or the fetish forests of
Benin, represent areas of considerable symbolic value to their custodians. Although
such values may be rather difficult for non-believers to understand or appreciate, they
should not be overlooked. As with the intrinsic values, symbolic values cannot be
quantified.
Taking into account the various types of values discussed above, we can now
attempt to assess the specific values of mangrove communities in terms of their
components, functions and attributes. Components have a direct use value which can
be assessed on the basis of costs (i.e. labour expended both as paid labour and as
labour time) and prices; functions are those indirect use values which support other
activities and which can be assessed on the basis of the replacement costs or the
value of the damage avoided; and attributes are those qualities that have a non-use
value even though they may also contribute to the direct and indirect use values.
Before proceeding further, however, it also should be noted that there is
increasing evidence that the coupling of such systems as mangroves with saltflats,
saltmarsh, seagrass or coral reef systems significantly enhances their ecological or
functional values (Fortes 1988, Yanez-Arancibia et al. 1993, Hemminga et at.
1994). For example, where mangroves and seagrasses and/or coral reefs occur
adjacent to each other, a dependence between these systems is generally apparent;
significant movements of fish and other organisms occur between the systems and
the sediment- and nutrient-retention functions of the mangroves provide waters
suitable for coral reef or seagrass development (Wolanski et al. 1997, 1998). While
the enhanced values resulting from the juxtaposition of these systems is hard to
quantify, the possibility of synergistic interactions between mangrove and other
adjacent systems, or even between different mangrove systems (such as dense, tall
and short, open-spaced communities) needs to be recognized and evaluated on a site-
by-site basis.
The Value of Mangroves 209

It must also be emphasized that not all mangrove systems provide all of the
goods and services described below (Ewel et al. I 998b ); understanding the diversity
that exists between mangrove forests should provide a sounder framework for
constraining uses to the type of forest where they are likely to be tolerated and even
sustained. As Ewel et al. (1998b:92) have aptly stated, 'being able to restrict
development by means of easily understood guidelines may be the first step not only
toward reducing the loss of mangrove forests in a region but also to reducing the rate
of loss of their goods and services'.

7.2 Components
7.2.1 Plant Resources
Mangrove systems are widely used for the plant products that they provide (Saenger
et al. 1983, Hamilton and Snedaker 1984, Weinstock 1993, Dahdouh-Guebas et al.
2(00), both at a subsistence and at a commercial level. Firewood collecting by local
communities is widespread. Collection of poles, firewood, and bark for tanning
appear to be the major uses of mangrove forest products.
In West Africa, for example, wood is collected for domestic purposes as well as
for salt production and fish smoking (Adam 1958, Kinako 1977, Egnankou 1985,
Bertrand 1991). Thus, Adam (1958) reported that 25-30 m high R. racenwsa are
heavily exploited in Gambia and that much of the wood and bark is marketed
through Dakar, Senegal. Similarly, Paradis (1989) described from aerial photograph
interpretation that between 1962-76, Rhizophora racenwsa areas around Fresco, Cote
d'Ivoire, were reduced by about 80% due to the collection of firewood, and tanbark
which is used to prolong the life of wood used for native canoes (pirogues). As a
consequence, extensive prairies of salt-tolerant grass Paspalum vaginatum have been
formed. In Cote d'Ivoire, Nicole et al. (1994) estimated that to smoke 30-40 kg of
fish required 60 kg of mangrove wood.
In Guinea, annual estimates of mangrove wood consumption (Diallo 1993) are
rural firewood 152,000 t y-I, urban firewood 54,000 t y-I, fish smoking 58,000 t y-I
and salt extraction (burning to evaporate saltwater) 93,000 t y-I.
In Benin, domestic firewood usage is also high while additional demands result
from 'acadja fishing' (brushpark fishing), which is common in the lagoons and
which requires additional mangrove wood for the stakes. The World Bank has
estimated firewood consumption of Cotonou and Porto Novo as follows:
1984-300,000 m 3 y-I actual; 2000-650,000 m 3 y-I actual; 2020-1,400,000 m 3 y-I
predicted. Such levels of demand cannot be met sustainably without large-scale
mangrove afforestation.
In the Niger Delta, R. racenwsa was heavily exploited for pitprops, for poles,
for firewood, and by the local fishermen for tanning their fishnets (Rosevear 1947).
The wood of Avicennia was rarely used, although there is some localized use of the
leaves for preparation of salt. Present day exploitation. however, is largely confined
to domestic firewood. and estimates of the standing volume of 283.2 x 106 m 3 are
considered to be capable of providing a sustainable timber yield of 0.6-1.6 x 106 t y-I
(Kinako 1977).
On the Pacific coast of Nicaragua, collection of poles, firewood, and bark for
tanning appears to be the major use of forest products in the Heroes y Martires re
Veracruz region. On the basis of the structural appearance of the forest of the region
and the number of boats carrying poles, pole extraction from Rhizophora mangle and
R. harrisonii is intense throughout the region. Most poles observed were relatively
210 Mangrove Ecology and Conservation

straight, 2-4 m long, and had a diameter of 5-10 cm, suggesting that young trees
about 10- I 5 years old are the prime targets. Most poles are apparently used for
construction purposes, although some are used for firewood. The bark is removed
from these poles by pounding (fig. 7.1), but this bark is not collected and used
further. Less use is made of the other species present, although Avicennia genninans
is used as a dry season firewood. At the same time, bark removal from large (> 10 m)
Rhizophora trees is also widespread, and dead debarked trees are common. Apparently
those trees selected for debarking are not cut down for firewood because of their size,
but are left standing and die within 2-3 weeks. The bark removed by this process
comes in sheets which are bundled and sold to local tanneries.

Fig. 7.1 Debarking of Rhizophora mangle and R. harrison;; near Corinto, on the Pacific
coast of Nicaragua. These poles are used for construction and the removed bark is not
used further.

Early commercial timber operations in India and Malaysia commenced in the


late 19 th century (see section 8.2.2) and cutch factories, to process tanbark, were
established in Sabah and Sarawak, Malaysia. in the early 1900s and continued to
operate until their closure was brought about in the late 1950s by fishermen
changing to nylon nets. In 1970, these small-scale operations were replaced by pulp
mills licensed to produce 6.1 million m3 of pulp chips annually (Phillipps 1984).
Wood-chipping was stopped in Malaysia in 1985 but it continues in the Indonesian
province of Kalimantan.
Similar developments also occurred in Africa. The tanbark industry commenced
in East Africa in 1900 (Grewe 1941) and the first systematic exploitation of
mangroves in Cameroon commenced at Manoka in 1919 when the 'Societe
Nationale du Cameroun' was granted logging concessions and built a timbermill
(Hedin, 1928). Considerable quantities of Rhizophora timber were extracted, initially
as railway sleepers, but subsequently for staves which were ultimately used for
barrels to export palm oil. A sizeable export of Rhizophora timber took place but
the proportion that this represents of the total timber extracted is not ascertainable.
The Value of Mangroves 211

Net primary production of mangrove forests is often high relative to upland


forests of equivalent latitude, but tree growth rates (and rates of biomass
accumulation) vary considerably, generally declining with increasing latitude
(Twilley et al. 1992, Saenger and Snedaker 1993). Mangrove forests achieve highest
productivity where there is no distinct dry season, and above-ground dry biomass can
reach 400 t ha- I (Komiyama et al. 1988). Riverine forests, where floods deposit
sediments periodically, generally have the best mangrove growth (Hussain and
Acharya 1994, Ewel et al. 1998b). Commercial harvesting is often most common,
however, in basin forests where large monospecific stands and less frequent flooding
make harvesting economically more attractive, and the risk of riverbank erosion is
avoided.
Estimates of annual wood production range from 3 to 28 t ha- l y-l (Saenger and
Siddiqi 1993, Saenger, 1994b) and in such areas, sustainable forest operations with
20-30 year rotation times are operating to produce construction timber, charcoal and
firewood (see section 8.7.6). Mangrove wood is generally dense (Table 7.1),
combining good appearance with strength, relatively easy workability, and
reasonable resistence to marine deterioration (Panshin 1932, Marco 1935). This
combination of desirable properties, together with its ready availability, makes
mangrove wood a valuable commodity.
Table 7.1 Mangrove wood density as dry weight per fresh volume.

Species Density (dry) Ref.


kg m·J
Aegiceras comiculatum 700 7
Avicennia alba 580 3
Avicennia germinans 868 2
Avicennia marina 650-880 1,6,7
Bruguiera cylindrica 810-890 7
Bruguiera gymnorhiza 665-975 1,7,8
Bruguiera parvijlora 650-930 7,8
Bruguiera sexangula 860-910 7
Camptostemon schultzii 445-500 5, 7
Cerbera manghas 600 7
Ceriops australis 752 8
Ceriops decandra 880-1070 7
Ceriops tagal 800-1070 1,7
Dolichandrone spatheata 500 7
Excoecaria agallocha 385-450 1,5,7
Heritiera fomes 1010 3
Heritiera littoralis 800-895 1.7
Laguncularia racemosa 759 2
Lumnitze ra littorea 600-680 7
Lumnitzera racemosa 650 7
Osbomia octodonta 850 7
Rhizophora apiculata 810-900 7,8
Rhizophora mangle 1011 2
Rhizophora mucronata 770-1130 1,7
Rhizophora stylosa 810-900 6,8
Scyphiphora hydrophyllacea 900 7
Sonneratia alba 590-850 7
Sonneratia apetala 570 4
Sonneratia caseolaris 700 7
Xylocarpus moluccensis 610-800 1,7
Xylocarpus granatum 486-800 1,7,8

lCause et aI. 1989; 2Rumbold and Snedaker 1993; JSattar and Bhattachaljee 1987; 4Sattar and
Bhattachaljee 1983; 5Phillips and Watson 1959; 6Boland et al. 1984; 1Panshin 1932; 8Clough and Scott
1989.
212 Mangrove Ecology and Conservation

Charcoal making, undertaken by small-scale producers, is an important cottage


industry in St. Lucia (Smith and Berkes 1993). Charcoal-makers in Mankote, a 40
ha mangrove area in the south-east of the island, work individually or in small
groups, helping one another on a reciprocal basis. Each producer uses one named
cutting area per season (two seasons per year, before and after the rains), and rotates
cutting areas, returning to a cut-over area after about two years. Strips of 10-20 m
are selectively cut in zig-zag patterns to provide access. Cut stems are placed in
rectangular pits about 4-6 m long, dug in the forest floor, partially covered with
grass or leaves and then with soil, and fired for about three days. The charcoal is
bagged, each sack holding about 22 kg. and sold in the town market for US$ll per
bag. Mean production is around 2.5 t mth- 1 from the 40 ha area (Smith and Berkes
1993).

Fig. 7.2 In St Lucia. small-scale charcoal production is an important cottage industry.


Cut stems of Rhizoplwra mangle are converted to charcoal in pits dug in the
mangroves_ (Photo_ lR. Clark)

In Bangladesh, extracted forest resources include saw-logs and such other timber
products as paper pulp and matchsticks from Excoecaria agallocha. Nypa fruticans
fronds for thatching, and Ceriops decandra as fuel wood (Hussain and Acharya 1994,
Siddiqi 2(01). In addition, however, considerable non-timber products are harvested
from the mangrove forests, including honey, wax and grass fodder.
Elsewhere, traditional usage includes the consumption of tender leaves and
macerated propagules of Avicennia mo.rina, Sonneratia caseolaris and Bruguiera
gymnorhiza as salads (Bandarananayake 1998), while the unrolled fronds of
Acrostichum spp. are eaten as cooked vegetatables; such healthy meals may be
supplemented by an alcoholic beverage fermented from the sap of Nypa fruticans
(Fong 1992).
As discussed in 8.2.2, plant products as well as the fisheries resources (see
7.2.2) associated with them, form important sources of commercial revenue in some
areas, whereas in other areas they underpin the very survival and livelihood of coastal
people.
The Value of Mangroves 213

7.2.2 Fisheries Resources


Mangrove systems, including their contained waterways, are an important fisheries
habitat which has been widely exploited throughout the world (Saenger et al. 1983,
Hamilton and Snedaker 1984, Matthes and Kapetsky 1988). For example, Pinto
(1987) has shown that fish catches in mangrove waterways of the Philippines ranged
from 1.3-8.8 kg hr- 1 throughout the year and that the fish biomass of the total catch
was positively correlated with the carbon content of the sediments and the mangrove
litterfall. In addition to fish, other species such as crustaceans and shellfish are also
commonly harvested.
Certain species of penaeid prawns, including Penaeus indicus. P. merguiensis,
P. monodon, P. subtilis. P. brasiliensis. and most species of Metapenaeus, are
dependent on mangrove forests for shelter during their juvenile stages. While it is
not possible to say 'no mangroves: no prawns', in some areas a good correlation
between prawn landings and mangrove areas holds; in other areas the relationship
does not hold (see 5.2.5).

Fig. 7.3 Cast netting for fish and prawns, as here in El Salvador. is a common
subsistence activity in most mangrove waterways around the world. (Photo. J.R. Clark)

Gill-netting for mullet, cast netting for shrimps, and digging for shellfish
(Anadara sp.) are important subsistence activities in many mangrove areas (fig. 7.3).
Brushpark fisheries, consisting of mangrove sticks or branches placed in shallow
waters in mangroves, have independently developed in many mangrove regions an:l
all present slight variations on the same theme (figs. 7.4 and 7.5). These small,
artifical mangrove thickets attract fish partly by the availability of food (organisms
attached to the plant material) and partly by the protection and shelter offered by the
thickets. In any case, when a sufficient number of fish have congregated in the
brushpark, it is enclosed with encircling nets and the fish harvested. Other designs.
such as the brushpark traps used in South Africa, guide the fish to a terminal trap
where they are easily harvested. Various forms of brushparks are known in Benin,
China, Cambodia, Bangladesh, Papua New Guinea, Mexico, Sri Lanka, Ecuador,
Brazil, Madagascar and South Africa.
214 Mangrove Ecology and Conservation

Fig. 7.4 This elaborate brushpark fishtrap from the Nypa swamps of Sulawesi,
Indonesia, consist.~ of fish guides to aggregate the fish and a fishing platform where,
with the use of lights, fish are netted or speared.

Fig. 7.5 Variations on a theme: brushpark fishtraps from mangrove areas in Kosi Bay,
South Africa (A); the coastal lagoon in Benin (B) (Photo. F. Bla~co); north-east of
Madang, Papua New Guinea (C); and in Negombo Lagoon. Sri Lanka (D).

What features of mangrove estuaries underpin the habitat value of these


mangrove-fringed waterways? Various features have been suggested including
protection from predators, the abundant availability of organic detritus (see 5.2.5),
and the generally sheltered and osmotically-favourable conditions (Blaber et al. 1989,
Sasekumar et al. 1992, Ruiz et al. 1993, Blaber 1997, Ley et al. 1999). Vidy (2000)
The Value of Mangroves 215

has investigated the fish habitat values of the Sine Saloum estuary in Senegal which
has been subject to drought since the 1970s. As a result of the drought, this estuary
has become a 'reversed estuary' with salinity increasing upstream, reaching maximal
levels of 100%0. A three-year study of the juvenile fish community was undertaken
to ascertain whether the estuary still functioned as a nursery area for important fish
stocks. The study also provided an opportunity to determine the relative importance
of estuarine processes (particularly freshwater mixing) and mangroves in the nursery
function. Vidy (2000:50) concluded 'that good estuarine conditions alone are
sufficient for good nursery function but mangrove alone is not'. The nursery role of
mangrove estuaries is depressed when insufficient freshwater inflow occurs.

7.2.3 Wildlife Resources


Mangrove systems support a range of wildlife resources including crocodiles, birds,
tigers, deer, monkeys, and bees for honey production (Saenger et al. 1983, Hamilton
and Snedaker 1984).
Many of these resources may be used directly (via hunting or gathering) or they
may contribute to the recreational resources of the area. For example, the Caroni
Swamp in Trinidad generates several millions of dollars from ecotourism, with boat
cruises to see the scarlet ibis (Eudocimus ruber) nesting in the mangroves.

7.2.4 Water Supply Resources


Aquifers of potable water commonly occur along the landward margins of mangrove
systems because the clayey sediments of the mangroves often restrict subsurface
groundwater flows. Many coastal communities use these lenses of water to
supplement rainwater supplies during the dry season. For example, such aquifers
exist and are utilized in and around Corinto on the Pacific coast of Nicaragua,
throughout Bangladesh, and in the Niger delta, Nigeria.

7.2.5 Agricultural Resources (including Salt Production and Aquaculture)


Small-scale use of landward mangrove lands for various agricultural activities is
common throughout the world (Hamilton and Snedaker 1984). Such small-scale
activities, which use mangrove lands without the complete conversion of mangrove
systems include, for example, salt production, salt-tolerant rice crops, oil palm
plantations, and aquacUlture ponds for shrimp and fish production. These activities
are distinct from the large-scale mangrove conversions for aquaculture (e.g. as in
Ecuador or the Philippines) or rice and sugar cane production (e.g. as in Indonesia).
Even these traditional agriCUltural activities, however, may not always be
sustainable. For example, in Benin the rate of mangrove wood consumed for salt
production is approximately 1 m 3 100 kg-· salt (Hachimou 1993). The progressive
introduction of more efficient methods of concentrating brines using solar energy or
LPG will increase the actual yield of salt while minimizing further mangrove losses.

7.2.6. Forage Resources


As mentioned in 5.2.3, mangrove foliage constitutes a nutritious food source for
herbivores and is widely used as fodder throughout parts of India, Bangladesh,
Pakistan, East Africa and the Middle East (Hamilton and Snedaker 1984, Hutchings
and Saenger 1987, Faye 1993, Fouda and AI-Muharrami 1995). In fact, Morton
216 Mangrove Ecology and Conservation

(1965) found that when used as cattle feed. the leaves of Rhizophora mangle
increased the yield of milk.

Fig. 7.6 Traditional salt production in the mangroves of Benin involves the leaching of
salt from mangrove sediments in large earthen containers. The brines are then further
concentrated over mangrove wood fires. (Photo. F. Blasco)

As elsewhere. extensive pastoralism is practised in Djibouti with camels and


goats. using A vicennia marina leaves are used as fodder; leaves of Ceriops tagal and
Rhizophora mucronata, although available. are not consumed. However. as Faye
(1993) emphasizes. Avicennia marina provides survival fodder only; it is inadequate
in relation to calcium. phosphorus and. above all. copper, zinc and manganese. Only
iodine and cobalt is present in sufficient concentrations to meet nutrientional
requirements. Permanent and massive exploitation of the mangroves to teed camels
is not sustainable, either for the mangroves or the camels. Disorders induced by a
diet only of mangrove foliage include lameness, weight loss, dermatitis. paralysis
and ectoparasitism. Complementary foods (e.g. cereal based concentrates) should be
used to keep mangrove exploitation at sustainable levels while safeguarding the
health of the camels (Faye 1993).

7.2.7 Water Transport Resources


Tidal flows in mangrove rivers and creeks is asymmetric. i.e. tidal velocities on the
ebb tide considerably exceed tidal velocities on the incoming tide (Wolanski et al.
1992). The extent of this asymmetry depends on the shape of the meandering
channels. and on the frictional drag of mangrove roots.
As a result, most mangrove-lined waterways are self-scouring (Wolanski et al.
1992). becoming important transport corridors because they are sheltered and usually
of sufficient depth to allow small boat movements on all states of the tide. In
addition, the absence of roads in many mangrove areas result in mangrove waterways
often providing the only means of communication and transportation between
coastal settlements. As such, they are a vital resource to these communities.
The Value of Mangroves 217

Fig. 7.7 Mangrove waterways serve a~ important aneries for the transport of people
and goods in areas where road networks are often poorly developed.

7.2.8 Recreational Resources


Many mangrove areas provide a range of recreational opportunities including
boating, fishing, sightseeing, swimming, birdwatching, and wilderness enjoyment
(Hamilton and Snedaker 1984). The consbUction of boardwalks, to allow easy access
to mangroves without sinking into the mud, has increased the recreational arxl
educational values of mangrove ecosystems, particularly those situated close to
urban settlements.

7.2.9 Energy Resources


Some mangrove areas may supply energy sources other than fuelwood. These
include peat deposits, hydrocarbons, or tidal energy which can be harnessed to
provide an alternative energy source. At present, few such sources other than
petroleum are utilized or identified.
The Niger delta has been recognized as a major oil and gas field since natural oil
seeps into the mangroves were reported by Hutter (1906). Now yielding huge
revenues (US$33.7 billion in 1998) to Nigerian and international oil companies, 23
out of 62 oilfields with some 1,800 oil wells occur within the mangroves of the
Niger delta (Commission of the European Communities 1987). Oil terminals are
situated at Bonny, Brass and Kanuskiri, while 8,000 km of seismic lines (20-3Om
wide) and oil pipelines criss-cross the mangroves.
The Nigerian mangroves also yield 20 billion m3 of natural gas each year. Until
recently, this was simply flared off. Much of it is now processed, adding to the
economic values of this resource.
Other countries earning large revenues from petroleum from in, or near,
mangrove areas include Angola, Brunei, Cameroon, Ecuador, Equatorial Guinea,
Gabon, Trinidad and Tobago, and Venezuela. Bangladesh has recently granted
petroleum exploration permits for the Sundarbans mangrove forests although no
hydrocarbon resources have been located to date.
218 Mangrove Ecology and Conservation

Fig. 7.8 Oil rigs amongst the mangroves of the Bonny River, Nigeria.

7.2.10 Pharmaceutical Resources


Numerous uses of mangroves for pharmaceutical benefits have been recorded. Walsh
(1977) and Bandaranayake (1998) have reviewed the folk medicinal use of
mangroves. However, as pointed out in 5.1.2 and 5.2.3, phytochemical screening of
mangroves is incomplete. Nevertheless, by way of example, E. agallocha is used for
the treatment of ulcers and as an aphrodisiac throughout south-east Asia and the
Pacific. The extract of this plant is also used to treat rheumatism, paralysis,
cutaneous infection, as a purgative, and as an abortificant. The wound healing
property of its latex has been investigated and found to be comparable with that of a
standard pharmaceutical preparation, furacin (Balu and Mathavan 1995).
The leaves and roots of Clerodendrum merme are used in Indian medicine as
therapy for rheumatism and for various skin diseases. It contains significant
quantities of the caffeic glycoside ester, verbascoside, which has known analgesic and
antimicrobial properties (Fauvel et al. 1989).
In his comprehensive (but poorly edited) review, Bandarananyake (1998)
provides an overview of the range of folkloric medical use of mangroves, including
their use as aphrodisiacs, blood purifiers, diuretics, purgatives, antiseptics,
anitfertility agents and anti-inflammatories. Their use to deal with snake bites, boils,
rheumatism, asthma, lice, smallpox, hepatitis, leprosy, syphilis, malaria,
gonorrhoea, diabetes. scabies, piles, typhoid, tuberculosis and elephantiasis are
particularly intriguing.
While most of these medicinal applications have yet to be pharmacologically
validated, coastal communities have derived considerable benefit (either real or
perceived) from various components of mangroves.
One proven use of mangroves by traditional fishers has involved the piscicidal
properties of some species which facilitate the capture of fish. The fish poisoning
properties of the climber Derris trifoliata were well known to the coastal people
from Asia to northern Australia. Similar properties with varying degrees of
effectiveness also occur in Aegiceras spp., Excoecaria agallocha. Heritiera littoralis,
Xylocarpus granatum and Barringtonia racemosa (Bandarananyake 1998).
The Value of Mangroves 219

7.3 Functions
7.3. J Shoreline Protection
As a result of their intricately entangled above-ground root systems, mangrove
communities protect shorelines during storm events by absorbing wave energy and
by reducing the velocity of water passing through the root barrier (Mazda et al.
I 997a, b). Many species of mangroves also have extensive cable root systems
(Saenger 1982) which assist in binding sediment particles. In this way, mangrove
covered shorelines are less likely to erode, or will erode significantly more slowly,
than unvegetated shorelines during periods of high wave energy.

7.3.2 Windbreak and Storm Protection


The frictional drag over a mangrove canopy is considerably higher than over the
water surface, causing a rapid decrease in wind speed inland of the mangrove
community. A mangrove shoreline can therefore act as an effective windbreak during
storm events (Oliver 1982), protecting leeward coastal settlements from intense
storm damage. In fact, in Bangladesh large mangrove afforestation areas were planted
so as to minimize the damage to coastal villages and agricultural land from the
frequent typhoons (hurricanes) originating in the Bay of Bengal (Saenger and Siddiqi
1993). Although the mangrove canopy may be damaged during such storm events,
the mangrove community is 'self-repairing', recovering rapidly from wind damage.
The effectiveness of mangroves in providing protection during cyclonic storms
was described by McConchie and Saenger (l991) as follows: 'The effect of cyclones
was well illustrated during work in the Sitakunda area of Bangladesh where the
results of one storm during the 1990 season led to some unexpected observations. In
March 1990 workers completed the construction of a 2 km seawall (about 10 m
high) using 2 t steel-reinforced concrete blocks. However, within 6 months the wall
had been in the path of a cyclone and about 24% of it had been smashed; some
concrete blocks had been broken while others had been moved up to 100m inland.
About 0.5 km south of the end of the wall, and also in the path of the cyclone, the
shoreline is occupied by a mangrove plantation with trees 3-5 m tall but during the
storm less than 1% of the trees were damaged and most of these had recovered within
6 months. Furthermore, nearby coastal sections which lacked expensive concrete
wall protection and were protected by mangroves alone had suffered no more storm
damage than areas protected by the wall. Our interpretation of these observations is
that the mangroves survived relatively undamaged because they possessed a degree of
flexibility and, unlike the concrete wall, did not offer a rigid barrier to the wave and
current action. Thus, it appears that mangrove forests can not only provide some
protection from storms and their associated tidal surges but when damaged by
storms, the damage is to a large extent self-repairing.'

7.3.3 Sediment Regulation


The erosive capacity of water is reduced on passing through the mangrove root
system, while the reduced velocity of the water leads to sediment deposition within
the root system (Table 4.5). In addition, as mangroves occur within the brackish
zones of estuaries and inlets, they are found in those areas where colloidal material
washed into the estuaries comes into contact with saltwater and flocculates.
Flocculation is the process whereby fine colloidal particles aggregate because of
surface charges, and clay particles form which tend to settle out of suspension. This
220 Mangrove Ecology and Conservation

process occurs in the mangrove zone of estuaries and causes clay deposits to fonn
within the mangrove systems. Once these clay deposits have fonned, they become
relatively resistant to re-suspension owing to electrostatic bonding within the clays
and the reduced wave energy conditions within the mangroves.
Thus, mangrove systems regulate sediment movement by reducing the erosive
tendency of water, by enhancing the fonnation of clay deposits, and by minimizing
the subsequent re-suspension of these clay deposits. In this way, mangroves improve
water clarity in nearshore waters and hence increase primary productivity by
phytoplankton in tidal creeks and coastal waters surrounding mangrove communities
(Wolanski et al. 1997). By simply leaving mangrove forests intact, society gains the
service of sediment trapping, because removal of the forest, particularly along the
banks of fringe and riverine mangrove forests, opens up vulnerable soils to erosion
and offshore sediment deposition. Mangrove forests can also be utilized for this
service when excess sediment generated by human activities such as road
construction and upland land clearance are prevented from washing out to offshore
seagrass beds and coral reefs.

7.3.4 Nutrient Retention


Because flocculated clays are largely deposited within the mangrove zones of
estuaries, many nutrients which are adsorbed onto the clay particles, are also retained
within the mangrove systems. This function of mangrove systems not only prevents
the loss of nutrients from the catchment area to the sea but also removes the
nutrients from the water column and stores them in the mangrove sediments (see
7.3.5).
Soluble phosphates are adsorbed onto the sediments and suspended particulate
matter as insoluble ferric phosphate, becoming stored in the sediments (Pailles et al.
1993). Microbial action reduces the insoluble ferric phosphate to soluble ferrous
phosphate which can then be absorbed and used by the mangrove plants. Organic
nitrogen brought down from the catchment area is similarly adsorbed onto the
sediments and suspended particulate matter and incoprporated into the mangrove
sediments. Bacterial activity can convert the organic nitrogen into ammonia which
can be used by some plants, or nitrite bacteria can convert the ammonia into nitrite.
The nitrite is then generally converted into nitrate by nitrifying bacteria. and these
nitrates are then actively cycled within the mangrove system (Rivera-Monroy arxl
Twilley 1996). By this mechanism, mangrove systems are able to scavenge
nutrients from oligotrophic tidal waters and maintain a productivity considerably
above that predicted from the nutrient concentrations in the water column (Lugo et
al. 1976).
In those areas where very low nitrogen levels occur, bacterial fixation of
atmospheric nitrogen in the mangrove sediments can subsidize the total nitrogen
availability (see 4.5.3).

7.3.5 Water Quality Maintenance


Apart from the nutrient retention function described above, mangrove systems have
an additional role in maintaining estuarine water quality. Suspended matter is reduced
by the mangrove root system and often deposited within the mangroves (Boto arxl
Patrick 1978), nutrient and heavy metal levels are reduced (Montgomery and Price
1979, Harbison 1981, Clark et aI., 1997, 1998), while dissolved oxygen
The Value of Mangroves 221

concentrations are raised. Where large inputs of organic material and nutrients (e.g.
from sewage discharges) are present, mangrove systems can also act as 'anaerobic
digestion' systems. removing organic material and reducing excessive organic
nitrogen concentrations. by direct conversion into ammonia which is then released
into the atmosphere (NedwellI975. Clough et al. 1983. Corredor and Morell 1994.
Rivera-Monroy and Twilley 1996).
Compared with the expense of constructing a wastewater treatment plant.
wetlands in general. and mangroves in particular. are still commonly selected as
receiving areas for effluent. Increasingly. however. the notion of specifically
constructed mangrove wetlands rather than natural wetlands is being adopted and used
for treatment of aquaculture (Robertson and Phillips 1995) and sewage effluents
(Nedwell 1975. Clough et al. 1983. Kelly 1995).

7.3.6 External Support


Mangrove systems provide considerable support to external systems such as the
lower estuaries. coastal lagoons and inlets. and the nearshore open sea. This support
is based on some of the functions described above (e.g. shoreline protection.
sediment and nutrient retention. and the maintenance of water quality) as well as the
organic productivity of the mangrove system.
Based on extrapolations of annual litter fall (Saenger and Snedaker 1993) and
wood production (Saenger 1994b). mangrove communities are likely to produce
around lOt ha-) y-) of litter and 17.5 t ha-) y-) of wood. On that basis. mangroves are
likely to produce about 4.3 t ha-) y-) and 7.3 t ha-) y-) of dissolved organic carbon
(DOC) and particulate organic carbon (POC) respectively. While a variable
proportion of this detrital production is likely to be cycled within the mangrove
systems (and this proportion will vary on a site-by-site basis). some will be
exported to support nearshore fisheries production (Lee 1999).
The relationship between nearshore fisheries and mangrove systems has been
studied at a number of localities (see review by Kapetsky 1985) and appears to be
particularly strong for penaeid shrimps (Pauly and Ingles 1999). certain shellfish.
and a number of finfish. Sometimes this relationship is based on tidal currents and
lateral trapping of larvae in mangrove-fringed channels (Chong et al. 1996). At other
times it is based on the nursery value of mangrove systems where juveniles of
important fisheries species feed (directly or indirectly) on the mangrove detritus -
thereby cycling detritus within the mangrove system. These species may be
harvested within the mangrove system (see under fisheries resources above) or they
may leave the mangrove systems at a certain stage in their life-cycle. thus
constituting a form of biotic export. It has been estimated that as much as 90% of
the US commercial catch. and 70% of the recreational catch in the Gulf of Mexico.
is made up of fish. shellfish and crustacean species that spend all or a vital part of
their life histories in estuarine areas where mangroves provide the major detrital
input (Caddy and Sharp 1986). In Australia. Pollard (1981) estimated that 70% of
the total commercial catch was estuary dependent. driven by mangrove productivity.
At other times. hydraulic export of mangrove detritus from mangrove systems
may support nearshore fisheries stocks. Kapetsky (1985) estimates a total fisheries
yield from mangrove-fringed estuaries and lagoons of 91 kg ha-) yolo In economic
terms. the most relevant figures are those from the Gulf of Panama, where a
US$94.628 km- l of mangrove shoreline value was found for shrimp. other
crustaceans and finfish combined (Kapetsky 1985).
222 Mangrove Ecology and Conservation

Of particular interest are the data from the Khlung District of Thailand where
the value of fisheries products captured inside the mangrove system was compared
with the value of mangrove-associated species caught elsewhere. On an area basis,
the respective values found were US$3,OOO km-2 for fisheries products captured inside
the mangrove system, and US$lO,OOO km- 2 for mangrove-associated species caught
elsewhere (Kapetsky 1985), suggesting that the nearshore catch of mangrove-
dependant species is approximately three times the catch from within the mangrove
system.

7.3.7 Groundwater Discharge and Recharge


The recharge of groundwater aquifers is commonly associated with wetland
communities but because of their position at the interface of land and sea, mangrove
communities are of little or no importance in this regard. In contrast, however,
mangrove systems regulate and often constrain the groundwater discharge at the
land/sea boundary and may be responsible for the accumulation of freshwater lenses
on their landward margins.

Fig. 7.9 An unsual groundwater discharge occurs in the porous limestone fonnations
of Celestun, Mexico, where crystal-clear groundwater is discharged through an
underground spring, providing a beautiful natural swimming pool amongst the
mangroves.

The regulated release of groundwaters of low salinity into mangrove waterways


is ecologically very significant (Semeniuk 1983, Santos et a!. 1997), particularly in
those areas where seasonal rainfall and high evaporation rates potentially lead to
hypersaline estuarine conditions (fig. 7.9). Groundwater discharge increases the
productivity of the mangroves, enhances species diversity of the mangroves aM
associated biota, and provides estuaries with characteristic chemical signatures which
are used as 'homing' signals by a range of juvenile fish and crustaceans (Odum
1970).
The dissolved organic content often associated with groundwater has been
shown to form organic flakes on contact with seawater, and these flakes may be
The Value of Mangroves 223

directly used as a food source by copepods, amphipods, isopods, crabs and shrimps
(Camilleri and Ribi 1986), all important components of mangrove food webs.

7.3.8 Local Microclimatic Stabilization


Like any other plant community, the mangrove community influences and stabilizes
the local microclimate. Particularly extensive and dense mangrove communities
provide a shaded habitat where day-time temperatures are lower, and night-time
temperatures are higher than in comparable open areas, thereby providing a
moderating influence. Humidity beneath the canopy also is generally higher than in
open situations. Evapotranspiration and long-wavelength reflectance from an
extensive mangrove canopy contributes to the regional humidity and cloud density,
which in tum may contribute to regional rainfall.

7.4 Attributes
7.4.1 Biodiversity
There is no doubt about the usefulness of genetic, biological and ecological diversity
in nature, and we can imagine the potential practicality of those parts which are not
currently used - the serendipity factor. Clearly, diversity in nature is of some
instrumental value in advancing human inerests and well-being, either now or in the
future. Equally, there are the unpredictable risks that biological impoverishment
could mean for the continuous dynamic functioning of biotic systems. This rationale
underpins our new found interest in preserving biodiversity both at the genetic and
the species level.
Biodiversity is usually expressed as some function of the number of species
contained within some specified area. The value of retaining natural and
representative plant and animal communities and their gene pools is difficult to
assess although attempts have been made on the basis of the density of various
organisms and the cost of those organisms if purchased from a Biological Supply
House (Lugo and Brinson, 1978). Other attempts have calculated the replacement
costs of the natural system on the basis of the cost of maintaining all of the various
component species in botanical or zoological gardens.
The simplest approach, though not necessarily the most accurate, might be to
assess the potential income that can be earned from areas of high biodiversity by
comparisons with the income generated by similar areas elsewhere (e.g. Everglades
National Park, Florida, or Caroni Swamp, Trinidad) and to adjust that potential
income for the actual number of visitors and the local economic conditions,
including local assessment of 'willingness to pay'.
While globally, angiosperm diversity is low in mangroves, this picture changes
drastically when associated plants and animals are included. For example, in a study
of benthic diatoms from a single mangrove lagoon in Mexico, 230 diatom taxa were
identified (Siqueiros Beltrones and Sanchez Castrej6n 1999).
Similarly, in a study of lichens epiphytic on mangroves along the eastern
Australian coast, Stevens (1979) recorded 56 species in the tropical region (north of
23° 30' S), 77 in the subtropics (between 23° 30' and 30° S), and 45 species in the
warm temperate zone (between 30° and 38° S), a total of 105 species.
Thirty taxa of bryophytes (mosses and liverworts) were collected from a variety
of soil and bark substrates in mangroves from Tampa Bay, Florida (TeStrake et at.
1986), while Windolf (1989) recorded 16 species from subtropical mangroves in
224 Mangrove Ecology and Conservation

eastern Australia. This is a surprisingly high level of diversity, particularly as


bryophytes are rarely associated with saline habitats.
Once invertebrates are considered, the biodiversity increases dramatically. For
example, Cantera et al. (1999) recorded 40 species of invertebrates from mangrove
branches and roots in Malaga and Buenaventura Bays on the Pacific coast of
Colombia. From the muddy substrates amongst the mangroves, they recorded a
further 70 species of invertebrates.
In an Australian mangrove community, 16 species of ants were recorded (Clay
and Anderson 1996), of which two species in the genus Polyrhachis appear to be
restricted to this habitat. Polyrhachis sokolova nests directly in the soft mud of the
mangroves (Nielsen 1997).
Even vertebrates can be regionally well represented in mangroves. For example,
Khan (1986) recorded around 400 species of vertebrates from the mangroves of
Bangladesh, including 8 species of amphibians, 50 species of reptiles, 261 species of
birds, and 49 species of mammals. Similarly, Lefebre and Poulin (1997) examined
two 4.5 ha mangrove sites in Panama and recorded 57 species of birds at their Pacific
coast site and 82 species at the Atlantic coast site.
These examples of the high biodiversity of plants and animals associated with
mangroves demonstrate the significant role mangroves play in maintaining coastal
biodiversity.

7.4.2 Uniqueness and Heritage


The attribute of 'uniqueness and heritage' covers the human cultural, scientific,
aesthetic, religious and historical perceptions of these systems and is thus highly
subjective. Nevertheless, examples of such values can be found in mangroves, as in
other places, although their significance may not be universally appreciated. For
example, in the mangroves of the delta surrounding the Pacific port of Corinto,
Nicaragua, lie the remains of the historic port of Realejo, established in the 16th
century as one of the earliest colonial gateways into central America. Similarly, the
site of the first landfall on the American mainland by Columbus in 1498 amongst
the mangroves of Venezuela, is significant, even though the exact location in the
Gulf of Paria is obscured by time and the motivation for the landing is questionable.
The mangroves of the Endeavour River, near Cooktown in north-eastern
Australia (fig. 7.10), have great scientific significance because of the early botanical
observations and collections made during the seven weeks that the Endeavour was
beached there for repairs. The mangroves were also mute witness to the conflict over
the land and its riches (turtles in this instance) between the newcomers and the
original inhabitants.
Relatively modem naval battles have been fought in the mangroves, bestowing
on them a special place in history. Thus, the First World War battle between the
German raider Konigsberg and two British monitors was fought entirely within the
mangroves of the Rufiji River delta in Tanzania (then German East Africa). This
surface raider had disrupted the important shipping lane via Suez and had hindered
strategic Allied operations in East Africa until it was entrapped in the Rufiji River
delta. Although military casualties of the protracted battle were surprisingly low, the
mangroves took a hammering from the high explosive shells used by both sides.
Mangroves have recolonized the mysterious ruins of Nan Madol (The Venice of
Micronesia'), dating back to the 13th century, on the eastern coast of Pohnpei,
Federated States of Micronesia. This group of 92 islands with their town and
The Value of Mangroves 225

ceremonial centre was abandoned as the religious centre for all of Pohnpei in the
early I 880s, and has since been reclaimed by the mangroves.

Fig. 7.10 The mangroves of the Endeavour River are typical of a drowned river valley
on the north-eastern coast of Australia. Where Cooktown is now was the spot where
Captain James Cook beached the Endeavour to make repairs after grounding on the
Great Barrier Reef. As the 'botanical gentlemen' found much to hold their interest,
these mangroves occupy an important place in AIL~tralia's scientific history.

While many other such examples undoubtedly exist, the few described here
highlight the range of 'heritage' values associated with some special mangrove places
which warrant recognition and appreciation.

7.5 An Economic Perspective


Having discussed the ecological, functional and other values of mangrove
ecosystems, let us briefly consider some economic aspects. From the infonnation
presented above, it is clear that mangroves are of economic value, even though
putting figures to that value is extremely difficult. The direct or indirect use values
can usually be quantified, even though there is no general agreement on a consistent
approach. Even less straightforward are non-use values of mangroves such as, for
example, carbon sequestration, biodiversity conservation or the opportunities they
offer for research and education. Nevertheless, assigning notional monetary values to
ecosystem goods and services is important in clarifying and guiding decisions
between two or more management alternatives within a particular mangrove system.
Nickerson (1999) has used such an approach to assess the relative benefits of:
(I) leaving the mangrove areas of Lingayan Gulf, Philippines, undeveloped; (2)
developing 468 ha of mangroves for polyculture (milkfish and prawns); or (3)
developing the same area for semi-intensive aquaculture (Penaeus monodon). Using a
multi-objective benefit analysis, the benefits were calculated per ha and then applied
to detennine the trade-offs to the larger Lingayan Gulf community as a whole.
The results of her analysis showed that the undeveloped option was superior in
tenns of total net benefits to the sectors examined, with the benefits obtained
226 Mangrove Ecology and Conservation

exclusively by the traditional sectors, predominantly the subsistence fishers. More


importantly, that option kept the diverse ecosystem functions available and provided
a more diverse range of seafood than either polyculture and semi-intense aquaculture
(Nickerson 1999). These considerations are significant in seeking to ensure
sustainable resource use and improvements in social welfare.
The Rufiji Delta and the Lone Raider Konigsberg
The Konigsberg was stationed in then German East Africa, partly to wave the
Imperial flag as well as to assist in quelling the numerous uprisings against German
colonial rule. Launched in December 1905 from the Kiel dockyards, and armed with
ten 10.5 em guns, it wa.~ both a formidable and an impressive ship for its day and
amply demonstrated the might of the German Imperial Navy. Refitted in 1911, she
arrived off Dar-es-Salaam in July 1914. Soon Commander Max Looff was parading
around the harbours of German Ea.~t Africa and the Konigsberg became known as the
'Manowari na Bomba Tatu' (the warship with three funnels). With the outbreak of
hostilities in early August, the Konigsberg commenced raiding merchant shipping, first
in the area around Aden and then around Madagascar. When in need of maintenance,
the Konigsberg was taken into the broad Rufiji Delta which offered an ideal
opportunity for protracted repairs to be undertaken in secret. Operational again by
September, the Konigsberg headed off to Zanzibar where she sank the unlucky HMS
Pegasus. Hiding again amongst the mangroves of the Rufiji Delta, the Konigsberg was
finally spotted by the searching British cruisers which, however, were unable to enter
the delta because of their greater draughts. Out of the range of British guns, this
standoff continued, with the Konigsberg remaining ensconced in the mangrove
swamps for 255 days. Christmas of 1914 wa.~ celebrated by cutting mangroves,
decorating these with Christmas baubles, and drinking the Ia.~t of their German beer
supplies. The mangrove confinement continued until II July 1915, when two shallow
draught monitors, HMS Severn and HMS Mersey, finally sank her after a fierce and
protracted battle. Each armed with two six-inch guns, these monitors had been
ordered by the Brazilian government as river patrol boats but were subsequently
purchased by the Admiralty, and towed from England via the Suez Canal to Mafia
Island, off the East African coast. Until it~ removal in 1962, the wreck of the
Konigsberg in the mangroves of the Rufiji was a reminder of the dark days of the First
World War. The mangroves bad barely recovered from their wartime bombardment
when a second episode of mangrove destruction from the tanbark industry took hold in
Tanganyika during 1923-37.

The monetary benefits derived from selected forms of harvesting or utilization


of mangrove systems are summarized in Table 7.2, where negative values indicate
that these benefits have been obtained after mangrove areas have been converted
irreversibly to another form of land use.
While monetary returns are highly variable due to the varying socio-economic
settings and the valuation techniques used, they suggest relatively low returns for
most traditional goods, for forest products, and for many fisheries products. The
monetary returns from mangrove conversion to extensive prawn and fish production
or agricultural production are not noticeably better. The only markedly enhanced
monetary returns seem to be associated with conversion to intensive aquaculture
(whereby ecosystem services are foregone) and the use of mangroves for such
ecosystem services as tertiary sewage treatment (whereby mangrove products are
largely foregone).
In their analysis of the value of the Pagbilao mangroves in the Philippines,
Gilbert and Janssen (1998) highlight one of the major problems with this type of
economic analysis. Because of the ease and precision with which total goods can be
valued compared with the valuation of environmental services. estimates of the value
of mangroves are often based on total goods rather than on the total services they
The Value of Mangroves 227

provide. This often leads to the attractive option to maximize the tangible values of
total goods rather than the intangible environmental services.
Table 7.2 Estimates of the monetary value of mangroves. No corrections for time-related fluctuations
in the US$ have been made. (Data from Christensen 1982, Saenger et al. 1983, Dixon 1989, Lal 1990,
Ruitenbeek 1994, Mastaller 1997, Gilbert and Janssen 1998, and Nickerson 1999).

Country Year Value Yield or Benefit


US$ ha') y.1
Traditional goods
PNG 1977 0.5 Crocodile skins, game
Benin 1989 12 Salt production
Indonesia 1990 33 Traditional product'>
Indonesia 1992 33 Traditional products
Thailand 1982 230 Traditional products
Philippines 1984 650 Alcohol from Nypa
Forest products
Fiji 1976 6 Timber and fuelwood
Indonesia 1990 66 Chipwood production
Indonesia 1992 67 Timber, fuelwood and chipwood
Trinidad 1974 70 Timber and fuelwood
Malaysia 1984 110 Fuelwood
Malaysia 1985 150 Charcoal
Philippines 1996 151 Timber and fuel wood
Thailand 1982 230 Nypa shingles
Thailand 1985 500 Timber and fuelwood
Malaysia 1986 566 Timber and fuel wood
Aquaculture after mangrove conversion
India 1986 -145 Extensive shrimp production
Philippines 1978 -180 Milkfish (Chanos chanos)
Thailand 1982 -206 Extensive shrimp production
Ecuador 1982 -390 Extensive shrimp production
Philippines 1979 -1,600 Intensive shrimp production
Thailand 1982 -2,106 Intensive shrimp production
Philippines 19% -7,124 Aquaculture
Fisheries products
Thailand 1982 30 Artesanal fisheries
Fiji 1976 100 Mangrove fisheries catch
Indonesia 1992 117 Mangrove fisheries catch
Trinidad 1974 125 Artesanal fisheries
Thailand 1977 130 Wild-caught shrimps
Fiji 1985 166 Artesanal fisheries
Indonesia 1977 1,010 Wild-caught shrimps
Malaysia 1982 2,770 Mangrove fisheries catch
Philippines 1996 60 Mangrove fisheries catch
Agriculture after mangrove conversion
India 1984 -35 Coconut
Fiji 1976 -52 Agricultural production
Senegal 1984 -80 Rice
Thailand 1983 -165 Rice
Indonesia 1992 -220 Rice
Ecosystem services
Indonesia 1992 3 Erosion control
Indonesia 1992 15 Biodiversity conservation
Trinidad 1974 200 Ecotourism park fees
Fiji 1976 5.820 Polishing treated sewage

However, the removal of goods (such as timber) from mangroves, particularly


when removal rates exceed sustainability, reduces the environmental services the
mangroves can perform. Mangrove utilization options which urgently need to be
explored and developed are those where the unimpeded ecosystem services can
continue to be provided in combination with the sustainable extraction of a range of
mangrove goods. Clearly, there is a trade-off between goods supplied and services
performed which requires better data on the montary values of the environmental
228 Mangrove Ecology and Conservation

services performed by mangroves if we are to make truly sustainable decisions


concerning the uses of mangroves.
8. Mangrove Silviculture and Restoration
In attempting to study the ecological development of the forest,
one must beware ofjumping to conclusions
based purely on present day evidence.
To avoid this,
one must keep in mind
some itka of the general trends
affecting the mangrove banks.
R.O. Dixon (1952:1)

8.1 Introduction
The above quote reveals an innate understanding of mangrove systems: they are
changeable, they are dynamic, they are unpredictable, they are subject to aperiodic
and periodic fluctuations of the extreme kind, and, as important as each of the above
is, each mangrove community has a history. Reading that history from the tell-tale
signs of today, is the artful skill of the silviculturalist or restoration ecologist who
is likely to succeed. Sound silviculture of mangroves, whether for commercial or
amenity planting, is based on an historical reading of the sites or areas involved. Not
only will this prevent the repetition of past mistakes, but it will provide the best
measure of the potentialities of any site or area for mangrove plantings for whatever
reasons it may be undertaken.
In today's world. mangrove planting may have an economic, aesthetic or
practical basis. Objectives include timber production, shoreline protection, channel
stabilization, fisheries and wildlife enhancement, legislative compliance, social
enrichment and ecological restoration (Field 1996, 1998a, b, Ellison 2(00). Many
readers could nominate additional reasons based on their own experience or
circumstances. In some countries, such as Bangladesh, Indonesia, Malaysia, Thailand
and Myanmar (Burma), there are large commercial mangrove forestry operations
which require the replenishment of harvested mangrove stands. In China, Vietnam
and Bangladesh, for example, mangroves are planted for purposes of stabilizing and
protecting the coastline and coastal towns and villages from cyclone damage and
seawater intrusion. Elsewhere, mangrove plantings are undertaken to provide amenity
values or to augment the ecological functioning of prawn (Philippines and Indonesia)
and fish ponds (India).
In this chapter, we will look at the various reasons for mangrove planting and
how those plantings might be undertaken in order to meet the original objectives.

8.2 Objectives for Mangrove Planting


The mere fact that a need exists to plant mangroves implies that some loss,
degradation or alteration has occurred in the past, sometimes from climatic events,
but more often from human activities. Given that those activities have resulted in an
area that does not comply with our defined management or conservation objectives,
some remedial action including mangrove planting may be feasible. At this juncture,
however, the quote at the head of this chapter, needs heeding: what is the history of
the site or area, or, more specifically, what prior activities have led to the present
conditions.
230 Mangrove Ecology, Silviculture and Conservation

8.2.1 Setting the Objectives


It is essential that objectives be clearly defined and prioritized (where more than one
objective is involved) as a first step in the planting process. For example, the coastal
afforestation project in Bangladesh has several objectives, the foremost of which are
the production of commercial timbers, acceleration of the rate of accretion of new
land areas, and improvement of the protection of nearshore agricultural and residential
lands from storm damage (Saenger 1986, McConchie 1990, Saenger and Siddiqi
1993). The project assisted in achieving all of these objectives but, in some
situations, success in achieving one objective meant less success in achieving
another. For example, in planting sites where very high sedimentation rates occurred,
trees were buried and timber production was negligible. In assessing the significance
of high sedimentation rates at particular sites, it is important to determine whether
the production of timber or the reclamation of new land is the highest priority for the
coastal afforestation program (McConchie 1990).

Fig. 8.1 Plantations of A vicenniu marina partially covered by sand deposition near
Chittagong, Bangladesh. Enrichment planting is probably the best management option.

Clearly. if timber production has priority then sites likely to experience high
sedimentation should be avoided, or the first planting should be considered as
sacrificial, and be planted solely to help raise the land to the point where another
more desirable species would survive. If stabilization and reclamation is the priority
then the burial of an initial plantation by sediment should be regarded as a success,
and replanting (,enrichment planting') should be undertaken to extend the gains
already made.
Prioritized objectives determine the planning process and help identify the
elements which must be included to provide the undertaking with a clear framework
for operation and implementation. As Field (l998a, b) has emphasized. the early
establishment of criteria for evaluating the success or otherwise of the planting
process is an essential. albeit difficult. task.
Generally, two main types of criteria are used in judging the success of a
mangrove planting program: (1) The effectiveness of the planting. which can be
Mangrove Silviculture and Restoration 231

considered as the closeness to which the new mangrove forest meets the original
objectives of the planting program; and (2) the efficiency of the planting, which can
be measured in terms of the amounts of labour, resources and material that were used
during the establishment and maintenance phases (Field 1998a, b). However, as Field
(1998a) has wrily noted, 'in the case of restoring a mangrove forest, what constitutes
a good restoration is often defined by the restorer'.

8.2.2 Timber Production


Mangrove silviculture and extraction has a long history, formalized by the colonial
forest departments of the European powers that usurped tropical lands. Thus, the
Dutch, Belgian, French, Spanish, German, Portuguese and British colonial forestry
agencies managed many of the mangroves of Africa, South America and Australasia,
generally on the management constructs of the spruce and oak forests of Europe, and
always with the need of colonial advancement firmly at the forefront of their
organisational objectives. As an example, in the Belgian Congo (now Zaire), tannin
extraction dates back to mangrove concessions granted by the Belgian Colonial
Forestry Service in 1908, which included permit conditions for: (1) riverine buffer
zones to prevent erosion and siltation; (2) concern over traditional user access to the
mangroves; and (3) a condition to replant extracted areas at 1,000 seedlings ha- 1
(Pynaert, 1933). While each of these requirements seems desirable in terms of our
modem perceptions of sustainable development, enforcement was virtually non-
existent. Nevertheless, Pynaert (1933) reported that the replanted mangroves had a
stem diameter (DBH) of 15 cm after 15 years and had become mature after 20-25
years.
Other examples of colonial forestry service management or mismanagement of
mangrove resources have been documented by Baillaud (1904), Brown and Fisher
(1918), Watson (1928), Hedin (1928), Heske (1937), Grant (1938), Grewe (1941),
Griffith (1950), Pellegrin (1952), Saenger and Bellan (1995) and Saenger (1998).
Despite these generally over-exploitative episodes, there were other colonial forestry
ventures that have been reasonably successful and two of these are briefly reviewed
on a case study basis_

The Matang Mangrove Forest


The Matang Mangrove Forest, in Perak State, Malaysia, covering just over 40,000
ha, is one of the outstanding examples of mangrove forestry. Gazetted as a mangrove
Forest Reserve in 1904, the first working plan for the 'Island Reserves' was
developed by A.E. Wells, followed by one for the 'Mainland Reserves' by J.P. Mead
in 1908. The plans were revised by J.P. Mead in 1915, A.E. Singer-Davies in 1924,
and J.G. Watson in 1925 (Watson 1928). The most comprehensive revision of the
plan, based on aerial surveys, was prepared for 1930-1949 by C.C. Durant.
During the Japanese occupation from 1942-45, however, the plan was not
followed and little or no supervision of forestry operations was undertaken; salt
production overtook many mangrove areas in the region as salt imports from
Thailand ceased. This hastily developed salt industry 'employed forced labour which
was remunerated by a miserable rice and fish ration only. These unfortunate workers
collected fuel for the evaporation process and, as might naturally be expected, they
stripped the bakau [Rhizophora] from the river frontages, allowing the lenggadai
[Bruguiera parviflora] to remain and prosper under this selective influence.' (Dixon
232 Mangrove Ecology, Silviculture and Conservation

1959:12). Immediately after the war, a critical renewal of management options was
required and in 1950, D.S.P. Noakes (Noakes 1952) devised a new management plan
for the reserve. This plan, a 100year plan, was revised in 1959 by R.O. Dixon, the
District Forestry Officer, Lamt and Matang (Dixon 1959). Since that time, 100year
working plans have been in place and form the basis for the subsequent management
of the Matang Mangrove Forest.
Dixon (1959:1), who modestly 'confesses to little mangrove forest experience
previously', provided the initial dynamic insights into the ecology of mangrove
communities, recognizing the sedimentological dynamics of the deltaic system he
was dealing with, as well as the longer term sealevel changes which had produced
this system.
In the Matang Forest Reserve, a total area of 40,711 ha is set aside for timber
production and about 993 ha (2.4%) is harvested annually for firewood and charcoal,
based mainly on Rhizophora apiculata and R. mucronata (Ong 1982). The pattern of
harvest is in alternating rectangular parcels so that each cleared area has mature
mangroves remaining on three sides. Each parcel itself is relatively small (300 x 50
m). Harvesting is not undertaken in areas where the remaining mangrove fringe
would be narrow after a 300 m wide parcel was taken out.

Fig. 8.2 In the Matang Forest Reserve. timber harvesting is in alternating rectangular
parcels to ensure mature mangroves surround the cleared area.
Mangrove Silviculture and Restoration 233

Management at Matang is based on to-year working plans, and a 30-year clear-


felling rotation is practised. Although this felling cycle was initially arbitrary and
intuitive, subsequent silvicultural studies (Watson 1928, Putz and Chan 1986, Gong
and Ong 1995) have largely vindicated its length (see 8.7.6).
After clear-felling, any dense infestations of Acrostichum ferns are eradicated
manually. Two years after felling, planting of Rhizophora is carried out in August
in poorly stocked areas. Seedlings of R. apiculata and R. mucronata are planted at
1.2 and 1.8 m spacings respectively. After 15 years, the first thinning using a 1.3 m
'spacer' stick is carried out and trees are removed for poles. There is a second
thinning at 20 years using a 1.9 m stick. Following the second thinning, no other
silvicultural treatment is undertaken until clear-felling after 30 years. During clear-
felling, the Forest Department imposes the retention of at least 7 seed-trees per ha
and a 3m wide river bank strip of mangrove trees.

Fig. 8.3 Post-harvesting sla.~h is left to decay for a period oftwo years by which time it
is sufficiently degraded not to damage newly established seedlings.

A number of villages are located within the mangrove forests, the population
being employed either in forestry or in the mangrove associated fishery (Ong 1982,
Tang et al. 1984). It has been estimated that the forestry industry employs a work
force of approximately 2,400 persons directly and indirectly. Of the 1,400 people
directly employed in the forest industry, around 150 contractors extract the timber
and produce the charcoal in about 300 kilns. The estimated total annual value of
mangrove products from the area is about US$9 million.
The fishing industry in the same area provides direct employment for an
estimated work force of about 2,500 and indirect employment for about another
7,500 (Ong 1982, Tang et al. 1984). Based on the value of prawns and cockles
alone, the total annual value of the Matang mangrove fisheries is estimated to be in
the order of US$33 million.
Clearly, the fishing industry provides employment for four times as many
persons as the forestry industry, and the return from fishing is three times that
234 Mangrove Ecology, Silviculture and Conservation

derived from forestry, demonstrating the value of the combined, multi-objective


management regime.

Fig. 8.4 Contractors fell their aJloted areas and, for charcoal production, poles 3-5 m
long are delivered by boat to the charcoal kilns where slow, controlled firing of the
wood produces approximately 0.5 m3 charcoal from each I m3 of green wood.

Despite this apparently sustainable operation, forest yields have declined with
each rotation from around 300 t ha" in the early 1900s to around 150 t ha- ' by the
early 1980s (Gong and Ong 1995); these declines are thought to be due to
silvicultural practices which are currently being reviewed (see 8.7.6).
Silvicultural practices similar to those at Matang were also introduced into
other areas of Malaysia. For example, in the Selangor mangroves, a 25-year cycle
has been adopted since 1957. No intermediate thinnings and Acrostichum eradication
is carried out. In the State of Johore, a 20 year clear-felling cycle is used. Only one
thinning (after 15 years using a 1.8 m spacer stick) is carried out. Apart from this,
no other silvicultural treatment is undertaken (Putz and Chan 1986).
Large mangrove areas in Sarawak have been clearfelled for the production of
chips for export, based on a 20-year rotation. Annual chip export is estimated at
US$2.6 million (Chai and Lai 1984). This large-scale exploitation has caused severe
damage in some areas. Thus, serious erosion of the surface mud in frequently
inundated areas has been observed, and at higher elevations there has been a rapid
drying out of the soil. Insufficient seed trees has resulted in poor regeneration of
mangroves and extensive Acrostichum infestations.
The main utilization of mangrove forests in Sabah is for chip production for
export. About 530,000 t of chips are produced annually by two chipmills,
generating annual export earnings of around US$5.5 million (Phillipps 1984).
Problems with regeneration, and degradation following large-scale exploitation, are
similar to those of Sarawak (FAO 1984, Phillipps 1984).
Mangrove Silviculture and Restoration 235

The Sundarbans Mangrove Forest


The Sundarbans forests were recognized as an important resource base about five
centuries ago. Scientific management was initiated more than one hundred and
twenty years ago, the area being set aside as a Reserved Forest in 1875. The first
ten-year plan, written by R.L. Heining, came into operation in 1893-94. In this
plan, the forests were divided into two felling series and 10 annual coupes
(Chowdhury and Ahmed 1994, Siddiqi 2(01).
Between 1906-12, a revised working scheme was developed by Sir Henry
Farrington, raising the exploitable girth of Heritiera fornes, the most desirable
timber from these forests, to 106.6 cm. This was followed by a working plan
designed by F. Trafford which introduced exploitable girths for H. fomes and S.
apetala. By the 1930s, standing stocks of other species such as E. agallocha, A.
officinalis, C. decandra and N. fruticans had been reduced because of their unregulated
felling and a new plan, devised by SJ. Curtis, came into force in April 1931.
At that time, volume functions were developed and the forests were divided into
management units called compartments and the felling cycle was reduced from a 40
year cycle to 20-years in the weakly saline areas and 30 years in the moderately
saline areas. This management regime remained in place with only minor
modification until a full forest inventory could be carried out in 1959.

Fig. 8.5 The mangroves of the Sundarbans display a high species diversity, forming a
rich, harvestable resource. Often characterized by a frontal zone of Nypa fruticans,
as shown here. the mangroves show dense development of Avicennia marina.
Heritierafomes, and Excoecaria agallocha on slightly higher ground.

Based on this 1959 inventory, a new working plan was developed by A.M.
Chowdhury to manage the forests on a sustainable yield basis with an accurate
projection of availability of timber, fuelwood and other forest produce. Three
working circles were established based on Heritiera, Excoecaria and S. apetala.
Prescriptions were also introduced for all other species and a permit system for
honey and wax was introduced. In addition, wildlife sanctuaries were established as
part of this plan.
236 Mangrove Ecology, Silviculture and Conservation

This working plan remained in operation until the war of liberation in 1971.
Government bans on exploitation were introduced in 1978-9 to allow the forests to
recover from over-exploitation during the period of the war and its immediate
aftermath. A second inventory was carried out by the Overseas Development
Administration (ODA) of the United Kingdom between 1981-85 using aerial
photography and modem inventory techniques. It provided detailed information on
forest types by species, stem numbers, standing volume, regeneration status, top-
dying of H. fomes and on other physical aspects of the forest (Chaffey et al. 1985).
Additionally, this inventory allowed comparisons with the earlier one completed in
1959, thereby providing a clear picture of the overall depletion of forest resources in
the period of 1959-83.
From a comparison of the 1959 and 1983 inventories, it was clear that there
had been a significant depletion in H. fomes and E. agallocha stems above a DBH of
7.5 cm, as well as a significant decline in average per hectare basal area of these two
species. Other species did not show such declines.

Forests of Sundari

Stretching for around 260 Ian along the Bay of Bengal from the Hooghly River
Estuary in India. to the Meghna River Estuary in Bangladesh, is the largest contiguous
tract of mangrove forests in the world, the Sundarbans. Reaching between 80-100 Ian
inland, it is traversed by numerous tidal creeks and rivers, and mangrove forests cover
a total area of around 10,000 km 2• The name Sundarbans is derived from the term
meaning 'forest of sUllliari or sunilri', a reference to the large mangrove tree,
Heritiera fomes, which can reach the impressive height of 25 m. Felled mainly for its
high value timber used for shipbuilding, electricity poles, railway sleepers and house
construction, it accounts for about 75% of total wood extraction and exports. Other
economic activities that take place in the Sundarbans include fishing, thatching, and
bees' wax and honey collecting. The fertility ofthe region is maintained by the annual
flood of the Ganges-Meghna-Brahmaputra Rivers which deposit huge amounts of
sediment, carried down from the Himalayas. With the advent of British rule in India.
control over the Sundarbans was assumed in 1828, and since 1875, the Forest
Department assumed responsibility for its management. The ftrst minimum exploitable
diameter for sunilri was established at this time. In 1879, the head office of the Forest
Department was established at Khulna to manage the entire Sundarbans areas, both
the Indian and Bangladesh portions as a single integrated unit. Since 1966 much of the
Bangladesh Sundarbans has been a wildlife sanctuary to protect the Royal Bengal
tigers and their principal prey, the spotted deer. Similarly, India established the
Sundarbans Tiger Reserve in 1973, covering an area of 2,500 km2, becoming the
Sundarbans Biosphere reserve in 1989. In Bangladesh, a \,400 km2 area was declared
as a World Heritage Site in 1997 in recognition of its outstanding biological value.
Thus, it is of particular concern that approval ha~ recently been given for gas and oil
exploration in the region.

At present, forest management of the Sundarbans mangroves consists of a


selection system to create and maintain uneven-aged stands which regenerate
naturally (Hussain and Acharya 1994, Siddiqi 2001). The areas due for felling are
divided into lots, generally 8-24 ha, and auctioned. Present annual timber yields are
around 250,000 m3, a significant quantity of timber in a country with few forests.
Despite this long record of management, there are problems in maintaining
sustainability in the Sundarban mangroves. Over-exploitation has resulted from poor
data on standing stock and from illicit cutting. Geomorphological and hydrological
changes have adversely affected the growth performance of the mangroves while
natural regeneration is limited. Top-dying of Heritiera fomes, a disorder causing
progressive die-back from the tops of trees downwards, currently affects 17% of the
Mangrove Silviculture and Restoration 237

trees (Siddiqi 2(01) and, until the exact cause is established, poses a significant
threat.
In general, then, what do these two examples show us? First, stability in
management and commerce (such as the price of timbers) is one of the major factors
involved. Second, the benign climate closely correlates with mangrove productivity
and, hence, timber production. Third, the harvesting regime must be ecologically
acceptable as well as economically workable. Finally, the silvicultural practices
must be suited to the conditions so that natural regeneration is the main driving
force behind the operation, as artificial replanting erodes the economic base
substantially.
Where these characteristics are not present, the stage is set for a non-sustainable
operation. Thus, for example, in the forest reserve of Guarapiche on the Rio San
Juan in Venezuela, a large concession of 21,680 ha was granted in the mid 1970s,
although the area for harvesting in the first stage was only 5,613 ha. Harvest was
proposed on a 30-year cycle. Thus, only about 187 ha, equivalent to less than 1% of
the total concession, was to be harvested at any given time. Nevertheless, the rate of
extraction appeared to be unsustainable and considerable concern was expressed
(Pannier 1979), particularly as silvicultural prescriptions were based only on data
derived from the mangroves of south-east Asia. Not surprisingly, the venture
ultimately failed, leaving a badly scarred swamp system (Alarcon and Conde 1993).

8.2.3 Shoreline Protection, Channel Stabilization and Storm Protection


Because of their intricate above-ground root systems, mangroves cause a reduction in
the velocity of waters flowing through and around them (see 7.3.1 and 7.3.2). Such
a reduction in water velocity gives them a significant role in coastal protection and
there are numerous examples where mangrove plantings have been specifically
undertaken for such purposes.
Late last century, there were proposals to stabilize the banks of the newly
constructed Suez Canal and the north African shorelines generally with Rhizophora,
based on these characteristics of mangroves (Riviere 1909) and, in 1905, mangroves
were planted on Molokai, Hawaii, to control shoreline erosion (MacCaughey 1917).
By 1911, demands were being made for the protection of mangroves and for the
regulation of their exploitation around Douala, Cameroon, in order to safeguard their
coastal protective function (Grewe, 1941).
In Bangladesh, a coastal afforestation program was initiated in 1966 specifically
to protect coastal villages from surge and storm damage during the frequent cyclones
that sweep into the Bay of Bengal during the monsoon season (Saenger and Siddiqi
1993, Siddiqi and Khan 1996). Most of the early plantations were established using
Sonneratia apetala because of the ease with which this species could be propagated.
The fruits of this species are collected in July to September and allowed to
decompose, releasing on average 50 seeds per fruit. These seeds are broadcast over
bunded areas of the intertidal zone where they germinate rapidly, generally within 4-7
days.
When the seedlings are around 15-25 cm high, they are gently removed and
planted out into the coastal areas where survival and growth is good (Saenger and
Siddiqi 1993, Siddiqi 2(01). The effectiveness of such mangrove plantations in
terms of stabilizing newly deposited sediments in the Bay of Bengal has been
confirmed by the 160,000 ha of mangrove plantations established over the past 35
years.
238 Mangrove Ecology, Silviculture and Conservation

Fig. 8.6 Intertidal nursery beds in Bangladesh for raising seedlings of Sonneratia
apetala for subsequent planting out in the coastal areas.

The effectiveness of mangroves to provide protection during cyclonic storms


was graphically demonstrated in September 1990 (see section 7.3.2) and apart from
Bangladesh, mangroves have been used specifically to stabilize estuarine channels or
to provide coastal protection in many other parts of the world, including the USA
(Teas et al. 1975, Gofortl) and Thomas 1980), the Virgin Islands (Lewis and Haines
1980), China (Lin and Xin-Meng 1983), Vietnam (Hong 1996), Australia (Saenger
1996), and Cuba (Padron 1996).

8.2.4 Fisheries and Wildlife Enhancement


Where mangroves have been depleted, there may be an associated decline in nearshore
productivity which, in tum, may prompt a mangrove replanting program as, for
example, in India (Untawale 1996), Pakistan (Qureshi 1990, 1996) and Colombia
(Bohorquez 1996). Bacon (1993) reports the increase of waterfowl usage of wetlands
in the Caribbean following mangrove enhancement through a planting program.
For the Indian Sundarbans, a silvopiscicuIture system has been established as a
land management system for rural development (Lahiri 1987). It combines fish
ponds with mangrove silviculture. Derelict forest land adjoining creeks was enclosed
with earthen bunds; each enclosure was 40 ha in size with about 30 ha of land area
and 10 ha of channels. Stocked with Mugil parsia and juvenile shrimp, the channels
Mangrove Silviculture and Restoration 239

yielded 150-200 kg ha- 1 after three months_ Upland species were planted on the top
of the bundwalls and mangroves at their base_ Good growth of the plants should
ensure sustained incomes over longer periods.
Similar shrimp-mangrove forestry farming systems are now being developed in
other parts of the world. particularly in Indonesia and Vietnam (Weinstock 1994.
Johnston et at. 1999. Clough 2(01).

8.2.5 Legislative Compliance


In several countries with mangroves. strict protective measures have been put in
place and in several instances. some compensatory requirements also exist (Snedaker
and Biber 1996). In most Australian States. mangroves are protected and the States
have adopted a procedural requirement that any mangrove losses resulting from
public works must be replaced by similar areas of mangrove wetlands. In Ballina.
New South Wales, for example, the re-alignment of a major arterial road required the
traversing of a mangrove community which caused a net loss of 7 hectares. In turn,
an equivalent area of coastal upland was set aside to be replanted with mangroves to
offset the ecological losses sustained. Such 'no net loss' policies are increasingly
being used by land-use management agencies.

.. BAlLiNA SHIRE COUNCil


-

I~C.)iC~=I·
. MANGROVE
REHABILITATION AREA
AUTHORISED ENTRY ONLY

Fig. 8.7 The Ballina 'Mangrove Rehabilitation Area' is not concerned with mangrove
rehabilitation but is an upland area which ha.~ been converted to an intertidal area and
planted with mangroves to offset the losses of natural mangroves during highway and
bridge construction.

The objectives for the Ballina planting were simply to comply with the
legislation and to duplicate, as far as practical (in terms of area, species composition,
structural complexity, ecological functions), the mangrove area lost during the bridge
construction. Compliance was assessed on the basis of an independent assessment
after an 8-year period (Latif 1996).
240 Mangrove Ecology. Silviculture and Conservation

Other examples of mangrove planting programs to comply with legislative


requirements include: replanting following an oil spill in Panama (Duke 1996); and
replanting following deforestation in Thailand (Aksornkoae 1996) and in the
Caribbean (Bacon 1993).

8.2.6 Social Enrichment


Planting for social enrichment covers a wide variety of objectives. ranging from
aesthetics to income generation through eco-tourism. For example. in the Middle
East where green vegetation is at a premium. the planting of mangroves is viewed as
one of the achievable means of beautification and 'greening' the landscape. In the
United Arab Emirates (fig. 8.8). mangrove planting programs have been ongoing
since the late I 970s. Plantations have been established on Sadiyat Island. Mubarraz
Island. Abu AI-Abyad Island. Al Sammaliah Island and at Ras Al Sidre. and. more
recently. there have been extensive municipal plantations around Dubai.

Fig. 8.8 In the United Arab Emirates. mangrove planting has been undertaken purely
for aesthetic purposes. Seedling nurseries are set up. a~ here on Abu AI-Abyad Island.
to harden the seedlings prior to outplanling.

Similarly in Qatar. in 1988 the Ministry of Municipal Affairs & Agriculture in


co-ordination with Japan International Co-operation Agency started a program to
plant mangroves. selecting a creek site. now within the limits of Ras Laffan
Industrial City (Tayab 1998). The intertidal zone of Ras Laffan Creek contains
diverse microtopographic features such as mudflats. calcium and bicarbonate rich
zones, salt marshes. and silty zones creating suitable conditions for mangrove
plantations. Management of the mangroves was assumed by the city administration
in 1995.
In these instances. the only objective has been to provide a green backdrop to
an otherwise stark landscape; the plantations have not only met that objective but
have provided additional benefits such as roosting and feeding sites for migratory
waders. flamingoes, and a range of other wildlife.
Mangrove Silviculture and Restoration 241

Rubin et al. (1998) present an interesting case where restoration is needed to


overcome mangrove exploitation, which has become severe because of dam
construction in the catchment of the Volta River in Ghana. The damming of the
Volta River at Akossombo resulted in reduced flooding and silt supply (Ly 1980),
the onset of coastal erosion, and a simultaneous increase in mangrove cutting, due to
the virtual collapse of agriculture and fishing in the estuary. Another ecological
consequence of the reduced level of flooding has been reduced dispersal of seedlings of
the principal mangrove, Rhizophora racemosa. Therefore, after cutting of this
species, recolonization is either by conspecifics growing at very high densities or by
any of a number of weed species. Environmental data collected in the vicinity of the
mangroves and each of the weed species have enabled suggestions to be made as to
whether R. racemosa or Avicennia germinans would be the more suitable mangrove
to replant. In addition, the introduction of the palm Nypa fruticans to the area is
being considered. Combined with the development of nature-based tourism, the
mangrove restoration program is expected to provide alternative means of income
generation, thereby reducing the need to cut remaining mangroves.
With any program aimed at social enrichment, it is important that
economically impoverished people who depend on the direct exploitation of
mangroves for their livelihood, must obtain benefits from the planting program. A
replanting program can be used to improve the economic opportunities and well-
being of rural people by providing direct employment for them in the planting
program, by restoring the productivity of lands and ecosystems that they use, and by
increasing the diversity of plants and animals that they harvest (Walters 1997). Such
benefits may include the change from an open-access to a communal property regime
as occurred in St. Lucia (Smith and Berkes 1993), where the wood products of an
area that used to be freely open to all potential users are now used and regulated by
an organized community co-operative for a 'cottage industry' in charcoal production.
Similarly in the Philippines, Walters (1997) describes the granting of 25-year private
leases over small plots of land for mangrove planting by people living near the shore
(,household planting').

8.2.7 Ecological Restoration


Although it has only recently become topical (Field 1996, 1998a, b, 1999, Kaly and
Jones 1998, Ellison 2(00), projects involving mangrove restoration in one form or
another have been around for many years. Some of the earliest mangrove harvesting
programs included replanting conditions (e.g. in the Belgian Congo, now Zaire;
Pynaert, 1933). Walters (2000) similarly reports on village-based mangrove
restoration programs in the Philippines going back for 90 years. Nevertheless, there
is an increasing awareness of the benefits of ecological restoration of mangroves
where these have been degraded or destroyed by human activity or natural causes.
One of the driving forces behind the restoration of mangrove ecosystems to
something like their presumed original states is the spectacular rise in environmental
consciousness over the past 30 years. The restoration of mangrove lands for
conservation remains a matter of choice and, while such choice should be determined
by the socio-economic priorities of the communities involved, more often it is
influenced by pressure from conservation organisations advocating the preservation
of nature (Field 1999).
Another important driving force for mangrove restoration is the worldwide
depletion of mangrove resources through increasing population (especially in coastal
242 Mangrove Ecology, Silviculture and Conservation

areas), food production, industrial and infrastructural development, excessive


extraction of materials, chemical contamination, and altered hydrology of mangrove
land (Hamilton and Snedaker 1984, Turner and Lewis 1997, Field 1996, Ellison
2000). Lewis (1982) was one of the first to espouse the principle of planting
mangroves specifically for the environmental services they can provide. One of the
necessities of such a program is to avoid the monocultures of mangroves that still
frequently characterize timber production. But, as Ellison (2000) points out in his
wide-ranging review, most restoration projects are still based mainly on one or two
species.
Nevertheless, there are a number of restoration programs that have achieved a
degree of ecological functioning similar to natural mangrove systems. Latif (1996)
reviewed mangrove restoration projects on the eastern Australian coast, ranging in
age from 1-18 years. He examined each of the restored areas in terms of survival rate
of plants, natural regeneration and general health of the plantation. For the 8-year old
Ballina plantations, consisting of 7 ha. additional detailed studies were carried out in
relation to fish usage at high tides, fish food availability, and short-term (marker
horizon) and long-term (sequential topographic surveys) accretion rates in planted and
control areas. It was found that within 4 years, the planted areas showed comparable
results to the control areas in all aspects.
McKee and Faulkner (2000) compared the structural development and
biogeochemical functioning of restored and natural mangrove forests within the same
physiographic setting in south-west Florida. One site had been replanted in 1982
while the second site had been replanted in 1990. Adjacent natural mangrove forests
which had been undisturbed for 50-60 years were used for comparisons. Despite
being planted initially with R. mangle, both sites had been naturally colonized by L.
racemosa and A. germinans. The restoration forests were dense but immature, with
lower basal area and stand height, and higher tree density when compared with
reference forests; at an age of 6-13 years, the restored stands were structurally still at
the pioneer stage (see fig. 5.6) but only a few biogeochemical differences were
apparent.
Morrisey et al. (in press) compared the faunal, floral and sedimentological
properties of A vicennia marina stands of two different ages in New Zealand. Older
(>60 y) and younger (3-12 y) stands showed a clear separation on the basis of
environmental characteristics and benthic macrofauna. Numbers of faunal taxa were
generally larger at younger sites, and numbers of individuals of several taxa were also
larger at these sites. As these mangrove stands mature, the focus of faunal diversity
shifted from the benthos to epiphytic fauna. such as insects and spiders. Differences
in the faunas were coincident with differences in the nature of the sediment; in older
stands the sediments were more compacted and contained more organic matter and
leaf litter. Interestingly, mangroves in the younger stands were able to take up more
Nand P than those in older stands, presumably because of less root competition in
the younger stands. Overall, these investigators concluded that younger and older
stands have significantly different functional characteristics which should be
considered in planning and management.
The major objective of any restoration program should be to facilitate natural
regeneration and to select target areas where some assisted regeneration is feasible and
needed. Secondary objectives might include the enhancement of growth of existing
mangroves and the introduction of selected additional species into the community.
Mangrove Silviculture and Restoration 243

Where local people are dependent on various mangrove resources, any


restoration program needs to consult with those communities, take their needs into
account, and involve them in the restoration project in a meaningful way. Walters
(1997) has proposed a set of human ecological questions that should be addressed at
the project formulation stage, as follows:
Economic impacts: How are local people likely to be economically affected by
the proposed restoration? Are there particularly vulnerable groups that may be
affected by the restoration? If so, how? Can the restoration be used directly to
economically benefit the local people?
Land use, resource management and tenure: How was the proposed restoration
area previously used? Who owns the land in the proposed restoration area and who
has rights to its use? What are the current uses of the area and how can they best be
managed for restoration? How can expected future uses of the area best be managed to
ensure long-term sustainablity?
Local knowledge, skills and custom: Have the local people done, or are they
doing, ecological restoration activities of any kind? Do local people have knowledge
that might be relevant, or skills that might be applicable, to the proposed
restoration? Are there any local customs that could benefit or conflict with the
proposed restoration?
Local social organization and institutions: How might local social groups and
networks contribute to, or impede, restoration efforts? What opportunities might
there be for collaboration with local, non-governmental organizations.
Government administration, policies and capacities: Are relevant government
laws and policies friendly to restoration? How committed are relevant government
administrations to environmental protection and restoration? What opportunities
might there be for collaboration with interested government agencies?
In summary, success with any restoration project is considerably enhanced
when (l) restoration is viewed by the local people as offering economic or other
tangible benefits to them; (2) the restoration is made compatible with local patterns
of resource use and land tenure; (3) local knowledge and skill relevant to restoration
are successfully embedded in the project; (4) local social groups and organizations are
effectively mobilized to support and implement restoration activities; and (5) relevant
policies and political factors are supportive of restoration efforts.

8.3 Macropropagation of mangroves


Once objectives have been established for mangrove planting in an area,
locating suitable germplasm and propagating it is the next phase. There are several
approaches that have been used for macropropagation of mangroves, including: direct
planting of propagules collected from the wild; outplanting of up to one-year-old
nursery-raised propagules; outplanting of small seedlings collected from the wild
after nursery-raising; direct transplanting of seedlings (<0.5 m) and shrubs (>0.5-1 m
high); raising of air-layered material; and the use of stem cuttings. Each of these
approaches is reviewed briefly below.

8.3.1 Direct Planting of Propagules collected from the Wild


As propagules are generally only available for 2-3 months of the year, direct seeding
needs to be scheduled according to their seasonal availability (Youssef 1997).
Spacing distance is difficult to control, but distances between 0.4-1.5 m have been
used (Lewis 1982, Saenger and Siddiqi 1993, Field 1996).
244 Mangrove Ecology, Silviculture and Conservation

The generally high susceptibility of propagules to desiccation, dislodgment by


waves and tides, and damage by predators and debris, make this material unsuitable
for sites of medium to high energy. It seems preferable that propagules are selected
for more protected sites or areas that already have some mangrove stands. Hamilton
and Snedaker (1984) suggested that planting of propagules should be carried out
during cool weather and on flood-free days.

8.3.2 Outplanting of up to One-Year-old Nursery-raised Propagules


Propagules of different species can be collected during the fruiting season and then
grown on freely drained sandy substrate. For most species, the soil should be kept
damp, preferably with 25-50% seawater (to reduce fungal infection). Relatively high
temperatures and humidity are essential for mangrove growth. Indirect sunlight of
about 400 ~mol m- 2 S-I photosynthetic flux density (PPF) is more effective in
producing a larger leaf area than direct sunlight (Youssef 1997). As mangroves
respond to fertilizers in early stages (Teas 1977), fertilizer can be 00ded once roots
have started to develop and the first pair of leaves have expanded. This results in
greater root growth and healthier shoots (Youssef 1997).
The benefit of raising seedlings in nurseries is that it provides a year-round
supply, which is of particular importance in large scale projects (Saenger and Siddiqi
1993, Field 1996, Youssef 1997). Nursery-grown seedlings also generally show
higher success rates (survival rates, increases in height, number of leaves) than
propagules (Saenger and Siddiqi 1993, Toledo et al. 2001), although higher nursery
costs and increased difficulties in planting when compared to propagules, may offset
these advantages in certain situations.

8.3.3 Outplanting of small Seedlings after Nursery-raising


Where propagules are unavailable. newly-established seedlings have been used to
establish nursery stock. The seedlings are dug out, washed free of soil. and replanted
in acid-washed coarse sand. Using this approach, Youssef (1997) found that survival
rate after one month was more than 90%. Seedlings also showed an exceptional root
recovery and strong shoot growth. After 12 months, the average height of the
seedling ranged between 50-70 cm.
With this approach the advantages of both nursery raised seedlings and small
seedlings «50 cm) transplantation have been combined. Damage to the donor site
has been avoided by using small seedlings. The technique also reduced the time
required for the propagules to grow to the required height (50-70 cm) - a height
where a high survival rate when transplanted into the field could be expected
(Youssef 1997).

8.3.4 Direct Transplanting of Seedlings and Shrubs


Where young seedlings are removed from the natural mangrove forest and
transplanted to sites under rehabilitation, the most common seedling height is <50
cm. Young seedlings are best removed by using a 10 cm PVC pipe as a corer,
pushed 20-25 cm into the substrate around the seedling. Transplants should be kept
moist and protected from direct heat and wind during collection and transportation.
Few reports have described trials of transplantation of young mangrove trees
from natural forests. The extensive root system of young mangrove trees offers the
promise of greater and faster success of establishment than could be expected from
seedlings. Therefore, these individuals would have a higher resistance to wave
Mangrove Silviculture and Restoration 24S

erosion and debris and can be transplanted into sites of higher energy where
propagules or seedlings are unlikely to survive.
Survival has generally been found to be inversely proportional to the size of
transplants and the length of time held in containers. Generally, for Avicennia and
Aegiceras less than 30 cm in height with a 20 cm diameter root-ball, survival rates
were over 80% for at least three months, while for plants more than SO cm tall,
survival rates fell below SO% within a month (Saenger 1996). Liao et al. (1998)
reported a 37% survival rate for transplanted seedlings and saplings of Kandelia
candel, less than 2 years old. For Rhizophora, most of the successful trials have used
saplings with heights ranging between SO-ISO cm (Pulver 1976, Evans 1977,
Goforth and Thomas 1980). Much taller transplants do not seem to have any
advantages. Consistent failures in transplanting A. genninans and L. racemosa trees
up to 6 m tall have been reported by Teas (1977).

8.3.5 Raising ofAir-layered Material


Air-layering involves the removal of short sections of bark and phloem of mature
lateral branches until the cambium becomes exposed. The injured area of the stern is
then wrapped with Sphagnum moss or similar material in aluminium foil to retain
moisture. After roots have developed, the stern is cut below the layering area and a
new plant is formed.
This technique has been applied successfully to various mangroves by Carlton
and Moffler (1978), Kathiresan and Ranvikumar (199S), Calderon and Echeverri
(1997), and Eganathan et al. (2000). Unlike raising young seedlings, the air-layering
method reduces the risk of root damage by insects and crustaceans, especially in the
early stages of establishment. However, the technique utilizes mature trees in the
field, and is thus subjected to many other biotic and abiotic variables that cannot be
controlled (e.g. fungal and bacterial infection, fluctuation in temperature and rainfall,
and tide height). Most importantly, the technique is relatively expensive (Carlton and
Moffler 1978).

8.3.6 Use of Stem Cuttings


Basak et al. (199S) reported auxin-induced rooting through stern cuttings in
Bruguiera parvijlora, Cynometra iripa, Excoecaria agallocha, Heritiera fomes and
Thespesia populnea. Das et al. (1997) reported on the induction of rooting in stern
cuttings treated with indoleacetic acid (IAA), indolebutyric acid (IBA), and
naphthalene acetic acid (NAA), both singly and in combination. They obtained
reasonable root development in Aegiceras comicuiatum, Avicennia officinalis,
Ceriops decandra, Excoecaria agallocha, Heritierafomes, Kandelia candel, Rhizophora
mucronata and Sonneratia apetala.
Youssef (1997) took soft-, medium- and hard-wood cuttings from five
mangrove species. Leaf area was reduced and cuttings were dipped into SO% liquid
rooting hormone (Rootex-L, Bass Lab., Vic). The cuttings were then planted in a
mixture of coarse sand and peatmoss (2: 1) and kept moist but well drained
throughout the experiment (fig. 8.9). For each species, fifteen cuttings for each wood
type were planted within 24 hours.
Of the five species, only Excoecaria agallocha and Hibiscus tiliaceus were
successful. Excoecaria agallocha gave a success rate of 90%, 60% and 30% with
medium-, soft-, and hard-wood respectively within 3-4 weeks. Hibiscus tiliaceus
gave a 100% success rate within 2-3 weeks with hard- and medium-wood but less
246 Mangrove Ecology, Silviculture and Conservation

than 50% with soft-wood. A similar approach has been used for propagating the
mangrove Scyphiphora hydrophyllacea at the Forest Research Institute at Los Banos,
Philippines.

Fig. 8.9 Stem cuttings of Excoecaria aga/locha grown on coarse sand and peatmoss.
The cuttings were prepared from soft-wood and pretreated with a rooting hormone.

Eganathan et al. (2000) reported the large scale propagation of Excoecaria


agallocha, Heritierafomes and lntsia bijuga using cuttings and air-layers. They found
maximal rooting of cuttings and air-layers after treatment with rnA alone, up to
2,500 ppm, in all three species.

8.3.7 Use of Propagule Segments


In the early I 980s, Thai foresters developed a method of cutting viviparous
propagules of Rhizophora apicu/ata into segments and planting each segment
separately. Generally cut into three segments (top, middle and bottom), many of
these segments formed roots and shoots, particularly when treated with auxins. This
cut-piece method offered the potential to increase the availability of planting material
from a limited number of propagules.
Using Kandelia candel, Ohnishi and Komiyarna (1998) evaluated the
effectiveness of this technique. The propagules of K. candel used ranged in length
from 20-28 cm and they were cut into segments of I, 2, 3, 5 and 10 cm, measured
from the top or the bottom with the remaining central part discarded. Thus, five
treatments and a control using intact seedlings were used for root and shoot
formation. Top segments of <3 cm resulted in virtually no root formation but in 35-
90% of segments >3 cm, root formation occurred. Bottom segments with or without
the radicle resulted in 40-100% root formation. In relation to shoot formation, top
segments with the plumule resulted in shoot formation in 55-95% of segments but
less than 20% in those segments not containing the plumule. For bottom segments
(except those of 10 cm), those containing a radicle did not initiate shoot formation
and only between 25-45% of bottom segments without the radicle produced shoots.
Mangrove Silviculture and Restoration 247

While it seems clear that the presence of the radicle has an inhibitory effects on
shoot formation in this cut-piece method, the results are promising. Ohnishi and
Komiyama (1998) estimated that using segments >3 cm, a 1.5-2-fold increase in
available planting material can be obtained for K. candeL The potential of this
approach for species with even longer propagules (e.g. Rhizophora mucronata with
propagule length of 40-75 cm) seems substantial.

8.4 Tissue Culture or Micropropagation of Mangroves


While considerable work has been carried out on terrestrial plants, very limited tissue
culture experimentation has been undertaken with mangroves to date. Maity (1994)
investigated the potential for tissue culture of the mangrove palm, Nypa fruticans.
and the high quality timber species Heritiera fomes from India. Leaf fragments from
young Nypa and Heritiera seedlings were used as explants and cultured aseptically in
Murashige & Skoog (MS) and SH media supplemented with vitamins and growth
hormones such as 2,4-0, BAP and kinetin. Sodium chloride (0.1-1 %) was also atJed
to some cultures with inhibitory effects. Undifferentiated callus tissue readily formed
around 60-70% of all explants but regeneration of plantlets was not successful with
either species.
Limited tissue culture work on the mangrove Bruguiera gymnorhiza has been
conducted by Satuwong et al. (1995). Using hypocotyls as explants, cultures were
established aseptically in MS and Woody Plant (WP) media supplemented with
sugar, hormones and vitamins. Callus was formed in around 6% of the cultures. On
sugar-free media, multiple shoot development occurred and some root initiation took
place with the addition of rnA and phloroglucinol while cultures were maintained in
darkness. However, no further root development could be induced and no successful
plantlet regeneration was achieved. Baba and Onizuka (1997) further refined the
techniques for callus induction and initiation of re-differentiation in callus cultures of
Bruguiera gymnorhiza. Kandelkl candel. Pemphis aciduJa and Rhizophora stylosa.
However, neither buds nor roots could be successfully induced. Akatsu et al. (1996)
were able to maintain a callus culture of Sonneratia alba using the pistils of
unopened flowers as the explant. Basal MS medium was used, supplemented with a
range of growth hormones and 3% sucrose.
At the 'International Symposium on Mangrove Ecology & Biology', held in
Kuwait in 1998, two research groups reported their progress in micropropagation of
mangroves, including A. marina. The first group, from the Kuwait Institute for
Scientific Research used embryo explants of Avicennia marina which were either
matured into seedlings on modified MS medium or into callus induction using
growth hormones (Sudhersan et al. 1998). Numerous meristematic nodules were
formed from the callus after 3 months and these could be used for plantlet
production. No details of the technique or of the results obtained were provided.
The second group from the M.S. Swaminathan Research Foundation in India
reported the development of micropropagation technique for Excoecaria agallocha,
Avicennia marina and Acanthus ilicifolius (Rao et al. 1998a) using a species-specific
mix of macronutrients, micronutrients, and iron and vitamin supplements for the
MS medium. Specific details concerning only Excoecaria agallocha were published
soon afterwards (Rao et al. 1998b). They reported using nodal segments for axillary
shoot proliferation; optimal axillary sprouting was obtained on a newly formulated
medium containing BA, zeatin and rnA in concentrations of 13.3 11M, 4.65 11M and
1.23 11M respectively. Repeated subculturing ofbinodal segments was possible from
248 Mangrove Ecology, Silviculture and Conservation

the axillary shoots and resulted in 10-12 shoots per explant in three months.
Rooting was initiated by growing the shoots in the medium with 0.23 ~M mAo
Regenerated plants were successfully acclimatized to the natural environment with an
8S% survival rate. Eganathan and Rao (2001) subsequently described the methods
used to culture Acanthus ilicifolius and Avicennia officinalis.
Working with Avicennia marina from eastern Australia (fig. 8.10), a tissue
culture protocol has been established (Cousins and Saenger, in press). Using actively
growing tips and the first binodal stem sections, six different chemical sterilants
were tried with a 0.1 % solution of mercuric chloride being the only successful one.
Spraying the foliage with a systemic fungicide (O.S g L- 1 Benlate) several days before
collection improved the results.

Fig. 8.10 Micropropagation of Avicennia marina: these plantlets have been


vegetatively propagated in half-strength MS medium through three generations in
aseptic culture.

Mineral salts of various concentrations were tested including MS, WP, Rao
medium 'X' (Rao et al. 1998b) and Gamborg BS (BS), of which O.5MS and BS were
found the most suitable. For mUltiplication, binodal explants were the most
responsive in a medium containing 112 MS macro salts + MS micro salts + 2 mg L-
1 Glycine + 1 mg L- 1 Thiamine + 400 mg L- 1 Inositol + 0.5 mg L- 1 Nicotinic acid +
O.S mg L-1 Pyridoxine + a combination of 3 mg L- 1 BAP, I mg L- 1 Zeatin, 0.2S mg
L- 1 mA and I.S mg L- 1 GA 3, using Gelcarin rather than Phytogel. Successful root
Mangrove Silviculture and Restoration 249

development was induced in the above medium using either 1 mg C 1 IBA or 1 mg L-


lNAA.
Despite the very limited studies on mangrove micropropagation to date, this is
a potentially significant area for mangrove germplasm improvement. For example,
Sharp (1980) concluded that high-frequency cloning may be useful for forest trees
characterized by poor seed set, absence of uniform seed production, and seed prone to
genetic damage or loss of viability during storage - features common in mangroves
(Saenger 1982). Of equal importance is that once a tissue culture propagation
technique has been established for a tree species, the cost per propagule is
considerably lower than that of rooted cuttings, although it would probably still
exceed the cost of seedlings derived from genetically improved seed. More
importantly, the availability of vegetatively propagated plants will also increase the
accuracy of field trials as the same genotypes can be replicated across sites.
With the successful introduction into in vitro culture, considerable experimental
screening of mangrove genotypes could also be undertaken as a long-term objective
e.g. for salinity and frost tolerance, or insect resistance (Kamosky 1981, Maity
1994). Consequently, cell screening may offer some opportunity to test a large
number of mangrove genotypes in shorter periods and in a smaller space.
Nevertheless, following any preliminary screening, batches of vegetatively cloned
plantlets will still be required for field testing, as cell-screening systems cannot
replace on-site plant genotype evaluations (Kamosky 1981).

8.5 Site-Species Matching


8.5.1 Some General Principles
From studies of their natural occurrences, we have gained some insights in terms of
the soil conditions under which specific mangroves perform optimally. Site selection
is of primary importance at the commencement of any planting program when the
most favourable sites should be the initial targets. Detailed surveys should be carried
out for potential planting sites, followed by consultation with other stakeholders to
ensure that other valuable ecosystems, activities or conservation initiatives are not
interfered with.
Site surveys should examine at least the following parameters to assess each
site in terms of its suitability (Saenger et al. 1996): (1) Exposure: Mangrove
seedlings will not survive in areas exposed to periodic high wave energy or persistent
wind-generated turbulence. Thus, sites should be sheltered from wave energy in semi-
enclosed embayments or lagoons, or where shelter is provided by coastal features
such as headlands or offshore reefs and sandbanks. (2) Stability of the substrate:
Substrate stability is another requirement for mangrove growth at least during the
establishment phase. Sheltered localities generally exhibit stable substrates but tidal
scour or littoral drift can occur in sheltered locations and cause erosion or accretion of
sediments. Such areas can usually be easily recognized by their sedimentology (e.g.
coarse gravelly sediments, pronounced sand ripples, absence of crab burrows etc) and
should be avoided. Where such areas cannot be avoided, temporary stabilizing
techniques (e.g. grass matting, chicken wire, tyre barriers) may be necessary. (3)
Shoreline morphology/sedimentology: Mangroves generally grow best in muds and
clays which have been deposited under sheltered conditions and which are regularly
inundated by tidal waters. Topographic height in relation to optimal growth is
generally in the top third of the tidal range, and to achieve this at any particular site
some recontouring may be required.
250 Mangrove Ecology, Silviculture and Conservation

-500
C/adium LaguncuJaria
-400
:>
g -300

I8..
-200
-100
..
0
0
Avicennia
""..,
100 Conocarpus

'" 200
300
~
40 AeglCeras
~ RhiZOPhorrlll
~

0
30
~
C
B
c:::J . l}"t=1l.umnitzera Bruguiera
c 20 AVlcenma

D
0
c.>
~

~ Osbomia
~
10
:aen 0
0 10 20 30 40 50 60 70 80 90 100
Soil salinity (gIL)

Fig. 8.11 The upper graph shows the distribution of dominant species against soil
salinity (mg L- I ) and soil Eh (mV) at 25 cm depth at Guadeloupe. (Data from Imbert
et al. 2000a). The lower graph shows selected species against soil salinity (mg L- I )
and soil water content (% fresh wt.) measured over 1.5 y at Proserpine, Queensland.
(Data from Hutchings and Saenger 1987)

Once site selection has been made, site characteristics can be used to identify
those species best suited to that site. The degree of waterlogging (Eh of the soil),
depth and frequency of tidal inundation, and the salinity of the porewater are probably
the main factors determining species suitablity (fig_ 8.11). Conversely, if a
plantation of a particular species is to be established, the requirements of the species
will determine the site for its establishment.

8.5.2 Nursery and Planting Techniques


Nursery and planting techniques vary considerably among the various mangrove
species (Siddiqi et al. 1993), and a summary of the techniques employed for some of
the major species is given below:
Avicennia spp. The cryptoviviparous propagules of species of Avicennia are
usually collected from around the base of mother trees when they have matured
(Tables 3.8 and 3.9). When kept in air, these propagules rapidly lose their viability
within a few days. These propagules may be directly planted into sheltered areas by
'dibbling' - where the propagule is gently pushed into the soft sediment until firmly
wedged. 'Dibbling' is usually undertaken during neap tide periods to allow the
seedling to develop roots. Pre-treatment of the propagules has sometimes been used
to decrease the establishment time. Such treatment consists of placing the
propagules in small nets and exposing them to daily tidal inundation to hasten the
decay of the pericarp. Removal of the pericarp by pre-treatment reduces the
establishment time to 2-3 days, while without pre-treatment, 5-6 days are required.
Alternatively, propagules may be raised in nursery beds that are exposed to daily tidal
inundation. Seedlings are raised for about 1-2 months after which they are gently
Mangrove Silviculture and Restoration 251

pulled out of the ground and packed for transportation to previously chosen
afforestation sites where they are usually planted out into holes of 3 cm diameter at a
spacing of 1.0 x 1.0 m.
Propagules for nursery raising are simply placed on the surface of freely
draining sand and vermiculite mixtures, kept in full sunlight, and watered once daily,
preferably using 25% seawater which suppresses fungal infections in the seedlings
and acclimates them to the saline conditions into which they are to be planted.
Growth and survival rates are significantly higher in A vicennia marina in 25%
seawater than in freshwater. Humid conditions and the once-off addition of a slow
release fertilizer such as 'Osmocote' (NPK 18:2.6:10 at approximately 3 g per 10 cm
pot) will enhance optimal growth.
Excoecaria agallocha. Seeds of this species ripen in August to September in
Bangladesh and February to March in the southern hemisphere (Table 3.9). The seeds
retain their viability for about one month and can be sown directly onto sheltered
intertidal areas, or may be grown in raised nursery beds. Although nursery raising is
not used for large plantings, nursery-grown seedlings are allowed to grow for about
12 months and seedlings of the desired height (10-25 cm) become available for the
next suitable planting season. At this time, the seedlings are gently pulled out of the
ground and packed for transport to previously chosen afforestation sites where they
are usually planted out at a spacing of 1.0 x 1.0 m. Similarly, one-year-old
seedlings may also be collected from the floor of natural forests.
Heritiera spp. In Bangladesh, H.fomes seeds are available in June to July, each
weighing around 10-12 g (Siddiqi et al. 1991b). Mean germination rates of 60% were
noted, with seeds sprouting within 6-45 days when sown onto moist soils in pots or
trays. Seed viability declines rapidly and no seeds germinated after 70 days of storage
under normal atmospheric conditions. In north-eastern Australia, H. littoralis
produces mature seeds from September to December (Duke et al. 1984), each
weighing around 10-15 g. Germination rates are similar to H. fomes.
Hibiscus tiliaceus. This species has a closed-grained, beautifully marked timber
that takes a good polish. The seeds are small (61,750 seeds kg· l ) and abundantly
produced. Germination rates are low, but cuttings strike easily.
Karufelia carufel In China, propagules are collected from February to April
(Liao et al. 1998). Propagules are simply pushed into the soft mud to two-thirds of
their length in the upper intertidal regions. Propagule spacing was either 0.5 x 1.0 or
1.0 x 1.0 m and resulted in survival rates of more than 88%.
Nypa fruticans. The propagule of this species is an aggregated fruit, consisting
of closely packed carpels, containing between 40-120 seeds (Siddiqi et al. 1991 a).
Each seed weighs between 80-120 g and, on average, I kg contains 10-12 seeds. The
mature seeds are buoyant and may sprout while floating. As they appear to have no
dormancy period, they can be designated as cryptoviviparous (Saenger 1982). The
plant commences flowering and fruiting when aged between 3-7 years and, in
Bangladesh, mature seeds are available mainly from February to April (Siddiqi et al.
1991a). After collection, seeds are generally pre-treated by placing them into a
shallow pond where brackish water inundation occurs for about 8 hours each day.
Such pre-treatment increased survival from around 25% to about 60%. Seeds have
visibly sprouted within 5 days and have 25 cm and 15 cm shoot- and root-extensions
within 60 days (Siddiqi et al. 199Ia).
Rhizophora spp. Propagules are available during most of the year although
maximal abundance is generally in late summer (Tables 3.8 and 3.9).
252 Mangrove Ecology, Silviculture and Conservation

Characteristically planted by inserting into the sediment for one-third to one-half of


the length of the propagule, germination is rapid (within 7-20 days) with around
100% germination rates (Siddiqi et aI. 1993).
Sonneratia spp. In Bangladesh and India, mature green fruits of keora
(Sonneratia apetala) are generally collected during August to September and they are
heaped for 20 days to allow the pericarps to decay. They are then rubbed and washed
in water to separate the small seeds from the rotted fleshy portion of the fruits.
About 1 kg of green fruit will yield approximately 275 g or 7,500 seeds (Islam and
Siddiqi 1987). These seeds maintain their viability for about one month (Siddiqi and
Islam 1988). Approximately 200 g of these seeds are broadcast onto intertidal
nursery beds 1.2 m wide and 12 m long, slightly raised above the surrounding
sediment. The beds are usually encircled by low (15 cm) earthen walls which retain
water pumped into the enclosures during unusually dry periods.
Germination onset and success is controlled by salinity which needs to be
maintained below 20%0, above which germination performance declines rapidly
(Siddiqi et al. 1989). Germination trials of seeds from plants introduced to China
indicated that optimal seed germination was obtained when temperatures were above
25°C and salinity was below 10 g L- 1 (Li et al. 1999).
The seedlings are allowed to grow for about 10 months and from each nursery
bed about 2,350 seedlings of the desired height (30-60 cm) become available for the
next suitable planting season, i.e. from July to August. At this time, the seedlings
are gently pulled out of the ground and packed for transport to previously identified
afforestation sites. Generally, one such nursery bed provides sufficient seedlings to
plant an area of 0.4 ha at the usual spacing of 1.2 x 1.2 m (Islam and Siddiqi 1987).
In Bangladesh, S. caseolaris is pretreated and raised in an identical manner to S.
apetala (Siddiqi et al. 1993).

8.5.3 Field Trials


Field trials in which different species are matched against different edaphic
conditions are essential prior to any planting program. For example, Li et al. (1996)
introduced Bruguiera sexangula, Sonneratia caseolaris and Rhizophora apiculata to
Shenzen and North Bays (-22°N), Guangdong, from material collected from
Dongzhai Harbour (_200 N) on Hainan Island in 1993. Three years later, both B.
sexangula and S. caseolaris were growing well (maximal height of 1.5 and 3.0 m
respectively) while R. apiculata was unable to survive at the more northerly site.
Although it is tempting to suggest that the colder conditions were responsible, the
latitudinal change is rather small and it seems more likely that there were some
specific soil conditions which suited some species but not others. Such site specific
characteristics can only be determined through field trials.

8.6 Silvicultural Management of Planted Areas


8.6.1 Objectives for Management
The major objective for managing any plantation or restored area should be to
facilitate natural regeneration, to enhance productivity through fertilization and weed
or herbivore suppression, and to select target areas where some assisted regeneration
is required. The introduction of selected additional species may also be considered to
either increase the biodiversity of the community or to enhance the standing stock
with suitable understorey species.
Mangrove Silviculture and Restoration 253

8.6.2 Site Management


Many mangrove plantations or restored areas require stringent site management,
including temporary regulation of access to minimize damage to the area through
human or animal trampling, grazing, motor vehicles, and so on. In addition, some
form of temporary shelter from current scouring may be needed to assist in
stabilizing the sediments after disturbance and allowing seedlings to establish
themselves. Once the site has been re-stabilized after construction, or once assisted
regeneration has been carried out, the mangroves are likely to require little in the way
of on-going management.

8.6.3 Natural and Assisted Regeneration


Self-regeneration is an important aspect of plantation or restoration project as costs
of establishing plantations or maintaining restored areas will vary significantly
depending on the extent of artificial planting that may be required.
Blanchard and Prado (1995) used ten 30 x 120 m clearcut strips in Rhizophora
mangle forests in Ecuador to study the factors involved in determining natural levels
of regeneration. Each strip was divided into 5 x 5 m subplots in which the abundance
and heights of seedlings were measured. Proximity to plot edges where mature trees
occurred had a highly significant effect on seedling density, particularly within 5 m
of the edges. DBHs of mature trees adjacent to plot edges were positively correlated
with seedling density, and mean seedling density was strongly inversely correlated
with plot elevation over the range studied (range, 2.7-3.6 m above MSL with a tide
range of 2.5-4.0 m above MSL). Interstitial soil salinities did not correlate with
seedling density and appeared to be related primarily to plot elevation. The amount of
logging slash or the abundance of Acrostichum speciosum in the subplots had no
statistically detectable effects on seedling density. On the other hand, seedling mother
trees left in the clearcut strips had more seedlings within close proximity than
occurred in equivalent areas in plots without seedling mother trees (fig. 8.12).
Based on these findings, Blanchard and Prado (1995) concluded that large trees
produce more propagules, resulting in more being available locally for recruitment
without any input from long-distance propagule dispersal. Given the degree of local
recruitment, they suggested that a 20 m wide clearcut strip would supply adequate
regeneration in most cases. On those sites with less tidal inundation «20 tides
month-I) or where bordering trees have a DBH <25 cm, well-shaped, reproductively
active seedling mother trees should be left during felling.
To enhance regeneration in natural sparsely-vegetated areas of the Indian
Sundarbans (Chaudhuri 1990), aerial seeding using Avicennia spp. and Sonneratia
apetala was tried in late August 1989. The trial was established because planting out
in the area was costly and dangerous because of man-eating tigers (Lahiri 1991).
Only barely adequate results were obtained.
Impediments to natural regeneration include infestation of cleared areas by the
mangrove fern Acrostichum. which Watson (1928) described as a 'vegetable pest'.
Acrostichum aureum has been shown to have a wide tolerance range in terms of light
and soil salinity (Medino et al. 1990), facilitated by its ecotypic plasticity on the one
hand, and its osmocompensation ability via cyclitols (Balasubramaniam et al. 1992)
on the other.
254 Mangrove Ecology, Silviculture and Conservation

15~~--------------------~

T
r]t:·
:::::::::::::: T
:.:.:.:.:.:.:. :...:.):.:.:. T
...
::::::::::::::. :Ato
............
.:.......... "."::.::.:.:::.::.::.:
.~ .~ .~j.~j.~j.~j.~j .............. t~~f~
:.':.":.:.:.:.':.':.":.:.':.':.'..: :.' :::.:::.:::.:::.:::.:::.:::.
:=:=:=:::=:=:: ..............
:.:.:.:.:.:.:.:.
.•:..:.. :•.:..:•.:.. ' :.:.:.:.:.:.:.:.:.:..:.:.:'." ::::::::::::::
.............. :..::..::....:..::••..:....:....
:::::::::::::: ::.::::.:::':::::::::.::::.::::::'

::::::::::::::
0~~-r~L-~~.~~~~~.~~
:::::::::::::: ::::::::::::::.
0-5 5-10 >10
Distance from Seedling-Mother-Tree (m)
Fig. 8.12 Rhizophora mangle seedling density (number ha· l ) in relation to distance
from seedling mother trees in strip c1earcuts in Ecuador. Vertical bars are standard
errors. (Data from Blanchard and Prado 1995)

By way of example, natural regeneration is adequate at the lower tide levels


where Acrostichum aureum does not occur (Sukardjo 1987) in the Tanjung Bungin
production forests in the Banyuasin District of southern Sumatra. These mangroves
(predominantly Rhizophora apicuiata and 8ruguiera gymrwrhizaJ have been exploited
for wood and tannin since Dutch colonization in the 16th century. At higher tide
levels, where A. aureum is abundant, seedling establishment of these species was
found to be inadequate. Similarly, Srivastava et al. (1987) found that Acrostichum
aureum and A. specfosum adversely affected natural regeneration of Rhizophora
apicuiata, R. mucronata and Bruguiera parviflora in the Matang Mangrove forest,
Malaysia, while Jawa and Srivastava (1989) found that Acrostichum reduced
regeneration of Rhizophora and Bruguiera in the Rejang Mangrove Forest Reserve in
Sarawak. Wherever the coverage of Acrostichum exceeded 60%, regeneration was
invariably inadequate because of the competitive suppression and subsequent
mortality of seedlings. Control measures (slashing, uprooting, burning and
chemical) were reviewed and, with the exception of chemical means, were found to
be ineffective. Chemical control measures (2-4-5 T, Velpar, Dybar) were effective,
but concern over downstream effects on mangrove biota led to their abandonment.
Strip logging was suggested as a possibility, reducing the non-shaded areas favoured
by Acrostichum.
Elsewhere, however, Acrostichum does not appear to be a problem. For
example, Blanchard and Prado (1995) found that Acrostichum aureum density hid
little effect on juvenile establishment and success on Rhizophora mangle in
experimental strip clearcuts in Ecuador. Chan (1989) also questioned the need to
eradicate Acrostichum speciosum prior to regenerating logged-over areas of
Rhizophora mucronata, suggesting that the growth of seedlings will gradually
suppress the fern. Indeed, Medina et al. (1990) showed that under a closed canopy,
clumps of Acrostichum aureum tended to be larger but were significantly fewer (0.8
as against 3.3 clumps m-2), concluding that A. aureum takes advantage of reduced
Mangrove Silviculture and Restoration 255

salt stress under shady conditions but requires full sun exposure to reach maximal
development, reproductive capacity, and productivity.
Natural regeneration of plantations may also be affected by fauna; quite apart
from propagule predation (see 3.5.2), grazing of established seedlings and saplings
may be locally significant. In Bangladesh, for example, spotted deer (Axis axis)
plays a significant tole (Siddiqi 1996). Ten-month old seedlings of Avicennia
officinalis, Bruguiera sexanguJa and Xylocarpus moluccensis were virtually
eliminated in unfenced experimental plots within two months, suggesting that these
species are unlikely to regenerate naturally where deer populations are significant.
For other species, the effects on survival and height increments were least on
Excoecaria agailoc/w < Heritierafomes < Ceriops decandra (Siddiqi 1996).
As already discussed in chapter 5, biotic interactions can also affect the
regeneration process. The climber Derris trifoliata may seriously affect growth and
survival in some areas (Srivastava et al. 1988) while insect infestations have been
reported to cause failures in recruitment in plantations of Avicennia germinans
(Elster et al. 1999a) and Rhizophora mucronata (Ozaki et al. 2(00). Similarly,
Siddiqi et al. (1995) found that in plantations established along the coastal zone of
Bangladesh, natural regeneration was limited (partly by edaphic changes and partly by
grazing deer) and only Excoecaria agalloc/w appeared to be able to develop a second
rotation crop by natural regeneration alone. In that sense, sustainability of
plantations was questionable.
If indeed there is a need for a program of assisted regeneration, the following
approach is suggested. Propagules need to be collected and either planted out or raised
in the nursery. For planting, the propagules can be scattered over the target areas or
alternately pushed into the substrate. The dispersion of propagules should take place
at a time when the area is not expected to be flooded for a few days, thereby allowing
the propagules to be frrmly established prior to subsequent inundation. It is to be
expected that some losses will occur and it is important to ensure that distribution of
propagules to the site is repeated as necessary. If losses are high, propagules may
need to be held in place temporarily using chicken wire, seagrass matting or encasing
the seedlings (Riley and Kent 1999). In some situations, there is greater survival and
growth of the propagules if they are nursery grown before planting out. This allows
the seedling to develop a healthy root system before implantation. Success rates of
establishment using propagules is clearly site specific but has been found to range
from 30-90%. Nursery-grown seedlings generally show higher success rates than
propagules.
To ensure maximal survival, planting out of seedlings should be done at the
most favourable growth period, i.e. middle summer. At other seasons, a decline of
10-20% in seedling survival can be expected.

8.6.4 Cost of Assisted Regeneration


Costs for any of the described approaches are difficult to quantify. The cost of
restoration of mangrove areas is highly variable depending on such factors as local
labour costs, the characteristics of the site (its accessibility and size), its proximity
to propagule sources, and whether propagules, seedlings or transplants are to be used.
The following costs (Table 8.1) were incurred by the Brisbane airport replanting
program and they provide a relatively reliable guide (Saenger 1996).
Transplantation involved the collection of suitable plants within the airport
site, and their transportation, planting, establishment and maintenance. Two species
256 Mangrove Ecology. Silviculture and Conservation

(Avicennia marina and Aegiceras corniculatum) were planted on 10 m wide benches


on both sides of the tloodway (approximately 6.2 kIn long). A plant density of 41
plants per 100 m2 was used, i.e. spacing of 1.5 x 1.5 m. The total cost (in 1980) of
transplanting small plants was $A228.271 or $4.50 per transplant.
Propagation. planting and establishment of one-year-old nursery-grown
seedlings on a prepared (contoured) site ranged from $A1.28-1.33 when done by a
commercial nursery. and cost $AO.74 when raised by the construction agency. No
reliable data are available for the costs of freshly collected propagules; however. it
would be considerably less than the costs of nursery-grown seedlings.
Table S.1 Costs per plant (in 1980 $AustraIian) for raising and planting of mangroves at Brisbane
International Airport. (Data from Saenger 1996).

Plant material Cost per plant Survival rate


Propagules <$0.50 30-90%
One year old seedlings $0.74 - $1.33 40-60%
Transplants $4.50 SO%

Similar costs for plant materials were reported by Snedaker and Biber (1996)
from the U.S.A. Depending on the density of planting, these costs translate into
approximately $US20.000 ha- I (Saenger 1996). Costs of mangrove replanting were
reported at $US70-122 ha- I in India (Untawale 1996). $USI40 ha- I in Thailand
(Aksomkoae et al. 1993) and $US314 ha- I in Malaysia (Chan 1996) and retlect the
much lower cost of labour.

8.7 Plantation Performance


8.7.1 Optimal Planting Season
In Bangladesh. planting during June to August appears to result in maximal survival
of newly planted seedlings of Sonneratia apetaJa (fig. 8.13). However, not only the
time of planting but the age of seedlings is a contributing factor. and seedlings older
than 12 months generally display low survival rates. In some areas, such as the
Chittagong and Noakhali Divisions. winter planting is considered to result in higher
survival rates but. to date. no experimental data support this view.

100 -r--------::-Su-rv~iv""':'a1~6-mo-n-:lh-.-:afIer:--p~lan-I""'ing---'

80 • Observed
_ Calculated
60

40

20

o
-20 ..I.r,-,r-r,-,r-r-r-r-T"'-T"'I-'-T""'lr-T"T""'lr-T",-,r-r-rl
AMI I AS ONDI F MAMI I AS OND I FM
Months
Fig. 8.13 Percent survival of seedlings of SOII1U!ratia apetalo 6 months after planting
out from the same, even-aged nursery stock over a 24-month period. Both the
seasonal changes in survival and the declining trend due to increasing age of seedlings
can be recognized in the fitted curve (r
= 0.7 P <0.(01). (Data from Islam et at.
1992)
Mangrove Silviculture and Restoration 257

8.7.2 Optimal Initial Spacing


A wide range of plant spacing has been used (see Table 16.6 in Field 1996) in a
variety of mangrove planting programs. Determining proper spacing should be based
on minimizing competition. Thinning will occur naturally and by proper spacing.
artificial thinning can be delayed to later stages in the development of the vegetation
(Islam et al. 1993). For A. marina, a spacing of 1.5 m was used at the Brisbane
airport site (Saenger 1996); after 4 years no evidence of self-thinning was apparent.
Similarly. Toledo et al. (2001) used a spacing of 1.0±0.2 m for Avicennia
germinans in Mexico. obtaining a survival rate of 74% after two years.
The growth performance of Sonneratia apetala planted at various densities is
given in Table 8.2. For timber production without thinning. an initial plant spacing
of 1.2 m appears optimal. However. for coastal protection where features such as
multiple stems or reduced height are desirable. initial plant spacing in excess of 2 m
might be more suitable.
Table 8.2 Mean (%SE) survival. H. DBH. stemwood production (V). and tree quality at different initial
spacings for 5-year-old Sonneratia operata plantations at Barisal. Bangladesh. from three replicated
random blocks. (Data from Siddiqi 1987. 1988. Saenger and Siddiqi 1993)
Spacing Survival H DBH V Unforked trees
(m) (%) (m) (cm) (m) ha· l yol) (%)
0.85 x 0.85 29%14 9.I±O.8 8.8%1.3 16.1%2.3 90%7
1.2 x 1.2 52%9 8.7±O.3 7.9±0.3 18.5%2.7 9Qj:2
1.7 x 1.7 42%15 8.9±0.2 9.7±O.5 11.8%2.2 84%2
2.4 x 2.4 37%21 7.8±O.5 13.6%2.5 5.6%1.2 73%6
3.4 x 3.4 39%21 7.9±0.8 12.5%3.1 2.0%1.0 38%16

Where the products of thinning are economic products (e.g. for charcoal
production in Matang Mangrove Reserve). higher initial planting densities are
generally used (Chan 1996). However. where thinning products have little or no
economic value. it is preferable to adjust initial spacing to reduce the need for
thinning and to optimize plant growth (Islam et al. 1993).

8.7.3 Survival
Because of the highly dynamic nature of the Bangladesh coastline. survival of
mangroves is generally poor and replacement planting often needs to be undertaken
for up to 3 years. In sheltered localities. however. survival is generally around 70%.
Long-term survival is also highly variable. but in experimental plots at Barisal.
survival in 5-year old Sonneratia apetala ranged from 29-52% (Table 8.2).
In mangrove replanting at Orissa, India. survival rates ranged from 25-82% after
two years (Das et al. 1997). Sonneratia apetala showed the highest survival rates in
both monocultures (76%) and mixed stands (82%).

8.7.4 Standing Stock


The general assessment of standing stock or biomass is similar in plantations and in
old-growth mangroves. and many of the techniques developed for natural mangrove
areas can be directly applied to the evaluation of planted mangroves.
The methods of biomass estimation fall into three broad. but overlapping,
categories: clear-felling, representative tree sampling, and the establishment of
allometric relationships.
Clear-felling involves the harvest of the entire above-ground biomass (AGB)
from a specified area. The plant material is usually sorted into the component parts
258 Mangrove Ecology, Silviculture and Conservation

(wood, branches, leaves, fruits and flowers) within the individual species comprising
the stand. The weight of each component and the summed total is converted into the
dry weight value on a unit area. Examples where this approach has been used for
mangrove old-growth forests include Lugo and Snedaker (1974) and Christensen
(1978), while its applicability to mangrove plantations was demonstrated by
Aksornkoae (1975). While efficient in young and small-sized stands, the clear-felling
method becomes more difficult to use for biomass estimation as the size of the
stands or the trees become larger.
Representative tree sampling is particularly suitable where the stands are even-
aged or consist of an obvious series of ages. Under those conditions, suitable
representative trees can be selected for felling and harvesting of the individual
components. The dry weight obtained per tree is then multiplied by the stand density
(trees ha- 1) to estimate the stand biomass. Thus, Briggs (1977) estimated the biomass
from even-aged natural mangrove stands using a limited number of representative
trees, and Choudhuri (1991) used ten representative trees to estimate biomass
production in 5-6 year-old mangrove plantations in the Indian Sundarbans.
Allometric techniques have been widely used to estimate stand biomass in
terrestrial forests (Brown et al. 1989). Such methods involve the establishment of a
relationship between the biomass of whole trees, or their component parts, and some
readily measured parameter such as diameter of the stem at breast height (DBH),
and/or height of the tree (H). Once the allometric relationships have been established,
the technique can be applied in a non-destructive way at other sites; in contrast, the
other two techniques are site-specific and are likely to introduce large errors when
applied at other localities. Allometric relationships between AGB and DBH (and/or
H) have been reported for a range of species of mangroves, including those of high
forestry interest (Golley et al. 1962, Cintr6n and Schaeffer-Novelli 1984, Ong et al.
1985, Putz and Chan 1986, Komiyama et al. 1988. Imbert and Rollet 1989, Clough
and Scott 1989, Sukardjo and Yamada 1992, Amarasinghe and Balasubrananiam
1992b, Fromard et al. 1998).
While various equations have been used to establish allometric relationships
(Woodroffe 1985a, Da Silva et al. 1993, Mackey 1993, Slim et al. 1996, Ross et al.
2001), the power curve:

Biomass =A.DBHB

where A and B are parameters has been most widely used. More often its linear
transformation as:

log (Biomass) =a + Blog(DBH)

where a = Log (A), provides a good description of the relationship between above-
ground biomass (AGB) and DBH in a wide range of forest types. Other workers
(Suzuki and Tagawa 1983, Cintr6n and Schaeffer-Novelli 1984, Komiyama et al.
1988,2000. Tam et al. 1995) have used relationships which include height (H) as a
variable in addition to DBH, usually in the form of:

or its linear transformation:


Mangrove Silviculture and Restoration 259

Log (Biomass) =a + Blog [(DBH2)H]

primarily because the equation used for the estimation of stem volume (V) by
foresters is generally of the form:

As Clough and Scott (1989) have pointed out, however, in dense mangroves
height is not a parameter that can be estimated rapidly for each tree over relatively
large mangrove areas and as the simple form of the relationship (using only DBH)
provides an accurate estimate, there is no need for additional input variables.
The general question of the reliability of allometric relationships for estimating
biomass and their application to forests of similar species in a range of
environmental settings, has been widely considered (Brown et al. 1989, Henry and
Aarssen 1999). Where sampling takes place from relatively dense, closed-canopy
forests, tree shape may be influenced by neighbour effects such as self-thinning of
lower branches. Similarly, the proportion accounted for by the crown may differ
between dense and more open forests where crown expansion is less restricted. Thus,
for example, a crowded tree would be expected to have a smaller DBH and crown than
an open-grown tree of the same height (H).
Theoretically of greater significance is the appropriateness of the statistical
techniques used to derived allometric relationships. For example, all allometric
relationships for mangroves (as for most forest systems; Brown et al. 1989) have
been derived by least square regressions where either DBH (and H) or AGB must be
selected as an independent variable and regressed against the other. The use of reduced
major axis regression, which does not assume either axis to be fixed, is more
appropriate than least square regression (McArdle 1988).
Despite the practical and theoretical shortcomings of allometric relationships,
they nevertheless provide useful and consistent results; internal consistency can be
evaluated for any given species by summing the estimates of biomass for individual
components (derived from their allometric relationships) at a particular DBH, and
comparing this value with the estimate obtained for total AGB at the same DBH. As
Clough and Scott (1989) showed, the independent estimates for total AGB thus
obtained were within 10% for most species over the size-classes sampled.
Selected allometric equations for major species for AGB or other components
(in g dry wt.) in relation to DBH (in cm) for individual trees are provided in Table
8.3. These include significant allometric relationships for the major old and new
world species of mangroves. The similarity between these variously derived
allometric equations can be seen in fig. 8.14.
Selected allometric equations for major species for AGB or other components
(in g, or leaf area in m2) in relation to DBH (in cm) and H (in m), for individual trees
are provided in Table 8.4. These include significant allometric relationships for the
major old and new world species of mangroves.
260 Mangrove Ecology. Silviculture and Conservation

Table 8.3 Allometric equations for various mangroves and their component parts based on OBH.

Species & component Equation Regression Coeff. Ref


A. gemlinans
Above-ground biomass AGB = 14O.00BH2.4 r = 0.97, n = 25 1
Above-ground biomass AGB = 94.20BH2.54 r = 0.99, n = 21 2
Trunk biomass TB = 9.790BH 3•2O r2= 0.97, n = 21 2
Branch biomass BB = 392.60BHI.44 r= 0.94, n = 21 2
Leaf biomass LB = 23.60BH 182 r= 0.95, n = 21 2
La. racemosa
Above-ground biomass AGB = 102.30BH25 r= 0.97, n = 70 I
Aboce-ground biomass AGB = 208.80BH2.24 r = 0.99, n= 17 2
Trunk biomass TB = 76.20BH2.51 r= 0.98, n = 17 2
Branch biomass BB = 224.30BHI.59 r= 0.83, n = 17 2
Leaf biomass LB = 1O.650BH2.o2 r= 0.92, n = 17 2
R. mangle
Above-ground biomass AGB = 177.90BH2.47 r = 0.98, n = 17 2
Trunk biomass TB = 0.9390BH4.03 r= 0.92, n = 16 2
Branch biomass BB = 749.90BH 139 r= 0.88, n = 17 2
Leaf biomass LB = 300.70BHI.25 r= 0.72, n = 17 2
Leaf biomass LB = 27.5510BHI.7914 r= 0.81, n = 26 4
Proproot biomass PB = 12.620BH297 r= 0.97, n = 17 2
Rhizophora spp.
Above-ground biomass AGB = 128.20BH26 r = 0.92, n = 9 1
Above-ground bioma.~s AGB = 105.00BH2.6848 r= 0.99, n = 23 3
B. gymnorhiza
Above-ground biomass AGB = 185.80BH2.3055 r=0.99,n= 17 3
B. parviflora
Above-ground biomass AGB = 167.90BH24167 r = 0.99, n = 16 3
C. australis
Above-ground biomass AGB = 188.50BHB379 r2= 0.99, n = 26 3
X. granatum
Above-ground biomass AGB = 82.30BH2.5883 r= 0.99, n = 15 3

IFromard et al 1998; 2Imbert and Rollet 1989; 3C1ough and Scott 1989; 4Cintffin and Schaeffer-Novelli
1984.

1000
o C&S

bD ¢ S&Y
C 750
Vl
Vl
ro
o Ong
S
0 ~ P&C
iii 500
-0
s=
;::l
m Fromard
0
....
,.,0.0 • Imber!
;> 250
0
.n
~

0
0 5 10 15 20 25 30 35
DBH(cm)
Fig. 8.14 Generalized curve for AGB (kg tree· l ) of Rhizophora against OBH, based
on data from R. apiculata (Clough and Scott 1989). The data for R. mucronata
(Sukardjo and Yamada 1992), R. apiculata (Ong et al. 1985, Putz and Chan 1986), R.
mangle (Imbert and Rollet 1989) and R. mangle and R. racemosa (Fromard et al.
1998) are superimposed to indicate the overall similarity between allometric equations
for Rhizophora.
Mangrove Silviculture and Restoration 261

Table 8.4 Allometric equations for various mangroves and their component parts based on 08H and H.

Species & component Equation Regression Coeff. Ref


All mangroves
Above-ground biomass AG8 = 116.6[(08H)2H]o.8877 r= 0.99, n = 20
Rhizophora spp.
Trunk biomass TB = 4O.36[(08H)2H]o.9660 r=O.99, n=9
8ranch biomass 88 = 1O.46[(08H)2H]o.9103 r=0.98, n=9
Leaf biomass L8 = 69.74[(08H)2H]o.m2 r= 0.89, n = 9
AlFr biomass F8 = 0.3128[(OBH)2H]o.8B2 r=0.90,n=9
Proproot biomass P8 = 3.625[(08H)2H]I.062 r=0.96, n =9
Leaf area LA = 0.3617[(08H)2H]o.5193 r= 0.90, n = 9
Non-Rhizophora species
Trunk biomass T8 = 24.11[(08H)2H]o.9982
88 = 25.63[(08H)2H]o.8534
r= 0.97, n = 11
8ranch biomass r=0.87,n= II
Leaf biomass L8 = 8.189[(08H)2H]o.8067 r= 0.81, n = 11
AlFr biomass F8 = 0.2913[(08H)2H]o.7725 r = 0.45, n = 11
Leaf area LA = 0.08892[(08H)2H]0.7758 r= 0.79, n = 11
R. mangle
Above-ground biomass AG8 = 23.6398[(08H)2HI:902 r = 0.99, n = 26 2
Proproot biomass P8 = 5.4406[(08H)2H]I. r = 0.97, n = 26 2
Leaf biomass L8 = 23.6398[(08H)2H]o.5902 r = 0.80, n = 26 2
C. tagal
Trunk biomass T8 = 26.17[(08H)2H]o.8n2 r= 0.99, n = 10 3
8ranch biomass 88 = 35.55[(08H)2H]o.!I062 r= 0.97, n = 10 3
Leaf biomass L8 = 24.09[(08H)2H]o.86114 r= 0.96, n = 10 3
A. comiculatum
Above-ground biomas.~ AG8 = 31.33[(08H)2H]o.465 r = 0.88, n = 15 4
K. candel
Above-ground biomass AG8 = 651.63[(08H)2H)'-053 r= 0.69, n = 19 4
A. marina
Above-ground biomass AG8 = 123.59[08H)2H]o.529 r=0.20, n =6 4

IKomiyama et aI. 1988; ZCintr6n and Schaeffer-Novelli 1984; lKomiyama et al. 2000; "Tam et al. 1995.

For species with shrubby growth or unusual growth features, allometric


relationships have been established using other measures such as diameters at certain
heights (other than 1.3 m) above the substrate (Slim et al. 1996, Ross et al. 2(01),
or crown vo]ume (Ross et at. 2(01). For well-established forests, however, the
allometric equations based on DBH and/or H, appear to give useful and consistent
results.
Given the above allometric equations, and the wood density data in Table 7.1,
wood volumes can easily be calculated. However, Da Silva et at. (1993) deve]oped a
volume equation for the stems, branches and above-ground roots >3 cm in diameter
for Avicennia schaueriana, Rhizophora mangle and IAguncularia racemosa. They
correlated the diameter above the highest aerial root (DAR in cm) and the height of
this measurement above the soil surface (HOAR in m) with volume (V in m 3) as
follows:

V = (0.02176DAR + 0.04274HDAR - 0.09729)1.86881


=
(~ 0.96, n 50). =
Saenger and Snedaker (1993) examined 43 AGB data sets and used these to
obtain correlations with mean height (H in m) and latitude (L in 0) of the stands to
allow global trends to be identified. Highly significant correlations were found and
these can be used as first approximations of AGB (in t ha- I ) on an area basis rather
than for individual trees or stands:
262 Mangrove Ecology, Silviculture and Conservation

AGB = 10.800H + 34.994


(rZ = 0.60, n = 43)

AGB = 244.994 - 5.570L


=
(rZ 0.47, n 43)=
AGB = 161.405 - 46.39310& (L H- I)
(rZ = 0.66, n = 43)

Annual Iitterfall (ALF in t ha- I y-I) data (91 data sets) were also examined
(Saenger and Snedaker 1993) to identify correlations with H and L of the stands,
allowing estimates to be made of the mangrove stands in terms of litter input to the
system on an area basis.

ALF = 0.342H + 5.976


(rZ = 0.20, n = 91)

ALF = 11.786 - O.I60L


(rZ=O.13, n=91)

ALF = 10.366 - 1.66910& (L H- I )


(rZ = 0.26, n = 91)

These area-based equations have been tested since they were published on a
number of occasions with newly established data sets and were found to provide
reasonable general approximations. Fromard et al. (1998), for example, found good
agreement between estimates derived from these equations and their allometric
estimates of AGB, and Conacher et al. (1996) found good agreement with ALF in
northern Australia. Similarly, Slim et aJ. (1996) found good agreement with their
AGB for Rhizophora mucronata forests. In their Ceriops tagal stands, however,
growing under suboptimal conditions, considerable differences in calculated and
actual AGB were found. Li and Lee (1997) also found reasonable agreement with
ALF for Chinese mangroves although the calculated AGB generally resulted in an
underestimate.
Estimates of trunk biomass from the allometric relationships given above, can
provide an indication of commercial biomass or merchantable volume by using the
ratio of the mean AGB (ha- I or tree-I) and the commercial biomass (ha- I or tree-I)
(Brown et al. 1989). This ratio is clearly dependent on several considerations,
including tree size and the definition of 'commercial biomass'. All trees have some
positive AGB; however, trees below a certain arbitrary commercial size are said to
have no commercial biomass. As the AGB is always greater than the commercial
biomass, the ratio approaches infinity as commercial biomass approaches zero.
However, as trees become larger, proportionately more of the biomass is commercial
and the ratio will decrease. When a tree reaches maximum height, further growth
takes place as diameter and crown increment and the ratio may well approach some
asymptotic value. Clearly, this ratio will depend on the size of the trees composing
the forest.
The quadratic stand DBH (QSDBH) is commonly used as an index for
describing the mean size of trees in a stand (Brown et al. 1989). It is easily calculated
Mangrove Silviculture and Restoration 263

where basal area (BA) and number of stems per unit area (D) are known. The
QSDBH (in m) is the quadratic mean of the diameter of the trees in a stand, or the
diameter of the tree of average BA:

where D is the density of stems (in n ha- I ) and BA is the basal area (in m2 ha- I ).
In their survey of tropical forest inventory data sets, Brown et al. (1989)
standardized the various inventories of commercial biomass according to the QSDBH
of the stands. They found that the ratio of AGB and the commercial biomass were
very high for forests with a low QSDBH, and then approached an asymptote of 1.50-
2.00 for forests with QSDBH ~ about 30 cm. They concluded that undisturbed
mature forests with QSDBH ~ 30 cm have a ratio around 1.50-2.00. The ratio will
vary for disturbed forests, depending on the method of and reason for disturbance. For
example, a forest which has been selectively logged will probably have a lower
QSDBH than before, so the ratio will increase. A recovering secondary forest or
plantation may initially have a high ratio which then decreases with time. Using the
data sets available to them, Brown et al. (1989) estimated the ratios of AGB to
commercial biomass for tropical forests as a function of QSDBH which, with
caution, should be applicable to mangrove forests:

Ratio for QSDBH < 30 cm = 1.75


Ratio for QSDBH ~ 30 cm = exp(5.7671 - 1.5309[lnQSDBH))
~=0.77
Table 8.5 Total AGB (kg tree-I), trunk biomass (TB in kg tree-I), ABGffB, and actual volume of timber
from selected species of mangroves of various DBH classes. (Data from Clough and Scott 1989)

Species (and DBH) AGB TB ABGffB Green


(kg) (kg) Volume
(m 3)
R. apiculatolstylosa
15 em 151 91 1.65 0.113
20 em 327 191 1.71 0.235
25 em 595 338 1.76 0.417
B parvijlora
15 em 117 91 1.28 0.141
20 em 234 182 1.28 0.281
25 em 401 312 1.29 0.480
B. gymnorhiza
15 em 96 74 1.29 0.114
20 em 186 137 1.35 0.232
25 em 310 221 1.41 0.402
C. australis
15 em 106 83 1.28 0.110
20 em 207 162 1.28 0.215
25 em 350 273 1.28 0.364
X. granatum
15 em 91 64 1.42 0.132
20 em 192 131 1.47 0.269
25 em 342 226 1.51 0.465

Using the data from Clough and Scott (1989), we can calculate the ratio of
total- to trunk-biomass which provides an (,optimistic') estimate of the commercial
biomass for the various species sampled. based on the assumption that the entire
trunk biomass is equivalent to commercial biomass (Table 8.5). While this
264 Mangrove Ecology, Silviculture and Conservation

assumption clearly overestimates the commercial biomass, the estimated ratios


suggest that the values of around 1.75 for mature tropical forests (Brown et al. 1989)
can also be applied to mangrove forests.
For more rapid stock assessment, other structural features can be used to
describe the standing stock, including DBH, QSDBH, BA, D and H. Selected
examples of these parameters from natural and planted mangrove stands are given in
Table 8.6.
Table 8.6 Structural characteristics of selected natural and planted (*) mangrove forests.

Location Mean Number Mean Mean Mean Mean Ref.


(Species) Age of density DBH Basal Height
(y) Species (stems ha· 1) (cm) Area (m)
(m 2 ha· 1)
Bermuda
RmlAg 2 2,500 9.6 26.5 6.6 7
RmlAg 2 4,200 5.7 23.6 4.7 7
Costa Rica
L racerrwsa >50 4 1,370 94.0 96.4 16.0
R. mangle >50 2 1,050 52.7 23.2 10.0
Cuba
*L racemosa 15 3,500 8.4 19.4 8.7 12
*R. rrwngle 5 3,167 1.9 0.9 1.8 12
French Guiana
R. mangle 60-70 3+ 780 18.5 33.6 22.7 3
L racerrwsa 5-6 2 11,944 4.7 20.6 7.7 3
L racerrwsa 3-4 2 41,11\ 2.1 13.7 3.5 3
Mexico
A. genninans >50 2 3,120 24.7 15.2 9.0 I
L racerrwsa >50 2-3 1,968 52.7 44.2 12.3 I
Mixed 3 1,407 10.1 32.7 6.7 6
USA, Florida
A. genninans -15 3 5,900 20.8 20.3 6.5 I
R. mangle -15 2 2.867 33.2 25.7 7.5 I
R. mangle >60 3 1,840 11.3 26.3 7.5 4
R. mangle >50 3 2,131 1\.4 28.2 7.4 4
*LalRm 13 3 6,830 2.3 3.2 3.6 4
*L racemosa 6 2 27,700 2.7 18.4 4.8 4
USA, Puerto Rico
L racerrwsa -20 3 2,237 33.3 19.4 13.3
A. genninans -20 3 1,380 39.2 16.9 16.0
Bangladesh
*S. apetala 5 3.61\ 7.9 17.7 8.7 13
Indonesia
Mixed 4+ 2,349 12.8 30.3 2
*R. mucronata 7 4+ 3,270 6.1 12.7 6.8 8
Malaysia
*R. apiculata 40 2+ 660 22.0 25.1 1\.3 9
*R. apiculata 15 2 2,200 11.0 20.9 6.2 9
*R. apiculata 12 1 4,181 6.9 15.6 12.6 10
*R. apiculata 9 1 4,661 5.5 11.1 11.0 10
*R. apiculata 6 1 8,371 3.3 7.2 6.3 10
Thailand
Mixed 10 1,821 35.9 21.8 17
Oman
A. marina 650 20.0 20.4 5.2 15
PNG
RhlBr 26 2,027 11.5 21.0 14
N. /ruticans 15 1,108 10.5 38.0 14
AvISo 23 7,036 6.5 22.0 14
F.S. Micronesia
Fringe - mixed 4 54 40.5 8.5 18.8 16
Riverine - mixed 3 87 61.8 14.5 23.4 16
Basin - mixed 3 55 65.8 11.2 27.2 16
Mangrove Silviculture and Restoration 265

Table 8.6 continued

Location Mean Number Mean Mean Mean Mean Ref.


(Species) Age of density DBH Basal Height
(y) Species (stems ha· l ) (cm) Area (m)
(m 2 ha· l )
Australia
R. apiculata/stylosa 2+ 1.600 20.6 59.0 18
R. apiculata/stylosa 2+ 2.725 16.6 69.0 18
Philippines
Mixed 14 4.210 17.3 33.2 9.4 5
*R. mucroTUlta 28 3+ 13.060 4.7 21.8 7.7 5
Vietnam
*R. apiculata 15 3.950 7.5 17.4 10.1 II
*R. apiculata 6 14,349 3.0 10.1 4.6 II

'pool et at. 1977; 2Atmadja and Soerojo 1991; 3Fromard et at. 1998; 4McKee and Faulkner 2000;
SWalters 2000; 6Ramfrez-Garcfa et aI. 1998; 7Ellison 1997; 8Sukardjo and Yamada 1992; 9Chan 1996;
IOSrivastava et at. 1988; "Hong 1996; 12Padron 1996; 13Siddi~i and Khan 1996; 14Robertson et aI. 1991;
15Fouda and AI-Muharrami 1995; 16Ewel et at. 1998a; I Chansang 1984; 18Clough 1992. RhlBr:
Rhizophora apiculata - Bruguiera parvijlora forests; AvISo: Avicennia - Sonneratia forests; La/Rm:
Laguncularia racemosa - Rhizophora mangle forests; RmlAg: Rhizophora mangle - Avicennia
germinans forests.

8.7.5 Mean Annual Increment


Assessing the perfonnance of any plantation depends largely on the objectives of the
planting exercise, and monitoring of parameters need to be adjusted accordingly.
Generally, however, assessment is usually done annually by monitoring the survival
rate and one or more of the structural characteristics of the stand, generally including
DBH and H, although other parameters such as BA, D, AGB and timber volume 01
in m 3 ha· l ) are commonly used in forest inventories. Thus, for example, the
composition and growth pattern of Rhizophora in variously aged stands in Matang
Mangrove Reserve, planted for timber production, were investigated by Srivastava et
al. (1988). In all, 10 stands aged 6 y (4), 9 y (3) and 12 y (3) were studied and on
average there were 8,371, 4,661 and 4,181 stems ha'\ of Rhizophora in 6, 9 and 12
year-old crops respectively. Mean DBH and H for the respective age groups were
3.26 cm and 6.34 m, 5.50 cm and 10.96 m, and 6.91 cm and 12.62 m. Highest
mortality occurred in 6-9 year-old age group.
In an ecological restoration at Laguna de Balandra, Baja California Sur, Mexico,
the following parameters were deemed to provide an assessment of perfonnance
(Toledo et al. 2001): survival rates of plantlets at weekly, monthly and then 6-
monthly intervals for 4 years, together with data on height increments and the
number of leaves per plant. As natural recruitment to the disturbed site was also
monitored and found to be negligible, a survival rate of planted seedlings of 74%
after 4 years was assessed to be a success. Mean height of the plants after 4 years
was 0.63 m.
Measuring the increasing AGB with variously aged mangrove plantations is
time-consuming and merely provides pseudo-time-series, although allowing some
comparison between the perfonnance of a particular plantation against other
plantations or old-growth mangroves in the region. Data from variously aged
plantations of Rhizophora from Thailand (Aksornkoae 1975, Christensen 1978),
Malaysia (Ong et al. 1981) and Indonesia (Sukardjo and Yamada 1992) are shown in
fig. 8.15, illustrating the generally rapid increase in AGB of plantations under
favourable tropical conditions.
266 Mangrove Ecology, Silviculture and Conservation

These data on AGB from variously aged plantations can be divided by the age of
the plantations to provide a time-averaged increase or mean annual increment (MAl)
in AGB. As shown in Fig. 8.16, these MAls vary from 3.2 t hao! y'! for 5-year old
plantations to 17.1 t ha-! yo! in IS-year old plantations in Malaysia (Ong et al.
1981).

y =234.687LOG(x) - 113.766
f1- =0.650
300

250

200
IX!

~ 150

100

50

0
1 10 100
Age(y)
Fig_ 8.15 Increase in above-ground biomass (in t ha with varying ages of
O
')

plantations_ (Data from Aksornkoae 1975, Christensen 1978, Ong et aL 1981, and
Sukardjo and Yamada 1992)

Although these time-averaged increases in AGB are useful, as fig. 8.16


indicates, there are changes in the MAl over the age of a stand, with an initial
increase up to around 15-20 years followed by a decline as the stand matures and
when biomass accumulation decreases on a yearly basis. Thus, measuring MAl at
shorter intervals will more accurately reflect the current annual increment (CAl) of
the stand at any particular time. While the terms MAl (mean growth rate to any age)
and CAl (instantaneous growth rate at any age) measure growth rates, CAl should be
used for shorter assessment intervals where the actual (presently occurring)
increments in AGB (or any other characteristic) are being measured.
As for AGB, MAl can also be calculated for any other parameters, such as
DBH, BA or H, resulting in DBHMAI, BAMAI or HMAI. These measures are
commonly used for rapid assessment as they are relatively easy to undertake.
In plantations where timber production is an objective, measurement of timber
volumes in permanent plots is used to monitor the stands (Saenger and Siddiqi 1993,
Devoe and Cole 1998). Such detailed forest inventories are costly and time-
consuming and are rarely carried out on an annual basis. Rather, they are performed
at longer intervals and the changes in volume between successive surveys are
expressed as mean (or current) annual increments in volume (VMAI or VCAI in m3
ha-! yo!).
Mangrove Silviculture and Restoration 267

20.0 - . - - - - - - - - - - - - - - - - ,
y = -0.041x2 + 1.342x + 1.101
r2 =0.356 0

15.0
0
0
r;FP
~
,:q
10.0
0
~
0
5.0

O.O~-~~-_r----r_--_r----~--~

o 5 10 15 20 25 30
Age
Fig. 8.16 Changes in AGB MAl (in t ha· 1 y.l) with age of variously aged mangrove
plantations in Indonesia. Malaysia and Thailand. (Data from Aksomkoae 1975.
Christensen 1978. Ong et at 1981. Aksomkoae et at 1993. and Sukardjo and Yamada
1992)

Volumes (or standing stock) of mangrove stands are generally derived from
permanent plots using allometric equations, and expressed as m 3 ha·'. Relatively few
data are available on volume standing stock and even fewer on volume or other
increments (VMAI, DBHMAI, HMAI); some of these have been collated in Table
8.7.
Table 8.7 Mean annual increment in volume. DBH. and height from natural and planted mangrove
forests.
Location Volume VMAI DBH DBHMAI H HMAI Ref.
m3 ha· 1 m3 ha· 1 y.1 cm cmy·1 m my·1
Cuba
R. mangle 16.6 L7 3.6 0.36 3.3 0.3 8
C. erectus 24.8 5.0 5.7 Ll3 5.5 LI 8
A. gemlinans 126.7 6.3 ILl 0.55 7.4 0.4 8
L. racemosa 87.5 5.8 8.4 0.42 8.7 0.4 8
Bangladesh
S. apetala 100.4 13.4 12.1 LI 10.6 LI
S. caseolaris 204.0 28.3 18.2 1.8 8.2 0.8
A. oflicinalis 74.8 8.1 11.2 0.9 6.3 0.5
A. marina 37.3 4.7 8.3 0.8 4.8 0.5
A. alba 15.1 2.3 6.1 0.7 4.4 0.4
B. gymnorhiza 5.5 0.6 3.5 0.3 3.4 0.3
B. sexangula 0.5 0.1 LI 0.1 2.0 0.2
E.agallocha 45.9 5.0 9.4 0.7 6.5 0.5
C. decandra 1.3 0.1
X. moluccensis 4.0 0.5 3.3 0.3 3.8 0.3
268 Mangrove Ecology, Silviculture and Conservation

Table 8.7 continued

Location Volume VMAI DBH DBHMAI H HMAI Ref.


m 3 ha· 1 m 3 ha· 1 y.1 cm cm y.1 m my·1
FS of Micronesia
Overall mean 95.9 4.5 2
Maximum 239.2 17.8 20.6 15.9 2
Chuuk 10.4 0.4 24.4 7.9 2
Kosrae 156.2 7.2 29.1 16.7 2
Pohnpei 76.3 3.5 21.8 11.9 2
Yap 130.2 6.8 23.7 13.5 2
Malaysia
Mean 102 5
Best 210 8.9-9.7 5
B. gymnorhiza 10-20 0.17 3
B. gymnorhiza 20-30 0.20 3
B. gymnorhiza 30-40 0.21 3
R. apiculata 10-20 0.26 3
R. apiculata 20-30 0.28 3
R. apiculata 30-40 0.29 3
R. apiculata 40-50 0.25 3
R. apiculata 50-60 0.24 3
Thailand
Seaward zone 30-35 4
Mid-zone 120 4
Landward zone 50-84 4
Vietnam
R. apiculata 2.6 0.9 1.9 0.62 2.4 0.8 7
R. apiculata 8.8 2.2 2.4 0.61 3.3 0.8 7
R. apiculata 16.2 3.2 3.0 0.61 4.1 0.8 7
R. apiculata 25.0 4.2 3.5 0.59 4.6 0.8 7
R. apiculata 44.1 6.3 4.6 0.66 5.8 0.8 7
R. apiculata 64.5 8.1 5.4 0.67 7.2 0.9 7
R. apiculata 67.4 7.5 5.9 0.65 8.1 0.9 7
R. apiculata 74.0 7.4 6.6 0.63 8.5 0.9 7
R. apiculata 82.3 7.5 6.5 0.59 9.0 0.8 7
R. apiculata 75.9 5.1 7.5 0.50 10.1 0.7 7
Australia
R. apiculata 20.6 0.51 6
R. apiculata 12.6 0.13 6

ISaenl%er and Siddiqi 1993; 2Devoe and Cole 1998; 3Putz and Chan 1986; 4Aksornkoae 1975; 5Noakes
1955; Clough 1992; 7Hong 19%; 8Padron 1996.

8.7.6 Rotation and Thinning Schedules


Determining rotation times and thinning schedules requires some data on the growth
rates (as VMAI) and standing stock (as V), as well as the effect of natural thinning in
the stand or plantation. More importantly, the main products to be obtained from the
stand will determine the rotation time in a commercial operation. Some examples of
various rotation times have been summarized in Table 8.8.
Gong and Ong (1995) have used a demographic approach to examine growth
performances of mangrove stands of different ages (5, 8, 13, 18, 23 and 28 years) in
the Matang Forest, Malaysia. They found that the high density of 15,030
Rhizophora apiculata trees ha- I in the 5 year-old stands with a sharp decline to 9,810
trees ha- I in the 8 year-old stands suggested that the initial stocking was too high.
They recommended that artificial regeneration should be carried out at 1.2 m spacing
only if natural regeneration is less than 50% (rather than 90% as is the current
practice). They also found that high mortalities occurred in the 23 year-old (43%) and
18 year-old stands (29%) and suggested that thinning should be carried out at 12-13
years rather than at 17-18 years which is the current practice. Finally, they found that
Mangrove Silviculture and Restoration 269

biomass of the trees did not increase from 23 (155 t ha- I ) to 28 years (153 t ha- I ) and
they suggested a rotation of 25 years be used instead of the present 30 years.
Table 8.8 Rotations (in years) for various forestry products
Country Fuelwood Poles Sawn timber Pulp
Fence posts Pilings Chips
Charcoal
(lOcm DBH) (2S cm DBH) (40cm DBH)
Bangladesh IS-20 20
FJji IS-2S 40 40
Gambia 30 30 30
India IS-20
Indonesia 20-3S 20
Malaysia IS-30 IS-30 20-2S
Micronesia 2S-S0 70-100 100-140
Myanmar 29S
Philippines 7-IS
Puerto Rico 30
Thailand IS-30 IS-30
Venezuela IS-30 30 30
Vietnam 20
Virgin Islands 2S

Based on Hamilton and Snedaker 1984. FAO 1984. Hussain and Acharya 1994. Gong and Ong 1995.
Devoe and Cole 1997. 1998. and Ak.~ornkoae 1997.

In Bangladesh, thinning is generally not required because of the relatively slow


rate of tree growth, the low returns from the products of thinning, and the loss of
trees in plantations due to stem borer attack. However, in some dense Sonneratia and
Avicennia plantations, thinning is carried out after 9-10 years when up to 50% of the
stems may be removed (Saenger and Siddiqi 1993). Thinning of these plantations
consists largely of removing stunted trees and cutting smaller stems from multi-
stemmed trees.

8.8 Indices of 'Health' in Mangrove Communities


Apart from monitoring the structural development of old-growth or plantation
mangroves as a measure of growth performance, it is also useful to assess the
'health' of the trees, stands or forests. Such assessments are commonly made on the
basis of indicators that have been found to co-exist with certain pathological
conditions or other forms of environmental stress.
A preliminary list of such indices of 'health' for mangroves (Table 8.9) has
been compiled from observations made on neotropical Avicennia and Rhizophora by
S.C. Snedaker, and on palaeotropical representatives of these genera by the author.
Although this listing is tentative, it is intended to provide an easy and rapid
means of assessing a mangrove stand in terms of the presence or absence of certain
characteristics symptomatic of pathological or other stressful conditions. In this
sense, this list may be used as an early diagnostic tool.
270 Mangrove Ecology, Silviculture and Conservation

Table 8.9 Indices of 'health' in mangrove trees, stands and forests: 'healthy' communities will not display
these features.

Aerial roots
proliferation of undersized proproots
twisting and curling of pneumatophores
adventitious aerial roots
death of proproot tips
fissuring or peeling of periderm
Trunks and branches
top-dying of uppermost and outermost sun branches
fissuring and cracking of bark
expanded and/or more numerous lenticels
shortened internode distances
cessation of terminal shoot growth
appearance of trunk sprouts from secondary meristems
Foliage
reduced leaf number per branch
reduced leaf size, twisting and curling
abscission of buds and immature leaves
altered leaf maturation sequence
spotty chlorosis or necrosis
change in leafing and shedding processes
reduced leaf area index
Reproductive structures
change in timing of flowering and fruit set
absent or grossly excessive flowering
deformed seeds and propagules
development failure of fruit
excessive abortion of immature fruit
Regeneration
failure to orient geotropically
seeds and propagules fail to establish primary root system
abnormal growth form.~ in established seedlings
failure to initiate primary branching
chlorosis or necrosis of propagules
9. Conservation and Management

In several parts of eastern tropical Africa


where the shores are mostly of upraised coral limestone.
I have noticed the effect of mangrove in eating away this rock.
but nowhere have I seen it so well
as in the island ofAldabra ... which I surveyed in 1878.
W.J.L. Wharton (1883:76)

9.1 The Need for Mangrove Management


In today's world of burgeoning human popUlations, few resources can be considered
as unlimited. Few, if any, forest communities can be ignored from the standpoint of
conservation. This is equally true of mangroves.
Mangroves have a long history of use by local communities for subsistence
purposes going back over thousands of years (Barrau and Montbrun 1978, Lacerda et
al. 1993b, Cormier-Salem 1994). Early views of mangroves during the modern em
were largely based on ignorance, although, exceptionally, there were some early calls
for their conservation. A phase of enlightenment is recognizable where mangroves
were viewed as scientific curiosities; this phase was soon followed, however, by a
period in which the earliest commercial forest operations became established.
This phase of ruthless exploitation, often through colonial administrations (e.g.
tan-bark in eastern Africa for the German, then the American shoe industry),
continued until around the 1930s when replacement products could be substituted and
when, in any case, the industrial world was again engrossed in war preparations.
Following the war, there was a period of rebuilding followed by a renewed phase of
environmental exploitation (e.g. the mangroves of Puerto Rico) until the mid 1970s,
when environmental impact assessment procedures became common (Heatwole
1985).
In the late 1970s to early 1980s, there came a gradual realization that
mangroves and associated ecosystems must be used on a sustainable basis and be
better conserved (Hegerl 1982, Saenger et al. 1983). Unfortunately, this realization
came at the same time as the large scale development of industrial aquaculture.
particularly for prawns.
As a result of past and present human activities, mangrove resources have been
depleted around the world. Keen competition for use of the remaining estuarine
resources is apparent, and it is likely to intensify further with an increasing
population. The considerable range of competing uses presents the decision-makers
with confusing options. Their task is made all the more difficult by (1) the strong
emphasis on economic evaluation of alternatives in an ecosystem whose biological
values have rarely been quantified, (2) the need to resolve conflicts in resource use in
the interests of a community which, by and large, fails to appreciate the value of the
resource, (3) the absence of a realistic ecological basis on which to evaluate and
manage the various forms of resource utilization, and (4) the temporal and spatial
variability of mangroves and other estuarine resources.
The decision-maker can manage the mangrove ecosystem as a renewable
resource producing fuel, construction material, dyes, tannins, honey and, not least,
providing fisheries products and a range of scientific, educational and recreational
opportunities. Alternatively, the mangrove ecosystem can be considered to be non-
272 Mangrove Ecology, Silviculture and Conservation

renewable and can be exploited for the space it occupies, for agriculture, buildings,
wharves, airports, marinas and roads. Somewhere between these two extremes lie
additional alternative uses for this ecosystem: mariculture, waste disposal, and wood
chips and other forestry products.
In theory then, the ideal decision-maker will manage the ecosystem so as to
leave open in the long term as many resource-use options as possible. In practice,
decisions are usually taken in the short-term interests of expediency, mostly
dominated by a desire for economic and political gains. Making decisions purely by
economic and political yardsticks is acceptable provided that the economic costs are
truly inclusive of all the elements involved (including economic losses owing to
permanent loss of resources for alternative uses). In addition, the community must
be sufficiently well informed to enable it to translate the misuse of a valuable
resource into a political weapon. In general, both the decision-makers and the public
at large have failed, generally, because of a lack of awareness and because of
inadequate public education.

'Spare the Mangrove'

'There has recently been presented to parliament a dispatch from Earl Derby
containing an interesting translation of a pamphlet by Senhor Pedro Soares Calderia,
of Rio de Janeiro, on the injurious effects on the climate produced by the destruction
of mangrove forests. It is stated that where the mangrove flourishes there is a
predominance of what is termed tannin which fosters quantities of fish and shellfish
and in fact every species of inhabitant of the sea. This tannin is said to be a powerful
antidote against putrefaction, and by its stringent nature it solidifies the surface of the
mud and raises it, convening the mud little by little into solid ground. From the leaves,
bark and seeds which fall from the mangrove the mud receives immense quantities of
tannin. By the destruction of the mangrove the mud becomes exposed to the heat of
the sun which encounters a good conductor of heat in the colour of the mud. The heat
that is thus thrown out few living things can bear. The water which the receding tide
leaves in pools on the mud banks is decomposed by excessive heat and with it
enormous quantities of fish, oysters, mussels, molluscs of various kinds. The water is
evaporated by the heat of the solar rays and the mud becomes entirely dry and cracks
on the surface and when the tide returns the layer thus cracked detaches itself floats
away and is deposited on the shore where it forms small hillocks and finishes the
incipient fermentation. In this way the deadliest malarias are engendered. The writer,
in proof of the climatic influence exercised by the destruction of mangrove, lays great
stress upon the fact that when this 'precious shrub' fringed the shores of the Bay of Rio,
yellow fever had never showed itself there, but since the devastation of the marine
forest, fevers and other disorders of an epidemic kind have been prevalent.'

'Pon Denison Times' 30th August 1884

The situation today is that mangrove ecosystems are still not generally
recognized as a valuable national asset by most decision-makers nor by the
community. In many parts of the world, mangroves are being destroyed at very rapid
rates (Saenger et al. 1983, Nurkin 1994, Mastaller 1997, Ellison and Farnsworth
1999, Blasco et al. 2(01). For example, the island of Puerto Rico originally had a
maximum area under mangroves of 26,300 ha. The rate of destruction was low in the
early history of the island (for example, only about 1.7% per decade between 1930
and 1960), but the overall destruction from the maximum extent was 28% by the
mid 1960s. The following decade, especially the five-year period between 1965 and
1970, saw vast destruction, with the largest single mangrove area (325 hal being
88% destroyed and showing little regeneration. Other large swamps were similarly
severely damaged and many of the smaller ones were completely destroyed (Heatwole
Conservation and Management 273

1985). The causes ofthis severe depletion of the mangroves were diverse. including
petrochemical pollution. cutting for charcoal. construction of marinas. and altering
drainage patterns from swamps by road construction and sand mining.
In Hong Kong. the largest mangrove stands were in Deep Bay and inside Tolo
Harbour (Tam et a1. 1997). At Deep Bay. prawn and fish ponds have reduced the
remaining mangrove area to only around 15% of the original area while more than
42% of the original mangroves of Tolo Harbour have been lost to urban arxl
infrastructural development. The new airport and associated facilities resulted in the
destruction of 50 ha of mangroves on the northern shore of Lantau Island.
Ellison and Farnsworth (1999) estimated a rate of annual mangrove loss
throughout the Caribbean of -1%. lower on the islands (-0.2% yo!) than on the
mainland (-1.7% yo!). At the Santiago River mouth in Mexico. Ramirez-Garcia et
al. (1998) have estimated a loss of mangroves of around 32% in a 23-year period due
to intense cattle grazing. This represents an annual loss of approximately 1.4%.
There are many other cases (Table 9.1). In many parts of the world the
destruction of mangroves is proceeding on a large scale, but has seldom been
documented in detail. Places in which massive damage and destruction are now
occurring are Indonesia. the Philippines. Myanmar. Gambia. Nigeria. Ecuador arxl
India (Saenger et a1. 1983, Farnsworth and Ellison I 997b. Blasco et a1. 2001).
Equally important is that many areas are being converted from dense to degraded
mangroves. Blasco et a1. (2001) have shown that dense natural mangroves have
almost totally disappeared in the Mekong delta, while in Myanmar 70% of all
mangroves are either degraded or very degraded.
Table 9.1 Mangrove losses in various parts of the world over differing time intervals.

Country/Region Year Area Year Area %Los.~ Ref.


(km 2) (km2)
Mozambique 1972 4.081 1990 3.%1 3 7
India to Vietnam 1996 19.190 1999 18.150 5 4
Cuba 1%9 4.760 1989 4,480 6 3
Caribbean -1980 14,844 -1990 13.501 9 2
Malaysia 1980 1990 12 1
Cambodia 1970 958 1992 830 13 5
Bangladesh 1%3 6,850 1990 5,870 14 3
Peninsular Malaysia 1979 1,130 1986 890 21 3
Myanmar 1954 2,347 1984 1.809 23 6
Lamu. Kenya 1981 305 1990 235 23 7
Ecuador 1966 2,350 1989 1.775 24 3
North coast. Haiti 1978 60 1989 42 30 8
USA 1958 260 1983 175 33 3
Vietnam 1%1 4,250 19% 2,525 40 1
Indonesia 1969 42,200 1986 21.760 48 3
Thailand 1%1 5,500 1986 2,470 55 1
Philippines 1968 4,480 1996 1.600 64 1
Singapore 1922 7 1989 2 75 3
Puerto Rico 1930 263 1985 30 89 3
Kerala, India 1911 700 1989 3 % 3
·Spalding et al. 1997; ~lison and Farnsworth 1999; 3Mastaller 1997; 4BIa.~ et al. 2001; sViboth and
Ashwell 1996; 6Htay 1996; 7Semesi 1998; KAubC and Caron 2001.

In view of these considerable losses. mangroves are clearly in need of programs


of conservation and management if they and the services they provide are not to be
lost. An economic rationale for mangrove management cannot be denied. Estuarine
wetlands are among the most productive natural systems in the world (Clark 1974);
274 Mangrove Ecology, Silviculture and Conservation

for example, the amount of organic material produced annually by certain temperate
salt marshes in North America exceeds that produced by the world's best strains of
wheat, com or sugarcane (Odum 1973). Because of their high productivity, estuaries
function as nursery and feeding grounds for a very large percentage of coastal fish
taken by commercial and amateur fishers (see 7.2.2 and 7.3.6). Prawn and oyster
production is also largely estuarine-dependent.
In addition, many species of local and migratory waterbirds breed, roost and feed
in estuarine areas. Finally, coastal wetlands store nutrients and regulate their passage
into the estuary and near-shore region (see 7.3.4). Wetlands also have the ability to
remove contaminants (various hydrocarbons and heavy metals) and suspended
sediments from estuarine waters (see 7.3.3 and 7.3.5).
The biological importance of estuaries and estuarine wetlands necessitates their
conservation and sustainable management, especially as humanity is now looking
increasingly to the shallow coastal seas and the estuaries to augment the world's
supply of protein.
Careful planning and sound management are essential to the proper use of any
resource, and this principle also must be applied to the use of coastal and estuarine
resources, including mangroves. The need for conservation and management of
estuaries near the larger cities is particularly urgent as these estuaries are subjected to
the greatest stresses (Bucher and Saenger 1991).

Fig. 9.1 Planning and sound management can provide opportunities that satisfy
conservation, recreational and aesthetic objectives, as does this 'watering hole' in a
mangrove lagoon near Tema, Ghana.

The main benefits from the planning and sound management of estuarine areas
include the following: the maintenance of attractive and readily accessible areas of
high scenic and aesthetic value, suitable for both passive and active recreational
pursuits by all members of the community (fig. 9.1); the conservation of important
wetland and estuarine habitats and of the breeding and nursery grounds of many
marine organisms and waterbirds; the conservation of feeding areas for migratory
birds; the retention of drought refuge habitats which can be used by inland
Conservation and Management 275

waterbirds in dry years; the continuing profitability of shellfish cultivation and of the
inshore and estuarine prawning and fishing industries; the maintenance of a range of
natural ecosystems which, on the one hand, are suitable for teaching and research
purposes and, on the other, constitute rich storehouses of genetic and biotic
diversity; and the reduction of the problems of sedimentation or erosion and the
consequent avoidance of expensive corrective engineering works. Many other direct
but lesser benefits from proper management of estuarine resources could undoubtedly
be added to this list (Cocks 1975, Lugo and Brinsen 1978).
In view of these benefits and the demonstrated extent of mangrove degradation,
mangrove systems clearly are in urgent need of more adequate protection through
management and preservation. From the benefits described above as emanating from
a properly managed resource, it is apparent that the aim should be to maintain the
use of mangroves as a renewable resource, providing fisheries and forestry products
and possessing an inherent amenity value based on their geomorphological,
recreational and scientific characteristics. Only the most pressing and essential
community demand should be considered to justify treatment of the mangrove
resource as non-renewable (Saenger et al. 1983, Ellison and Farnsworth 1999). Canal
estates, garbage tips, industrial land, playing fields or other uses requiring
reclamation would need to be justified in the light of the fact that a valuable
renewable resource was being permanently destroyed.

9.2 Management Frameworks


Having decided that mangrove systems are worthy of prudent management, It IS
essential to determine who should have responsibility for the design, implementation
and ongoing evaluation of mangrove management plans. As mangrove management
and conservation clearly involve land-use allocations, the ultimate responsibility
must reside with the owner. In most jurisdictions this is the state, be it national,
provincial or local. In some jurisdictions, traditional or communal ownership is
paramount, e.g. Samoa (Zann 1999); northern Australia (Bergin 1993); and Tanzania
(Semesi 1998), although the state will generally be involved to some degree. State
responsibility is generally allocated to a particular agency, often with little referral to
the environmental requirements of the resource. Thus, around the world, mangroves
are usually managed by fisheries and/or forestry agencies while other agencies may
have regulatory control over some specific activities.
Mangroves are rarely managed by a single agency as are terrestrial forests,
national parks, mining and fishing. In the absence of a coherent attitude towards this
resource, management decisions in relation to mangroves are often taken in a
piecemeal fashion. The development area is seen in isolation, and the regional
context or the catchment area of the particular site is conveniently ignored. The
result is a constant gnawing away of the resource without taking into account the
full implications of the impact on the resource (Odum 1982). Not only is this a
wasteful process in relation to the resource, but it may give rise to problems that
with proper consideration would not have arisen. A more balanced assessment of
management options for mangrove systems could be provided by a decision-making
unit which had the sole responsibility for them.
The picture is further complicated by a range of international treaties and
conventions that relate to mangroves, either directly or, more often, indirectly. For
example, Ellison and Farnsworth (1999) have found that at least eleven international
treaties and conventions could be applied to conserve or sustainably use these forests
276 Mangrove Ecology, Silviculture and Conservation

in the Caribbean. Furthennore, numerous bilateral or regional agreements also exist.


For example, bilateral agreements between Australia and Japan (JAMBA; Japan-
Australia Migratory Bird Agreement) and China (CAMBA; China-Australia
Migratory Bird Agreement) to protect the stop-over locations for migratory birds
shared by these countries has led to the reservation of large mangrove areas (and other
estuarine habitats) in Australia.
To develop effective management plans for mangrove resources, it is necessary
to relate them to management problems of the adjoining tidal lands and estuarine
waters. Mangroves must be viewed as a part of a complex estuarine system of inter-
related habitat and dependent biota which, in turn, is maintained by natural drainage
patterns and rates of freshwater discharge from the catchment on the one hand and the
natural tidal and salinity regimes on the other. It is the natural movement of water
that provides the essential linkage of the terrestrial and aquatic elements in these
coastal ecosystems (Clark 1974, Howarth et al. 1991). Thus, in planning the
management of estuaries it is important to recognize that some activities in the
catchment can have far-reaching effects on associated near-shore regions through their
influence on the quality of the water in the catchment streams. Clearly, then, the
catchment of an estuary should be considered as part of the estuarine ecosystem and
land use in the catchment must be coordinated with the overall aims of estuarine
planning (Duke and Wolanski 2(01).
Management within the physical boundaries established above must proceed
primarily on an ecological basis. Estuaries and mangrove systems cannot be assessed
by a cursory reconnaissance, as Wharton's quote at the head of this chapter
illustrates, nor is the simple application of forestry principles, largely developed to
manage temperate trees, an adequate foundation for the good management of
mangroves. In other words, management planning requires sufficient field data for
each specific mangrove system to enable recognition of those processes, qualities and
organisms which are in need of protection and specifically how these may be
vulnerable to human activities. Although different estuaries and shorelines share
similar geomorphological and hydrological characters and have similar sorts of
wetlands, individual estuaries and shorelines do possess unique characteristics which
further complicate assessment. Social characteristics and their significance also must
be evaluated. It is only after the ecological factors have been adequately assessed,
however, that other factors, be they economic, social or political, should be brought
into focus. Any attempt to work outside an ecological framework must ultimately
meet with difficulties and involve remedial expense or irreversible losses.
Management of mangroves must be based on a philosophy of conservation
which, as a first step. seeks to prevent further destruction of existing mangrove
ecosystems (fig. 9.2). Most importantly. it should recognize the need to devise
management practices which optimize the conservation of mangrove resources in
such a way as to provide for traditional and contemporary human needs. while
ensuring adequate provision of reserves suitable for protection of the diversity of
plant and animal life within them. Being a renewable resource, mangrove ecosystems
must be managed on a sustainable basis (Saenger et al. 1983. Saenger 1987. Saenger
and Bilham 1996). The concept of sustainable use involves sustainable harvest and
economic benefit, and sustainable economic returns. and at the same time
maintaining the ecosystem as close to its natural or original state as possible. This
is difficult to attain except in a few cases such as use for tourism. Consequently,
sustainable use often does not mean the original natural system in its pristine
Conservation and Management 277

condition; a compromise may be reached which allows sustainable yield and


reasonable resemblance to an undisturbed or non-harvested system. However,
preservation or maintaining a completely undisturbed or unexploited state may be a
desirable management policy for certain localities or for some parts of extensive
mangroves (fig. 9.11). Such unexploited areas serve as a refuge for fauna and flora
and as a biological resource for restoring areas in which management policies have
failed or accidents have occurred. Preservation of some proportion of a mangrove area
can buffer the area generally and can be an advantageous part of an overall sustained-
use management plan.

Department of Mines,
Forest Branoh,
Sydney, 22nd Maroh,1887.
EXEKPTION OF THE MANGROVE TREE FROM
THE OPERATION OF TIMBER L ICENSE!J OR
PERHITS.
N OTICE i, ilereby ginn thai; und.or the pro.vitions .ot the
2nd and 63rd Timber Regulations, of 18th August, 1884,
the Mangrove 1:'ree ise:rompted Crom the olleration of Wood-
cutter', Licensee or ;Permits, and the felling of Iluch tr.::ea is
prohibited.
FRA.NCIS ABIGAIL.

Fig. 9.2 In the Australian colony of New South Wales. timber was initially managed by
the Forest Branch of the Department of Mines. This included mangroves which were
protected in 1887. Responsibility for mangroves was subsequently transferred to the
Fisheries Department.

The potential for implementation of conservation and management strategies


differs within existing patterns of legislation and governmental organizations
(Saenger et al. 1983, Semesi 1998). A higher potential is discernible in those
administrative systems in which mangroves are regarded as an integral component of
coastal regions and not as botanical curiosities. Decisions concerning the use of
mangroves can then be made in the proper context of mangrove dependency on land-
use in the adjacent water catchment and on the important inter-relationships of
mangroves with estuaries, lagoons and coral reefs.

9.3 Some Specific Management Issues


Many coastal urban centres already have had detrimental effects on nearby estuaries
and their mangroves. Management plans which inventory existing wetlands, seek to
eliminate or reduce the stresses already imposed by urban centres, and prevent similar
conditions developing in other areas should be developed by appropriate authorities.
Such plans should involve the creation of specific management schemes for each of
the stress-producing factors discussed below.

9.3.1 Excessive Extractive Use


In many parts of the world, excessive extractive use is the major stress on mangrove
systems. In Cambodia, it was estimated that an average coastal household consumed
around 2 m 3 y.1 of mangrove firewood for cooking purposes (Viboth and Ashwell
278 Mangrove Ecology, Silviculture and Conservation

1996). Such unsustainable wood extraction for fuel and charcoal production is
widespread and often species-specific (Farnsworth and Ellison 1997b, Semesi 1998,
Aure and Caron 2001, Allen et al. 2(01). For example, Rasolofo (1993) reported the
overexploitation of Rhizophora mucronata, Bruguiera gymnorhiza, Ceriops tagal and
Heritiera littoralis for construction purposes on Madagascar. Avicennia marina is
also extensively cut for firewood to the point where around Tulear, it has been
virtually eliminated. Elsewhere, overexploitation has markedly reduced the
distribution of selected species such as Rhizophora racemosa in West Africa,
Rhizophora mucronata in the Arabian Gulf, and Ceriops tagal ir. parts of
Madagascar.
Heavy exploitation is evident also in fisheries products from mangrove areas.
For example, fishing in mangrove areas is traditional and intense in Madagascar
(Rasolofo 1993). Prawns are captured from the mangroves at annual rates of 3,000 t
by artisanal, and 7,000 t by commercial operations. The mudcrab Scylla serrata is
estimated to have a stock size of 7,500 t of which 1,200 t is harvested annually.
Prawn culture is just starting in Mahajamba Bay on the north-western coast of
Madagascar and some mangrove areas have been lost (Blasco et al. 1998). Elsewhere
in eastern Africa similar practices occur (Semesi 1998).
Whether these rates of harvesting are sustainable, given concomitant mangrove
losses, is doubtful. Certainly the industrial woodchipping of mangroves as occurred
until recently in Sabah and Sarawak (Ong 1995) were not. Ellison and Farnsworth
(1999) found that with the 10% decline of mangrove areas in the Caribbean from the
early 1980s to the early 1990s, the total marine fish catch has shown a similar
decline.

9.3.2 Discharges o/Wastes


The utilization of estuaries and mangrove areas as sinks for the discharge of liquid
wastes is well-established and still-growing. In most instances, regulatory controls
merely license discharges and endeavour to keep them to levels which are assumed
capable of absorption by the water body without any permanent deleterious change.
The work of Nedwell (1975) in Fiji showed that mangroves appear to have the
capacity to absorb high levels of nutrients, particularly those contained in sewage.
He suggested that a suitable tertiary treatment for sewage may well be attained by
simply discharging secondary effluent into shallow retaining ponds in mangrove
areas and allowing the overflow to discharge into the mangroves. This is feasible
only where no industrial wastes are included in the sewage; where there is the
possibility of toxic bioaccumulation because of industrial contamination of domestic
sewage, such use of mangrove areas is risky (Mandura 1997). With only one
exception, Fiji's municipal sewage discharges are associated with mangroves (Jaffar
1993). Treatment ponds are built within mangroves to attain secondary treatment
whereupon the discharge is either dispersed into the mangroves or allowed to flow
into mangrove creeks.
Kelly (1995) investigated the impact of the disposal of secondary treated sewage
effluent on subtropical mangroves dominated by Avicennia marina in Moreton Bay,
Queensland. He compared the sediment and foliar concentrations of nutrients (N and
P) at Tingalpa Creek (27 0 29' S, 153 0 12' E) with a similar control area with no
outfall at Hilliards Creek (27 0 30' S, 153 0 15' E). The Thorneside Sewage Treatment
Plant uses an activated sludge method of treatment and discharges around 5,000-
6,000 kL effluent each month. At the impact site, sediment total N and ammonia N
Conservation and Management 279

were significantly higher than at the control site while no differences were found in
total P. At both sites significant positive linear relationships were found between
sediment concentrations and concentrations in roots, pneumatophores, and young and
mature leaves for total N and total P.
While it seemed that Avicennia marina simply increased its absorption of
nutrients when these were present at high concentrations, no clear growth enhancing
effects were noted. Clough et al. (1983) had found that sewage effluent was beneficial
in terms of biomass but the only enhancement of growth observed by Kelly (1995)
was the extensive growth of fibrous roots throughout the top 12 cm of sediments.
Studies in other experimental and natural systems support the generally useful
role that mangroves can play in terms of sewage treatment (Corredor and Morell
1994, Tam and Wong 1995, Chen and Wong 1996, Wong et al. 1995, 1997a, b).
Other studies have suggested that mangroves may play a similarly useful role in the
treatment of aquaculture effluents (Robertson and Phillips 1995, Trott and Alongi
2(00).
Given that organic wastewaters have beneficial, or at least no harmful, effects
on mangroves, the concerns over industrial wastes, particularly heavy metals,
nevertheless remain. Montgomery and Price (1979) and Tam and Wong (1997)
showed that heavy metals contained in sewage sludge are indeed released and are
largely immobilized as sulfides in mangrove sediments. Small amounts of heavy
metals may be accumulated by species in the mangrove and seagrass ecosystems to
which sewage is discharged (Clark et al. 1997, 1998, Lacerda 1998). Concentration
factors for various elements found in Rhizophora mangle are given in Table 9.2 ..
Concentration factors are usually higher with low external concentrations but they
may indicate the potential for enrichment of selected elements. Excessive uptake or
accumulation of heavy metals by mangroves may cause damage at the cellular level,
or lead to reduced growth and increased mortality. However, evidence to date seems to
indicate that mangroves exhibit high tolerance to heavy metal exposure and
accumulation (Lacerda 1998).
From laboratory-based studies, the seedlings of Avicennia alba, A. marina,
Aegiceras corniculatum, KaruJelia caruJel, Rhizophora mucronata and R. mangle do
not appear to be adversely affected by copper, zinc, lead, cadmium and mercury at
sediment concentrations up to around 500 J,lg g-l (Walsh et al. 1979, Thomas and
Ong 1984, Chiu et al. 1995, Chen et al. 1995, Wong et al. 1997a, b, MacFarlane
2(00); accumulation of these metals occurs mainly in the roots, with only limited
translocation of some metals (Cu and Zn) to the aerial parts of the plants (Tam and
Wong 1997, MacFarlane 2(00). Despite the limited translocation of heavy metals to
the aerial parts of the plants, the metal content of leaves of Avicennia marina in
Australia increases consistently with increasing sediment concentrations (fig. 9.3).
Zn and Cu concentrations show these trends particularly well; the trend is less
apparent for Pb.
The root epidermis of A vicennia marina provides a major barrier to the
transport of Pb only (MacFarlane and Burchett 1999, 2(00). The endodermal
casparian strip was shown to provide a barrier to movement of all three metals into
the stele. Washings from mature leaves contained significantly higher amounts of Zn
and Cu than control plants after one month, suggesting extrusion of both metals
from the glandular trichomes. In addition, salt crystals extruded from the glands onto
the adaxial surface of mature leaves were composed of alkaline metals; Zn in Zn-
treated plants, and Cu in Cu-treated plants. Leaf tissue in seedlings dosed with 4g L- 1
280 Mangrove Ecology, Silviculture and Conservation

Zn showed a decreasing gradient of the metal from xylem tissue, through


photosynthetic mesophyll, to hypodermis, with a subsequent increase in
concentration in the glandular tissues.
Table 9.2 Elemental composition (on dry weight basis) of roots, stems and leaves of Rhizoplwra
mangle, together with the concentration factor (CF) of leaves in relation to the artificial seawater
medium used to culture the piInts. (Data from Jayasekera 1991)

Element mgL· I Mean Concentration <F


in growth Roots Stems Leaves
medium
Metals
Al 1.4 130 mg kg· 1 35 mgkg·1 30 mgkg·1 21
Ca 190 0.63% 1.03% 1.38% 73
K 320 0.57% 0.34% 1.78% 56
Mg 443 0.30% 0.13% 0.41% 9
Na 6,476 3.85% 1.63% 1.96% 3
Sr 0.67 38mgkg·1 57.2mgkg· 1 57.7 mgkg· 1 86
Heavy metals
Co 0.0001 13mgkg·1 7.35 mgkg· 1 7.05 mgk§·1 100,714
O! 0.032 36.2 mg kf"1 12mgkg-1 11.1 mg kg- 347
Fe 0.18 1,100 mg kg- 37.3 mgkg- I 52.9 mgkf"1 302
Mn 0.02 55.4mgkg- 1 117.8 mg kg-I 298mgkg- 14,900
Pb 0.01 15.0 mgkg- I 1,500
Zn 0.007 159 mgkg· 1 9.4mgkg- 1 15.6 mg kg· 1 2,229
Non-metals
N 12 0.8% 1.7% 0.40% 333
P 0.217 685 mgkg- I 817 mgkg- I 1,500 mg kg- I 7,000

A similar gradient was observed across leaf tissue in seedlings dosed with Cu at
4 g L-t, but there was no subsequent increase in Cu concentration in glandular
tisssue. For both metals, leaf cell wall metal concentrations were consistently higher
than intracellular concentrations (MacFarlane and Burchett 1999, 2000). On
individual trees, young leaves generally had higher metal contents than old leaves
(Saenger and McConchie, unpublished data).
Similar findings have been made from field studies of more mature trees (Chiu
and Chou 1991, Sadiq and Zaidi 1994, Che 1999, MacFarlane 2000). The anoxic
soils of mangroves precipitate most heavy metals as virtually insoluble sulfides
which are not generally bioavailable (Silva et aI. 1990, Saenger et al. 1991, Clark et
al. 1997, 1998, Lacerda 1998, Lacerda et aI. 2001). In addition, the formation of
iron-plaques on tree root surfaces as a result of the leakage of oxygen from the
underground roots, provides an efficient barrier to trace metals entering mangrove
roots (Lacerda 1997).
Once accumulated in mangrove tissues, heavy metals may be passed along the
food chain via leaves or leaf detritus, ultimately with a potential for attaining
unacceptably high levels in higher organisms in estuarine food chains. For example,
Nye (1990) estimated that 1,063 kg Cu and 412 kg Zn were exported annually from
mangrove forests in south-east Florida through leaf litter and detritus. In Brazil, on
the other hand, the export of metals through leaf fall was found to be small,
representing less than 0.01 % of the total sediment reservoir (Silva et al. 1998).
Conservation and Management 281

90p-~~~~~~ ____ ~ __ ~ __ ~ __ ~ __--+


6
80
6

50 100 150 200 250 300 350 400


Zn-S

8~--~~~~------------------------+
Y =0.059x + 1.36
7 r2 = 0.46 D

O~--r---r---r---r---r---r---r---r---+
o to 20 30 40 50 60 70 80 90
Pb-S
45~ __ ~~~~~~ __ ~ __ ~ __ ~ __ ~~ __ ~

40 o
35

25 50 75 100 125 150 175 200 225 250


Cu-S
Fig. 9.3 Regression of zinc. lead and copper concentrations (mg kg· l ) in sediments
(Zn-S. Pb-S and Cu-S) and young leaves of Avicennia mo.rina (Zn-L. Pb-L and Cu-L)
in eastern Australia. (Data from Saenger and McConchie. unpublished)
282 Mangrove Ecology, Silviculture and Conservation

Given the propensity of mangroves and their sediments to act as a sink for
heavy metals, from a management perspective it becomes imperative to minimize or
avoid heavy metal input into mangrove systems (fig. 9.4). In particular, planners
should avoid municipal garbage dumps being sited near mangroves (Clark 1998,
Shriadah 1999) and keep industrial effluents separate from sewage: well-treated
sewage effluent with a low metal content does not appear to have an adverse effect on
mangroves (Nedwell 1975, Clough et at. 1983, Kelly 1995).

Fig. 9.4 Municipal garbage dumps. such as the one here in the mangroves of the
Bonny River at Port Harcourt. Nigeria. pose a serious risk of heavy metal
contamination to adjacent mangrove systems.

Thermal wastes also are often discharged into nearshore environments


containing mangroves. Numerous studies of the effects of thermal wastes on
estuaries and mangroves have been reported (Roessler 1971. Thorhaug et at. 1973,
Kolehmainen et at. 1974, Canoy 1975, Saenger 1988). In all of these studies,
marked reductions in the invertebrate fauna was reported, at least in the zone where
mixing of the heated water occurred. Seagrasses also were affected adversely,
although the mangroves themselves appeared to be able to cope with the increased
temperatures. What these changes mean to the ecosystem in the long term is not
known, nor have investigations been undertaken to determine the reversibility of
such induced changes once a power station ceases operation.

9.3.3 Hydrocarbon Contamination


Oil can damage mangroves both through direct toxicity (particularly by the lighter
hydrocarbon fractions, i.e. hydrocarbons below CIS) and by interfering with the root
ventilation system of mangroves whereby oxygen is delivered to their roots.
The light hydrocarbon fraction, representing between 20-50% of most light
crudes, normally evaporates within 1-3 days depending on the thickness of the oil
film, temperature, light intensity, wind speed and tidal movement. Low molecular
weight aromatic compounds such as benzene, xylenes and toluene are considered the
primary toxic components responsible for the initial toxicity of crude oil (Durako et
at. 1993).
Conservation and Management 283

Grant et al. (1993) exposed seedlings of Avicennia marina to fresh oils (light
Arabian and Bass Strait crudes), which were found to cause greater mortality than
aged oils. Youssef and Ghanem (in press) exposed 6-month old seedlings of
A vicennia marina to fumes of the volatile fraction of light Arabian crude oil under
controlled conditions and investigated their response in relation to stomatal
behaviour and salt secretion. Plants were fumigated for 0, 3 and 6 hours at the
beginning of the photoperiod and monitored for total salt extrusion, stomatal
resistance and transpiration over 48 hours following fumigation.
At all salinity levels used, salt extrusion declined significantly with increasing
duration of fumigation. This decline in salt extrusion follows the induced shrinkage
of the central extrusion cells of the salt glands (fig. 9.5) of the attached leaves. While
shorter periods of fumigation appear to be reversible (Youssef and Ghanem, in
press), exposure beyond 12 hours is not, resulting in complete inhibition of salt
extrusion activity.

Fig. 9.5 Shrinkage of the central extrusion cell of the salt glands from attached leaves
of Avicennia marina seedlings fumigated for 0 hours (A). 12 hours (8) and 24 hours
(C). Scale bars 1O~. (Photo. T. Youssef)

Stomatal resistance in the Avicennia marina seedling leaves increased


significantly with increasing fumigation while transpiration correspondingly declined
(Youssef and Ghanem, in press). Fumigation also induced high amplitude stomatal
oscillations in proportion to the duration of fumigation. Stomatal recovery was
variable, depending on the length of fumigation treatment. Seedlings fumigated for 3
hours showed recovery within 48 hours after exposure while those fumigated for 6
hours, took almost twice that time.
These physiological responses of mangrove seedlings to the fumes of the
volatile fraction of light Arabian crude oil suggest that the volatile components of
unweathered oil will significantly inhibit the growth performance of these plants.
Two forms of damage have generally been noted in mangroves coated with oil.
Where oil is present in large amounts, trees usually are defoliated within 1-2 months
which, in most cases, leads to mortality and deforestation that may continue for
several years (Fagbami et al. 1988, Lai et al. 1993, Lamparelli et al. 1997, Duke et
al. 1998). Where the quantities of deposited oil are lower, sublethal effects are more
usual; these include partial defoliation and open canopies, reduced growth rates, and
changes in species composition (Duke et al. 1997). Duke et al. (1997) found that
only 18% of trees oiled in Bahia Las Minas on the Caribbean coast of Panama
eventually died.
Youssef et al. (2000) reported on the effects on mangroves of a 4,000 t spill of
crude oil in January 1998 at Umm-AI-Qwain Bay in the United Arab Emirates.
Some 20 km of monotypic stands of Avicennia marina were coated with crude oil up
to a height of 1.2 m (mean tidal range); seedlings and saplings were completely
covered. Massive defoliation commenced within a month, followed by the formation
284 Mangrove Ecology, Silviculture and Conservation

of new leaves in the following 8-10 months. Reproductive output during the
following flowering and fruiting season (September - October) was significantly
reduced as a result of the resource diversion towards vegetative growth. The mean
fresh weight of propagules produced by oiled trees was half that of those produced by
non-oiled trees while the number of propagules produced was one-sixth that of the
control trees. When grown under non-oiled conditions, however, propagules from
oiled and non-oiled trees showed no differences in growth performance.
Finally, as has been noted earlier (Snedaker et al. 1981, BOer 1993), the
frequency of anomalous root growth increased significantly 7 months after the spill.
Multiple branching of the pneumatophores was the most frequently observed root
anomaly.
During 1982-83, a series of oil spills contaminated stands of Avicennia marinLJ
in the northern Red Sea (Dicks 1986). The oil was viscous and weathered, and
formed a complete coating over pneumatophores. Although death of the mangroves
had been anticipated, the majority of the trees survived. Field observations suggest
that sediment drainage characteristics influence pneumatophore density for this
species and playa major role in determining the ability of the trees to survive heavy
oiling. Dicks (1986) suggested that pneumatophore density may provide a simple
index for identifying areas of high susceptibility to oil spill damage.
With the destruction of over 730 oil wells by the retreating Iraqi forces in
February 1991, huge volumes of hydrocarbons were released directly into the marine
environment while additional volumes entered the marine environment indirectly as
fall-out from the numerous oil fires. Together with shipping losses and the
destruction of oil processing facilities, it is estimated that around 10.8 million
barrels of oil were released into Arabian Gulf waters over the period from January to
June 1991, oiling the shorelines of Kuwait and part of Saudi Arabia, at least as far
south as Abu Ali (Saenger 1994a). Large-scale mortality of mangroves was expected.
and data from Qurma Island (27 0 08' N) showed mangrove survival after 2 years to be
around 30% of those mangroves present at the time of oiling. A similar survival rate
was noted in experimental Avicennia marina plantations (fig. 9.6) even further north
at Kbafji (28 0 25' N), and surviving seedlings were about 30-60 cm high two years
later.
Oil spill effects may be localized or extensive, depending on the amount and
extent of oil drifting into mangroves. Bums and Codi (1998) investigated a localized
spill in north-eastern Queensland, concluding that, despite the locally high
hydrocarbon concentrations in the mangrove sediments, no large-scale biological
damage occurred.
Because mangrove sediments are usually anoxic, microbial breakdown of
hydrocarbons is generally slow (Scherrer and Mille 1989, Oudot and Dutrieux 1989).
On the other hand, dissolution processes are fairly rapid, particularly for removing
the lighter hydrocarbon fractions (Le. the C 1rC 26 n-alkanes). For example, Bums and
Codi (1998) showed that within one month, hydrocarbons in mangrove sediments
had lost a considerable proportion of the more soluble light aromatics, including
virtually all n-alkanes and polyaromatic hydrocarbons (PARs). They recommended
that if the oil spill is small enough not to cause tree death, removal of the source of
contamination followed by minimal intervention is the best management option.
Oudot and Dutrieux (1989) similarly advised against the disturbance of the surface
layers of oiled mangrove sediments. If the trees die, then degradation of oil in the
sediments is severely slowed. and microbial decay of oil may take around 20 years
Conservation and Management 285

before non-toxic concentrations are attained (Bums et al. 1993). The benefits of
'bioremediation', usually the addition of fertilizers to enhance microbial activity, are
not universally accepted. However, the addition of oleophilic (hydrophobic) fertilizers
enhanced biodegradation of n-alkanes (Scherrer and Mille 1989) and allowed newly
recruited seedlings to develop roots (Scherrer and Blasco 1989).

Fig. 9.6 Experimental plantations at Khor AI-Maqtq, Khafji, on the Arabian Gulf coast
of northern Saudi Arabia two years after oiling during the Gulf War.

Long-term studies have indicated that a series of phases can be recognized in


terms of the mangrove communty's response and recovery. The initial phase, lasting
up to a year, is characterized by defoliation, death of seedlings and saplings, and
decreases in associated fauna such as crabs and molluscs (Getter et al. 1985, Grant et
al. 1993, Lamparelli et al. 1997). A second phase where mortality of adults becomes
apparent, lasts from 1-3 years (Lamparelli et al. 1997). A stabilization phase
generally follows with little further structural change or tree mortality. This phase
can vary in duration depending on the specific site characteristics, but has been
reported to range from 5-20 years. Finally, a recovery phase follows during which
time measurable improvements occur in the structural parameters of the stand,
including plant densities, basal area and biomass, as well as in the recruitment
success (Lamparelli et al. 1997, Duke et al. 1997).
Given the effects described above, it is generally agreed that mangroves are
particularly sensitive to oil and that they warrant high priority for protection. The
optimal management strategy would appear to be the avoidance of oil spills in or
near mangrove areas. Nevertheless, when spills occur, the following protective
measures seem prudent: (1) recovery of spilt oil by mechanical means offshore from
the mangroves; (2) dispersal (using oil spill dispersants such as Corexit 9527) of oil
offshore from the mangroves; and (3) booming of mangrove areas and embayments
to curtail incursion of oil. While the use of dispersants remains controversial,
experimental evidence indicates that mangrove trees tolerate dispersed oil better than
untreated oil (Teas et al. 1987, Wardrop et al. 1987, Ballou et al. 1989, Duke et al.
1998).
286 Mangrove Ecology, Silviculture and Conservation

If oil enters mangroves, limited clean-up options are available, including


booming and skimming of oil on the water surface in mangrove creeks, removal of
pooled oil from the sediment surface, depressions and channels, water flushing to
concentrate oil where it might be collected. and the use of absorbant materials for
subsequent collection and disposal. Most other forms of intervention are likely to
cause considerable physical damage, particularly in dense mangroves, so should be
avoided.
Rehabilitation of oil-damaged mangroves normally cannot be undertaken
immediately as hydrocarbons remain in the sediments. Time for residual
hydrocarbons to degrade is variable and depends on the Eh of the sediments and the
microorganisms present. 'Bioremediation' may increase the rate of microbial oil
degradation in the sediments (Scherrer and Mille 1989, Scherrer and Blasco 1989).
The time required after an oil spill for the sediments to become sufficiently non-toxic
for the survival of propagules (natural or planted) depends on the type of oil, the type
of sediments, the degree of tidal flushing, and rainfall. However, mangroves can
become established and apparently grow normally at oil spill sites where there is
some residual weathered oil (Duke 1996, Youssef et al. 2(00).

9.3.4 Reclamation and Foreshore Development


Much of the world's population lives in close proximity to the coast; in south-east
Asia alone, with 500 million people, 90% of the population lives within 100 km of
the coast (fig. 9.7). Around the world, numerous urban centres are situated directly at
the mouths of estuaries or adjacent to lagoons. Thus, it is not surprising that the
most widespread destruction of both mangroves and tidal marshes has resulted from
filling the wetlands to create dry land for industrial purposes, airports, port facilities,
and places for people to live and play.
Perhaps some social justification exists for the use of estuaries and mangroves
for essential purposes such as harbours and loading facilities, but destruction of
mangroves appears less justified in the case of airports, and totally unjustified for
housing and playing fields. The practice of dredging artificial canals in estuarine
wetlands and using dredge spoils for land-fill to create canal or key-type residential
subdivisions is accompanied by many insidious effects (Odum 1970, Lindall and
Trent 1975, Cosser 1989, Morton 1989, 1992).
One of the reasons for the decline in the extent of wetlands is the practice of
draining them under the pretext of flood mitigation. The high fertility of floodplain
soils encourages land developments and investments which are often at risk because
of the flood-prone nature of these areas. Future town-planning practice should
prevent inappropriate developments (especially residential development) in flood-
prone areas.
Although it is recognized that some flood mitigation works may be necessary
for the protection of existing urban areas, the need for such schemes should not be
used merely as a justification for draining of swamps for agricultural land. Future
flood mitigation practice should not isolate wetlands from the estuary. Where
wetlands are already cut off from tidal influence by floodgates, a revised schedule of
flood-gate operations should be introduced to allow for tidal flushing of wetlands. As
the natural pattern of drainage provides for the optimum function of the ecosystem,
land-use planners and civil engineers should seek to retain the natural drainage pattern
of the land, especially at the wetland's boundary with the land.
Conservation and Management 287

Fig. 9.7 Urban development, as here in Naha, Okinawa, poses many problems for
wetlands, particular mangroves. Tidal constriction and increased stormwater flows
significantly alter the hydrological regime and ecosystem processes.

The sediments in estuaries (particularly sand) are important to society either


positively as a resource, or negatively as an accumulation that hinders navigation or
causes river flooding. Dredging in estuaries, whether for resource recovery or removal
of 'spoils', is acommon activity which unfortunately has many adverse effects on
estuarine ecosystems (Clark 1974).
Interference with natural vegetation in the catchment through forestry,
agriCUlture, mining, or urban development often leads to an accelerated rate of
sedimentation in the estuary (Duke and Wolanski 2(01). The accumulated sediments
can cause problems which sometimes make it necessary to remove them. Disposal
of the unwanted dredge 'spoil', either in deeper parts of the estuary or on estuarine
wetlands, should not be allowed. Preferably, planners should set aside non-tidal areas
(not wetlands) near estuaries and rivers for the disposal of 'spoils' from such
operations. Once disposed of in this way, steps should be taken to ensure minimal
return to the estuary by soil erosion.
Washing of the sand to remove the fines and shellgrit is a common activity
associated with commercial dredging for sand or gravel. This activity, unless
properly controlled, can cause serious problems in estuaries by decreasing
photosynthesis of aquatic plants (Wolanski and Collis 1976), by smothering benthic
organisms, by reducing the level of dissolved oxygen, and by releasing toxic
substances which have accumulated in the sediments. If the sediment is discharged
over the roots of mangroves, it generally leads to their death.
Dredging at the mouth of a river or lagoon is often required when training its
opening or in cutting or maintaining navigation channels. When such dredging is
undertaken without adequate knowledge of the dynamics of sand movement in the
area, the resulting instability of nearby channels, sand-bars or beaches can have
undesirable effects which may be costly to rectify. Such dredging programmes should
always be preceded by an analysis of the estuary's sand budget.
288 Mangrove Ecology, Silviculture and Conservation

It must also be remembered that the provision of a navigable channel will open
an area which was previously less accessible. The detrimental effect of excessive
human activity can be lessened by controlling or restricting access to certain areas.
When estuarine inventories have been compiled, priorities for wetland and waterway
use can be determined, and these priorities implemented by controlling the types and
extent of access available.
Large-scale alteration of the floor of the estuary resulting from dredging can
produce far-reaching effects on natural flow patterns, turnover rates, and the salinity
regime. Alterations of this type can induce changes in the kinds, abundance and
distribution of estuarine organisms and thus in the fisheries. A study of the hydraulic
consequences of large-scale dredging should precede any major dredging works.
Tin-mining in and adjacent to mangrove areas in Thailand resulted in marked
changes in microtopography which, in tum, led to changes in soil hardness
(Komiyama et al. 1996). This uneven microtopography and the altered soil particle
distribution affected seedling establishment, causing a mortality rate in seedlings of
around 50% over three years, and generally low growth rates in the surviving
seedlings. These effects were ascribed to the high resistance to root penetration on
the one hand, and the low nutrient status of the soils on the other.

Fig. 9.8 A large forest of Bruguiera gymnorhiza near Port Moresby, Papua New
Guinea, has been killed by unintended ponding of water during road construction.

Similarly. the construction of sea walls. revetments. causeways and bridges


may cause changes to the shoreline far distant from the actual structure owing to
reflection of wave energy (Vousden and Price 1985). The gentle slopes found on
most beaches lead to the dissipation of almost all the energy of waves incident upon
them. The steep sides of a sea wall or revetment can easily produce an increase of an
order of magnitude in the energy of the reflected waves. Thus. areas which previously
were not SUbjected to wave activity may now be affected. and shoreline erosion will
inevitably follow.
Bund walls are sometimes constructed around mangrove and saltmarsh areas to
prevent tidal inundation. or may be incidentally built by the construction of
Conservation and Management 289

roadways (Patterson-Zucca 1982, Elster et al. 1999b). With waterlogging, the


mangroves within the bund walls generally are killed within approximately six
weeks (fig. 9.8), and the clearing of dead mangroves usually follows. Gordon (1987)
has described the mangroves of the wet-dry tropical Pilbara region of Western
Australia. During the dry season, groundwater attains salinities of around 90%0.
Because of road construction and other developmental activities, tidal circulation has
been restricted, leading to extreme hypersalinity and widespread mortality.
These bunded areas can then be put to some other use - to provide effluent
ponds for industry, to be filled for playing fields, and so on. The bund walls in the
majority of cases, however, are vulnerable to breaching by floods, or are permeable
to the effluents they contain, or may be designed to overflow after a period of
retention.
Increasingly, around the world, bund walls have been constructed in mangrove
areas for aquaculture projects (Macnae 1968, Ong 1982, Primavera 1995). This
activity has increased dramatically in the past 10-20 years, particularly for high value
crops such as penaeid prawns (fig. 9.9). As Ong (1982:255) has succinctly pointed
out, 'the mangrove is nature's own aquaculture system, with a number of advantages:
an artificial system enjoys relatively easier harvest and selection of particular species,
but the natural system is vastly more stable and less susceptible to disease and
epidemics. Unless the artificial ponds can very significantly surpass the natural
ecosystem, the establishment of aquaculture ponds may be a case of robbing Peter to
pay Paul - with the possible IrlJed cost of having to compensate Peter later'.
Nickerson (1999) has demonstrated the actual amounts taken from Peter to pay Paul
(see 7.5).

Fig. 9.9 Aquaculture development near Surat Thani. Thailand. has been responsible
for significant losses of mangroves along the Gulf of Thailand coast. Disease
outbreaks and acid sulfate soils have eroded much of the economic strength of the
industry.

Despite these equity considerations, prawn farming can be highly profitable and
because of its export earning potential, is often encouraged by government agencies.
For example, from the late 1970s, the Government of Bangladesh encouraged the
290 Mangrove Ecology, Silviculture and Conservation

conversion of the reserved mangrove forests of the Chokoria Sundarbans, near Cox's
Bazaar on the east coast of Bangladesh, to prawn ponds. Today, the entire 8,000 ha
of mangroves of the Chokoria Sundarbans have been replaced by a conglomeration of
prawn ponds but prawn yields are a mere 180 kg hal y.l (Siddiqi 2001).
In peninsular Malaysia, 20-25% of the mangroves area has been given over to
fish and prawn culture (Ong 1982), while for Indonesia, Thailand, Philippines and
Ecuador the figures are even higher (Table 9.3). At present, aquaculture is not
widespread in West or East Africa but it can be expected to increase in the near
future. In those localities where aquaculture is already practised in mangrove areas,
problems have been noted. For example, in Nigeria aquaculture is being actively
encouraged in mangrove areas (Wokoma and Ezenwa, 1982) even though problems
of acidification and land degradation are associated with it (Ajao, 1993). Similarly at
Makali, in northern Sierra Leone, problems have been noted with intense aquaculture
in mangrove areas (Johnson and Johnson, 1993). The first prawn (Penaeus indicus)
mariculture attempts commenced in Kenya in 1978 in a joint FAOIUNDP and
Fisheries Department pilot project. Average production levels only achieved 525 kg
ha- l y-l but the project was responsible for the destruction of 60 ha of mangroves in
Ngomeni Creek (Ruwa 1993). More recently, large mangrove areas have been
allocated to prawn culture in the Rufiji Delta in Tanzania despite strong opposition
from local communities (Semesi 1998).
Table 9.3 Prawn production in major producing countries for 1994. (Data from Tobey et al. 1998)

Country Pond Area Number of Farms Production


(ha) (t)
Indonesia 350,000 60,000 167,410
India 200,000 10,000 91,974
Vietnam 200,000 2,000 36,000
Bangladesh 140,000 13,000 28,763
Ecuador 130,000 1,200 98,731
China 127,000 6,000 63,872
Thailand 70,000 16,000 267,764
Philippines 60,000 1,000 92,647
Mexico 14,000 240 13,454

While aquaculture has enormous potential for production of food, alleviation of


poverty, and generation of wealth for impoverished people living in coastal areas,
there are significant problems associated with it. These include destruction of
productive mangrove areas, poor production levels, susceptibility to poor water
quality and aquatic pollution, poor disease and stock control, and the inequitable
distribution of the benefits derived from the industry (de Graaf and Xuan 1998). From
a broad ecological perspective, aquaculture practices must augment and not endanger
production from natural stock.
Conversion of mangroves for agriculture is also widespread and includes the
cultivation of rice, sugarcane, coconuts and oil palm. All of these forms of
cultivation are constrained by high soil salinities and by the development of acid
sulfate soils. In parts of West Africa and south-east Asia, conversion of mangrove
lands to rice still occurs despite the problems with salinity intrusion (Wolanski and
Cassagne 2000) and acid sulfate soils (Feller et al. 1989). For example, in the
Konkoure River delta in Guinea, rice farms have been established in mangrove areas
(fig. 6.1) of the Konkoure, Sankine and Soumba Rivers. High salinity and slow
flushing (residence time of around 14-30 days) in the Sankine and Soumba River has
caused these ventures to fail. The Konkoure River, with its higher freshwater
Conservation and Management 291

discharge, allows rice to be grown with some success (Wolanski and Cassagne
2(00). Rice cultivation on the landward edge of the mangroves has been attempted
(e.g. in the delta of the Tsiribihina, Madagascar) but it is now largely abandoned
because ofthe salinization of the soil (Rasolofo 1993). The use of mangrove land for
sugarcane production is also widespread, particularly in Australia, Indonesia, Fiji,
Trinidad, Martinique and Puerto Rico.
Conversion of mangroves to salt pans occurs in some areas but, more often,
vegetation-free salt flats behind the mangroves are used. Solar salt production is
largely confined to seasonally arid climates such as north-eastern and north-western
Australia, the north-west coast of India, Senegal, Kenya, Tanzania, Mozambique, El
Salvador, north-eastern Brazil, and Colombia. The traditional salt production in
Benin and neighbouring countries was described earlier (7.2.5).

9.3.5 Mangroves and Global Climate Change


Global climate change is the predicted result of anthropogenic release of CO 2 into
the atmosphere. Natural and anthropogenic release of CH4 and NzO also contributes;
both of these 'greenhouse' gases are released from mangrove sediments (Lu et al.
1999, Corredor et al. 1999). Two major outcomes need to be considered in relation
to mangroves. First, how will increased levels of atmospheric CO 2 affect mangrove
growth and performance? And, second, how might mangroves respond to changes in
sealevel as a result of altered atmospheric CO 2 concentrations?
Flowering plants evolved during the Cretaceous when atmospheric CO 2 was
considerably higher than at the present time. In fact, Beerling and Woodward (1997)
have suggested t.'lat, from a geological perspective, current or future increases in the
concentration of atmospheric CO 2 might be considered as restoring plant function to
that more typically experienced by plants over the majority of their evolutionary
history.
The general plant responses to increased CO 2 have recently been reviewed by
Murray (1995) who noted that there was a narrowing of the stomata leading to an
increased water use efficiency, no consistent changes in stomatal numbers or
densities, a small inhibition of respiration, increased root development, and a slight
increase in flower production with earlier onset of flowering. For most species it
appears probable that the improved water use efficiency exhibited by plants growing
under elevated CO2 was brought about predominantly by the direct effects of CO 2 on
stomatal aperture, rather than by a significant morphogenetic effect on the relative
abundance of stomata. The responses of the root system of plants growing under
elevated CO 2 include a lowered respiration rate, which may confer early growth
advantages. The root system of plants growing under elevated CO 2 may also
contribute to an improved resistance to water stress.
How closely do mangroves conform to these general plant responses?
Farnsworth et al. (1996) grew seedlings of Rhizophora mangle for one year under
ambient (350 J.IL L- 1) and double-ambient (700 J.1L L- 1) CO 2 concentrations. The
seedlings had greater biomass, longer stems, more branching, and greater leaf area
under elevated CO 2 concentrations. Enhanced plant biomass under high CO 2 was
associated with higher root:shoot ratios, relative growth rates, and net assimilation
rates. Maximal rates of photosynthesis were enhanced at high CO 2 levels while
stomatal conductances were lower. Stomatal density decreased as epidermal cells
enlarged in elevated CO 2• Interestingly, after only one year of growth, seedlings
grown under high CO2 were reproductive, produced aerial roots, and showed extensive
292 Mangrove Ecology, Silviculture and Conservation

lignification of the main stem, features not shown by the seedlings grown under
ambient CO2 •
The effects of increased CO 2 may vary with other physical and chemical
conditions under which a species is growing (Ball and Munns 1992). Thus, both
Rhizophora apiculata and R. stylosa both benefit from increased CO 2 via higher
photosynthetic rates, water use efficiencies, and growth rates, but the stimulatory
effect is much greater under low salinity than under high salinity conditions (Ball et
al. 1997).
The effects of increased CO 2 may also differ among habitats and species.
Snedaker and Araujo (1998) found that Florida mangroves responded to increasing
levels (between 6-34%) of atmospheric CO 2 with a highly significant decrease in
stomatal conductance and transpiration, and an increase in water use efficiency
(measured as instantaneous transpiration efficiency - ITE) in the four species tested.
However, there was no significant change in net primary productivity in Rhizophora
mangle, Avicennia genninans and Conocarpus erectus, whereas there was a
significant decrease in Laguncularia racemosa. These responses suggest that with
increasing CO2 levels, Laguncularia might be competitively disadvantaged relative to
the other three species.
Because of the increase in water use efficiency, it also seems that mangrove
vegetation in arid regions of the world might be enhanced in that CO 2 uptake will be
accompanied by a decreased water loss via transpiration. However, this advantage
may be lost if aridity is accompanied by increasing salinity (Ball and Munns 1992).
Should significant global warming occur, changes in sealevel due to oceanic
thermal expansion and subsidence of ice sheets are predicted. Because of their
occurrence at the land-sea interface, mangroves will be among the first ecosystems to
be subjected to such change. We can attempt to predict how mangroves are likely to
respond by (1) experiments on individual plants to measure responses to artifically
altered tidal conditions, (2) measuring changes in mangrove communities subjected
to contemporary changes in tidal regimes, and (3) examining how mangroves have
responded to past changes in sealevel.
Ellison and Farnsworth (1997) grew seedlings of Rhizophora mangle under
controlled conditions for one year after which they were subjected to different levels
of regular tidal inundation in a tank system. Ambient, raised and lowered sealevel
treatments were used so that growth performance of raised and lowered treatments
could be compared with those plants grown at ambient levels. Plants grown under
lowered sealevel conditions were shorter and narrower and produced fewer branches
and leaves than control plants. Plants grown under raised sealevel conditions intially
grew faster than the control plants but growth slowed dramatically once they reached
the sapling stage and, by the end of 30 months in the tanks, the control plants were
larger and growing more rapidly. Other subtle changes were also noted with the
various treatments, demonstrating a range of potential impacts that might result
from both lowered and raised sealevels. Ellison and Fansworth (1997) concluded that
for several key physiological and growth parameters, when changing sealevels and
effects of elevated CO2 were combined, CO 2 enhancement will be offset by negative
responses to changing sealevel.
Relatively little information is available on the response of mangrove
communities subjected to changing tidal regimes over recent decades (Semeniuk
1994, Blasco et al. 1996). Saintilan and Hashimoto (1999) described the
encroachment of mangroves into saltmarsh communities as a result of the abrupt
Conservation and Management 293

increase in sealevel in eastern Australian in the last 40 years. Mangrove


communities subjected to seasonal changes in mean sealevel (see 4.4.1) would repay
detailed study in this regard.
Even small changes in sealevel are likely to have a marked effect on
mangroves, both at the species and the community level; while some have suggested
ecosystem collapse (Ellison and Stoddart 1991, Ellison 1993), others have criticized
these predictions (Snedaker et al. 1994) and offered other scenarios (Pernetta 1993,
Snedaker 1995). Ellison and Stoddart (1991) suggested that mangroves become
stressed by sealevel rises between 9-12 cm 100 y.l, and that rises beyond those rates
would lead to ecosystem collapse. Citing examples of mangrove expansion in
Florida with a sealevel rise nearly twice that predicted by Ellison and Stoddart (1991)
to collapse the system, Snedaker et al. (1994) suggested that mangroves have
survived in Florida with a sealevel rise of 23-27 cm 100 yolo
Analyses of stratigraphic sequences from peat cores indicate that mangroves
gradually moved inland while their seaward fringes died back as sealevels rose during
Holocene transgressions (Woodroffe 1981, 1987, 1990, 1993a, b, Ellison 1989,
1993, 1994, Woodroffe and Grinrod 1991, Fujimoto et al. 1996). Today, inland
migration of mangroves during sealevel rise can occur only where a elevational
gradient exists and where mangrove colonization is not constrained by other land
uses (fig. 9.10).

Fig. 9.10 A seawall with a prote(:tive mangrove fringe of Kandelia candel encloses
arable land adjacent to the Juilong estuary in Fukien Province, southern China. What
protection such walls will offer if sealevels rise as predicted is unknown.

Regional differences in the effects of sealevel rise are also likely (Semeniuk
1994, Ong 1995). For example, after investigating peat accretion in Florida,
Parkinson et al. (1994) concluded that rates of peat accretion in carbonate settings are
around 1.3 mm yol and that with a sealevel rise of as much as 8 mm yol, mangrove
forests are likely to be submerged throughout the wider Caribbean. Similarly,
mangrove communities with high allochthonous sediment inputs are likely to do
better with sealevel rise than those with low external inputs (Pernetta 1993).
294 Mangrove Ecology, Silviculture and Conservation

At the community level, changes due to sealevel rise may be difficult to


distinguish from those caused by anthropogenic factors (Kjerfve and Macintosh
1997). For example, in the Florida Everglades, where sealevels have increased by 9.9
cm at Key West (Maul and Martin 1993), changes in hydrology (McIvor et al. 1994)
and vegetational sequences have been documented (Ross et al. 2(00). Thus, since the
mid-1940s, the boundary of the mixed mangrove-grass- and the Cladium-zones have
shifted inland by 3.3 Ian while the 'white zone' (Rhizophora scrub) has moved inland
by about 1.5 Ian. Although water management practices had local effects, the overall
vegetation changes were ascribed to sealevel rise since 1940, with an extension of
marine and brackish water conditions into formerly freshwater wetlands (Ross et al.
2000).
Other environmental changes associated with global warming include regional
increases or decreases in precipitation and increased storm frequency and intensity
(Pernetta 1993, UNEP 1994). These environmental changes are likely to lead to
increased erosion, particularly along the seaward margin, and increased incidence of
storm and surge damage. Mangrove communities are likely to show a blend of
positive (e.g. from higher levels of CO 2 and temperature) and negative effects (e.g.
increased saline intrusion and erosion) which will largely depend on site specific
factors.
It must also be stressed that anthropogenic climate changes are far from certain.
Nevertheless, prudent management will incorporate mitigative measures to reduce the
build-up of greenhouse gases such as CO 2, as well as adaptive measures for coping
with changes due to global warming, to offset costs and increase benefits. Many
adaptive strategies are effectively the same as those which constitute sound
environmental management, wise resource use, and appropriate responses to present-
day climate variability. These strategies are often in policies and plans for
sustainable development. Adaptive strategies are beneficial even if the climate does
not change as predicted. Strategies which are beneficial for reasons other than climate
change and which are presently justified, should be selected first in developing
responses to climate change. This results in a 'no regrets' strategy. Clearly for
mangroves, such a strategy should include the establishment or extension of
mangrove green belts to provide a natural, self-repairing barrier for coastal
protection, and the afforestation of suitable coastal areas with mangroves.

9.4 Management Tools


A number of management tools have been widely used in mangrove conservation
programs, including reservation and zoning of mangrove areas. Other management
tools have been less widely used but hold considerable promise. Some of these are
briefly reviewed.

9.4.1 The Reserves


There can be little doubt that the use of public lands as reserves is a frontline
weapon for conservation of natural resources including mangroves and has been
widely used. The difficulties with tidal limits of governmental authority can
generally be overcome where strong public support is evident. The Virgin Jungle
Reserve within the Matang Mangrove Forest Reserve (fig. 9.11) is an excellent
example of a reserve established in conjunction with a timber production scheme.
While it fulfils conservation objectives, it also supports the forest operation as a
source of propagules for regeneration programs.
Conservation and Management 295

Another outstanding example is the Ranong Biosphere Reserve, encompasssing


30,309 ha, only one of four such reserves in Thailand. It is the only one to include
coastal habitats such as mangrove forests, seagrass beds, evergreen hill forests and
agricultural land. Extensive mangroves associated with the Kra-buri River, on the
border of Thailand and Myanmar, are included in the reserve. These forests are mainly
secondary forests which have previously been harvested for charcoal. The remaining
old growth forests, including an ancient stand at Had Sai Khao estimated to be 200
years old, are now reserved for research. The reserve contains more than 27 species of
mangrove trees, shrubs and vines. Through Biosphere reservation, the Thai
Government works closely with the local community to find solutions to the
problems facing the reserve, and encourages community involvement in its
management so as to ensure their continued livelihood and economic well-being. In
this way, the reserve links ecology with economics, sociology and politics.

Fig. 9.11 The use of reserves in combination with commercial forest production is
well demonstrated by the Matang Virgin Jungle Reserve shown here. Set aside in
1904, it serves the conservation of genetic and biotic diversity. Structural development
of parts of the reserve can be appreciated by the two figures on the path.

Numerous other reserves have been established around the world (IUCN 1994,
Spalding et al. 1997, Li and Lee 1997), some with unusual features. For example,
the Farasan Island group was proclaimed a protected area in 1989 as part of the
network of conservation areas being established in Saudi Arabia. Although the exact
boundaries have not been defined, the protected area includes all terrestrial and
296 Mangrove Ecology. Silviculture and Conservation

intertidal communities. However. those mangroves in Saudi Arabia not formally


reserved should not be seen as 'sacrificial' areas - rather they are managed on a low-
key basis through public education campaigns and community involvement and
ownership.
It should be emphasized that conservation and management is not invariably
dependent on public acquisition of land. Community-based management of
conservation areas has been shown to be highly effective. and given the long
traditional use of himas reserves managed by local communities in Saudi Arabia
(Grainger and GanadeUy 1984). such approaches should find ready acceptance within
local communities elsewhere.
Other non-legislated forms of reservation include specific areas set aside and
protected for religious or ritual purposes. Thus. sacred groves are tracts of
undisturbed forests in India protected by local communities to avoid the perceived
wrath of the resident God. Grazing and hunting are prohibited and. sometimes. even
removal of dead wood is restricted (Rajendraprasad et al. 1998). Such protected areas
can form important conservation reserves managed by local communities. Similarly.
'fetish' forests in Benin (fig. 9.12) still exist and offer some protection for
mangroves and. in fact. contain some of the last known occurrences of rare species
(e.g. Rhizophora harrisonii in the fetish forest at Adounko) (Akoegninou et al.
1997).

Fig. 9.12 In Benin, West Africa, some mangrove forests (,fetish' forests) are protected
for their religious significance. In these forests, access is very restricted and many
activities are prohibited.

In St. Lucia. in the West Indies. for example. a commercial fuelwood


plantation (using Leucaena leucocephala) and agriculture project was established to
provide local charcoal producers with alternatives to cutting mangrove trees. At the
same time. a co-operative of charcoal-producers was set up to sustainably use and
regulate the communual Mankote mangrove area (Smith and Berkes 1993).
Conservation and Management 297

9.4.2 Zoning of Mangrove Areas


Zoning of mangroves has a long history, going back to the inundation classes used
by Watson (1928) to underpin harvesting prescriptions. Its modern equivalent can be
found in the zoning used in the 'Man and the Biosphere' program which utilizes 'core'
and 'buffer' areas. This zoning approach has been applied to mangrove forests in
Thailand (Aksornkoae 1993, Aksornkoae et al. 1993) which utilizes the three
following zones.
Preservation Zones are protected areas for nature conservation and include areas
for the preservation of economically important flora and fauna; areas of value for
reproduction of flora and faunal breeding areas; areas subject to soil erosion and land
degradation, such as beaches and sand bars, mud flats, islets and coral reefs; areas of
historical or archaeological value; areas that represent a local symbol; national parks
and wildlife sanctuaries; areas for wind protection; areas for environmental and
ecosystem preservationm; and all areas within 2 metres of the natural riverside and
within 75 metres of the coastline.
Economic Zone A covers mangrove areas that are available for utilization of
forest resources and are to be managed for sustainable long-term yield and includes
concession forests; public forests other than concession forests for local use; and
state and public forest plantations.
Economic Zone B is for mangrove areas in which other development is
allowable after careful consideration of the impact on the environment. This zone
includes areas to be converted to agriCUlture, including crops, animal husbandry,
fisheries and salt production; and industrial development including mining, urban and
residential development, seaboard development and other industrial and commercial
activities.
The allocation of mangrove areas to the various zones (Table 9.3) shows that
Economic Zone A accounts for the greatest amount and the Preservation Zone
represents just over 10% of the total mangrove area.
Table 9.3 Mangrove area land use zones in Thailand. (Data from Aksomkoae et a\. 1993)

Land use Zone Area % Total Area


(ha)
Preservation 42,678 11.5
Economic A 199,689 53.6
EconomicS 130,081 34.9
Total 372,448 100

Similar zoning schemes have been used or recommended elsewhere, such as


Brazil (Kjerfve and Lacerda 1993). When combined with remote sensing and
geographical information systems (GIS), such zoning schemes can be readily
monitored and/or adjusted accordingly. To establish a zoning scheme for mangroves,
data is compiled and some ranking is superimposed. Thus, Bacon and AlIeng (1992)
used a ranking system with biological and social attributes to evaluate conservation
priorities of mangroves on islands in the Caribbean. Yipp et al. (1995) and Tam et
al. (1997) also used ranking systems comprising various measures of plant
community structure and socicreconomic attributes to prioritize conservation of
mangroves in Hong Kong.
298 Mangrove Ecology. Silviculture and Conservation

9.4.3 Ecosystem Modelling


Ecosystem modelling and simulation is widely used in natural resource management
and some notable mangrove examples exist. The classical model regarding nutrients
and mangroves in Florida has already been described (see 4.5.3). Such models are
valuable tools to predict mangrove responses to perturbations and to tests various
management actions in terms of outcomes.
Chen and Twilley (1998) recently used a gap dynamic model, which simulates
demographic processes and tree growth in forest gaps, to examine the response of
mangroves to gradients of soil salinity and nutrient availability. Calibrated with data
from the mangroves of the Shark River. Florida. the model was used to predict forest
development and species dominance. The model accurately predicted total basal area
and species-specific contributions to total basal area. as well as the species-specific
size-class distributions. Response curves for each species showed that Laguncu/aria
racemosa dominated in fertile soils with low salinity at early stages of growth, but
its abundance decreased over time while Avicennia genninans increased. Rhizophora
mangle, on the other hand. was dominant in areas of low nutrient availability and
low salinity.
Another recent modelling study by Twilley et al. (1998) indicated that
maintenance of mangrove stands in and around prawn ponds in Ecuador could lessen
the increased eutrophication and turbidity in estuaries cleared of mangroves for prawn
pond construction. Calibrated with water quality data collected over a 14 month
period, the model allows the evaluation of various scenarios for conversion of
mangroves to prawn ponds in relation to changes in environmental quality.
Scenarios tested included conversion of mangrove to prawn ponds in three regions of
the estuary, and the construction of a dam which would change river discharge to the
Guayas River estuary. The use of this model allowed the interaction between land
use practices, ecological functioning of mangroves, and prawn pond operations to be
clearly demonstrated to planners. conservationists and pond operators.
The use of economic models has been discussed in 7.5. All of these modelling
approaches are extremely useful in terms of scenario testing, evaluating the effects
of interventionist management. or to compress long time-scales of successional
forest changes into manageable modules. Physical simulation can also be used to
study ecosystem responses. although it is considerably more costly. time-consuming
and less flexible. Finn (1996) described the establishment of a mangrove mesocosm
incorporated in Biosphere 2. a specially constructed 1.25 ha greenhouse in southern
Arizona. USA. Growth performance and Iitterfall data for the first three years
approximated that for field mangroves in south-west Florida. Comparing the
mangrove mesocosm with natural stands provides insights into the functioning of
these systems, offers complexity, sustainability and experimental condition control,
and provides an educational tool because of its public accessibility.

9.4.4 Raising Public Awareness


As mentioned earlier, lack of public awareness and appreciation of the values of
mangroves is a major contributing factor to their loss in many parts of the world
(Table 9.1). Programs to raise public awareness include facilitated access and
interpretation of mangroves, particularly near urban centres. One of the means
whereby this has been achieved is through mangrove boardwalks (fig. 9.13), which
allow easy access to mangrove forests and enhance the recreational and educational
experience of visitors.
Conservation and Management 299

Fig. 9.13 The provision of boardwalks, such as this one in Cairns, Queensland,
facilitates access into mangrove areas. It is a form of site management that will
probably bring many benetits in terms of greater public appreciation of mangroves.

Boardwalks of various designs and with differing features can now be found in
many mangrove areas. If constructed with appropriate materials, the costs of their
upkeep is small while their life expectancy is around 10-15 years. Visitor
satisfaction is usually high and their environmental impact is likely to be small
(Kelaher et al. 1998a, b).
Another means of raising public awareness which has proven highly successful
in relation to other issues but has hardly been used in relation to mangrove
conservation - featuring mangroves on postage stamps. While mangroves have been
shown incidentally on stamps, only two instances are known where mangroves have
been explicitly illustrated. One is a mini-sheet featuring the mangroves (Rhizophora
stylosa and Hibiscus tiliaceus) and associated fauna of the Pacific Ocean island of
Palau which was produced for World Stamp Expo '89 in 1989 (fig. 9.14) and, thus,
probably of limited postal use. The other, produced by Thailand in 1997, features
Acanthus ilicifoIius as part of a block of four floral designs for domestic postage.

9.S Future Prospects of Mangroves


Around the world, mangroves span 70° of latitude. As a result of this wide
geographic range, these communities extend over tropical and temperate regions, and
possess a concomitant diversity of species and patterns. However, being largely
linear systems occupying habitats with high levels of stress and disturbance,
mangroves are particularly susceptible to degradation and fragmentation.
Threats to mangroves emanate from the immediate pressures of subsistence
users, but particularly from the failure of the general community and of decision-
makers specifically to appreciate the values of this community. Failure to recognize
the ecological values in terms of direct products, indirect products and the
environmental services mangroves provide, has resulted in an attitude that these
communities are wastelands. which, in turn, has led to their exploitation as a non-
300 Mangrove Ecology, Silviculture and Conservation

renewable resource. This tendency has been compounded by the failure both of
developers and of legislators to recognize the extended boundaries of mangrove
communities, with the consequent, but nevertheless frequent, unintended destruction
or fragmentation of them. The failure of ecologists to undertake appropriate research
and to communicate the results to the public and the decision-makers must also
receive its share of the blame.

Fig. 9.14 Philatelic minisheet from the Pacific Ocean island of Palau, featuring an
artistic version of the mangroves and their associated fauna. Each individual stamp
features one species of which the scientific name appears on the back of the sheet. ©
Republic of Palau Postal Service.

Whatever the reason, the exploitation and conversion of tidal wetlands,


particularly of mangroves, continues with undiminished intensity, largely because of
economic and population pressures (Hamilton and Snedaker 1984). Until the
ecological services and capabilities of mangroves are fully appreciated by all levels of
societies, the outlook for these communities appears bleak.
Mangroves are increasingly recognized as appropriate candidates for restoration
or phytostabiIization programs (Salt et al. 1995). Mangrove communities are
effective traps to immobilize water-borne heavy metals, with both field and
laboratory trials concluding that sediments and plants are often successful in
removing all metals from leachates (Saenger et al. 1991, Tam and Wong 1995,
1999, Wong et al. 1997a, b, Clark et al. 1997, 1998). Mangrove sediments possess
the ability to immobilize metals in unavailable forms through formation of
complexes with organic matter, and precipitation with iron and sulfides under
permanently reducing conditions. The plants have the capacity to be extremely
tolerant of metals, which largely accumulate in root tissue (Lacerda et al. 1985,
1993a). Mangrove systems thus possess the ability to have a potential role in trace
metal contamination control via phytostabilization strategies (Salt et al. 1995) with
Conservation and Management 301

minimal export of metals to adjacent ecosystems (Silva et al. 1998, MacFarlane


2000]. Moreover. mangroves can produce a high above- and below-ground biomass
in contaminated areas. thereby maintaining productivity. Specifically. Avicennia
marina may also be employed as a possible biological indicator of environmental
heavy metal concentrations for Pb and Zn in roots and Zn in leaves due to the
proportional dose response relations with sediment concentrations (MacFarlane
2(00).
Mangroves have a range of mechanisms which enable them to survive and grow
at high salt concentrations. The physiological and biochemical basis of salt tolerance
is only gradually being elucidated. We are only just beginning to appreciate some of
the genetic complexities involved. such as the oxygen evolving enhancer proteins
(Sugihara et al. 2(00). The possibility of using mangrove proteins or genes to
enhance the salt resistance of our crop plants is promising.

Fig. 9.15 Drying fronds of Acrostichum aureum by villagers on Hainan Island, China.
Rather than cut mangroves for firewood. the villagers have discovered that the dried
fronds of the mangrove fern have a high heat output and are very convenient for
domestic cooking fires.

Mangroves have an important role in the global carbon cycle, particularly in the
sequestration of CO 2 , They are among the most productive ecosystems and their
carbon stock per unit area can be enormous (Twilley et al 1992, Ong 1993). while
their rate of carbon fixation in tropical areas ranges from 29-56 t C ha· 1 y.1 (Clough
et al. 1997. Clough 1998). The prudent management of mangroves offers the
prospect of a measurable gain in CO2 sequestration.
At the 'Ecotone IV' conference, held in Surat Thani, Thailand, in 1995, one of
the concluding statements was that the best way to protect mangrove resources is
through sustainable use (fig. 9.15). If 'sustainable use' is not confined to direct and
indirect products, but includes sustainable ecosystem and biotechnological services,
then fostering sustainable use of mangroves is the safest approach to conserve
mangroves. While mangrove reservation and zoning will retain an important role in
mangrove conservation, developing the biotechnology to allow us to make full use
302 Mangrove Ecology, Silviculture and Conservation

of the capabilities of mangroves is undoubtedly the best means to develop ani


enhance humanity'S genuine interest in mangroves and their conservation.
10. References
Wir waren am 16. Juli mit Tagesanbruch auf dem Ankerplatz.
gegenuber der Mundung des Rio Manizanes ... Cumana.
die Hauptstadt von Neuandalusien.
liegt 4.5 Kilometer yom Landungsplarz .. .
Wir hatten uber eine weite Ebene zu gehen .. .
Die erste Pjlanze. die wir aUf dem amerikanischen Festland pjliickten.
war die Avicennia tomentosa.
die hier kaum 60 Zentimeter hoch wird.
Alexander von Humboldt (1985:68)

Abe. K.. 1988. Forest as an ecosystem. its structure and function. 6. Arboreal arthropod community of
mangrove forest in Halmahera. Indonesia. In: Ogino. K. and M. Chihara (Eds.). Biological
system of mangroves - a report of east Indonesian mangrove expedition. 1986. Ehime
University. Ehime. pp. 141-151.
Abeysinghe. P.O .• L. Triest. B. De Greef. N. Kaedam and S. Hettiarachi. 1999. Genetic differentiation
between Bruguiera gymnorrhiza and B. sexangula in Sri Lanka. Hydrobio10gia 413:11-16.
AboEI-Nii. M.M .• 2001. Growth and establishment of mangrove (Avicennia marina) on the coastlines
of Kuwait. Wetlands Ecology and Management 9:421-428.
Achmadi. S.• G. Syahbirin. E.T. Choong and R.W. Hemingway. 1994. Catechin-3-0-rhamnoside chain
extender units in polymeric procyanidins from mangrove bark. Phytochem. 35:217-219.
Acosta. C.A. and M.J. Butler. 1997. Role of mangrove habitat as a nursery for juvenile spiny lobster.
Panulirus argus. in Belize. Mar. Freshw. Res. 48:721-727.
Adam. J-G .• 1958. Floristique des paturages sal~s (halophytes et subhalophytes) et v~g~tation des
rizieres du sine-Saloum (S~n~gal). Journ. d·Agric. Trop. et de Botan. Appl. 5:505-541.
Adam. J-G .• 1965. La v~g~tation du delta du S~n~gal en Mauritanie - Ie cordon littoral et l'ile de
Thiong. Bulletin de 1'I.F.A.N. 27:121-138.
Adams. C.D .• 1957. Observations on the fern flora of Fernando Po. I. A description of the vegetation
with particular reference to the Pteridophyta. J. Ecol. 45:479-494.
Adams. E.S .• 1994. Territory defense by the ant Azteca trigona: maintenance of an arboreal ant
mosaic. Oecologia 97:202-208.
Airy-Shaw. H.K.. 1947. The vegetation of Angola. J. Ecol. 35:23-48.
Ajao. EA. 1993. Mangrove ecosystems in Nigeria. In: Diop. E.S .• C.D. Field and M. Vannucci (Eds.).
Proceedings of a Workshop on Conservation and Sustainable Utilization of Mangrove Forests
in Latin America and Africa Regions. Dakar. 20-22 January. 1993. ITTO/ISME Project
PO II 4/90 (F). pp. 13-14.
Akatsu. M.• Y. Hosoi. H. Sasamoto and H. Ashihara. 1996. Purine metabolism in cells of a mangrove
plant. Sonneratia alba. in tissue culture. J. Plant Physiol. 149: 133-137.
Akihisa. T .• P. Ghosh. S. Thakur. H. Nagata. T. Tamura and T. Matsumoto. 1990. 24.24-dimethyl-25-
dehydrolophenol, a 4a-methylsterol from Clerodendrum inerme. Phytochem. 29:1639-1641.
Akoegninou. A .• L.M. Oyede and D.M. Toffi. 1997. La mangrove du ~nin: environnement physique.
v~g~tation et essais de gestion. In: Kjerfve. B.. L.D. Lacerda and E.H.S. Diop (Eds.).
Mangrove ecosystem studies in Latin America and Africa. UNESCO. Paris. pp. 292-306.
Aksornkoae, S.• 1975. Structure. regeneration and productivity of mangroves in Thailand. Unpublished
Thesis. Michigan State University. Michigan.
Aksomkoae. S.• 1993. Ecology and management of mangroves. IUCN. Bangkok. Thailand.
Aksomkoae. S.• 1997. Scientific mangrove management in Thailand. In: Aksornkoae. S.• L. Puangchit
and B. Thaiutsa (Eds.). Tropical Forestry in the 21" Century. Vol. 10: Mangrove Ecosystems.
Kasetsart University. Bangkok. pp. 118-126.
Aksornkoae. S .• N. Paphavasit and G. Wattayakorn. 1993. Mangroves of Thailand: present status. use
and management. ISME Mangrove Ecosystems Technical Reports No. I, pp. 83-133.
Alarc6n, C. and J.E. Conde. 1993. Mangroves of Venezuela. In: Lacerda. L.D. (Ed.). Conservation and
sustainable utilization of mangrove forests in Latin America and Africa regions. Part I - Latin
America. ISME. Okinawa. pp. 211-244.
Albright, LJ .• 1976. In situ degradation of mangrove tissues. N.Z. J. Mar. Freshw. Res. 10:385-389.
Allaway, W.G .• M. Curran. L.M. Hollington. M.e. Ricketts and N.J. Skelton, 2001. Gas space and
oxygen exchange in roots of Avicennia marina (Forssk.) Vierh. var. australasica (Walp.)
Moldenke ex N.C. Duke. the grey mangrove. Wetlands Ecology and Management 9:211-218.
Allen, J .A .• K.C. Ewel and J. Jack. 200 I. Patterns of natural and anthropogenic disturbance of the
mangroves on the Pacific island of Kosrae. Wetlands Ecology and Management 9:279-289.
Allen. O.N. and E. K. Allen, 1981. The Leguminosae. A source book of characteristics, uses and
nodulation. Univ. Wisconsin Press. Madison.
304 Mangrove Ecology, Silviculture and Conservation

Alongi. D.M .• K.G. Boto and A.I. Robertson. 1992. Nitrogen and phosphorus cycles. In: Robertson. A.1.
and D.M. Alongi (Eds.). Tropical mangrove ecosystems. American Geophysical Union.
Washington. pp. 251-292.
Alongi. D.M .• 1998. Coastal ecosystem processes. CRC Press. Boca Raton.
Alongi. D.M .• T. Ayukai. GJ. Brunskill. B.F. Clough and E. Wolanski. 1998. Sources. sinb. and export
of organic carbon through a tropical. semi-enclosed delta (Hinchinbrook Channel. Australia).
Mangroves and Salt Marshes 2:237-242.
Aluri. RJ.S .• 1990. Observations on the floral biology of certain mangroves. Proc. Indian Natl. Sci.
Acad. (Ser. B. BioI. Sci.) 56:367-374.
Alvarez-Lopez. M .• 1990. Ecology of Pterocarpus officinalis forested wetland in Puerto Rico. In: Lugo.
A.E .• M. Brinson and S. Brown (Eds.). Forested wetlands: ecosystems of the world. Vol. 15.
Elsevier Science Publishers. Amsterdam. pp. 251-265.
Alvi. K.A.. P. Crews. B. Aalbersberg and R. Prasad. 1994. Limonoids from the Fijian medicinal plant
Dabi (Xylocarpus). Tetrahedron Letts. 47:8943-8948.
AI-Wetaid. A.H .• A.M. AI-Mansi and C. Miyamoto. 2000. Observations on the distribution and
environmental features of Rhizophora mucronata stands along the Red Sea coast of Saudi
Arabia. Paper presented at the Second Arab International Conference & Exhibition on
Environmental Biotechnology and Coastal Habitats. Abu Dhabi. May 2000.
Amarasinghe. M.D. and S. Balasubrananiam. I 992a. Structural properties of two mangrove forest
stands on the northwestern coast of Sri Lanka. Hydrobiologia 247: 17-27.
Amarasinghe. M.D. and S. Balasubrananiam. 1992b. Net primary productivity of two mangrove forest
stands on the northwestern coast of Sri Lanka. Hydrobiologia 247:37-47.
Andersen. F.O. and E. Kristensen. 1988. Oxygen microgradients in the rhizosphere of the mangrove
Avicennia marina. Mar. Ecol. Progr. Ser. 44:201-204.
Anderson. C. and S.Y. Lee. 1995. Defoliation of the mangrove Avicennia marina in Hong Kong: causes
and consequences. Biotropica 27:218-226.
Andersson. B.• C. Critchley. I.J. Ryme. e. Jansson. E. Larsson and I.M. Anderson. 1984. Modification
of the chloride requirements for photosynthetic O2 evolution: the role of the 23kD polypeptide.
FEB S Letters 168:113-117.
Andrews. TJ. and GJ. Muller. 1985. Photosynthetic gas exchange of the mangrove. Rhizophora stylosa
Griff.• in its natural environment. Oecologia 65:449-455.
Andrews. TJ .• B.F. Clough and GJ. Muller. 1984. Photosynthetic gas exchange and carbon isotope
ratios of some mangroves in North Queensland. In: Teas HJ. (Ed.). Physiology and
Management of Mangroves. Tasb for Vegetation Science. Dr W. Junk. The Hague. pp. 15-
23.
Andriesse. J.P.• N. van Breeman and W.A. Blokhuis. 1973. The influence of mudlobsters (Thalassina
anomala) on the development of acid sulphate soils in mangrove swamps in Sarawak (East
Malaysia). In: Dost. H. (Ed.). Acid sulphate soils. vol. 2. Intern. Inst. Land Reclamation and
Improvement. Wageningen. pp. 11-32.
Anthony. E .• 1989. Chenier plain development in northern Sierra Leone. West Africa. Mar. Geol.
90:297-309.
Aragon. G.T .• A.R.C. Ovalle and J.-P. Carmouze. 1999. Porewater dynamics and the formation of iron
sulfides in a mangrove ecosystem. Sepetiba Bay. Brazil. Mangroves and Salt Marshes 3:85-93.
Araujo. RJ .• I.e. Jaramillo and S.C. Snedaker. 1997. LAI and leaf size differences in two red
mangrove forest types in south Florida. Bull. Mar. Sci. 60:643-647.
Arentz. F .• 1988. Stand-level dieback etiology and its consequences in the forests of Papua New
Guinea. Geo 1. 17:209-215.
Areschoug. F.W.e.. 1902. Untersuchungen tiber den Blattbau der Mangrove-Pflanzen. Bibl. Botanica
56:1-90.
Armstrong. J .• W. Armstrong and P.M. Beckett. 1988. Phragmites australis: a critical appraisal of the
ventilating pressure concept and an analysis of restance to pressurized gas flow and gaseous
diffusion in horizontal rhizomes. New Phytol. 110:383-389.
Arnold. e.A .• 1952. Tertiary plants from North America. I: a Nypa fruit from the Eocene of Texas.
Palaeobotanist 1:73-74.
Ashihara. H .• K. Adachi. M. Otawa. E. Yasumoto. Y. Fukushima. M. Kato, H. Sano, H. Sasamoto and
S. Baba, 1997. Compatible solutes and inorganic ions in the mangrove plants Avicennia marina
and their effects on the activities of enzymes. Z. Naturforsch. 52:433-440.
Assemien, P., 1971. Etude comparative des flores actuelles et quaternaires recent de quelques
paysages vegetaux d'Afrique de rOuest. Unpublished Thesis, University of Abidjan, Abidjan.
Atkinson, M.R., G.P. Findlay, A.B. Hope, M.G. Pitman, H.D.W. Saddler and K.R. West, 1967. Salt
regulation in the mangroves Rhizophora mucronata Lam. and Aegialitis annulata R. Br. Aust. 1.
BioI. Sci. 20:589-599.
References 305

Atmadja. W.S. and Soerojo, 1991. Structure and potential net primary production of mangrove forests
at Grjagan and Ujong Kulon, Indonesia. In: Alcala, A.C. (Ed.), Proceedings of the regional
symposium on living resources in coastal area.~. University of the Philippines, Quezon City, pp.
441-451.
Atta-Ur-Rehman, S. Begum, S. Saied, M.1. Choudhary and F. Akhtar, 1997. A steroidal glycoside from
Clerodendrum inerme. Phytochem.45:1721-1722.
Attiwill, P.M. and B.F. Clough, 1980. Carbon dioxide and water vapour exchange in the white
mangrove. Photosynthetica 14:40-47.
Aube, M. and L. Caron 2001. The mangroves of the north coast of Haiti. Wetlands Ecology and
Management 9:271-278.
Aucan, J. and P.V. Ridd, 2000. Tidal asymmetry in creeks surrounding saltflats and mangroves with
small swamp slopes. Wetlands Ecology and Management 8:223-231.
Ayukai, T., D. Miller, E. Wolanski and S. Spagnol, 1998. Fluxes of nutrients and dissolved and
particular organic matter in two mangrove creeks in northeastern Australia. Mangroves and
Salt Marshes 2:223-230.
Baba, S. and R.I. Onizuka, 1997. Callus induction of five mangrove tree species. In: Kjerfve, B., L.D.
Lacerda and S. Diop (Eds.), Mangrove ecosystem studies in Latin America and Africa.
UNESCO, Paris, pp. 339-347.
Bacon, P.R., 1993. Wetland restoration and rehabilitation in the insular Caribbean. In: Moser, M., R.C.
Prentice and 1. van Vessem (Eds.), Waterfowl and wetland conservation in the 1990s - a
global perspective. IWRB, St. Petersburg (Fla.), pp. 206-209.
Bacon, P.R. and G.P. Alleng, 1992. The management of insular Caribbean mangroves in relation to site
location and community type. Hydrobiologia 247:235-241.
Baijnath, H. and L.M. Charles, 1980. Leaf surface structures in mangroves. I. The genus Rhizophora L.
Proc. Electr. Microscopy Soc. S. Afr. 10:37-38.
Baillaud, E., 1904. La question des paletuviers. Journ. d'Agric. trop. 37:200-206.
Baksha, M.W., 19%. Beehole borer infestation in coastal mangrove plantations in Bangladesh and
possible management options. Wallaceana 77:17-20.
Balasubramaniam, S., K.B. Ranawana and M. Popp, 1992. Mineral content and low molecular weight
carbohydrates of leaflets of the mangrove fern Acrostiehurn aureurn L. Aspects of Plant
Sciences 14:445-450.
Baldwin, A.H., WJ. Platt, K.L. Gathen, 1.M. Lessmann and TJ. Rauch, 1995. Hurricane damage and
regeneration in fringe mangrove forests of southeast Florida, USA. J. Coastal Res. 21:169-183.
Ball, M.e., 1980. Patterns of secondary succession in a mangrove forest of South Florida. Oecologia
44:226-235.
Ball, M.C., 1986. Photosynthesis in mangroves. Wetlands 6:12-22.
Ball, M.C., 1988a. Ecophysiology of mangroves. Trees 2: 129-142.
Ball, M.C., 1988b. Salinity tolerance in the mangroves, Aegiceras eomieulaturn and Avieennia marina.
I. Water use in relation to growth, carbon partitioning and salt balance. Aust. J. PI. Physiol.
15:447-464.
Ball, M.e., 1996. Comparative ecophysiology of tropical lowland moist rainforest and mangrove forest.
In: Mulkey, S.S., R.L. Chazdon and A.P. Smith (Eds.), Tropical forest plant ecophysiology.
Chapman and Hall, New York, pp. 461-496.
Ball, M.C., 1998. Mangrove species richness in relation to salinity and waterlogging: a case study along
the Adelaide River floodplain, northern Australia. Global Ecol. Biogeogr. Lett. 7:73-82.
Ball, M.C. and J.M. Anderson, 1986. Sensitivity of photosystem II to NaCI in relation to salinity
tolerance. Comparative studies with thylakoids of the salt-tolerant mangrove, Avieennia
marina, and the salt-sensitive pea, Pisurn sativum. Aust. 1. PI. Physiol. 13:689-698.
Ball, M.C. and C. Critchley, 1982. Photosynthetic responses to irradiance by the grey mangrove,
Avieennia marina, grown under different light regimes. Plant Physiol. 70:1101-1106.
Ball, M.C. and G.D. Farquhar, 1984a. Photosynthetic and stomatal responses of two mangrove species,
Aegieeras eomieulatum and Avieennia marina to long term salinity and humidity conditions.
Plant Physiol. 74:1-6.
Ball, M.C. and G.D. Farquhar, I984b. Photosynthetic and stomatal responses of grey mangrove,
Avieennia marina to transient salinity conditions. Plant Physiol. 74:7-11.
Ball, M.C. and R. Munns, 1992. Plant responses to salinity under elevated atmospheric concentrations of
CO2. Aust. J. Bot. 40:515-525.
Ball, M.C. and J.B. Passioura, 1994. Carbon gain in relation to water use: photosynthesis in mangroves.
In: Schulze, E.D. and N.M. Caldwell (Eds.), Ecophysiology of photosynthesis. Springer, Berlin,
pp.247-258.
Ball, M.C. and S.M. Pidsley, 1988. Establishment of mangrove seedlings in relation to salinity. In:
Larson, H.K., R. Hanley and M. Michie (Eds.), Darwin Harbour. ANU NARU Mangrove
Monograph 4:123-134.
Ball, M.C. and S.M. Pidsley, 1995. Growth responses to salinity in relation to distribution of two
mangrove species, Sonneratia alba and S. laneeolata, in northern Australia. Functional Ecology
9:77-85.
306 Mangrove Ecology, Silviculture and Conservation

Ball, M.C. and M.A. Sobrado, 1998. Ecophysiology of mangroves: challenges in linking physiological
processes with patterns in forest structure. In: Press, M.C., J.D. Scholes and M.G. Barker
(Eds.), Physiological plant ecology. Chapter 17. Blackwell Science, London, pp. 331-346.
Ball, M.e., W.S. Chow and J.M. Anderson, 1987. Salinity-induced potassium deficiency causes loss of
functional photosystem II in leaves of the grey mangrove, Avicennia marina through depletion
of atrazine-binding polypeptide. Plant Physiol. 14:351-361.
Ball, M.C., MJ. Cochrane and H.M. Rawson, 1997. Growth and water use of the mangroves
Rhiwphora apiculata and R. stywsa in response to salinity and humidity under ambient and
elevated concentrations of atmospheric CO2 , Plant, Cell and Environ. 20: 1158-1166.
Ball, M.C., J.R. Cowan and G.D. Farquhar, 1988. Maintenance ofleaftemperature and the optimisation
of carbon gain in relation to water loss in a tropical mangrove forest. Aust. J. PI. Physiol.
15:263-276.
Ball, M.e., S.E. Taylor and N. Terry. 1984. Properties of thylakoid membranes of the mangroves
Avicennia genninans and Avicennia marina and the sugar beet, Beta vulgaris, grown under
different salinity conditions. Plant Physiol. 76:531-535.
Balling. A. and U. Zimmermann, 1990. Comparative measurements of the xylem pressure of Nicotinana
plants by means of the pressure bomb and pressure probe. Planta 182:325-338.
Bailment, E.R., TJ. Smith and I.A. Stoddart, 1988. Siblings species in the mangrove genus Ceriops
(Rhizophoraceae), detected using biochemical genetics. Aust. Syst. Bot. 1:391-397.
Ballou, T.G., S.e. Hess, R.E. Dodge, A.H. Knap and T.D. Sleeter, 1989. Effects of untreated and
chemically-dispersed oil on tropical marine communities: a long-term field experiment. In:
Proceedings 1989 Oil Spill Conference, API Publications No. 4479, American Petroleum
Institute, Washington, DC, pp. 447-454.
Balsamo, R.A. and W.W. Thomson, 1993. Ultrastructural features associated with secretion in the salt
glands of Frankenia grandifolia (Frankeniaceae) and Avicennia genninans (Avicenniaceae).
Amer. J. Bot. 80:1276-1283.
Balsamo, R.A. and W.W. Thomson, 1995. Salt effects on membranes of the hypodermis and mesophyll
cells of Avicennia genninans (Avicenniaceae): a freeze-fracture study. Amer. J. Bot. 82:435-
440.
Baltzer, F. and B.H. Purser, 1990. Modem allivial fan and deltaic sedimentation in a foreland tectonic
setting: the lower Mesopotamian Plain and the Arabian Gulf. Sedimentary Geology 67:175-
197.
Balu, S. and S. Mathavan, 1995. Wound healing property of Excoecaria agallocha L. J. Econ. Tax. Bot.
19:571-575.
Bandarananayake, W.M., 1998. Traditional and medicinal uses of mangroves. Mangroves and Salt
Marshes 2:133-148.
BaneJji. J .• 1958. The mangrove forests of the Andamans. Trop. Silviculture 20:319-324.
Barrau. J. and e. Montbrun. 1978. La mangrove et !'insertion humaine dans les ecosystemes insulaires
des Petites Antilles: Ie cas de la Martinique et de Ia Guadeloupe. Social Science Information
17:897-919.
Barth, H.• 1982. The biogeography of mangroves. In: Sen, D.N. and K.S. Rajpurohit (Eds.).
Contributions to the Ecology of Halophytes, Tasks for Vegetation Science. Vol. 2. Dr W.
Junk. The Hague, pp. 85-1 10.
Bartlett, A.S. and E.S. Barghoorn, 1973. Phytogeographic history of the isthmus of Panama during the
last 12.000 years. In: Graham, A. (Ed.), Vegetation and vegetational history of northern Latin
America. Elsevier. Amsterdam. pp. 203-99.
Barusseau. J-P., E.H.S. Diop. P. Giresse. J. Monteillet and J-L. Saos. 1986. Consequences
sedimentologiques de I'evolution c1imatique fini-Holocene dans Ie delta du Saloum (Senegal).
Oceanogr. Trop. 21:89-98.
Basak. U.e.. A.B. Das and P. Das. 1995. Metabolic changes during rooting in stem cuttings of five
mangrove species. Plant Growth Regulation 17:141-148.
Basak. U.C .• A.B. Das and P. Das, 1998. Seasonal changes in organic constituents in leaves of nine
mangrove species. Mar. Freshw. Res. 49:369-372.
Basey, J.M. and S.H. Jenkins. 1993. Production of chemical defenses in relation to plant growth rate.
Oikos 68:323-328.
Bashan. Y., M.E. Puente, D.O. Myrold and G. Toledo, 1998. In vitro transfer of fixed nitrogen from
diazotrophic filamentous cyanobacteria to black mangrove seedlings. FEMS Microbiol. Ecol.
26: 165-170.
Bastow, W.J. and A. Chiarucci. 2000. Do plant communities exist? Evidence from scaling-up local
species-area relations to the regional level. J. Veg. Sci. 11:773-775.
Baylis, G.T.S .• 1940. Leaf anatomy of the New Zealand mangrove. Trans. Roy. Soc. N.Z. 70:164-170.
Beale. E .• 1977. Kennedy of Cape York.. Seal Books. Rigby, Brisbane.
Beard. J.S., 1967. An inland occurrence of mangroves. West. Aust. Nat. 10:112-115.
Becker, P.• A. Asmat, J. Mohamad, M. Moksin and M.T. Tyree. 1997. Sap flow rates of mangrove trees
are not unusually low. Trees II :432-435.
Beeckman. H., E. Gallin and E. Coppejans, 1990. Indirect gradient analysis of the mangal formation of
Gazi Bay (Kenya). Silva Gandavensis 54:57-72.
References 307

Beever, J.W., D. Simberloff and L.L. King, 1979. Herbivory and predation by the mangrove crab,
Aratus pisonii. Oecologia 43:317-328.
Benson, A.A. and M.R. Atkinson, 1967. Choline sulphate and phosphate in salt excreting plants. Fed.
Proc. 26: 394.
Benzing, D.H., 1980. The biology of the bromeliads. Mad River Press, Eureka, California.
Bergin, A., 1993. A rising tide of Aboriginal sea claims: implications of the Mabo case in Australia. Int.
J. Marine and Coastal Law 8:359-371.
Berry, E.W., 1924. The middle and upper Eocene floras of southern North America. US Geological
Survey Professional Papers No. 92. 206 pp.
Berry, E.W. 1930. Revision of the lower Eocene Wilcox Flora of the southern United States. US
Geological Survey Professional Papers No. 156.
Bertrand, F., 1991. Contribution a I'etude de I'environnement et de la dynamique des mangroves de
Guinee: donnees de terrain et apport de la teledetection. Unpublished Thesis, Universite de
Bordeaux, Bordeaux.
Bessedik, M., 1981. Une mangrove aAvicennia L. en Mediterranee occidentale au Miocene inferieur
et moyen. Implications paleogeographiques. C.R. Acad. Sc. (Paris) 293:469-472.
Bessedik, M., 1985. Reconstruction des environnments miocenes des regions nord-ouest
mediterraneenes a partir de la palynologie. Unpublished Thesis, L'Universite des Sciences et
Techniques du Languedoc, MontpelIier.
Betoulle, J .L., F. Fromard, A. Fabre and H. Puig, 200 1. Caracterisation des chutes de Iitiere et des
apports au sol en nutriments dans une mangrove de Guyane franc;:aise. Can. J. Bot. 79:238-249.
Bhosale, L.1. and L.S. Shinde, 1983. Significance of cryptovivipary in Aegiceras comiculatum (L.)
Blanco. In: Teas, H.1. (Ed.), Tasks for vegetation science, Vol. 8. Dr W. Junk, The Hague, pp.
123-129.
Biebl, R. and H. Kinzel, 1965. Blattbau und Salzhaushalt von Laguncularia racemose (L.) Gaertn. f.
und anderer Mangrovenbaume auf Puerto Rico. Ost. Bot. Zeit. 112:56-93.
Bird, E.C.F., 1971. Mangroves as land-builders. Vict. Nat. 88:189-197.
Bird, E.C.F., 1972. Mangroves and coastal morphology in Cairns Bay, north Queensland. J. Trop.
Geogr. 35:11-16.
Bird, E.C.F. and M.M. Barson, 1977. Measurement of physiographic changes on mangrove-fringed
estuaries and coastlines. Mar. Res. Indonesia 18:73-80.
Bjorkman, 0., B. Demmig and T.1. Andrews, 1988. Mangrove photosynthesis: response to high
irradiance stress. Aust. J. PI. Physiol. 15:43-61.
Blaber, S.1.M., 1997. Fish and fisheries of tropical estuaries. Chapman & Hall, London.
Blaber, S.1.M., D.T. Brewer and J.P. Salini, 1989. Species composition and biomasses of fishes in
different habitats of a tropical northern Australian estuary: their occurrence in the adjoining
sea and estuarine dependence. Est. Coastal Shelf Sci. 29:509-531.
Blake, S.T. and C. Roff, 1972. The honey flora of Queensland. Govt. Printer, Brisbane.
Blanchard, 1. and G. Prado, 1995. Natural regeneration of Rhizophora mangle in strip c1earcuts in
northwestern Ecuador. Biotropica 27:160-167.
Blasco, F., 1975. The mangroves of India. Institut Franc;:ais de Pondicbery, Pondichery.
Blasco, F., 1984. Taxonomic considerations of the mangrove species. In: Snedaker, S.c. and J.G.
Snedaker (Eds.), The mangrove ecosystem: research methods. UNESCO. Paris, pp. 81-90.
Blasco, F., 1985. Mangroves du Benin. CNRS/Paul Sabatier University, Toulouse.
Blasco, F. and M. Aizpuru, 1997. Classification and evolution of the mangroves of India. Trop. Ecol.
38:357-374.
Blasco, F., M. Aizpuru and C. Gers, 2001. Depletion of the mangroves of continental Asia. Wetlands
Ecology and Management 9:245-256.
Blasco, F., Saenger, P. and E. Janodet, 1996. Mangroves as indicators of coastal change. Catena 149:1-
12.
Blasco, F., T. Gauquelin, M. Rasolofoharinoro, J. Denis, M. Aizpuru and V. Caldairou, 1998. Recent
advances in mangrove studies using remote sensing data. Mar. Freshw. Res. 49:287-296.
Boer, B., 1993. Anomalous pneumatophores and adventitious roots of Avicennia marina (Forssk.)
Vierh. mangroves two years after the 1991 Gulf War oil spill in Saudi Arabia. Mar. Poll. Bull.
27:207-211.
Bohorquez, c., 1996. Restoration of mangroves in Colombia: a case study of Rosario's Coral Reef
National Park. In: Field, C. (Ed.), Restoration of mangrove ecosystems. ISMElITTO, Okinawa,
pp. 189-196.
Boland, 0.1., M.I.H. Brooker, G.M. Chippendale, N. Hall, B.P.M. Hyland, R.D. Johnston, D.A. Kleinig
and J.D. Turner, 1984. Forest trees of Australia. 4th Edition, Thomas Nelson and CSIRO,
Melbourne.
Bond, G., 1956. A feature of root nodules of Casuarina. Nature 177:192.
Bond, G., 1963. The root nodules of non-leguminous angiosperms. Symp. Soc. Gen. Microbiol. \3:315-
347.
Boon, P.1. and W.G. Allaway, 1982. Assessment of leaf-washing techniques for measuring salt
secretion in Avicennia marina (Forsk.) Vierh. Aust. J. PI. Physiol. 9:725-734.
308 Mangrove Ecology, Silviculture and Conservation

Boon, P.1. and W.G. Allaway, 1986. Rates and ionic specificity of salt secretion from excised leaves of
the mangrove, Avicennia marina (Forsk.) Vierh. Aquat. Bot. 26:143-153.
Bostrom, T.E. and C.D. Field, 1973. Electrical potentials in the salt gland of Aegiceras. In: Anderson,
W.P. (Ed.), Ion transport in plants. Academic Press, London, pp. 385-92.
Boto, K.G., 1982. Nutrient and organic fluxes in mangroves. In: Clough, B.F. (Ed.), Mangrove
ecosystems in Australia. Structure, function and management. ANU Press, Canberra, pp. 239-
259.
Boto, K.G., 1983. Nutrient status and other soil factors affecting mangrove productivity in northeast
Australia. Wetlands 3:45-49.
Boto, K.G., 1984. Waterlogged saline soils. In: Snedaker, S.C. and J.G. Snedaker (Eds.), The mangrove
ecosystem: research methods. UNESCO, Paris, pp. 114-130.
Boto, K.G., 1991. Nutrients and mangroves. In: Connell, D.W. and D.W. Hawker (Eds.), Pollution in
tropical aquatic systems. CRC Press, Boca Raton, pp. 129-145.
Boto, K.G. and 1.S. Bunt, 1981. Tidal export of particulate organic matter from a northern Australian
mangrove system. Est. Coastal Shelf Sci. 13:247-257.
Boto, K.G. and W.H. Patrick, 1978. Role of wetlands in the removal of suspended sediments. In:
Greeson, P.E., J.R. Clark and J.E. Clark (Eds.), Wetland functions and values. The state of our
understanding. Amer. Wat. Res. Assoc., Technical Report. No. TPS 79-2, pp. 479-489.
Boto, K.G. and A.I. Robertson, 1990. The relationship between nitrogen fixation and tidal exports of
nitrogen in a tropical mangrove system. Est. Coastal Shelf Sci. 31:531-540.
Boto, K.G. and J.T. Wellington, 1983. Phosphorus and nitrogen nutritional status of a northern
Australian mangrove forest. Mar. Ecol. Progr. Ser. 11:63-69.
BolO, K.G. and 1.T. Wellington, 1984. Soil characteristics and nutrient status in northern Australian
mangrove forests. Estuaries 7:61-69.
Boto, K., D.M. Alongi and A.L.J. Nott, 1989. Dissolved organic carbon-bacteria interactions at
sediment-water interface in a tropical mangrove system. Mar. Ecol. Progr. Ser. 51 :243-251.
Bousquet-Melou, A., 19%. Biosystematique de Paletuviers du genre Avicennia: recherche de
nouveaux marqueurs taxonomiques. Unpublished Thesis, Universite Paul Sabatier, Toulouse.
BouziIle, J.-B., E. Kerneis, A. Bonis and B. Touzard, 2001. Vegetation and ecological gradients in
abandoned salt pans in western France. J. Veg. Sci. 12:269-278.
Bowman, H.H.M., 1917. Ecology and physiology of the red mangrove. Proc. Amer. Phil. Soc. 56:589-
672.
Bowman, H.H.M., 1921. Histological variations in Rhizophora mangle L. Pap. Mich. Acad. Sci. 22: 129-
134.
Boye, M., F. Baltzer, C. Caratini and J.F. ViIliers, 1975. Mangrove of the Wouri Estuary, Cameroon. In:
Walsh, G.E., S.C. Snedaker and HJ. Teas (Eds.), Proceedings of the international symposium
on biology and management of mangroves. Vol. 2. University of Rorida, Gainesville, pp. 431-
455.
Braun-Blanquet, J., 1964. Pflanzensoziologie: Grundziige der Vegetationskunde, 3,d Edition. Springer
Verlag, Vienna.
Breen, C.M. and BJ. Hill, 1969. A mass mortality of mangroves in the Kosi Estuary. Trans. Roy. Soc. S.
Afr. 38:285-303.
Breteler, FJ., 1969. The Atlantic species of Rhizophora. Acta Bot. Need. 18:434-441.
Bridgewater, P.B., 1975. Peripheral vegetation at Westernport Bay. Proc. Roy. Soc. Vic. 87:69-78.
Bridgewater, P.B., 1985. Variation in the mangal along the west coastline of Australia. Proc. Ecol. Soc.
Aust. 13:243-256.
Bridgewater, P.B., 1989. Syntaxonomy of the Australian mangal refined through iterative ordinations.
Vegetatio 81:159-168.
Bridgewater, P.B. and I.D. Cresswell, 1999. Biogeography of mangrove and saltmarsh vegetation:
implications for conservation and management in Australia. Mangroves and Salt Marshes
3: 117-125.
Briggs, S.V., 1977. Estimates of biomass in a temperate mangrove community. Aust. J. Ecol. 2:369-373.
Brinkman, R.M., S.R. Massel, P.V. Ridd and K. Furukawa, 1997. Surface wave attenuation in mangrove
forests. In: Proceedings of the Combined Australasian Coastal Engineering and Ports
Conference, Christchurch, pp. 941-946.
Brouat, C. and D. McKey, 2000. Origin of caulinary ant domatia and timing of their onset in plant
ontogeny: evolution of a key trait in horizontally transmitted ant-plant symbioses. BioI. 1. Linn.
Soc. 71:801-819.
Brown, S., AJ.R. Gillespie and A.E. Lugo, 1989. Biomass estimation methods for tropical forests with
applications to forest inventory data. For. Sci. 35:881-902.
Brown, W.H. and A.F. Fisher, 1918. Philippine mangrove swamps. Bulletin No. 17, Bureau of Forestry,
Department of Agriculture and Natural Resources, Manila.
Brunnich, J.C. and F. Smith, 1911. Some Queensland mangrove barks and other tanning materials. Qld
Agric. J. 27:86-94.
Bucher, D. and P. Saenger, 1991. An inventory of Australian estuaries and enclosed marine waters: an
overview of results. Aust. Geogr. Studies 29:370-381.
References 309

Buckley R, 1983. Biogeography of mangroves on northern Great Barrier Reef islands. Bull. Mar. Sci.
33:767-776.
Bunt, J.S., 1996. Mangrove zonation: an examination of data from seventeen riverine estuaries in
tropical Australia. Ann. Bot. 78:333-341.
Bunt, J.S., 1999. Overlap in mangrove species zonal patterns: some methods of analysis. Mangroves
and Salt Marshes 3:155-164.
Bunt, J.S. and T. Stieglitz 1999. Indicators of mangrove zonality: the Normanby River, N.E. Australia.
Mangroves and Salt Marshes 3:177-184.
Bunt, J.S. and W.T. Williams, 1980. Studies in the analysis of data from Australian tidal forests
('mangroves'). I. Vegetational sequences and their graphic representation. Aust. J. Ecol.
5:385-390.
Bunt, J.S. and W.T. Williams, 1981. Vegetational relationships in the mangroves of tropical Australia.
Mar. Ecol. Progr. Ser. 4:349-359.
Bunt, J.S. and E. Wolanski, 1980. Hydraulics and sediment transport in a creek - mangrove swamp
system. Proceedings 7th Australasian Hydraulics and Ruid Mechanics Conference Brisbane,
pp. 492-495.
Bunt, J.S., W.T. Williams and E.D. Bunt, 1985. Mangrove species distribution in relation to tide at the
seafront and up rivers. Aust. J. Mar. Freshw. Res. 36:481-492.
Bunt, J.S., W.T. Williams and H.J. Clay, 1982a. River water salinity and the distribution of mangrove
species along several rivers in North Queensland. Aust. J. Bot. 30:401-412.
Bunt, J.S., W.T. Williams and N.C. Duke, 1982b. Mangrove distributions in northeast Australia. J.
Biogeogr.9:111-120.
Burbidge, N.T., 1960. The phytogeography of the Australian region. Aust. J. Bot. 8:75-211.
Burchett, M.D., C.D. Field and A. Pulkownik, 1984. Salinity, growth and root respiration in the grey
mangrove Avicennia marina. Physiol. Plant. 60: 113-118.
Burchett, M.D., CJ. Clarke, C.D. Field and A. Pulkownik, 1989. Growth and respiration in two
mangrove species at a range of salinities. Physiol. Plant. 75:299-303.
Burns, K.A and S. Codi, 1998. Contrasting impacts of localised versus catastrophic oil spills in
mangrove sediments. Mangroves and Salt Marshes 2:63-74.
Bums, K.A., S. Levings and S. Garrity, 1993. How many years before mangrove ecosystems recover
from a catastrophic oil spill? Mar. Poll. Bull. 26:239-248.
Burrows, D.W., 2001. A re-evaluation of insect leaf herbivory in Australian mangroves: is it higher
than previously thought? In: Abstracts of the International Symposium on Mangroves:
Evolution, Physiology and Conservation. RCAST, University of Tokyo, pp. 45-46.
Bush, R.T. and L.A. Sullivan, 1999. Pyrite micromorphology in three Australian Holocene sediments.
Aust. J. Soil Res. 37:637-653.
Butler, AJ., AM. Depers, S.c. McKiliup and D.P. Thomas, 1977. Distribution and sediments of
mangrove forests in South Australia. Trans. Roy. Soc. Sth Aust. 101:35-44.
Butler, V. and T.D. Steinke, 1976. Ultrastructural studies on Avicennia marina propagules. Proc. Electr.
Microscopy Soc. S. Afr. 6:67-68.
Caddy, J.F. and G.D. Sharp, 1986. An ecological framework for marine fisheries investigations. FAO
Fish. Tech. Pap. 283:1-152.
Cahoon, D.R. and J.C. Lynch, 1997. Vertical accretion and shallow subsidence in a mangrove forest of
southwestern Florida, USA. Mangroves and Salt Marshes I: 173-186.
Calderon, D.G. and B.R. Echeverri, 1997. Obtaining Rhizophora mangle seedlings by stimulating of
adventitious roots using an air-layering technique. In: Kjerfve, B., L.D. Lacerda and E.H.S.
Diop (Eds.), Mangrove ecosystem studies in Latin America and Africa. UNESCO, Paris, pp.
98-107.
<;aIis, I., M. Hosny and A Yiiriiker, I 994a. Inerminosides A and B, two novel complex iridoid
glycosides from Clerodendrum inerme. J. Nat. Prod. 57:494-500.
<;alis, I., M. Hosny and A. Yiiriiker, I 994b. Inerminosides AI, C and 0, three iridoid glycosides from
Clerodendrum inerme. Phytochem. 37: 1083-1085.
Camilleri, J.c., 1989. Leaf choice by crustaceans in a mangrove forest in Queensland. Mar. BioI.
102:453-459.
Camilleri, lC. and G. Ribi, 1983. Leaf thickness of mangroves (Rhizophora mangle) growing in
different salinities. Biotropica 15:139-141.
Camilleri, J.C. and G. Ribi, 1986. Leaching of dissolved organic carbon (DOC) from dead leaves,
formation of flakes from DOC, and feeding on flakes by crustaceans in mangroves. Mar. BioI.
91:337-344.
Canoy, MJ., 1975. Diversity and stability in a Puerto Rican Rhizophora mangle L. forest. In: Walsh,
G.E., S.c. Snedaker and HJ. Teas (Eds.), Proceedings of the International Symposium on
Biology and Management of Mangroves. Vol. I, University of Rorida, Gainesville, pp. 344-
356.
Cantera, J.R., B.A. Thomassin and P.M. Arnaud, 1999. Faunal zonation and assemblages in the Pacific
Colombian mangroves. Hydrobiologia 413: 17-33.
Cardale, S. and C.D. Field, 1971. The structure of the salt gland of Aegiceras comiculatum. Planta
99:183-191.
310 Mangrove Ecology, Silviculture and Conservation

Cardale, S. and C.D. Field, 1975. [on transport in the salt gland of Aegiceras. In: Walsh, G.E., S.c.
Snedaker and HJ. Teas (Eds.), Proceedings of the International Symposium on Biology and
Management of Mangroves. Vol. 2, University of Florida, Gainesville, pp. 608-614.
Cardona, P. and L. Botero, 1998. Soil characteristics and vegetation structure in a heavily deteriorated
mangrove forest in the Caribbean coast of Colombia. Biotropica 30:24-34.
Carey, G., 1934. Further investigations on the embryology of viviparous seeds. Proc. Linn. Soc. NSW
59:392-410.
Carey, G. and L. Fraser, 1932. The embryology and seedling development of Aegiceras majus Gaertn.
Proc. Linn. Soc. NSW 57:341-360.
Carlquist, S., 1975. Ecological strategies of xylem evolution. University of California Press, Berkeley.
Carlson, P.R., L.A. Yarbro, C.F. Zimmermann and J.R. Montgomery, 1981. Pore water chemistry of an
overwash mangrove island. Estuaries 4:282-287.
Carlton, 1. M. and M. D. Moffler, 1978. Propagation of mangroves by air-layering. Environ. Conserv.
5:147-150.
Cause, M.C., EJ. Rudder and W.T. Kynaston, 1989. Queensland Timbers - their nomenclature, density
and Lyctid susceptibility. QDPI Tech. Pamphlet No.2, pp. 1-126.
Cavagnetto, C. and Anad6n, P. 1995. Une mangrove complexe dans Ie Bartonien du Bassin de l'Erbe
(NE de I'Espagne). Paleontographica, B 236:147-165.
Chaffey, D.R., F.R. Miller and J.H. Sandom, 1985. A forest inventory of the Sundarbans, Bangladesh.
Overseas Development Administration, London.
Chai, P.P.K. and K.K. Lai, 1984. Management and utilisation of mangrove forests in Sarawak. In:
Soepadmo, E., A.N. Rao and DJ. Macintosh (Eds.), Proceedings of the Asian symposium on
mangrove environment - research and management. Unversity of Malaya, Kuala Lumpur, pp.
785-795.
Champagne, D.E., O. Koul, M.B.lsman, G.G.E. Scudder and G.H.N. Towers, 1992. Biological activity
of limo noids from the Rutales. Phytochem. 31:377-394.
Chan, H.T., 1989. A note on the eradication of Acrostichum aureum ferns in the Matang mangroves,
Perak, Peninsular Malaysia. J. Trop. For. Sci. 2:171-173.
Chan, H.T., 1996. Mangrove reforestation in Peninsular Malaysia: a ca.~e study of Matang. In: Field, C.
(Ed.), Restoration of mangrove ecosystems. ISMElITTO, Okinawa, pp. 64-75.
Chan, H.T., J.E. Ong, W.K. Gong and A. Sasekumar, 1993. The socio-economic, ecological and
evironmental values of mangrove ecosystems in Malaysia and their present state of
conservation. ISME Mangrove Ecosystems Technical Reports No.1, pp. 41-81.
Chanda, S., 1977. An eco-f1oristic survey of the mangrove of Sundarbans, West Bengal, India. Trans.
Bose Res. Inst. 40:5-14.
Chandler, M.EJ., 1951. Note on the occurrence of mangroves in the London Clay. Proc. Geol. Ass.
London 62:217-272.
Chansang, H., 1984. Compatibility of alternative uses of mangrove and estuarine resources. In:
Soepadmo, E., A.N. Rao and DJ. McIntosh (Eds.), Proceedings of the Asian Symposium
Mangrove Environments - Research and Management, UNESCO, pp. 821-828.
Chapman, V.I., I 944a. The 1939 Cambridge University Expedition to lamaica. I. A study of the
botanical processes concerned in the development of the Jamaican shoreline. Bot. 1. Linn. Soc.
52:407-447.
Chapman, VJ. 1944b. The 1939 Cambridge University expedition to lamaica. III. The morphology of
Avicennia nitida lacq. and the function of its pneumatophores. Bot. 1. Linn. Soc. 52:487-533.
Chapman, V.J., 1960. Salt marshes and salt deserts of the world. Hill, London.
Chapman, V.I., 1970. Mangrove phytosociology. Trop. Ecol. 11:1-19.
Chapman, V.J., 1975. Mangrove biogeography In: Walsh, G.E., S.C. Snedaker and HJ. Teas (Eds.),
Proceedings of the International Symposium on Biology and Management of Mangroves. Vol.
I, University of Florida, Gainesville, pp. 3-22.
Chapman, V.J., 1976. Mangrove vegetation. Cramer, Vaduz.
Chapman, V J., 1977. Introduction. In: Chapman, V.J. (Ed.), Ecosystems of the world. I. Wet coastal
ecosystems. Elsevier Sci. Publ. Co., Amsterdam, pp. 1-29.
Chapman, VJ. and J.W. Ronaldson, 1958. The mangrove and salt-marsh flats of the Auckland isthmus.
N.Z. Dept. Sci. Indust. Res. Bull. 125:1-79.
Chaudhuri, P.K.R., 1990. Artificial regeneration of Sonneratia apetala (Buch.-Ham.) in Sundarbans,
West Bengal. Indian Forester 116:773-779.
Che, RG.O., 1999. Concentration of 7 heavy metals in sediments and mangrove root samples from Mai
Po, Hong Kong. Mar. Poll. Bull. 39:269-279.
Cheeseman, 1.M., 1994. Depressions of photosynthesis in mangrove canopies. In: Baker, N.R (Ed.),
Photoinhibition of photosynthesis: from molecular mechanisms to the field. BIOS Scientific
Publishers, Oxford, pp. 377-389.
Cheeseman, 1.M., L.B. Herendeen, A.T. Cheeseman and B.F. Clough, 1997. Photosynthesis and
photoprotection in mangroves under field conditions. Plant, Cell and Environ. 20:579-588.
Cheeseman, J.M., B.F. Clough, D.R Carter, C.E. Lovelock, J.E. Ong and R.G. Sim, 1991. The analysis
of photosynthetic performance in leaves under field conditions: a case study using Bruguiera
mangroves. Photosyn. Res. 29: 11-22.
References 311

Chen, G.Z., S.Y. Miao, N.F.Y. Tam, Y.S. Wong, S.H. Li and C.Y. Lan, 1995. Effects of synthetic
wastewater on young Kandelia candel plants growing under greenhouse conditions.
Hydrobiologia 295:263-273.
Chen, R. and R.R. Twilley, 1998. A gap dynamic model of mangrove forest development along
gradients of soil salinity and nutrient resources. 1 Ecol. 86:37-51.
Chen, X.R. and Y.S. Wong, 1996. Effect of wastewater discharge on nutrient contamination of
mangrove soils and plants. In: Khemnark, C. (Ed.), Proceedings of Ecotone IV: ecology and
management of mangrove restoration and regeneration in east and southeast Asia. Surat Thani,
Thailand, pp. 269-283.
Chen, X.Y., 2000. Effects of plant density and age on the mating system of Kandelia candel Oruce
(Rhizophoraceae), a viviparous mangrove species. Hydrobiologia 432: 189-193.
Chiu, C.Y. and C.H. Chou, 1991. The distribution and influence of heavy metals in mangrove forests of
the Tamshui estuary in Taiwan. Soil Sci. Plant Nutr. 37:659-669.
Chiu, C.Y., F.S. Hsiu, S.S. Chen and C.H. Chou, 1995. Reduced toxicity of Cu and Zn to mangrove
seedlings in saline environments. Bot. Bull. Acad. Sin. 36:19-24.
Chong, V.c., A. Sasekumar and E. Wolanski, 1996. The role of mangroves in retaining penaeid prawn
larvae in Klang Strait, Malaysia. Mangroves and Salt Marshes 1:11-22.
Chong, C.V., A. Sasekumar, M.U.C. Leh and R. D'Cruz, 1990. The fish and prawn communities of a
Malaysian coastal mangrove ecosystem, with comparisons to adjacent mudflats and inshore
waters. Est. Coastal Shelf Sci. 31 :703-722.
Choong, M.F., P.W. Lucas, I.S.Y. Ong, B. Pereira, H.T.W. Tan and I.M. Turner, 1992. Leaf fracture
toughness and sclerophylly: their correlations and ecological implications. New Phytol.
121:597-610.
Choudhuri, P.K.R., 1991. Biomass production of mangrove plantation in Sundarbans, West Bengal
(India) - a case study. Indian Forester 117:3-12.
Chowdhury, M.A.M., 1996. Notes on natural regeneration potential of major mangrove species in
degrading Sundarbans ecosystem of Bangladesh. In: Khemnark, C. (Ed.), Proceedings of
Ecotone IV: ecology and management of mangrove restoration and regeneration in east and
southeast Asia. Surat Thani, Thailand, pp. 104-118.
Chowdhury, R.A. and I. Ahmed, 1994. History of forest management. In: Hussain, Z. and G. Acharya
(Eds.), Mangroves of the Sundarbans, Vol. 2: Bangladesh. IUCN, Bangkok, pp. 155-179.
Christensen, B., 1978. Biomass and primary production of Rhizophora apiculata BI. in a mangrove in
southern Thailand. Aquat. Bot. 4:43-52.
Christensen, B., 1982. Management and utilisation of mangroves in Asia and the Pacific. FAO
Environment Paper No.3. FAO, Rome.
Christensen, B. and S. Wium-Anderson, 1977. Seasonal growth of mangrove trees in southern Thailand.
I. The phenology of Rhizophora apiculata BI. Aquat. Bot. 3:281-286.
Churchill,O.M., 1973. The ecological significance of tropical mangroves in the early tertiary floras of
southern Australia. Geol. Soc. Aust. Spec. Publ. 4:79-86.
Cintr6n, G. and Y. Schaeffer-Novelli, 1984. Methods for studying mangrove structure. In: Snedaker,
S.C. and I.G. Snedaker (Eds.), The mangrove ecosystem: research methods. UNESCO, Paris,
pp.91-113.
Cintron, G., A.E. Lugo, O.J. Pool and G. Morris, 1978. Mangroves of arid environments in Puerto Rico
and adjacent islands. Biotropica 10: 110-121.
Clark, J., 1974. Coastal ecosystems: ecological considerations for management of the coastal zone. The
Conservation Foundation, Washington.
Clark, M.W., 1998. Management implications of metal transfer pathways from a refuse tip to mangrove
sediments. Sci. Tot. Environ. 222:17-34.
Clark, M.W., D. McConchie, D.W. Lewis and P. Saenger, 1998. Redox stratification and heavy metal
partitioning in Avicennia dominated mangrove sediments: a geochemical model. Chem. Geol.
149:147-171.
Clark, M.W., D. McConchie, P. Saenger and M. Pillsworth, 1997. Hydrological controls on copper,
cadmium, lead and zinc concentrations in an anthropogenically polluted mangrove ecosystem,
Wynnum, Brisbane, Australia. J. Coastal Res. 13:1150-1158.
Clark, R.L. and lC. Guppy, 1988. A transition from mangrove forest to freshwater wetland in the
monsoon tropics of Australia. 1. Biogeogr. 15:665-684.
Clarke, D.E., O.J. Macintosh and S. Havanond, 1997. Study to assess the possibility of using waste from
shrimp ponds as a medium to grow three species of mangrove trees. In: Aksornkoae, S., L.
Puangchit and B. Thaiutsa (Eds.), Tropical Forestry in the 21" Century. Vol. 10: Mangrove
Ecosystems. Kasetsart University, Bangkok, pp. 92-103.
Clarke, L.O. and N.J. Hannon, 1967. The mangrove swamp and saltmarsh communities of the Sydney
district. I. Vegetation, soils and climate. J. Ecol. 55:753-771.
Clarke, L.O. and N.J. Hannon, 1969. The mangrove swamp and saltmarsh communities of the Sydney
district. II. The holocoenotic complex with particular reference to physiography. J. Eeol.
57:213-234.
Clarke, L.D. and N.J. Hannon, 1970. The mangrove swamp and saltmarsh communities of the Sydney
district. III. Plant growth in relation to salinity and waterlogging. J. Eeol. 58:351-369.
312 Mangrove Ecology, Silviculture and Conservation

Clarke, L.D. and NJ. Hannon, 1971. The mangrove swamp and saltmarsh communities of the Sydney
district. IV. The significance of species interaction. J. Ecol. 59:535-553.
Clarke, PJ., 1992. Predispersa1 mortality and fecundity in the grey mangrove (Avicennia marina) in
southeastern Australia. Aust. J. Ecol. 17:161-168.
Clarke, P.l, 1993. Dispersal of grey mangrove (Avicennia marina) propagules in south-eastern
Australia. Aquat. Bot. 45:195-204.
Clarke, PJ., 1994. Baseline studies of temperate mangrove growth and reproduction; demographic and
litterfall measures of leafing and flowering. Aust. J. Bot. 42:37-48.
Clarke, PJ., 1995. The popUlation dynamics of the mangrove Avicennia marina: demographic synthesis
and predictive modelling. Hydrobiologia 295:83-88.
Clarke, PJ. and W.G. Allaway, 1993. The regeneration niche of the grey mangrove Avicennia marina
- effects of salinity, light, and sediment factors on establishment, growth and survival in the
field. Oecologia 93:548-556.
Clarke, PJ. and R.A. Kerrigan, 2000. Do forest gaps influence the population structure and species
composition of mangrove stands in northern Australia? Biotropica 32:642-652.
Clarke, PJ. and PJ. Myerscough, 1991a. Buoyancy of Avicennia marina propagules in south-eastern
Australia. Aust. J. Bot. 39:77-83.
Clarke, PJ. and PJ. Myerscough, 1991b. Floral biology and reproductive phenology of Avicennia
marina in south-eastern Australia. Aust. J. Bot. 39:283-293.
Clarke, PJ. and PJ. Myerscough, 1993. The intertidal distribution of the grey mangrove (Avicennia
marina) in southea~tern Australia: the effects of physical conditions, interspecific competition
and predation on propagule establishment and survival. Aust. 1 Ecol. 18:307-315.
Clarke, PJ., R.A. Kerrigan and CJ. Westphal, 2001. Dispersal potential and early growth in 14 tropical
mangroves: do early life history traits correlate with patterns of adult distribution? J. Ecol.
89:648-659.
Clay, R.E. and A.N. Anderson, 19%. Ant fauna of a mangrove community in the Australian seasonal
tropics, with particular reference to zonation. Aust. J. Zool. 44:521-533.
Clough, B.F., 1992. Primary productivity and growth of mangrove forests. In: Robert.~on, A.1. and D.M.
Alongi (Eds.), Tropical mangrove ecosystems. American Geophysical Union, Washington, pp.
225-249.
Clough, B.F., 1998. Mangrove forest productivity and biomass accumulation in Hinchinbrook Channel,
Australia. Mangroves and Salt Marshes 1:191-198.
Clough, B.F., 2001. Mangrove-based small-scale shrimp aquaculture. In: Utilizing different aquatic
resources for livelihoods in Asia: a resource book. IIRR, IDRC, FAO, NACA and ICLARM,
Manila, pp. 340-347.
Clough, B.F. and P.M. Attiwill, 1975. Nutrient cycling in a community of Avicennia marina in a
temperate region of Australia. In: Walsh, G.E., S.C. Snedaker and HJ. Teas (Eds.),
Proceedings of the international symposium on biology and management of mangroves. Vol. 1.
University of Florida, Gainesville, pp. 137-146.
Oough, B.F. and K. Scott. 1989. Allometric relationships for estimating above-ground bioma~s in six
mangrove species. For. Ecol. Manage. 27:117-127.
Clough, B.F. and R.G. Sim, 1989. Changes in gas exchange characteristics and water-use efficiency of
mangroves in response to salinity and vapour pressure deficit. Oecologia 79:38-44.
Clough, B.F., TJ. Andrews and I.R. Cowan. 1982. Physiological processes in mangroves. In: Clough,
B.F. (Ed.), Mangrove ecosystems in Australia - structure, function and management. ANU
Press, Canberra, pp. 193-210.
Clough, B.F., K.G. Boto and P.M. Attiwill, 1983. Mangroves and sewage: a re-evaluation. In: Teas, HJ.
(Ed.), Biology and ecology of mangroves. Tasks for vegetation science. Vol. 8. Dr W. Junk,
The Hague, pp. 151-161.
Clough, B.F., lE. Ong and W.K. Gong, 1997. Estimating leaf area index and photosynthetic production
in canopies of the mangrove Rhizophora apiculata. Mar. Ecol. Progr. Ser. 159:285-292.
Clough, B.F., D.T. Tan, D.x. Phuong and D.C. Buu, 2000. Canopy leaf area index and litter fall in
stands of the mangrove Rhizophora apiculo.ta of different age in the Mekong Delta, Vietnam.
Aquat. Bot. 66:311-320.
Cliisener, M. and S.W. Breckle, 1987. Reasons for the limitation of mangrove along the west coast of
northern Peru. Vegetatio 68:173-177.
Cocks, K.D., 1975. The social functions of aquatic ecosystems. Proc. Ecol. Soc. Aust. 8: 167-173.
Coleman, J.M. and L.D. Wright, 1975. Modem river deltas: variability of processes and sand bodies. In:
Broussard, M.L. (Ed.), Deltas, models for exploration. Houston Geological Society, Houston,
pp.99-150.
Coley. P.D., 1983. Intraspecific variation in herbivory on two tropical tree species. Ecology 64:426-
433.
Coley, P.D., 1988. Effects of plant growth rate and leaf lifetime on the amount and type of anti-
herbivore defense. Oecologia 74:531-536.
Collins, J.P., R.c. Bertkelhamer and M. Mesler, 1977. Notes on the natural history of the mangrove
Pelliciera rhizophorae Tr. & PI. (Theaceae). Brenesia II :17-29.
References 313

Commission of the European Communities, 1987. Mangroves d'Afrique et de Madagascar. Protection et


mise en valeur (Vol. 1-3). Leiden Univ. (Netherlands), Centre for Environmental Studies in
cooperation with Societe d'Ecoamenagement.
Compere, P., 1963. The correct name of the Afro-American black mangrove. Taxon 12:150-152.
Conacher, C.A., C. O'Brien, J.L. Horrocks and R.K. Kenyon, 1996. Litter production and accumulation
in stressed mangrove communities in the Embley River Estuary, north-eastern Gulf of
Carpentaria, Australia. Mar. Freshw. Res. 47:737-744.
Connor, OJ., 1969. Growth of grey mangrove (Avicennia marina) in nutrient culture. Biotropica 1:36-
40.
Conti, E., A. Litt and KJ. Sytsma, 1996. Circumscription of MyrtaIes and their relationships to other
rosids: evidence from rbcL sequence data. Amer. J. Bot. 83:221-233.
Cookson, I.C. and A. Eisenack, 1967. Some early Tertiary microplankton and pollen grains from a
deposit near Strahan, Western Tasmania. Proc. Roy. Soc. Vic. 88:131-140.
Corlett, R.T., 1986. The mangTQve understory: some additional observations. J. Trop. Ecol. 2:93-94.
Cormier-Salem, M.C., 1994. A la decouverte des mangroves: regards multiples sur un objet de
recherche mouvant. In: Cormier-Salem, M.e. (Ed.), Oynamique et usages de la mangrove
dans les pays des rivieres du sud (du senegal a la Sierra Leone). ORSTOM Editions, Paris, pp.
11-24.
Corredor, J.E. and J.M. Morell, 1994. Nitrate depuration of secondary sewage effluents in mangrove
sediments. Estuaries 17:295-300.
Corredor, J.E., J.M. Morell and J. Bauza, 1999. Atmospheric nitrous oxide fluxes from mangrove
sediments. Mar. Poll. Bull. 38:473-478.
Cosser, P.R., 1989. Water quality, sediments and macroinvertebrate community of residential canal
estates in south-east Queensland: a multivariate analysis. Water Res. 23:1087-1097.
Cousins, J.M. and P. Saenger, in press. Developing a protocol for in vitro propagation of the grey
mangrove A vicennia marina. International Association for Plant Tissue Culture &
Biotechnology Australian Branch, 71h Meeting, Plant tissue culture - its importance in biology,
ecology and agriculturelhorticulture. 20-23 January 2002, University of New England,
Armidale.
Cowan, I.R., 1978. Stomatal responses in mangroves. Paper presented at the 19th General Meeting
Australian Society Plant Physiologists, Sydney, Australia.
Cowan, I.R., 1986. Economics of carbon fixation in higher plants. In: Givnish. TJ. (Ed.), On the
economy of plant form and function. Cambridge University Press, Cambridge, pp. 133-170.
Cox, P.A. and C.J. Humphries, 1993. Hydrophilous pollination and breeding system evolution in
seagrasses: a phylogenetic approach to the evolutionary ecology of the Cymodoceaceae. Bot.
J. Linn. Soc. 113:217-226.
Cram, J., O. Rose and P. Torr, 2001. Contributions of leaf development and leaf fall to salt disposal in
mangroves. In: Abstracts of the International Symposium on Mangroves: Evolution. Physiology
and Conservation. RCAST, University of Tokyo, pp. 41.
Crawford, R.M.M., 1978. Metabolic adaptations to anoxia. In: Hook, D.O. and R.M.M. Crawford
(Eds.), Plant life in anaerobic environments. Ann Arbor Science, Michigan, pp. 119-36.
Crawford, R.M.M., 1992. Oxygen availability as an ecological limit to plant distribution. Adv. Ecol. Res.
23:93-185.
Crawford, R.M.M. and P.O. Tyler, 1969. Organic acid metabolism in relation to flooding tolerance in
roots. J. Ecol. 57:234-244.
Cribb, A.B., 1979. Algae associated with mangroves in Moreton Bay, Queensland. In: Bailey, A. and
N.C. Stevens (Eds.). Proceedings Northern Moreton Bay symposium. Royal Society of
Queensland, Brisbane. pp. 63-69.
Cridland, A.A.. 1964. Amyelon in American coal-balls. Palaeontology 7:186-209.
Crisp, P., L. Daniel and P. Tortell, 1990. Mangroves in New Zealand: trees in the tide. IGP Books,
Auckland.
Critchley, C., 1982. Stimulation of photosynthetic electron transport in a salt-tolerant plant by high
chloride concentrations. Nature 298:483-485.
Critchley, C., 1983. Further studies on the role of chloride in photosynthetic 02 evolution in higher
plants. Biochim. Biophys. Acta 724:1-5.
Critchley. C., I.C. Baianu, Govindjee and H.S. Gutowsky, 1982. The role of chloride in 02 evolution by
thylakoids from salt-tolerant higher plants. Biochim. Biophys. Acta 682:436-445.
Curran. M., 1985. Gas movement~ in the roots of Avicennia marina (Forsk.) Vierh. Aus!. J. PI. Physiol.
12:97-108.
Curran, M., M. Cole and W.G. Allaway, 1986. Root aeration and respiration in young mangrove plant~
(Avicennia marina (Forsk.) Vierh.). J. Exptl. Bot. 37:1225-1233.
Curran, M., P. James and W.G. Allaway, 1996. The measurement of ga.~ spaces in the roots of aquatic
plants - Archimedes revisited. Aquat. Bot. 54:255-261.
Curtiss. A.H .• 1888. How the mangrove forms islands. Garden and Forest 1:100.
Oa Silva, J.A.A., M.R.C.S. De Melo and B.E. Borders. 1993. A volume equation for mangrove trees in
northeast Brazil. For. Ecol. Manage. 58:129-136.
314 Mangrove Ecology, Silviculture and Conservation

Dahdouh-Guebas. F .• C. Mathenge. J.G. Kairo and N. Koedam. 2000. Utilization of mangrove wood
products around Mida Creek (Kenya) amongst subsistence and commercial users. Econ. Bot.
54:513-527.
Dahdouh-Guebas. F .• M. Verneirt. J.F. Tack and N. Koedam. 1997. Food preferences of Neosarmatium
meinerti de Man (Decapoda: Sesarminae) and its possible effect on the regeneration of
mangroves. Hydrobiologia 347:83-89.
Dahdouh-Guebas. F .• M. Verneirt. J.F. Tack. D. Van Spreybroeck and N. Koedam. 1998. Propagule
predators in Kenyan mangroves and their possible effect on regeneration. Mar. Freshw. Res.
49:345-350.
Darwin. c.. 1836. The voyage of the Beagle. Reprinted in 1988 by Mentor. New York.
Das. A.B .• A.K. Mukheljee and P. Das. 2001. Molecular phylogeny of Heritiera Aiton (Sterculiaceae).
a tree mangrove: variations in RAPD markers and nuclear DNA content. Bot. J. Linn. Soc.
136:221-229.
Das. P.• U.C. Basak and A.B. Das. 1997. Restoration of the mangrove vegetation in the Mahanadi delta.
Orissa. India. Mangroves and Salt Marshes 1: 155-161.
Das. P.K.. V. Chakravarti. A. Dutta and S. Maity. 1995. Leaf anatomy and chlorophyll estimates in
some mangroves. Indian Forester 121:289-294.
Davey. J.E .• 1975. Note on the mechanism of pollen release in Bruguiera gymnorhiza. 1. Sth. Afr. Bot.
41:269-272.
Davidson. D.W. and D. McKey. 1993. The evolutionary ecology of symbiotic ant-plant relationship. 1.
Hymenopterist Res. 2:13-83.
Davis. 1.H.• 1940. The ecology and geological role of mangroves in Aorida. In: Carnegie Inst.
Washington Pub!. No. 517. Papers from the Tortugas Lab. 32: 303-412.
Davis. R.A .• 1995. Geologic impact of Hurricane Andrew on Everglades coast of southwest Florida.
Environ. Geol. 25:143-148.
Davis. S.E.• D.L. Childers. 1.W. Day. D.T. Rudnick and F.H. Sklar. 2001. Wetland-water column
exchanges of carbon. nitrogen. and phosphorus in a southern Everglades dwarf mangrove.
Estuaries 24:610-622.
Debenay. J.P .• 1.1. Guillou. J. Pages. M. Ba. G. Moguedet. J.P. Perthuisot and C. Ponthoreau-Granet.
1997. L'ecosysteme de mangrove de la Casamance (au Senegal). In: Kjerfve. B.• L.D.
Lacerda and E.H.S. Diop (Eds.). Mangrove ecosystem studies in Latin America and Africa.
UNESCO. Paris. pp. 224-240.
Delibrias. G .• L. Ortlieb and N. Petit-Maire. 1976. New C J4 dates for the Atlantic Sahara. Tentative
interpretations. J. Human Evo!. 5:535-546.
Den Hartog. C .• 1970. The sea grasses of the world. North Holland Publ. Co .• Amsterdam.
Devoe. N.N. and T.G. Cole. 1997. Diameter and volume increment~ in mangrove forests of the
Federated States of Micronesia. In: Aksornkoae. S.• L. Puangchit and B. Thaiutsa (Eds.).
Tropical Forestry in the 21" Century. Vol. 10: Mangrove Ecosystems. Kasetsart University.
Bangkok. pp. 57-71.
Devoe. N.N. and T.G. Cole. 1998. Growth and yield in mangrove forests of the Federated State of
Micronesia. For. Ecol. Manage. 103:33-48.
Diallo. A .• 1993. The mangrove swamps of Guinea. In: Diop. E.S .• C.D. Field and M. Vannucci (Eds.).
Proceedings of a Workshop on Conservation and Sustainable Utilization of Mangrove Forests
in Latin America and Africa Regions. Dakar. 20-22 January. 1993. ITIO/ISME Project
PDlI4190 (F). pp. 6-7.
Dick. T .• 1915. Origin of the heliman or shield of the New South Wales coast Aborigines. J. Proc. Roy.
Soc. NSW 49:282-288.
Dicks. B .• 1986. Oil and the black mangrove. Avicennia marina in the northern Red Sea. Mar. Poll. Bull.
17:500-503.
Digby. MJ .• P. Saenger. M.B. Whelan. D. McConchie. B. Eyre. N. Holmes and D. Bucher. 1999. A
physical classification of Australian estuaries. National River Health Program. Urban Sub
Program. Report No.9. LWRRDC Occa.~ional Paper 16/99. pp. 1-41.
Din. N.• 1991. Contribution a l'etude botanique et ecologique des mangroves de l'estuaire du Cameroun.
Unpublished Thesis, Universite de Yaounde, Yaounde.
Ding Hou. 1958. Rhizophoraceae. Flora Malesiana (Ser. I) 5:429-493.
Diop. E.S .• A. Soumare. N. Diallo and A. Guisse. 1997. Recent changes of the mangroves of the
Saloum River Estuary. Senegal. Mangroves and Salt Marshes 1:163-172.
Dittmar. T. and R.J. Lara. 2001. Driving forces behind nutrient and organic matter dynamics in a
mangrove tidal creek in North Brazil. Est. Coastal Shelf Sci. 52:249-259.
Dixon. J.A .• 1989. Valuation of mangroves. Tropical Coastal Area Management 4:1-6.
Dixon. R.G .• 1959. A working plan for the Matang Mangrove Forest Reserve. Perak (First revision
1959). Forest Department. Perak, Federation of Malaya.
Dodd. R.S .• Z.A. Rafii and A. Bousquet-Melou. 2000. Evolutionary divergence in the pan-Atlantic
mangrove Avicennia germinans. New Phytol. 145:115-125.
Dodd. R.S .• F. Blasco. Z.A. Rafii and E. Torquebiau. 1999. Mangroves of the United Arab Emirates:
ecotypic diversity in cuticular waxes at the bioclimatic extreme. Aquat. Bot. 63:291-304.
References 315

Dodd, R.S., F. Fromard, Z.A. Rafii and F. Blasco, 1995. Biodiversity among West African Rhizophora:
foliar wax chemistry. Biochem. Syst. Ecol. 23:859-868.
Dodd, R.S., Z.A. Rafii, F. Fromard and F. Blasco, 1998. Evolutionary diversity among Atlantic coast
mangroves. Acta Oceol. 19:323-330.
Dodd, R.S., Z.A. Rafii, N. Kashani and J. Budrick, in press. Gene diversity and population structure in
Avicennia germinans L. (Avicenniaceae) using AFLP molecular markers. Mol. Ecol.
Dolianiti, E., 1955. Frutos de Nipa no Paleoceno de Pernambuco, Brasil. Dep. Nac. Prod. Min., Div.
Geol. e Min. Bol. 158:1-36.
Dominguez, C.A., L.E. Eguiarte, 1. Nunez-Farfan and R. Dirzo, 1998. Flower morphometry of
Rhizophora mangle (Rhizophoraceae): geographical variation in Mexican populations. Amer.
J. Bot. 85:637-643.
Doskotch, P.W., H.Y. Cheng, T.M. Odell and L. Girard. 1980. Nerolidol: an antifeeding sesquiterpene
alcohol for gypsy moth larvae from Melaleuca leucodendron. J. Chern. Ecol. 6:845-851.
Downton, W.1.S., 1982. Growth and osmotic relations of the mangrove Avicennia marina, as influenced
by salinity. Aust. 1. PI. Physiol. 9:519-528.
Draz, 0., 1956. Improvement of animal production in Yemen. Bull. Inst. Desert d'Egypt 6:79-95.
Drennan, P. and P. Berjak, 1979. Degeneration of foliar glands correlated with the shift to abaxial salt
excretion in Avicennia marina (Forsk.) Vierh. Proc. Electron Microsc. Soc. South. Afr. 9:83-
84.
Drennan, P., P. Berjak and N.W. Pammenter, 1992. Ion gradients and adenosine triphosphatase
localization in the salt glands of Avicennia marina (Forssk.) Vierh. Sth Afr. J. Bot. 58:486-490.
Dromgoole, F.I., 1988. Carbon dioxide fixation in aerial roots of the New Zealand mangrove Avicennia
marina var. resinifera. N.Z. J. Mar. Freshw. Res. 22:617-619.
Du Puy, 0.1., 1993. Christmas Island. In: Flora of Australia, Vol. 50 Oceanic Islands 2, Australian
Government Publishing Service, Canherra, pp. 1-30.
Ducker, S.C. and R.B. Knox, 1976. Submarine pollination in seagrasses. Nature 263:705-706.
Dudley, S.A., I 996a. Differing selection on plant physiological traits in response to environmental
water availability: a test of adaptive hypotheses. Evolution 50:92-102.
Dudley, S.A., 1996b. The response to differing selection on plant physiological traits: evidence for local
adaption. Evolution 50: 103-110. .
Duke, N.C., 1984. A mangrove hybrid, Sonneratia x gulngai (Sonneratiaceae) from north-ea~tern
Australia. Austrobaileya 2: 103-105.
Duke, N.C., 1988. An endemic mangrove species, Avicennia integra sp. nov. (Avicenniaceae), in
Northern Australia. Aust. Syst. Bot. 1:177-180.
Duke, N.C., 1990. Phenological trends with latitude in the mangrove tree Avicennia marina. J. Ecol.
78:113-133.
Duke, N.C., 1991a. Nypa in the mangroves of Central America: introduced or relict? Principes 35:127-
132.
Duke, N.C., 1991b. A systematic revision of the mangrove genus Avicennia (Avicenniaceae) in
Australasia. Aust. Syst. Bot. 4:299-324.
Duke, N.C., 1992. Mangrove floristics and biogeography. In: Rohert.~on, A.1. and D.M. Alongi (Eds.),
Tropical mangrove ecosystems. American Geophysical Union, Washington, pp. 63-100.
Duke, N.C., 1995. Genetic diversity, distributional barriers and rafting continents - more thoughts on
the evolution of mangroves. Hydrobiologia 295:167-181.
Duke, N.C., 1996. Mangrove reforestation in Panama: an evaluation of planting in areas deforested by
a large oil spill. In: Field, C.D. (Ed.), Restoration of mangrove ecosystems. ISMElITIO,
Okinawa, pp. 209-232.
Duke, N.e., 2001. Gap creation and regenerative processes driving diversity and structure of
mangrove ecosystems. Wetlands Ecology and Management 9:257-269.
Duke, N.C. and 1.S. Bunt, 1979. The genus Rhizophora (Rhizophoraceae) in north-eastern Australia.
Aust. J. Bot. 27:657-678.
Duke, N.C. and B.R. Jackes, 1987. A systematic revision of the mangrove genus Sonneratia
(Sonneratiaceae) in Australia. Blumea 32:277-302.
Duke, N.C. and E. Wolanski, 2001. Muddy coastal waters and depleted mangrove coastlines - depleted
seagrass and coral reefs. In: Wolanski, E. (Ed.), Oceanographic processes of coral reefs:
physical and biological links in the Great Barrier Reef. CRC, Boca Raton, pp. 77-91.
Duke, N.e., M.e. Ball and J.e. Ellison, 1998. Factors influencing biodiversity and distributional
gradient~ in mangroves. Global Ecol. Biogeogr. Lett. 7:27-47.
Duke, N.C., J.S. Bunt and W.T. Williams, 1984. Observations on the floral and vegetative phenologies
of north-eastern Australian mangroves. Aust. J. Bot. 32:87-99.
Duke, N.C., Z.S. Pinz6n and M.C. Prada, 1997. Large-scale damage to mangrove forests following two
large oil spills in Panama. Biotropica 29:2-14.
Duke, N.C., J.A.H. Benzie, J.A. Goodall and E.R. Bailment, 1998. Genetic structure and evolution of
species in the mangrove genus Avicennia (Avicenniaceae) in the Indo-West Pacific. Evolution
52:1612-1626.
Duke, N.C., K.A. Bums, J.C. Ellison, RJ. Rupp and O. Dalhaus, 1998. Effects of oil and dispersed-oil
on mature mangroves in field trials at Gladstone. J. APPEA 1:637-645.
316 Mangrove Ecology, Silviculture and Conservation

Durako, MJ., WJ. Kenworthy, S.M.R. Fatemy, H. Valavi and G.W. Thayer, 1993. Assessment of the
toxicity of Kuwait crude oil on the photosynthesis and respiration of seagrasses of the Northern
Gulf. Mar. Poll. Bull. 27:223-227.
Eganathan, P. and e.S. Rao, 2001. Manual on vegetative and micropropagation of mangroves. M.S.
Swaminathan Research Foundation, Chennai.
Eganathan, P., C.S. Rao and A. Anand, 2000. Vegetative propagation of three mangrove tree species
by cuttings and air layering. Wetland~ Ecology and Management 8:281-286.
Egnankou, W.M., 1985. Etude des mangroves de C6te d'lvoire: aspect ecologique et recherches sur les
possibilities de leur amenagement. Unpublished Thesis, Paul Sabatier University, Toulouse.
El-Shourbagy, M.N., M. El-Amry and A.R. Anwahi, 1995. Implementation of some controlled and
natural environmental conditions. J. Fac. Sci., U.A.E. Univ. 8:78-89.
Ellison, A.M., 2000. Mangrove restoration: do we know enough? Restoration Ecol. 8:219-229.
Ellison, A.M. and E.J. Farnsworth, 1993. Seedling survivorship, growth, and response to disturbance in
Belizean mangal. Amer. J. Bot. 80:1137-1145.
Ellison, A.M. and E.1. Farnsworth, 1996. Anthropogenic disturbance of Caribbean mangrove
ecosystems: past impacts, present trends, and future predictions. Biotropica 28:549-565.
Ellison, A.M. and EJ. Farnsworth, 1997. Simulated sea-level change alters anatomy, physiology,
growth, and reproduction of red mangrove (Rhizophora mangle L.). Oecologia 112:435-446.
Ellison, A.M., E.1. Farnsworth and R.E. Merkt, 1999. Origins of mangrove ecosystems and the
mangrove biodiversity anomaly. Global Ecol. Biogeogr. Lett. 8:95-115.
Ellison, A.M., E.1. Farnsworth and R.R. Twilley, 1996. Facultative mutualism between red mangroves
and root-fouling sponges in Belizean mangal. Ecology 77:2431-2444.
Ellison, A.M., B.B. Mukherjee and A. Karim, 2000. Testing patterns of zonation in mangroves: scale-
dependence and environmental correlates in the Sundarbans of Bangladesh. J. Ecol. 88:813-
824.
Ellison, J.C., 1989. Pollen analysis of mangrove sediments as a sea-level indicator: assessment from
Tongatapu, Tonga. Palaeogr. Palaeoclimatol. Palaeocol. 14:327-341.
Ellison, J.C., 1991. The Pacific palaeogeography of Rhizophora mangle L. (Rhizophoraceae). Bot. J.
Linn. Soc. 105:271-284.
Ellison, J.C., 1993. Mangrove retreat with rising sea-level. Est. Coastal Shelf Sci. 37:75-87.
Ellison, J.C., 1994. Palaeo-lake and swamp stratigraphic records of Holocene vegetation and sea-level
changes, Mangaia, Cook Islands. Pac. Sci. 48:1-15.
Ellison, J.e., 1997. Mangrove community characteristics and litter production in Bermuda. In: Kjerfve,
B., L.D. Lacerda and E.H.S. Diop (Eds.), Mangrove ecosystem studies in Latin America and
Africa. UNESCO, Paris, pp. 8-17.
Ellison, J.e. and D.R. Stoddart, 1991. Mangrove ecosystem collapse during predicted sea-level rise:
Holocene analogues and implications. J. Coastal Res. 7:151-165.
Elmqvist, T. and P.A. Cox, 1996. The evolution of vivipary in flowering plants. Oikos 77:3-9.
Elouard, P., 1968. Le Nouakchottien, etage quaternaire de Mauretanie. Ann. Fac. Sci., Dakar 22:121-
137.
Elsol, J.A. and P. Saenger, 1983. A general account of the mangroves of Princess Charlotte Bay with
particular reference to zonation of the open shoreline. In: Teas, H.1. (Ed.), Tasks for
vegetation science, Vol. 8. Dr W. Junk, The Hague, pp. 37-46.
Elster, C., L. Perdomo and M.-L. Schnetter 1999b. Impact of ecological factors on the regeneration of
mangroves in the Cienaga Grande de Santa Marta, Colombia. Hydrobiologia 413:35-46.
Elster, C., L. Perdomo, J. Polania and M.-L. Schnetter, I 999a. Control of Avicennia germinans
recruitment and survival by Junonia evarete larvae in a disturbed mangrove forest in
Colombia. J. Trop. Eco!. 15:791-805.
Emmerson, W.O. and L.E. McGwynne, 1992. Feeding and assimilation of mangrove leaves by the crab
Sesarma meinerti de Man in relation to leaf-litter production in Mgazana, a warm-temperate
southern African mangrove swamp. J. Exp. Mar. BioI. Ecol. 157:41-53.
Emery, K.O. and D.G. Aubrey, 1991. Sea levels, land levels and tide gauges. Springer Verlag, New
York.
Essig, p.B., 1973. Pollination in some New Guinea palms. Principes 17:75-83.
Evans, R.K., 1977. Techniques and seasonal growth rate of transplanted white mangroves. In: Lewis,
R.R. and D.P. Cole (Eds.), Proceedings of the fourth annual conference on restoration of
coastal vegetation in Florida. Hillsborough Community College, Florida, pp. 77-105.
Everett, J., 1994. New combinations in the genus Avicennia. Telopea 5:627-629.
Everist, S.L., 1974. Poisonous plant.~ of Australia. Angus & Robertson, Sydney.
Ewel, K.C., J.A. Bourgeois, T.G. Cole and S. Zheng, I 998a. Variation in environmental characteristics
and vegetation in high-rainfall mangrove forests, Kosrae, Micronesia. Global Eco!. Biogeogr.
Lett. 7:49-56.
Ewel, K.C., R.R. Twilley and J.E. Ong, 1998b. Different kinds of mangrove forest.~ provide different
goods and services. Global Ecol. Biogeogr. Lett. 7:83-94.
Ewel, K.C., S. Zheng, ZS. Pinwn and I.A. Bourgeois, 1998c. Environmental effects of canopy gap
formation in high-rainfall mangrove forests. Biotropica 30:510-518.
References 317

Exell. A.W. and Stace. e.A.. 1972. Patterns of distribution in the Combretaceae. In: Valentine. D.H.
(Ed.). Taxonomy. phytogeography and evolution. Academic Press. New York. pp. 307-323.
Exell. A.W .• 1945. Catalogue of the vascular plants of S. Tome (with Principe and Annobon). British
Museum of Natural History. London.
Exell. A.W .• 1973. Angiosperms of the islands of the Gulf of Guinea (Fernando Po. Sao Tome.
Annobon). Bull. Br. Mus. Nat. Hist.. Botany 4:325-411.
Fabre. A.• F. Fromard and V. Trichon. 1998. Fractionation of phosphate in sediments of four
representative mangrove stages (French Guiana). Hydrobiologia 392:13-19.
Faegri. K. and L. van der Pijl. 1971. The principles of pollination ecology. Pergamon. New York.
Fagbami. A.A .• E.J. Udo and C.T.1. Odu. 1988. Vegetation damage in an oil field in the Niger Delta of
Nigeria. J. Trop. Ecol. 4:61-75.
FAO. 1984. Mangrove forests of the Asia-Pacific region: a summary of available information. FAO.
Rome.
Farnsworth. E.J. and A.M. Ellison. 1991. Patterns of herbivory in Belizean mangrove swamps.
Biotropica 24:555-567.
Farnsworth. E.J. and A.M. Ellison. 1993. Dynamics of herbivory in Belizean mangal. l Trop. Ecol.
9:435-453.
Farnsworth. E.J. and A.M. Ellison. 1996. Sun-shade adaptability of the red mangrove. Rhizophora
mangle (Rhizophoraceae): changes through ontogeny at several levels of biological
organisation. Amer. J. Bot. 83:1131-1143.
Farnsworth. E.J. and A.M. Ellison. 1997a. Global patterns in pre-dispersal propagule predation in
mangrove forests. Biotropica 29:318-330.
Farnsworth. E.J. and A.M. Ellison. 1997b. The global conservation status of mangroves. Ambio 26:328-
334.
Farnsworth. E.J. and J.M. Farrant. 1998. Reductions in abscisic acid are linked with viviparous
reproduction in mangroves. Amer. J. Bot. 85:760-769.
Farnsworth. E.J .• A.M. Ellison and W.K. Gong. 1996. Elevated CO2 alters anatomy. physiology. growth
and reproduction of red mangrove (Rhizophora mangle L.). Oecologia \08:599-609.
Farquhar. G.D .• M.H. O'Leary and J.A. Berry. 1982. On the relationship between carbon isotope
discrimination and their intercellular carbon dioxide concentration of leaves. Aust. J. PI.
Physiol. 9:121-137.
Farrant. J.M .• N.W. Pammenter and P. BeJjak. 1992. Development of the recalcitrant (homoiohydrous)
seeds of Avicennia marina (Forssk.) Vierh.: anatomical. ultra~tructural and biochemical events
associated with the development from histodifferentiation to maturation. Ann. Bot. 70:75-86.
Farrell. M.J .• 1973. Studies on the ecology of Victorian mangrove and saltmarsh communities.
Unpublished Thesis. University of Melbourne. Melbourne.
Faure. H. and L. Hebrard. 1977. Variations des lignes de rivage au Senegal et en Mauretanie au cours
de I'Holocene. Studia Geologica Polonica 52: 143-157.
Fauvel. M.-T .• l Gleye and e. Andary. 1989. Verbascoside: a constituent of Clerodendrum inerme.
Planta Medica 55:577.
Fauvel. M.-T .• A. BouSQuet-Melou. e. Moulis. J. Gleye. and S.R. Jensen. 1995. lridoid glucosides from
Avicennia germinans. Phytochem. 38:893-894.
Faye. B .• 1993. Mangrove. secheresse and dromadaire. Secheresse 4:47-55.
Feller. e.. J. Triche!, lC. Fontes and e. Marius. 1989. Sur Ie role de la vegetation dans Ie stockage du
soufre dans les sols de mangrove (Senegal). Resultats preliminaires. Soil BioI. Biochem.
21 :947-952.
Feller. I.C.• 1995. Effects of nutrient enrichment on growth and herbivory of dwarf red mangrove
(Rhizophora mangle). Ecological Monographs 65:477-505.
Feller. I.e.. 1996. Effects of nutrient enrichment on leaf anatomy of dwarf Rhizophora mangle L. (red
mangrove). Biotropica 28:13-22.
Feller. I.e. and W.N. Mathis. 1997. Primary herbivory by wood-boring insects along an architectural
gradient of Rhizophora mangle. Biotropica 29:440-451.
Feller. I.C. and K.L. McKee. 1999. Small gap creation in Belizean mangrove forests by a wood-boring
insect. Biotropica 31 :607-617.
Feller. I.C .• D.F. Whigham. lP. O'Neill and K.L. McKee. 1999. Effects of nutrient enrichment on
within-stand cycling in a mangrove forest. Ecology 80:2193-2205.
Fernandes. M.E.B .• 1999. Phenological patterns in Rhizophora L .• Avicennia L. and Laguncularia
Gaertn. f. in Amazonian mangrove swamps. Hydrobiologia 413:53-62.
Fernando. E.S. and J.V. Pancho. 1980. Mangrove trees of the Philippines. Sylvatrop. Philipp. For. Res.
5:35-54.
Field. e.D .• 1984. Ions in mangroves. In: Teas. H.J. (Ed.). Tasks for vegetation science. Vol. 9. Dr W.
Junk. The Hague. pp. 43-48.
Field. e.D .• 1995. Journey amongst mangroves. International Society for Mangrove Ecosystems/lITO.
Okinawa. Japan.
Field. C.D. (Ed.). 1996. Restoration of mangrove ecosystems. International Society for Mangrove
Ecosystems/lITO. Okinawa. Japan.
Field. e.D.. 1998a. Rationales and practices of mangrove afforestation. Mar. Freshw. Res. 49:353-358.
318 Mangrove Ecology, Silviculture and Conservation

Field, C.D., 1998b. Rehabilitation of mangrove ecosystems: an overview. Mar. Poll. Bull. 37:383-392
Field, C.D., 1999. Mangrove rehabilitation: choice and necessity. Hydrobiologia 413:47-52.
Field, C.D., B.G. Hinwood and \. Stevenson, 1984. Structural features of salt gland of Aegiceras. In:
Teas, HJ. (Ed.), Physiology and management of mangroves, Vol. 8. Dr W. Junk, The Hague,
pp.37-42.
Finn, M., 1996. The mangrove mesocosm of Biosphere 2: design, establishment and preliminary results.
Ecological Engineering 6:21-56.
Fisher, P. and M.D. Spalding, 1997. Protected areas with mangrove habitat. Draft Report. World
Conservation Monitoring Centre, Cambridge.
Fitzgerald, M.A. and W.G. Allaway, 1991. Apoplastic and symplastic pathways in the leaf of the grey
mangrove Avicennia marina (Forsk.) Vierh. New Phytol. 119:217-226.
Fitzgerald, M.A., D.A. Orlovich and W.G. Allaway, 1992. Evidence that abaxial leaf glands are the
sites of salt secretion in leaves of the mangrove Avicennia marina (Forsk.) Vierh. New Phytol.
120:1-7.
Heming, A.I., E.R. Williams and J.W. Turnbull, 1988. Growth and nodulation of provenances of
Casuarina cunnillghamiana inoculated with a range of Fraflkia sources. Aust. J. Bot. 36:171-
181.
Hores-Verdugo, FJ., J.W. Day and R. Briseno-Duenas, 1987. Structure, litter fall, decomposition, and
detritus dynamics of mangroves in a Mexican coastal lagoon with an ephemeral inlet. Mar.
Ecol. Progr. Ser. 35:83-90.
Fong, F.W., 1992. Perspectives for sustainable resource utilization and management of Nipa vegetation.
Econ. Bot. 46:45-54.
Fortes, M.D., 1988. Mangrove and seagrass beds of east Asia: habitats under stress. Ambio 17:207-213.
Fosberg, p.R., 1961. Vegetation-free zone on dry mangrove coasts. US Geol. Soc. Prof. Pap. No.
4240:216-218.
Fouda, M.M. and M. AI-Muharrami, 1995. An initial assessment of mangrove resources and human
activities at Mahout Island, Arabian Sea. Oman. Hydrobiologia 295:353-362.
France, R., 1998. Estimating the assimilation of mangrove detritus by fiddler crabs in Laguna Joyuda,
Puerto Rico, using dual stable isotopes. I. Trop. Ecol. 14:413-425.
Fromard, F., H. Puig, E. Mougin, G. Marty, I.L. Betoulle and L. Cadamuro, 1998. Structure, above-
ground biomass and dynamics of mangrove ecosystems: new data from French Guiana.
Oecologia 115:39-53.
Fuchs, H.P., 1970. Ecological and palynological notes on Pelliciera rhizophorae. Acta Bot. Neerl.
19:884-894.
Fuentes-Yaco, c., D.A. Slas de Leon, M.A. Monreal-G6mez and F. Vera-Herrera, 2001.
Environmental forcing in a tropical estuarine ecosystem: the Palizada River in the southern
Gulf of Mexico. Mar. Freshw. Res. 52:735-744.
Fujimoto, K., 2000. Belowground carbon sequestration of mangrove forests in the Asia-Pacific region.
In: Asia-Pacific cooperation on research for conservation of mangroves: proceedings of an
international workshop, Okinawa, Japan, 26-30 March, 2000. United Nations University,
Tokyo, pp. 87-95.
Fujimoto, K., T. Miyagi, T. Kikuchi and T. Kawana, 1996. Mangrove habitat formation and response to
Holocene sea-level changes on Kosrae Island, Micronesia. Mangroves and Salt Marshes 1:47-
57.
Fujimoto, K., T. Miyagi, T. Kikuchi and T. Kawana, 1999. Below-ground carbon storage of
Micronesian mangrove forests. Ecol. Res. 14:409-413.
Furukawa, K. and E. Wolanski, 1996. Sedimentation in mangrove forests. Mangroves and Salt Marshes
1:3-10.
Furukawa, K., E. Wolanski and H. Mueller, 1997. Currents and sediment transport in mangrove forests.
Est. Coastal Shelf Sci. 44:301-310.
Gallin, E., E. Coppejans and H. Beeckman, 1989. The mangrove vegetation ofGazi Bay (Kenya). Bull.
Soc. Roy. Bot. Belg. 122:197-207.
Galloway, R.W., 1982. Distribution and physiographic patterns of Australian mangroves. In: Clough,
B.F. (Ed.), Mangrove ecosystems in Australia - structure, function and management.
Australian Institute of Marine Science with ANU Press, Canberra, pp. 31-54.
Gara, R.I., A. Sarango and P.G. Cannon, 1990. Defoliation of an Ecuadorian mangrove forest by the
bagworm, Oiketicus kirbyi (Lepidoptera: Psychidae). J. Trop. For. Sci. 3:181-186.
Ge, X.-J. and M. Sun, 1999. Reproductive biology and genetic diversity of a cryptoviviparous
mangrove Aegiceras comiculatum (Myrsinaceae) using allozyme and intersimple sequence
repeat (ISSR) analysis. Mol. Ecol. 8:2061-2069.
Ge, X.-J. and M. Sun, 2001. Population genetic structure of Ceriops fagal (Rhizophoraceae) in Thailand
and China. Wetlands Ecology and Management 9:203-209.
Gee, C.T., 1990. On the fossil occurrence of the mangrove palm Nypa. In: Knobloch, E. and Z. Kvacek
(Eds.), Proc. Symp. Paleofloristic and Paleoclimatic Changes in the Cretaceous and Tertiary.
Geological Survey, Prague, pp. 315-319.
Gee, C.T., 2001. The mangrove palm Nypa in the geologic past of the New World. Wetlands Ecology
and Management 9: 181-194.
References 319

Gentry, A.H., 1982. Phytogeographic patterns as evidence for a Choco refuge. In: Prance, G.T. (Ed.),
Biological diversification in the tropics. Columbia University Press, New Yor, pp. 112-136.
Germeraad, J.H., C.A. Hopping and J. Muller, 1968. Palynology of Tertiary sediments from tropical
areas. Rev. Paleobot. Palynol. 6: 189-348.
Gessner, F., 1967. Untersuchungen an der Mangrove in Ost-Venezuela. Int. Rev. ges. Hydrobiol.
52:769-781.
Getter, C.D., T.G. Ballou and C.B. Koons, 1985. Effects of dispersed oil on mangroves, synthesis of a
seven year study. Mar. Poll. Bull. 16:318-324.
Ghosh, A., S. Misra, A.K. Dutta and A. Choudhury, 1985. Pentacyclic triterpenoids and sterols from
seven species of mangroves. Phytochem. 24: 1725-1271.
Gilbert, AJ. and R. Janssen, 1998. Use of environmental functions to communicate the values of a
mangrove ecosystem under different management regimes. Ecol. Econ. 25:323-346.
Gill, A.M., 1971. Endogenous control of growth-ring development in Avicennia. For. Sci. 17:462-65.
Gill, A.M., 1975. Australia's mangrove enclaves: a coastal resource. Proc. Ecol. Soc. Aust. 8:129-146.
Gill, A.M. and P.B. Tomlinson, 1969. Studies on the growth of red mangrove (Rhizophora mangle L.). I.
Habit and general morphology. Biotropica 1:1-9.
Gill, A.M. and P.B. Tomlinson. 1971a. Studies on the growth of red mangrove (Rhizophora mangle L.).
II Growth and differentiation of aerial roots. Biotropica 3:63-77.
Gill. A.M. and P.B. Tomlinson. 1971b. Studies on the growth of red mangrove (Rhizophora mangle L.).
III Phenology ofthe shoot. Biotropica 3:109-124.
Gill. A.M. and P.B. Tomlinson, 1975. Aerial roots: an array of forms and functions. In: Torrey. J.G. and
D.T. Clarkson (Eds.). The development and function of roots. Academic Press. London. pp.
237-260.
Gill. A.M. and P.B. Tomlinson. 1977. Studies on the growth of red mangrove (Rhizophora mangle L.).
IV. The adult root system. Biotropica 9:145-155.
Gilmore, A.M. and O. Bjl:\rkman. 1994. Adenine nucleotides and the xanthophyll cycle in leaves. I.
Effects of CO 2- and temperature-limited photosynthesis on adenyl ate energy charge and
violaxanthin de-epoxidation. Planta 192:526-536.
Givnish. TJ .• 1988. Adaptation to sun and shade: a whole-plant perspective. Aust. J. PI. Physiol. 15:63-
92.
Goforth. HW. and Thomas. J.R.. 1980. Planting of red mangrove R. mangle L. for stabilisation of the
shoreline in the Florida Keys. In: Cole. D.P. (Ed.). Proceedings of the 6th annual conference
on the restoration and creation of wetlands coastal vegetation in Florida. Hillsborough
Community College. Florida. pp. 207-230.
Goldstein, G., F. Rada. L. da S.L. Sternberg, J.L. Burguera, M. Burguera, A. Orozco. M. Montilla. O.
Zavala. A. Azocar. MJ. Canales and A. Celis. 1989. Gas exchange and water balance of a
mistletoe species and its mangrove hosts. Oecologia 78:176-183.
Golley. F.B., H.T. Odum and R.F. Wilson. 1962. The structure and metabolism of a Puerto Rican red
mangrove forest in May. Ecology 43:9-19.
Gonr,;ales-Alvim, S.1., M.C.F. Vaz dos Santos and G.W. Fernades, 2001. Leaf gall abundance on
Avicennia germinans (Avicenniaceae) along an interstitial salinity gradient. Biotropica 33:69-
77.
Gong. W.K. and J.E. Ong. 1995. The use of demographic studies in mangrove silviculture.
Hydrobiologia 295:255-261.
Gordon, D.M., 1987. Disturbance to mangroves in tropical-arid Western Australia: hypersalinity and
restricted todal exchange as factors leading to mortality. Technical Series No. 12,
Environmental Protection Authority. Perth.
Gore, R., 1977. Wild nursery of the mangroves. Natl. Geog. 151:669-689.
Gossweiler. J. and F.A. Mendonr,;a. 1939. Carta Fitogeognifica de Angola. Republica Portuguesa
Ministerio das Col6nias. Lisbon.
Gowthorpe. P. and B. Lamarche. 1993. Les mangroves de la Mauritanie. In: Diop. E.S .• C.D. Field and
M. Vannucci (Eds.). 1993. Conservation et utilisation rationelle des fOrets de mangrove de
l'Amerique Latine et de l'Afrique. Dakar. 20-22 January. 1993. ITTOIISME Project PDl14190
(F). pp. 3-21.
Graaf, G.1. de. and T.T. Xuan, 1998. Extensive shrimp farming. mangrove clearance and marine
fisheries in the southern provinces of Vietnam. Mangroves and Salt Marshes 2:159-166.
Graham, A., 1975. Late Cenozoic evolution of tropical inland vegetation in Veracruz, Mexico.
Evolution 29:723-735.
Graham. A.. 1977. New records of Pelliciera (Theaceae/Pellicieriaceae) in the Tertiary of the
Caribbean. Biotropica 9:48-52.
Graham, A.. 1995. Diversification of Gulf/Caribbean mangrove communities through Cenozoic time.
Biotopica 27:20-27.
Graham, S.A., H. Tobe and P. Baa.~, 1986. Koehneria, a new genus of Lythraceae from Madagascar.
Ann. Missouri Bot. Gardens 73:788-809.
Grainger. J. and A. Ganadelly, 1984. Hemas: an investigation into a traditional conservation ethic in
Saudi Arabia. J. Saudi Arabian Nat. Hist. Soc. 2:28-32.
320 Mangrove Ecology, Silviculture and Conservation

Grant, D.K.S., 1938. Mangrove woods of Tanganyika Territory, their structure and dependent
industries. Tanganyika Notes and Records, April 1938.
Grant, D.L., PJ. Clarke and W.G. Allaway, 1993. The response of grey mangrove (Avicennia marina
(Forsk.) Vierh.) seedlings to spills of crude oil. J. Exp. Mar. BioI. Ecol. 171 :273-295.
Green, E.P., PJ. Mumby, AJ. Edwards, C.D. Clark and A.C. Ellis, 1998. The a.~sessment of mangrove
areas using high resolution multispectral airborne imagery. J. Coastal Research 14:433-443.
Grewe, F., 1941. Afrikanische Mangrovelandschaften. Verbreitung und wirtschafts-geographische
Bedeutung. Wiss. Veroffentlichungen des Deut~chen Museum flir Liinderkunde (Leipzig) N.F.
9:105-177.
Griffith, A.L., 1950. Working scheme for mangroves of the Zanzibar protectorate. Government
Printer, Zanzibar, 42 pp.
Grime, J.P., 1973. Competition and diversity in herbaceous vegetation - a reply. Nature 244:310-311.
Grime, J.P., 1977. Evidence for the existence of three primary strategies in plants and its relevance to
ecological and evolutionary theory. Amer. Natur. Ill: 1169-1194.
Grime, J.P., 1979. Plant strategies and vegetation processes. John Wiley, Chichester.
Grime, J.P. and B.D. Campbell, 1991. Growth rate, habitat productivity, and plant strategy as predictors
of stress response. In: Mooney, H.A., W.E. Winner and EJ. Pell (Eds.), Responses of plants to
multiple stresses. Academic Press, San Diego, pp. 143-159.
Grosse, W. and H. Mevi-Schutz, 1987. A beneficial gas transport system in Nymphoides peltata. Amer.
J. Bot. 42:92-98.
Gruas-Cavagnetto, C., 1987. Nouveaux elements megathermes dans la palynotlore Eocene du ba.~sin
parisien. Mem. Trav. E.P.H.E., Inst. Montpellier 17:207-233.
Gruas-Cavagnetto, C., Y. Tambareau, and 1. Villatte, 1988. Donnees paleoecologiques nouvelles sur Ie
Thanetien et l'I1erdien de l'avant-pays pyreneen et de la Montagne Noire. Inst. fro Pondichery,
Trav. Sec. Sci. Tech. 25:219-235.
Guppy, H.B., 1906. Observations of a naturalist in the Pacific between 1896 and 1899. Vol. 2, Plant
dispersal. Macmillan, London.
Hachimou, I., 1993. The mangrove swamps of Benin. In: Diop, E.S., C.D. Field and M. Vannucci
(Eds.), Proceedings of a workshop on conservation and sustainable utilization of mangrove
forests in Latin America and Africa regions. Dakar, 20-22 January, 1993. ITTOIISME Project
PDII4I9O(F),pp.II-12.
Halliday, LA. and W.R. Young, 19%. Density, biomass and species composition of fish in a subtropical
Rhizophora stylosa mangrove forest. Mar. Freshw. Res. 47:609-615.
Hamilton, L.S. and S.C. Snedaker (Eds.), 1984. Handbook for mangrove area management.
Environment and Policy Institute, East-West Centre, Hawaii.
Hanagata, N., T. Takemura. I. Karube and Z. Dubinsky, 1999. Salt/water relationships in mangroves.
Israel J. PI. Sci. 47:63-76.
Handler, S.H. and HJ. Teas, 1983. Inheritance of albinism in the red mangrove, (Rhizophora mangle
L.). In: Teas, HJ. (Ed.), Tasks for vegetation science. Vol. 8, Dr W. lunk, The Hague, pp.
117-21.
Hara, T., 1984. Modelling the time course of self-thinning in crowded plant populations. Ann. Bot.
53: 181-188.
Harbison, P., 1981. The case for the protection of mangrove swamps - geochemical considerations.
Search 12:273-276.
Harbison, P., 1986. Mangrove muds - a sink or source for trace metals. Mar. Poll. Bull. 17:246-250.
Hatton, J.C. and A.L. Conto, 1992. The effect of coastline changes on mangrove community structure,
Portuguese Island, Mozambique. Hydrobiologia 247:49-57.
Havanond. S., 1987. Effects of mud lobsters (Thalassina anomala Herbst.) on plant succession in
mangrove forests, Thailand. Bull. Mar. Sci. 41:635-636.
Heads, M., 2001. Regional patterns of biodiversity in New Guinea plant~. Bot. J. Linn. Soc. 136:67-73.
Heald, EJ. 1969. The production of organic detritus in a south Florida estuary. Unpublished Thesis,
University of Miami, Coral Gables, Florida.
Heald, EJ., 1971. The production of organic detritus in a south Florida estuary. Univ. Miami Sea Grant
Tech. Bull. No.6, pp. I 10.
Heatwole. H., 1985. Survey of the mangroves of Puerto Rico ... a benchmark study. Carib. J. Sci.
21:85-99.
Hedin, L., 1928. L'exploitation du paletuvier dans la baie de Manoka (Cameroun). Revue de Botanique
appliquee 8:623-626.
Heger!, EJ. 1982. Mangrove management in Australia. In: Clough, B.F. (Ed.), Mangrove ecosystems in
Australia - structure, function and management. Australian Institute of Marine Science with
ANU Press, Canberra. pp. 275-88.
Heger!, EJ .• P.I.F. Davie, G.F. Claridge and A.G. Elliott. 1979. The Kakadu National Park mangrove
forests and tidal marshes. Vol. 1, A review of the literature and results of a field
reconnaissance. Report prepared for the Aust. Nat. Parks and Wildl. Servo Brisbane,
Australian Littoral Society, Brisbane.
References 321

Hemminga, M.A., Fl. Slim, J. Kazunga. G.M. Ganssen, J. Nieuwenhuize and N.M. Kruyt, 1994.
Carbon outwelling from a mangrove forest with adjacent sea grass beds and coral reefs (Gazi
Bay, Kenya). Mar. Ecol. Progr. Ser. 106:291-301.
Henry, H.A.L. and L.W. Aarssen, 1999. The interpretation of stem diameter-height allometry in trees;
biomechanical constraints, neighbour effects, or biased regressions. Ecol. Lett. 2:89-97.
Heske, F., 1937. Kolonialwald und Kolonialwaldwirtschaft. D. Forstb.-Zeit. 3:632-657.
Hesse, P.R., 1961a. Decomposition of organic matter in a mangrove swamp soil. Plant and Soil 14:249-
263.
Hesse, P.R., 1961b. Some differences between the soils of Rhizophora and Avicennia mangrove
swamps in Sierra Leone. Plant and Soil 14:335-346.
Hicks, D.B., and L.A. Bums. 1975. Mangrove metabolic response to alterations of natural freshwater
drainage to southwestern Florida estuaries. In: Walsh, G.E., S.C. Snedaker and H.J. Teas
(Eds.), Proceedings International Symposium Biology and Management of Mangroves. Vol. I.
University of Florida, Gainesville, pp. 238-255.
Hogarth, P.J., 1999. The biology of mangroves. Oxford University Press, Oxford.
Hogg, R.W., and F.T. Gillan, 1984. Fatty acids, sterols and hydrocarbons in the leaves from eleven
species of mangrove. Photochem. 23: 93-97.
Holguin, G., P. Vazquez and Y. Bashan, 200 I. The role of sediment microorganisms in the productivity,
conservation, and rehabilitation of mangrove ecosystems: an overview. BioI. Fert. Soils 33:265-
278.
Holldobler, B., 1983. Territorial behavior in the green tree ant (OecophyUa smaragdina). Biotropica
15:241-250.
Holldobler, Band E.O. Wilson, 1978. The multiple recruitment systems of the African weaver ant
Oecophylla longinoda (Latreille) (Hymenoptera: Formicidae). Behav. Ecol. Sociobiol. 3:19-60.
Holldobler, B. and E.O. Wilson, 1990. The ants. Belknap Press of Harvard University Press,
Cambridge.
Hollins, S. and P.V. Ridd, 1997. Evaporation over a tropical tidal salt flat. Mangroves and Salt Marshes
1:95-102.
Hollins, S.E., P.V. Ridd and W.W. Read, 2000. Measurement of the diffusion coefficient for salt in salt
flat and mangrove soils. Wetlands Ecology and Management 8:257-262.
Hong, P.N., 1996. Restoration of mangrove ecosystems in Vietnam: a case study of Can Gio District,
Ho Chi Minh City. In: Field, C.D. (Ed.), Restoration of mangrove ecosystems. International
Society for Mangrove Ecosystems, Okinawa, pp. 76-96.
Hong, P.N. and H.T. San, 1993. Mangroves of Vietnam. IUCN, Bangkok.
Hosokawa, T., H. Tagawa and V.J. Chapman, 1977. Mangals of Micronesia, Taiwan. Japan, the
Philippines and Oceania. In: Chapman, V.J. (Ed.), Ecosystems of the world. I. Wet coastal
ecosystems. Elsevier, Amsterdam, pp. 271-92.
Houston, W.A., 1999. Severe hail damage to mangroves at Port Curtis, Australia. Mangroves and Salt
Marshes 3:29-40.
Hovenden, M.J. and W.G. Allaway, 1994. Horizontal structures on pneumatophores of Avicennia
marina (Forsk.) Vierh. - a new site of oxygen conductance. Ann. Bot. 73:377-383.
Hovenden, M.J., M. Curran, M.A. Doyle, P.F.E. Goulter, N.J. Skelton and W.G. Allaway, 1995.
Ventilation and respiration in roots of one-year-old seedlings of grey mangrove, A vicennia
marina (Forsk.) Vierh. Hydrobiologia 295:23-29.
Howarth, R.W., J.R. Fruci and D. Sherman, 1991. Inputs of sediment and carbon to an estuarine
ecosystem: influence of land use. Ecological Applications 1:27-39.
Hseu, Z. and Z. Chen, in press. Characteristics of mangrove soils in Taiwan. In: Lin, c., M.D. Melville
and L.A. Sullivan (Eds.), Acid sulfate soils in Australia and China. Science Press, Beijing.
Hsueh, M.-L. and H.-H. Lee, 2000. Diversity and distribution of the mangrove forests in Taiwan.
Wetlands Ecology and Management 8:233-242.
Htay, D.N., 1996. Problem of mangrove degradation in Myanmar. In: Khemnark, C. (Ed.), Proceedings
of Ecotone IV: ecology and management of mangrove restoration and regeneration in east and
southeast Asia. Surat Thani, Thailand, pp. 154-160.
Huang, S. and Y. Chen, 1997. The genetic differentiation in Kandelia candel along Ryukyu archipelago
and Taiwan. In: Aksornkoae, S., L. Puangchit and B. Thaiutsa (Eds.), Tropical Forestry in the
21" Century. Vol. 10: Mangrove Ecosystems. Kasetsart University, Bangkok, pp. I-II.
Huang, S., 1994. Genetic variation of Kandelia candel (L.) Druce (Rhizophoraceae) in Taiwan. In:
Drysdale, R.M., S.E.T. John and A.C. Yapa (Eds.), Proc. International Symposium on genetic
conservation and production of tropical forest tree seed. ASEAN-Canada Forest Tree Seed
Centre Project, Muak-Lek, Saraburi, Thailand, pp. 165-172.
Hughes, R.H., and Hughes, J.S. 1992. Repertoire des zones humides d'Afrique. IUCNIUNEP/wCMC.
Cambridge, 808 pp. ..
Humboldt, A. von, 1985. Vom Orinoko zum Amazonas: Reise in die Aquinoktial-Gegenden des neuen
Kontinents. F.A. Brockhaus. Wiesbaden. Original edition, Voyage aux regions equinoctiales du
Nouveau Continent, fait en 1799-1804, published 1808-1834.
Humm, H.J., 1944. Bacterial leaf nodules. J. N.Y. Bot. Gdn. 45: 193-99.
322 Mangrove Ecology, Silviculture and Conservation

Hussain, Z. and G. Acharya (Eds.), 1994. Mangroves of the Sundarbans. Vol. II. Bangladesh. lUCN,
Bangkok, Thailand.
Hutchings, P. and P. Saenger, 1987. Ecology of mangroves. Queensland University Press, Brisbane.
Hutchinson, J., J.M. Dalziel, R.W.J. Keay and F.N. Hepper, 1954n2. Flora of Tropical West Africa.
2nd Edition, 3 volumes.
Hutter, H., 1906. 1m Olgebiet von Kamerun. Globus 89:1-7.
Huxley, C.R., 1978. The ant-plants Myrmecodia and Hydnophytum (Rubiaceae) and the relationships
between their morphology, ant occupants, physiology and ecology. New Phytol. 80:231-268.
Huxley, C.R. and D.F. Cutter (Eds.), 1991. Ant-plant interactions. Oxford University Press, Oxford.
Hyde, K.D. and S.Y. Lee, 1995. Ecology of mangrove fungi and their role in nutrient cycling: what
gaps occur in our knowledge? Hydrobiologia 295:107-118.
Ibrahim, S. and I. Hashim, 1990. Classification of mangrove forest by using I :40,000-scale aerial
photographs. For. Ecol. Manage. 33:583-592.
Imbert, D. and S. Menard. 1997. Structure de la vegetation et production primaire dans la mangrove de
la Baie de Fort-de-France. Martinique (F.W.I.). Biotropica 29:413-426.
Imbert. D. and B. Rollet. 1989. Phytomass aerienne et production prilJlaire dans la mangrove du Grand
Cul-de-sac Marine (Guadeloupe. Antilles fran~aises). Bull. Ecol. 20:27-39.
Imbert. D.• P. Labbe and A. Rousteau. 19%. Hurricane damage and forest structure in Guadeloupe.
French West Indies. 1. Trop. Ecol. 12:663-680.
Imbert. D.• A. Rousteau and P. Scherrer. 2000a. Ecology of mangrove growth and recovery in the
Lesser Antilles: state of knowledge and basis for restoration projects. Restoration Ecol. 8:230-
236.
Imbert. D.• I. Bonheme, E. Saur and C. Bouchon, 2000b. Floristics and structure of the Ptercocarpus
ofjicinalis swamp forest in Guadeloupe. Lesser Antilles. J. Trop. Ecol. 16:55-68.
Inouye. D. W.• 1980. The terminology of floral larceny. Ecology 61: 1251-1253.
Inouye. D.W., 2000. The ecological and evolutionary significance of frost in the context of climate
change. Ecol. Lett. 3:457-463.
Ish-Shalom-Gordon. N. and Z. Dubinsky, 1990. Possible modes of salt secretion in A vicennia marina in
the Sinai. Plant Cell Physiol. 31 :27-32.
Ish-Shalom-Gordon, N. and Z. Dubinsky. 1992. Ultra.~tructure of the pneumatophores of the mangrove
Avicennia marina. S. Afr. J. Bot. 58:358-362.
Ish-Shalom-Gordon. N. and Z. Dubinsky, 1993. Diurnal pattern of salt secretion in leaves of the black
mangrove, Avicennia marina, on the Sinai coast of the Red Sea. Pac. Sci. 47:51-58.
Islam, M.R. and N.A. Siddiqi, 1987. Seed requirement and method of sowing keora (Sonneratia apetala
Buch.-Ham.) seeds in the nursery. Bano Biggyan Patrika 16:73-77.
Islam, M.R., M.A.S. Khan, N.A. Siddiqi and P. Saenger, 1992. Optimal planting season for keora.
Bangladesh J. Forest Sci. 19:1-9.
Islam, M.R., M.A.S. Khan, N.A. Siddiqi and P. Saenger, 1993. Effect of thinning on keora survival and
growth. Bangladesh J. Forest Sci. 20:8-24.
Itai, C .• A. Richmond and Y. Vaadia. 1968. The role ofroot cytokinins during water and salinity stress.
Israel 1. Bot. 17: 187-95.
lUCN, 1994. 1993 United Nations list of national parks and protected areas. Prepared by WCMC and
CNPPA. lUCN. Gland. Switzerland and Cambridge, UK.
Jaffar. M.• 1993. Country report of Fiji on the economic and environmental value of mangrove forest
and present state of conservation. ISME Mangrove Ecosystems Technical Reports No.1. pp.
135-170.
Janssonius, H.H .• 1950. The vessels in the wood of Javan mangrove trees. Blumea 6:464-469.
Janzen. D.H., 1974. Epiphytic myrmecophytes in Sarawak: mutualism through the feeding of plants by
ants. Biotropica 6:237-259.
Janzen, D.H., 1985. Mangroves: where's the understory? J. Trop. Ecol. 1:89-92.
Jaramillo, C. and G. Bayona, 2000. Mangrove distribution during the Holocene in Tribuga Gulf,
Colombia. Biotropica 32:14-22.
Jawa, R.R. and P.B.L. Srivastava, 1989. Dispersal of natural regeneration in some piai-invaded areas of
mangrove forests in Sarawak. For. Ecol. Manage. 26:155-177.
Jayasekera, R., 1991. Chemical composition of the mangrove, Rhizophora mangle L. 1. Plant Physiol.
138:119-121.
Jenik, J., 1970. Root system of tropical trees 5. The peg-roots and the pneumathodes of Laguncularia
racemosa Gaertn. Preslia (Prague) 42:105-113.
Jensen, J.R., H. Lin, X. Yang, E. Ramsey, B.A. Davis and C.W. Thoemke, 1991. The measurement of
mangrove characteristics in southwest Florida using SPOT multispectral data. Geocarto Int.
6:13-22.
Jimenez, J.A., 1984. A hypothesis to explain the reduced distribution of the mangrove PeLliciera
rhizophorae Tr. & PI. Biotropica 16:304-308.
Jimenez, J.A., 1987. A clarification on the existence of Rhizophora species along the Pacific coast of
Central America. Brenesia 28:25-32.
Jimenez, J.A., 1988. Floral and fruiting phenology of trees in a mangrove forest on the dry Pacific coast
of Costa Rica. Brenesia 29:33-50.
References 323

Jimenez. J.A .• A.E. Lugo and G. Cintron. 1985. Tree mortality in mangrove forests. Biotropica 17:177-
185.
Jimenez. lA .• R Martinez and L. Encarnacion. 1985. Massive tree mortality in a Puerto Rican
mangrove forest. Carib. J. Sci. 21 :75-78.
John. D.M. and G.W.A. Lawson. 1990. Review of mangrove and coastal ecosystems in West Africa
and their possible relationships. Est. Coast. Shelf Sci. 31 :505-518.
Johns. R.I .• 1986. The instability of the tropical ecosystem in New Guinea. Blumea 31:341-371.
Johnson. Rand R Johnson. 1993. Mangroves of Sierra Leone. In: Diop. E.S .• C.D. Field and M.
Vannucci (Eds.). Proceedings of a Workshop on Conservation and Sustainable Utilization of
Mangrove Forests in Latin America and Africa Regions. Dakar. 20-22 January. 1993.
ITTOIISME Project PDI14190 (F). pp. 7-9.
Johnston. D.• B. Clough. T.T. Xuan and M. Phillips. 1999. Mixed shrimp-mangrove forestry farming
systems in Ca Mau Province. Vietnam. Aquaculture Asia 4:6-12.
Johnstone. I.M .• 1981. Consumption of leaves by herbivores in mixed mangrove stands. Biotropica
13:252-259.
Johnstone. I.M .• 1983. Succession in zoned mangrove communities: Where is the climax? In: Teas. H.I.
(Ed.). Tasks for vegetation science. Vol. 8. Dr W. Junk. The Hague. pp. 131-140.
Jones. T.G.H. and J.M. Harvey. 1936. Essential oils from the Queensland flora.
Jones. W.T.• 1971. The field identification and distribution of mangroves in eastern Australia. Qd. Nat.
20: 35-51.
Joshi. A.C.. 1933. A suggested explanation of the prevalence of vivipary on the seashore. J. Ecol.
21 :209-212.
Joshi. G.V .• B.B. Jamale and L. Bhosale. 1975a. Ion regulation in mangroves. In: Walsh. G.E .• S.c.
Snedaker and H.I. Teas (Eds.). Proceedings of the International Symposium on Biology and
Management of Mangroves. University of Florida. Gainesville. pp. 595-607.
Joshi. G.V .• M. Pimlaskar and L.I. Bhosale. 1972. Physiological studies in germination of mangroves.
Bot. Mar. 15:91-95.
Joshi. G.V .• L.I. Bhosale. B.B. JamaJe and B.A. Karadge. 1975b. Photosynthetic carbon metabolism in
mangroves. In: Walsh. G.E .• S.C. Snedaker and H.J. Teas (Eds.). Proceedings of the
International Symposium on Biology and Management of Mangroves. Vol. 2. University of
Florida. Gainesville. 2. pp. 579-94.
Joshi. G.V .• M.D. Karekar. C.A. Jowda and L. Bhosale. 1974. Photosynthetic carbon metabolism and
carboxylating enzymes in algae and mangroves under saline conditions. Photosynthetica 8:51-
52.
Kalck. Y.. 1978. Evolution des zones Ii mangroves du Senegal au Quaternaire recent: Etudes
geologiques et geochimiques. Unpublished Thesis. University of Louis Pasteur. Strasbourg.
Kaly. U.L. and G.P. Jones. 1998. Mangrove restoration: a potential toni for coastal management in
tropical developing countries. Ambio 27:656-661.
Karmarkar. S.M .• 1982. Senescence in mangroves. In: Sen. D.N. and K.S. Rajpurohit (Eds.). Tasks for
vegetation science. Vol. 2. Dr W. Junk. The Hague. pp. 173-187.
Kangas. P.c. and A.E. Lugo. 1990. The distribution of mangroves and saltmarsh in Florida. Tropical
Ecol. 31:32-39.
Kapetsky.l.M .• 1985. Mangroves. fisheries and aquaculture. FAO Fish. Rep. 338 (Suppl.):17-36.
Karnosky. D.F.. 1981. Potential for forest tree improvement via tissue culture. BioScience 31:114-120.
Karstedt. P. and N. Parameswaran. 1976. Beitrag zur Anatomie und Systematik der atlantischen
Rhizophora-Arten. Bot. Jahrb. Syst. Pflangzengesch. Pflanzengeogr. 97:317-338.
Kassas. M.• 1957. On the ecology of the Red Sea coasta1land. 1 Ecol. 45:187-203.
Kathiresan. K.• 1992. Foliovory in Pichavaram mangroves. Environment and Ecology 10:988-989.
Kathiresan. K.• 1995. Rhizophora anMmalayaM: a new species of mangrove. Environment and
Ecology 13:240-241.
Kathiresan. K. and B.L. Bingham. 2001. Biology of mangroves and mangrove ecosystems. Adv. Mar.
BioI. 40:81-251.
Kathiresan. K. and M.X. Ramesh. 1990. Establishment of seedlings of a mangrove. Indian Forester
117:225.
Kathiresan. K. and Ravikumar. S.• 1995. Vegetative propagation through air-layering in two species of
mangroves. Aquat. Bot. 50: 107-1 10.
Kathiresan. K. and T.S. Thangam. 1989. Effect of leachates from mangrove leaf on rooting in
Rhizophora seedlings. Geobios 16:27-29.
Kathiresan. K.• P. Moorthy and N. Rajendran. 1993. Promotory effect of some chemicals on seedling
growth of Rhizophora apiculata Blume. Environment and Ecology 11 :716-717.
Kathiresan. K.. P. Moorthy and S. Ravikumar. 1996. A note on the influence of salinity and pH on
rooting of Rhizophora mucroMta Lamk. Seedlings. Indian Forester 122:763-764.
Kato. A.. 1975. Brugine from 8ruguiera cyclindrica. Phytochem. 14:1458.
Keay. R.W.J .• 1953. Rhizophora in West Africa. Kew Bull. 8:121-127.
Keene. A.F. and Melville. M.D .• 1999. Role of nutrients in the delimitation of the grey mangrove.
Avicennia mariM. in estuarine sediments. In: Kesby. J.A .• 1M. Stanley. RF. McLean and L.I.
Olive (Eds.). Geodiversity: readings in Australian geography at the close of the 20th century.
324 Mangrove Ecology, Silviculture and Conservation

Special Publication Series No.6. Canberra. ACT. School of Geography and Oceanography.
University College. Australian Defence Force Academy. pp. 45-53.
Kehar. N.D. and S.S. Negi. 1953. Mangrove (Avicennia officimdis Linn.) leaves as cattle feed. Science
and Culture 18:382-383.
Kelaher. B.P.• M.G. Chapman and AJ. Underwood. 1998a. Changes in benthic a.~semblages near
boardwalks in temperate urban mangrove forests. J. Exp. Mar: BioI. Ecol. 228:291-307.
Kelaher. B.P.• AJ. Underwood and M.G. Chapman. 1998b. Effect of boardwalks on the semaphore
crab Heloecius cordiformis in temperate urban mangrove forests. J. Exp. Mar. BioI. Ecol.
227:281-300.
Kelly. T .• 1995. Effects of long-term discharges of treated sewage on the nutrient status of adjacent
mangrove communities. Unpublished Thesis. Southern Cross University. Lismore.
Kemis. J.R .• 1984. Petiolar glands in Combretaceae: new observations and an anatomical description of
the extrafloral nectary of buttonwood (Collocarpus erectus). Amer. J. Bot. (Abstracts) 71 :34.
Kemis. J.R. and N.R. Lersten. 1984. Petiolar glands in Combretaceae: review of past ambiguities and an
anatomical description of the sunken gland of white mangrove (LagU1lcularia racemosa).
Amer. J. Bot. (Abstract~) 71:34-35.
Khan. M.A. and I. Aziz. 2001. Salinity tolerance in some mangrove species from Pakistan. Wetlands
Ecology and Management 9:219-223.
Khan. M.A.R .• 1986. Wildlife in Bangladesh mangrove ecosystem. J. Bombay Nat. Hist. Soc. 83:32-48.
Kienholz. R.. 1926. An ecological anatomical study of beach vegetation in the Philippines. Proc. Amer.
Phil. Soc. 65 (Suppl.):58-100.
Kimani. E.N .• G.K. Mwatha. E.O. Wakwabi. J.M. Ntiba and B.K. Okoth. 1996. Fishes of a shallow
tropical mangrove estuary. Gazi. Kenya. Mar. Freshw. Res. 47:857-868.
Kimball. M. and HJ. Teas. 1975. Nitrogen fixation in mangrove areas of South Florida. In: Walsh.
G.E .• S.c. Snedaker and HJ. Teas (Eds.). Proceedings of the International Symposium on
Biology and Management of Mangroves. Vol. 2. University of Florida. Gainesville, pp. 654-
661.
Kimura. M. and H. Wada. 1989. Tannins in mangrove tree roots and their role in the root environment.
Soil Sci. Plant Nutr. 35: 101- 108.
Kinako. P.D.S .• 1977. Conserving the mangrove forest of the Niger Delta. Bioi Conserv. 11:35-39.
Kingsley, M., 1897. Travels in West Africa. Macmillan Co. Ltd .• London.
Kitamura. S. and S. Baba, 1997. Phenological studies on mangroves in Indonesia. I. Flowering and
fruiting seasons of Rhizophora apiculata on Bali Island. In: Aksornkoae. S .• L. Puangchit and 8.
Thaiutsa (Eds.). Tropical Forestry in the 21" Century. Vol. 10: Mangrove Ecosystems.
Kasetsart University. Bangkok. pp. 13-20.
Kitaya. Y .• K., K. Yabuki. A. Tani, M. Kiyota. T. Hirano and I. Aiga, 2001. Gas exchange and oxygen
concentrations in pneumatophores of four mangrove species. In: Abstracts of the International
Symposium on Mangroves: Evolution. Physiology and Conservation. RCAST. University of
Tokyo. pp. 31.
Kitheka. J.U., B.M. Mwashote. B.O. Ohowa and J. Kamau, 1999. Water circulation. groundwater
outflow and nutrient dynamics in Mida Creek. Kenya. Mangroves and Salt Marshes 3:135-146.
Kjerfve. B. and L.D. Lacerda. 1993. Mangroves of Brazil. In: Lacerda. L.D. (Ed.). Conservation and
sustainable utilization of mangrove forersts in Latin America and Africa regions. Part I: Latin
America. ITIO/ISME Project PDI1419O(F). Okinawa. pp. 245-272.
Kjerfve. 8. and DJ. Macintosh, 1997. The impact of climatic change on mangrove ecosystems. In:
Kjerfve. 8.. L.D. Lacerda and E.H.S. Diop (Eds.). Mangrove ecosystem studies in Latin
America and Africa. UNESCO, Paris. pp. 1-7.
Klekowski. E.J., R. Lowenfeld and P.K. Hepler. 1994a. Mangrove genetics: II. Outcrossing and lower
spontaneous mutation rates in Puerto Rican Rhizophora. Int. J. Plant Sci. 155:373-381.
Klekowski. EJ .• J.D. Corredor. J.M. Morelli and C. Del Castillo. I994b. Petroleum pollution and
mutation in mangroves. Mar. Poll. Bull. 28:166-169.
Ko. W.C., 1985. Notes on the genus Sonneratia (Sonneratiaceae) in south-east Asia. Acta
Phytotaxonomica Siniea 23:311-314.
Ko. W.C .• 1993. Notes on some Sonneratia (Sonneratiaceae) of China. Journal of Tropical and
Subtropical Botany 1:11-13.
Koch. M.S .• 1997. Rhizophora mangle L. seedling development into the sapling stage across resource
and stress gradients in subtropical Florida. Biotropica 29:427-439.
Koch, M.S. and S.c. Snedaker. 1997. Factors influencing Rhizophora mangle L. seedling development
in Everglades carbonate soils. Aquat. Bot. 59:87-98.
Koch. V. and M. Wolff. 1996. The mangrove snail Thais kiosquiformis Duclos: a case of life history
adaptation to an extreme environment. J. Shellfish Res. 15:421-432.
Kogo. M.. 1988. Natural environmental factors affecting mangrove growth in the early stages - a
study from the experimental cultivation in Saudi Arabia, Abu Dhabi and Pakistan. Proceedings
of the UNDPIUNESCO Regional Mangrove Project. Colombo. 11-14 Nov. 1986, pp. 165-179.
Koizumi, M., K. Takahashi, K. Mineuchi. T. Nakamura and H. Kano. 1998. Light gradients and the
transverse distribution of chlorophyll fluorescence in mangrove and Camellia leaves. Ann. Bot.
81:527-533.
References 325

Kokpol, U., W. Chavasiri, V. Chittawong and D.H. Miles, 1990. Taraxeryl cis-p-hydroxycinnamate, a
novel taraxeryl from Rhizophora apiculata. J. Nat. Prod. 53:953-955.
Kolehmainen, S.E., F.D. Martin and P.B. Schroder. 1974. Thermal studies on tropical marine
ecosystems in Puerto Rico. Symposium on the Physical and Biological Effects of the
Environment of Cooling Systems and Thermal Discharges at Nuclear Stations, Oslo, 26-30
Aug. 1974.
Komiyama, A., V. Chimchome and J. Kongsangchai, 1992. Dispersal patterns in mangrove propagules
- a preliminary study on Rhizophora mucronatu. Res. Bull. Fac. Agric. Gifu Univ., pp. 27-34.
Komiyama, A., K. Ogino, S. Aksornkoae and S. Sabhasri, 1987. Root biomass of a mangrove forest in
southern Thailand. J. Estimation by the trench method and the zonal structure of root biomass.
J. Trop. Ecol. 3:97-108.
Komiyama, A., H. Moriya, S. Prawiroatmodjo, T. Toma and K. Ogino, 1988. Forest as an ecosystem, its
structure and function. 2. Primary productivity of mangrove forest. In: Ogino, K. and M.
Chihara (Eds.), Biological system of mangroves. A report of east Indonesian mangrove
expedition 1986. Ehime University, Ehime, pp. 97-117.
Komiyama, A., T. Santiean, M. Higo, P. Patanaponpaiboon, 1. Kongsangchai and K. Ogino, 1996.
Microtopography, soil hardness and survival of mangrove (Rhizophora upicuiatu BI.) seedlings
planted in an abandoned tin-mining area. For. Ecol. Manage. 81:243-248.
Komiyama, A., S. Havanond, W. Sisawatt, Y. Mochida, K. Fujimoto, T. Ohnishi, S. Ishihara and T.
Miyagi, 2000. Top/root biomass ratio of a secondary mangrove (Ceriops tugal (Perr.) C.B.
Rob.) forest. For. Ecol. Manage. 139:127-134.
Kondo, K., T. Nakamura, K. Tsuruda, N. Saito and Y. Yaguchi, 1987. Pollination in Bruguiera
gymnorrhiza and Rhizophoru mucronatu (Rhizophoraceae) in Ishigaki Island, The Ryukyu
Islands, Japan. Biotropica 19:377-380.
Kostermans, A.J.G.J., 1959. Monograph of the genus Heritiera Aiton (Sterculiaceae). Reinwardtia
4:465-483.
Kovacs, J.M., M. Blanco-Correa and F. Flores-Verdugo, 2001. A logistic regression model of hurricane
impacts in a mangrove forest of the Mexican Pacific. J. Coastal Res. 17:30-37.
Kumar, A. and K. Gurumurthi, 1999. Effect of Frankia on growth and nodulation of Casuarina
equisetifolia. I!.Idian Forester 125:490-495.
Kunkel, G., 1966. Uber die Struktur und Sukzession der Mangrove Liberias und deren
Randforrnationen. Ber. Schweiz. Bot. Ges. 75:20-40.
Kuraishi, S., H. Miyauchi, N. Sakurai and K. Supappibul, 1985a. Diurnal changes in transpiration and
stomatal aperture of mangroves grown in Trat, Thailand. In: Mangrove estuarine ecology in
Thailand. Thai-Japanese co-operative research project on mangrove productivity and
development, 1983-1984, pp. 5-19.
Kuraishi, S., K. Kojima, H. Miyauchi, N. Sakurai and H. Tsubota, 1985b. Brackish water and soil
components of mangrove forests on Iriomote Island. Japan. Biotropica 17:277-286.
Kwok, P.W. and S.Y. Lee, 1995. The growth performances of two mangrove crabs, Chiromanthes
bidens and Parasesurma plicata under different leaf litter diets. Hydrobiologia 295: 141-148.
Kytii, M., P. Niemela and S. Larsson, 1996. Insects on trees: popUlation and indivdual response to
fertilization. Oikos 75:148-159.
Lacerda, L.D., 1997. Trace metals in mangrove plants: why such low concentrations? In: Kjerfve, 8.,
L.D. Lacerda and E.H.S. Diop (Eds.), Mangrove ecosystem studies in Latin America and
Africa. UNESCO, Paris, pp. 171-178.
Lacerda, L.D., 1998. Biogeochemistry of trace metals and diffuse pollution in mangrove ecosystems.
ISME Mangrove Ecosystems Occasional Papers 2:1-65.
Lacerda, L.D., V. Ittekkot and S.R. Patchineelam, 1995. Biogeochemistry of mangrove soil organic
matter: a comparison between Rhizophora and Avicennia soils in south-eastern Brazil. Est.
Coastal Shelf Sci. 40:713-720.
Lacerda, L.D., C.E. de Rezende, D.V. Jose and M.C.F. Francisco. 1986a. Metallic composition of
mangrove leaves from the southeastern Brazilian coast. Rev. Bras. BioI. 46:395-399.
Lacerda, L.D., C.E.V. Carvalho, K.F. Tanizaki, A.R.C. Ovalle and C.E. Rezende, 1993a. The
biogeochemistry and trace metal distribution of mangrove rhizospheres. Biotropica 25:252-
257.
Lacerda, L.D., L.F.F. Silva, R.V. Marins, S. Mounier, H.H.M. Paraquetti and 1. Benaim, 2001.
Dissolved mercury concentrations and reactivity in mangrove waters from the ltacurussa
Experimental Forest. Sepetiba Bay. SE Brazil. Wetlands Ecology and Management 9:323-331.
Lacerda, L.D., J.E. Conde. C. Alarcon, R. Alvarez-Le6n, P.R. Bacon, L. D'Croz. B. Kjerfve, 1. Polania
and M. Vannucci, I993b. Mangrove ecosystems of Latin America and the Caribbean: a
summary. In: Lacerda, L.D. (Ed.), Conservation and sustainable utilization of mangrove
forest~ in Latin America and Africa regions. Part I - Latin America. ITTO/ISME, Okinawa,
pp. 1-42.
Lacerda, L.D., C.E. de Rezende, D.V. Jose, J.C. Wasserman and M.C.F. Francisco, 1985. Mineral
concentrations in leaves of mangrove trees. Biotropica 17:260-262.
326 Mangrove Ecology, Silviculture and Conservation

Lacerda. L.D .• D.V. Jose. C.E. de Rezende. M.C.F. Francisco. lC. Wasserman and J.C. Martins.
1986b. Leaf chemical characteristics affecting herbivory in a new world mangrove forest.
Biotropica 18:350-355.
Laegdsgaard. P. and c.R. Johnson 1995. Mangrove habitats as nurseries: unique assemblages of
juvenile fish in subtropical mangroves in eastern Australia. Mar. Ecol. Progr. Ser. 126:67-81.
Lahiri. A.K .• 1987. Silvopisciculture as land management system for rural development in mangrove
area. Indian Agric. 31 :305-311.
Lahiri. A.K.. 1991. Aerial seeding in mangrove swamps. Indian Forester 117: 159-228.
Lahmann. EJ .• 1988. Effects of different hydrological regimes on the productivity of Rhizophora
mangle L. A case study of mosquito control impoundments at Hutchinson Island. Saint Lucie
County. Florida. Unpublished Thesis. University of Miami. Miami.
Lai. C.H .• H.J. Teas. F. Pannier and 1.M. Baker. 1993. Biological impacts of oil pollution: mangroves.
International Petroleum Industry Environmental Conservation Association Report No.4,
London.
Lakshmi. M.• S. Rajalashmi. M. Parani. C.S. Anuratha and A. Parida. 1997. Molecular phylogeny of
mangroves. I. Use of molecular markers in assessing the intraspecific genetic variability of the
mangrove species Acanthus ilicifolius Linn. (Acanthaceae). Theor. Appl. Genet. 94:1121-
1127.
Lakshmi. M .• M. Parani. R. Nivedita and A.K. Parida. 2000. Molecular phylogeny of mangroves. VI.
Intra-specific genetic variation in mangrove species Excoecaria agallocha L.
(Euphorbiaceae). Genome 43:110-115.
Lal. P.N .• 1990. Conservation or conversion of mangroves in Fiji. Occasional Paper of the East-West
Environment and Policy Institute. paper No. II.
Lamparelli. C.C .• P.O. Rodrigues and D.O. de Moura. 1997. Long-term assessment of an oil spiU in a
mangrove forest in Sao Paulo. Brazil. In: Kjerfve. B .• L.D. Lacerda and E.H.S. Diop (Eds.).
Mangrove ecosystem studies in Latin America and Africa. UNESCO. Paris. pp. 191-203.
Lang. J. and G. Paradis. 1977. Un example d'environnement sedimentaire. biodetritique. non carbonate
marin et continental. holocene en c1imat intertropical: Ie domaine margino-littoral du Benin
(ex-Dahomey). Rev. Geogr. Phys. Geol. Dynam. 19:295-312.
Lang. J. and G. Paradis. 1984. Le Quaternaire margino-Iittoral beninois (Afrique de rOues!). Synthese
des datations au carbone 14. In: Coetzee. J.A. and E.M. van Zinderen Bakker (Eds.).
Palaeoecology of Africa and the surrounding islands. Balkema. Rotterdam. pp. 65-76.
Lange. W.P. de and P.J. de Lange. 1994. An appraisal of factors controlling the latitudinal distribution
of mangrove (Avicennia marina var. resinifera) in New Zealand. J. Coastal Res. 10:539-548.
LaRue. C.D. and TJ. Muzik. 1951. Does the mangrove really plant its seedlings? Science 114:661-662.
Latif. M.A.. 1996. An evaluation of mangrove restoration projects in eastern Australia. Unpublished
Thesis. Southern Cross University, Lismore.
Lawson. G.W .• 1966. The littoral ecology of West Africa. Oceanogr. Mar. BioI. Ann. Rev. 4:405-448.
Lebigre. J-M .• 1983. Les mangroves des rias du littoral gabonais. Essai de cartographic typologique.
Rev. Bois et Focets des Tropiques 199:3-28.
Lee. S. Y .• 1989a. The importance of Sesarminae crabs Chiromanthes spp. and inundation frequency on
mangrove (Kandelia candel (L.) Druce) leaf litter turnover in a Hong Kong tidal shrimp pond.
J. Exp. Mar. Bioi Ecol. 131:23-43.
Lee. S.Y.• 1989b. Litter production and turnover of the mangrove Kandelia candel (L.) Druce in a
Hong Kong tidal shrimp pond. Est. Coastal Shelf Sci. 29:75-87.
Lee. S.Y .• 1990. Primary productivity and particulate organic matter flow in an estuarine mangrove-
wetland in Hong Kong. Mar. BioI. 106:453-463.
Lee. S.Y .• 1991. Herbivory as an ecological process in a Kandelia candel (Rhizophoraceae) mangal in
Hong Kong. 1. Trop. Ecol. 7:337-348.
Lee. S.Y., 1995. Mangrove outwelling: a review. Hydrobiologia 295:203-212.
Lee. S.Y .• 1998. The ecological role of grapsid crabs in mangrove ecosystems: implications for
conservation. Mar. Freshw. Res. 49:335-343.
Lee, S.Y .• 1999. Tropical mangrove ecology: physical and biotic factors influencing ecosystem
structure and function. Aust. J. Ecol. 24:355-366.
Lefebvre. G. and B. Poulin. 1997. Bird communities in Panamanian black mangroves: potential effects
of physical and biotic factors. J. Trop. Ecol. 13:97-113.
Leichhardt. L.. 1847. Journal of an overland expedition in Australia from Moreton Bay to Port
Essington a distance of upwards of 3000 miles. during the years 1844-1845. T. & W. Boone.
London.
LePage. B.A .• R.S. Currah. R.A. Stockey and G.W. Rothwell. 1997. Fossil ectomycorrhizae from the
middle Eocene. Amer. J. Bot. 84:410-412.
Lersten. N.R .• and H.T. Homer. 1976. Bacterial leaf nodule symbiosis in angiosperms with emphasis on
Rubiaceae and Myrsinaceae. Bot. Rev. 42:145-214.
Leshem. Y. and E. Levison. 1972. Regulation mechanisms in the salt mangrove Avicennia marina
growing on the Sinai littoral. Oecol. Plant. 7:167-176.
Lewis. D. and A.N. Rao. 1971. Evolution of dimorphism and population dimorphism in Pemphis acidula
Forst. & Forst. F. Proc. Roy. Soc. London 178:79-94.
References 327

Lewis, O.A.M. and G. Naidoo, 1970. Tidal influence on the apparent transpirational rhythms of the
white mangrove. S. Afr. J. Sci. 66: 268-270.
Lewis, R. R., 1982. Mangrove forests. In: Lewis, R. R. (Ed.), Creation and restoration of coastal plant
communities. CRC Press: Florida, pp. 153-171.
Lewis, R.R. and K.C. Haines, 1980. Large scale mangrove restoration on St. Croix, US Virgin Islands.
In: Cole, D.P. (Ed.), Proceedings seventh annual conference on the restoration and creation of
wetlands. Hillsborough Community College, Tampa, pp. 137-148.
Ley, J.A., C.C. McIvor and C.L. Montague, 1999. Fishes in mangrove prop-root habitats of northeastern
Florida Bay: distinct assemblages across an estuarine gradient. Est. Coa.~taI Shelf Sci.48:70 1-
723.
Lezine, A.M., 1988. New pollen data from the Sahel, Senegal. In: Ward, J.V. and K.-B. Lui (Eds.),
Quaternary palynology of tropical areas. Rev. Palaebot. Palynol. 55:141-154.
Lezine, A.M., 1996. La mangrove ouest africaine, signal des variations su niveau marin et des
conditions regionales du c1imat au cours de la derniere deglaciation. Bull. Soc. Geol. France
167:743-752.
Li, M.S. and S.Y. Lee, 1997. Mangroves of China: a briefreview. For. Ecol. Manage. 96:241-259.
Li, Y., D.Z. Zheng, B.W. Liao, S.F. Zheng and X.G. Song, 1999. Effect of salinity and temperature on
seed germination of mangrove Sonneratia opetala Buch. Ham. Forest Research 10: 137 -142.
Li, Y., D.Z. Zheng, B.W. Liao, S.F. Zheng, Y.J. Wang and Z.T. Chen, 1996. Preliminary report on
introduction of several superior mangroves. Forest Research 6:652-655.
Liao, B.W., D.Z. Zheng, S.F. Zheng, Y. Li, X.G. Chen and Z.T. Chen, 1998. A study on the
afforestation techniques of Kandelia candel mangrove. Forest Research 9:586-592.
Lin, C., G. Lancaster, L.A. Sullivan, D. McConchie, P. Saenger and M.D. Melville, 2000. Actual acidity
and its assessment in acid sulfate soils. In: Siavich, P.G. (Ed.), Proceedings of workshop on
remediation and assessment of broadacre acid sulfate soils. NSW Acid Sulfate Soil
Management Advisory Committee, Wollongbar, pp. 22-28.
Lin, G. and L. da S.L. Sternberg, 1992a. Comparative study of water uptake and photosynthetic gas
exchange between scrub and fringe red mangroves, Rhizophora mangle L. Oecologia 90:399-
403.
Lin, G. and L. da SL. Sternberg, 1992b. Effect of growth form, salinity, nutrients and sulfide on
photosynthesis, carbon isotope discrimination and growth of red mangrove (Rhizophora mangle
L). Aust. J. PI. Physiol. 19:509-517.
Lin, G. and L. da S.L. Sternberg, 1993. Effects of salinity fluctuation on photosynthetic gas exchange
and plant growth ofthe red mangrove (Rhizophora mangle L.). J. Exptl. Bot. 44:9-16.
Lin, G. and L. da SL. Sternberg, 1994. Utilization of surface water by red mangrove (Rhizophora
mangle L.): an isotopic study. Bull. Mar. Sci. 54:94-102.
Lin, G. and L. da S.L. Sternberg, 1995. Variation in propagule mass and its effect on carbon
assimilation and seedling growth of red mangrove (Rhizophora mangle) in Florida, U.S.A. J.
Trop. Ecol. 11:109-119.
Lin, P. and W. Xin-Men, 1983. Ecological notes on the mangroves of Fujian, China. In: Teas, H.J.
(Ed.), Ta.~ks for vegetation scinence, Vol. 8. Dr W. Junk, Amsterdam, pp. 31-36.
Lindall, W.N. and L. Trent, 1975. Housing development canals in the coastal zone of the Gulf of
Mexico: ecological consequences, regulations and recommendations. Mar. Fish. Rev. 37:19-
24.
Lockhart, C.S., 1996. Aquatic heterophylly as a survival strategy in Melaleuca quinquenervia
(Myrtaceae). Can. J. Bot. 74:243-246.
Loder, J.W. and G.B. Russell, 1969. Tumor inhibitory plants. The alkaloids of Bruguiera sexangula and
Bruguiera exaristata (Rhizophoraceae). Aust. J. Chem. 22:1271-1275.
Loehle, C., 1996. Optimal defensive investment~ in plants. Oikos 75:299-302.
Lokkers, C., 1986. The distribution of the weaver ant, Oecophylla smaragdina (Fabricus)
(Hymenoptera: Formicidae) in northern Australia. Aust. J. Zoo!. 34:683-687.
Loneragan, N.R., S.E. Bunn and D.M. Kellaway, 1997. Are mangroves and seagrasses sources of
organic carbon for penaeid prawns in a tropical Australian estuary? A multiple stable-isotope
study. Mar. BioI. 130:289-300.
Long, B.G. and T.D. Skewes, 1995. A technique for mapping mangroves with Landsat TM satellite data
and geographic information system. Est. Coa.~taI Shelf Sci. 43:373-381.
Longhurst, A.R., 1962. A review of the oceanography of the Gulf of Guinea. Bull. de I'I.F.A.N. 24:633-
663.
L6pez-Portillo, J. and E. Ezcurra, 1989a. Zonation in mangrove and salt marsh vegetation at Laguna de
Mecoacan, Mexico. Biotropica 21:107-114.
L6pez-Portillo, J. and E. Ezcurra, I 989b. Response of three mangroves to salinity in two geoforms.
Functional Ecology 3:355-361.
Lot, H.A., C. Vazquez-Yanes and F. Menendez, 1975. Physiognomic and floristic changes near the
northern limit of mangroves in the gulf coast of Mexico. In: Walsh, G.E., S.c. Snedaker and
H.J. Teas (Eds.), Proceedings of the international symposium on biology and management of
mangroves. Vol. I. University of Florida, Gainesville, pp. 52-61.
328 Mangrove Ecology, Silviculture and Conservation

LOtschert, W. and F. Liemann, 1%7. Die Salzspeicherung im Keimling von Rhizoplwra mangle L.
wlihrend der Entwicklung auf der Mutterpflanze. Planta 72:142-156.
Lovelock, C.E. and B.F. Clough, 1992. Influence of solar radiation and leaf angle on leaf xanthophyll
concentrations in mangroves. Oecologia 91:518-525.
Lovelock, C.E. and I.C. Feller, in press. Photosynthetic performance and resource utilization of two
mangrove species coexisting in a hypersaline scrub forest. Wetlands Ecology and
Management.
Lovelock, C.E., B.F. Clough and I.E. Woodrow, 1992. Distribution and accumulation of ultraviolet-
radiation-absorbing compounds in leaves of tropical mangroves. Planta 188: 143-154.
Lowenfield, R. and E.J. Klekowski, 1992. Mangrove genetics. I. Mating system and mutation rates of
Rhizoplwra mangle in Florida and San Salvador Island, Bahamas. Int. J. PI. Sci. 153:394-399.
Lu, c.Y., Y.S. Wong, N.F.Y. Tam, Y. Ye and P. Lin, 1999. Methane flux and production from
sediments of a mangrove wetland on Hainan Island, China. Mangroves and Salt Marshes 3:41-
49.
Ludbrook, N.H., 1963. Correlations of the tertiary rocks of South Australia. Trans. Roy. Soc. Sth Aust.
87:5-15.
Lugo, A.E., 1981. The inland mangroves of Inagua. J. Natural Hist. 15:845-852.
Lugo, A.E., 1985. Nutrient cycling in mangroves. Estuaries 8: 106.
Lugo, AE., 1986. Mangrove understory: an expensive luxury. J. Trop. Ecol. 2:287-288.
Lugo, AE., and M.M. Brinson. 1978. Calculations of the value of saltwater wetlands. In: Greeson, P.E.,
J.R. Clark and J.E. Clark (Eds.), Wetland functions and values: the state of our understanding.
American Water Resource Association TPS 79-2, pp. 120-130.
Lugo, AE. and C. Patterson-Zucca, 1977. The impact of low temperature stress on mangrove structure
and growth. Trop. Ecol. 18:149-161.
Lugo, AE. and S.C. Snedaker, 1974. The ecology of mangroves. Ann. Rev. Ecol. Systematics 5:39-64.
Lugo, A.E. and S.c. Snedaker. 1975. Properties of a mangrove forest in southern Florida. In: Walsh,
G.E., S.C. Snedaker and H.J. Teas (Eds.), Proceedings of the international symposium on
biology and management of mangroves. Vol. 1. University of Florida, Gainesville, pp. 170-212.
Lugo, A.E., M. Sell and S.c. Snedaker. 1976. Mangrove ecosystem analysis. In: Patten, B.C. (Ed.),
Systems analysis and simulation in ecology. Academic Press, New York, pp. 113-145.
Lugo, A.E., G. Evink, M.M. Brinson, A. Broce and S.C. Snedaker. 1975. Diurnal rates of
photosynthesis, respiration and transpiration in mangrove forests in South Florida. In: Golley,
F.B. and E. Medina (Eds.), Ecological studies II. Tropical ecological systems. Springer Verlag,
New York, pp. 335-350.
Ly, C.K., 1980. The role of the Akosombo Dam on the Volta River in causing coastal erosion in central
and eastern Ghana (West Africa). Mar. Geo!. 37:323-332.
Lynch, J.c., J.R. Meriwether, B.A McKee, F. Vera-Herrera and R.R. Twilley, 1989. Recent accretion
in mangrove ecosystem based on 137es and 2 IOPb. Estuaries 12:284-299.
Mabberley, OJ., 1995. Meliaceae. In: Dassanayake, M.D., F.R. Fosberg and W.O. Clayton (Eds.), A
revised handbook to the flora of Ceylon. Vol. IX. A.A. Balkema, Rotterdam.
MacCaughey, V., 1917. The mangrove in the Hawaiian Islands. Hawaii Forest. Agric. 14:361-366.
MacFarlane, G.R., 2000. Biological indication of heavy metals in mangrove ecosystems. Unpublished
Thesis, University of Technology. Sydney.
MaqFarlane, G.R. and M.D. Burchett, 1999. Zinc distribution and excretion in the leaves of the grey
mangrove, Avicennia marina (Forsk.) Vierh. Environ. Exptl. Bot. 41: 167-175.
MacFarlane, G.R. and M.D. Burchett, 2000. Cellular distribution of copper, lead and zinc in the grey
mangrove, Avicennia marina (Forsk.) Vierh. Aquat. Bot. 68:45-59.
MacIntosh, S.B., 1987. The potential of Merope angulata (Willd.) Swingle, a wild relative of the citrus
found in the Malaysian mangroves. In: Sa.~ekumar A., S.M. Phang and E.L. Chong (Eds.),
Proceedings of the 10'" annual seminar, 28 March 1987: Towards conserving Malaysia's
marine heritage. Malaysian Society of Marine Sciences, Kuala Lumpur, pp. 63-68.
Mackey, A.P., 1993. Biomass of the mangrove Avicennia marina (Forsk.) Vierh. near Brisbane, south-
eastern Queensland. Aust. J. Mar. Freshw. Res. 44:721-725.
Macnae, W., 1%3. Mangrove swamps in South Africa. J. Ecol. 51:1-25.
Macnae, W., 1966. Mangroves in eastern and southern Australia. Aust. J. Bot. 14:67-104.
Macnae, W., 1967. Zonation within mangroves associated with estuaries in North Queensland. In:
Lauff, G.H. (Ed.). Estuaries. Amer. Assoc. Adv. Sci. Publ. 83, pp. 432-441.
Macnae, W., I %8. A general account of the fauna and flora of mangrove swamps and forests in the
Indo-West Pacific region. Adv. Mar. BioI. 6:73-270.
Maguire, T.L. and P. Saenger, 2000. The taxonomic relationships within the genus Excoecaria L.
(Euphorbiaceae) based on leaf morphology and rONA sequence data. Wetlands Ecology and
Management 8: 19-28.
Maguire, T.L., R. Peakall and P. Saenger, 2001. Comparative analysis of genetic diversity in the
mangrove species Avicennia marina (Forsk.) Vierh. (Avicenniaceae) detected by AFLPs and
SSRs. Theor. Appl. Genet. In press.
References 329

Maguire T.L.. KJ. Edwards. P. Saenger and R. Henry. 2000. Characterisation and analysis of
microsatellite loci in a mangrove species Avicennia marina (Forsk.) Vierh. (Avicenniaceae).
Theor. Appl. Genet. 101:279-285.
Maguire. T.L.. P. Saenger. P. Baverstock and R. Henry. 2000. MicrosatelIite analysis of genetic
structure in the mangrove species Avicennia marina (Forsk.) Vierh. (Avicenniaceae). Mol.
Ecol. 9:1853-1862
Maity. S .• 1994. Biotechnology in conservation of mangroves - a case study. Proc. Natl. Symp. Cons. &
Management of Living Resources. Bangalore. India. 1990. pp. 1-10.
Malaviya. C.V. 1963. On the distribution. structure and ontogeny of stone cells in Avicennia ojJicina/is
L. Proc. Indian Acad. Sci. 58:45-50.
Mall. L.P .• S.K. Billore and D. Amritphale. 1982. Certain ecological observations on mangroves of the
Andarnan Islands. Trop. Ecol. 23:225-233.
Mandura. A.S .• 1997. A mangrove stand under sewage pollution stress: Red Sea. Mangroves and Salt
Marshes 1:255-262.
Mann. F.D. and T.D. Steinke. 1989. Biological nitrogen fixation (acetylene reduction) associated with
blue-green algal (cyanobacterial) communities in the Beachwood Mangrove Nature Reserve.
I. The effect of environmental factors on acetylene reduction activity. S. Afr. 1. Bot. 55:438-
446.
Manson. FJ .• N.R. Loneragan. I.M. McLeod and R.A. Kenyon. 2001. Assessing techniques for
estimating the extent of mangroves: topographic maps. aerial photographs and Landsat TM
images. Mar. Freshw. Res. 52:787-792.
Marco. H.F.• 1935. Systematic anatomy of the woods of the Rhizophoraceae. Trop. Woods (Yale
University) 44: 1-26.
Markley. J.L .• C. McMillan and GA Thompson. 1982. Latitudinal differentiation in response to chilling
temperatures among populations of three mangroves. Avicennia germinans. Laguncularia
racemosa and Rhizophora mangle. from the western tropical Atlantic and Pacific Panama.
Can. J. Bot. 60:2704-2715.
Marshall. A.G. amd Lord Medway. 1976. A mangrove community in the New Hebrides. south-west
Pacific. BioI. J. Linn. Soc. 8:319-336.
Martin. H.A .• 1982. Changing Cenozoic barriers and the Australian palaeobotanical record. Ann.
Missouri Bot. Gardens 69:625-667.
Mastaller. M .• 1997. Mangroves - the forgotten forest between land and sea. Tropical Press Sdn. Bhd.•
Kuala Lumpur.
Matsui. N .• 1998. Estimated stocks of organic carbon in mangrove roots and sediments in Hinchinbrook
Channel. Australia. Mangroves and Salt Marshes 2:199-204.
Matthes. H. and J.M. Kapetsky. 1988. Worldwide compendium of mangrove-associated aquatic species
of economic importance. FAO Fish. Circ. 814. pp. 236.
Matthijs. S .• J. Tack. D. von Spreybroeck and N. Koedam. 1999. Mangrove species zonation and soil
redox state. sulphide concentration and salinity in Gazi Bay (Kenya). a preliminary study.
Mangroves and Salt Marshes 3:243-249.
Maul. G. and D.M. Martin. 1993. Sea level rise at Key West. Florida. 1846-1992: America's longest
instrument record? Geophys. Res. Lett. 20:1955-1958.
Maxwell. G.S., 1995. Ecogeographic variation in Kant/elia candel from Brunei, Hong Kong and
Thailand. Hydrobiologia 295:59-65.
Mazda. Y .• M. Magi. M. Kogo and P.N. Hong. I 997a. Mangroves as a coastal protection from waves in
the Tong King delta. Vietnam. Mangroves and Salt Marshes 1:127-\35.
Mazda. Y .• E. Wolanski. B. King. A. Sasa. D. Ohtsuka and M. Magi. I 997b. Drag force due to
vegetation in mangrove swamps. Mangroves and Salt Marshes 1:193-199.
McArdle. B.H.• 1988. The structural relationship: regression in biology. Can. J. Zool. 66:2329-2339.
McConchie. D.M .• 1990. Report on land stability problems affecting coastal plantations. F.A.O.
Technical Paper No. 25. FAO.• Rome.
McConchie. D.M. and P. Saenger. 1991. Mangrove forests as an alternative to civil engineering works
in coastal environments of Bangladesh: lessons for Australia. In: Arakel. A.V. (Ed.).
Proceedings of 1990 Workshop on Coastal Zone Management. Yeppoon. Queensland.
Queensland University of Technology, Brisbane. pp. 220-233.
McCoy. E.D. and K.L. Heck. 1976. Biogeography of corals. seagrasses. and mangroves: an alternative
to the center of origin concept. Syst. Zool. 25:201-210.
McCoy. E.D .• H.R. Mushinsky. D. Johnson and W.E. Meshaka. 1996. Mangrove damage caused by
Hurricane Andrew on the southwestern coast of Florida. Bull. Mar. Sci. 59: 1-8.
McDade. L.A. and M.L. Moody. 1999. Phylogenetic relationships among Acanthaceae: evidence from
non-encoding TRNL-TRNF chloroplast DNA sequences. Amer. J. Bot. 86:70-80.
McGuinness. K.A .• 1997a. Dispersal, establishment and survival of Ceriaps tagal propagules in a north
Australian mangrove forest. Oecologia 109:80-87.
McGuiness. K.A .• I 997b. Seed predation in a tropical mangrove forest: a test of the dominance-
predation model in northern Australia. J. Trop. Ecol. 13:293-302.
Mcivor. C.C.• JA Ley and R.D. Bjork. 1994. Changes in freshwater inflow from the Everglades to
Florida Bay including effects on biota and biotic processes: a review. In: Ogden. 1. and S.
330 Mangrove Ecology, Silviculture and Conservation

Davies (Eds.). Everglades: the ecosystem and its restoration. St Lucie Press. Delray Beach.
pp. 117-146.
McKee. K.L .• 1993. Soil physicochemical patterns and mangrove species distribution - reciprocal
effects? 1. Ecol. 81:477-487.
McKee. K.L .• I 995a. Mangrove species distribution patterns and propagule predation in Belize: an
exception to the dominance-predation hypothesis. Biotropica 27:334-345.
McKee. K.L.. I 995b. Seedling recruitment patterns in a Belizean mangrove forest: effects of
establishment ability and physico-chemical factors. Oecologia 101 :448-460.
McKee. K.L.. I 995c. Interspecific variation in growth. biomass partitioning. and defensive
characteristics of neotropical mangrove seedlings: response to light and nutrient availability.
Amer.1. Bot. 82:299-307.
McKee. K.L .• 1996. Growth and physiological responses of neotropical mangrove seedlings to root
zone hypoxia. Tree Physiol. 16:883-889.
McKee. K.L. and P.L. Faulkner. 2000. Restoration of biogeochemical function in mangrove forests.
Restoration Ecol. 8:247-259.
McKee. K.L. and I.A. Mendelssohn. 1987. Root metabolism in the black mangrove (Avicennia
germinans (L.) L): response to hypoxia. Environ. Exptl. Bot. 27:147-156
McKee. K.L.. I.A. Mendelssohn and M.W. Hester. 1988. Re-examination of pore water sulfide
concentrations and redox potentials near the aerial root~ of Rhizophora mangle and A vicennia
germinans. Amer. 1. Bot. 75:1352-1359.
McMillan. C .• 1971. Environmental factors affecting seedling establishment of the black mangrove on
the central Texas coast. Ecology 52: 927-930.
McMillan. c.. 1975a. Adaptive differentiation to chilling in mangrove populations. In: Walsh. G.E .• S.c.
Snedaker and H.1. Teas (Eds.). Proceedings of the International Symposium on Biology and
Management of Mangroves. Vol. I. University of Florida. Gainesville. pp. 62-70.
McMillan. C .• 1975b. Interactions of soil texture with salinity tolerances of A vicennia germinans (L.)
Lam. and Laguncularia racemasa (L.) Gaertn f. from North America. In: Walsh. G.E .• S.c.
Snedaker and H.1. Teas (Eds.). Proceedings of the International Symposium on Biology and
Management of Mangroves. Vol. 2. University of Florida. Gainesville. pp. 561-568.
McMillan. c.. 1986. Isozyme patterns among populations of black mangrove. Avicennia germinans.
from the Gulf of Mexico-Caribbean and Pacific Panama. Contr. Mar. Sci. 29: 17-25.
Medina. E., A.E. Lugo and A. Novelo. 1995. Contenido mineral del tejido foliar de especies de manglar
de la Laguna de Sontecomapan (Veracruz, Mexico) y su relaci6n con la salinidad. Biotropica
27:317-323.
Medina, E .• E. Cuevas, M. Popp and A.E. Lugo. 1990. Soil salinity. sun exposure and growth of
Acrostichum aureum, the mangrove fern. Bot. Gaz. 151:41-49.
Mepham, R.H., 1983. Mangrove floras of the southern continents. Part 1. The geographical origin of
Indo-Pacific mangrove genera and the development and present status of the Australian
mangroves. S. Afr. 1. Bot. 2: 1-8.
Mepham, R.H. and 1.S. Mepham, 1985. The flora of tidal forests - a rationalisation of the use of the
·mangrove·. S. Aft. Tydkr. Plantk. 51:77-99.
Merrill, E.D .• 1945. Plant life in the Pacific world. London: Macmillan.
Michel, P. and P. Assemien. 1970. Etudes sedimentologiques et palynologiques des sondages de Bogue
(Basse vallee du Senegal) et leur interpretation morphoclimatique. Rev. Geomorph. dyn.
19:98-113.
Micheli, F., 1993a. Effect of mangrove litter species and availability on survival. mOUlting. and
reproduction of the mangrove crab Sesamla messa. 1. Exp. Mar. BioI. Ecol. 171:149-163.
Micheli, F.• 1993b. Feeding ecology of mangrove crabs in North Eastern Australia: mangrove litter
consumption by Sesarma messa and Sesarma smithii. J. Exp. Mar. BioI. Ecol. 171: 164-186.
Micheli. F .• F. Gherardhi and M. Vannini, 1991. Feeding and burrowing ecology of two East African
mangrove crabs. Mar. BioI. 111:247-254.
Mildenhall. D.C.. 1994. Early to mid Holocene pollen samples containing mangrove pollen from Sponge
Bay, East Coast. North Island. New Zealand. 1. Roy. Soc. N. Z. 24:219-230.
Mildenhall, D.C. and LJ. Brown 1987. An early holocene occurrence of the mangrove Avicennia
nlarina in Poverty Bay. North Island, New Zealand: its climatic and geological implications.
N.Z. J. Bot. 25:281-294.
Miller. I.M .• I.C. Gardner and A. Scott. 1983. The development of marginal leaf nodules in Ardisa
crispa (Thunb.) A.DC. (Myrsinaceae). Bot. 1. Linn. Soc. 86:237-252.
Miller. P.C .• 1972. Bioclimate. leaf temperature and primary production in red mangrove canopies in
south Florida. Ecology 53:22-45.
Miller, P.c., 1975. Simulation of water relations and net photosynthesis in mangroves in southeru
Florida. In: Walsh. G.E .• S.c. Snedaker and HJ. Teas (Eds.), Proceedings of the International
Symposium on Biology and Management of Mangroves. Vol. 2. University of Florida.
Gainesville, pp. 615-631.
Miller, P.C .• 1. Hom and D.K. Poole. 1975. Water relations of three mangrove species in south Florida.
Oecol. Plant. 10:355-367.
References 331

Minocha, P.K. and K.P. Tiwari, 1981. A triterpenoidal saponin from roots of Acanthus iUicifolius.
Phytochem.2O:135-137.
Miyawaki, A., K. Suzuki, S. Suzuki, Y. Nakamura. Y. Murakami, Y. Tsukagoshi and E. Nakata, 1983.
Pflanzensoziologische Untersuchung der Mangrove Vegetation in Japan. 1. Mangrove
Vegetation auf der lriomote-Insel. Bull. Inst. Env. Sci. Technol. Yokohama Natl. University
9:77-89.
Mizrachi, D., R. Pannier and F. Pannier, 1980. Assessment of salt resistance mechanisms as
determinant physio-ecological parameters of zonal distribution of mangrove species. I. Effect
of salinity stress on nitrogen metabolism balance and protein synthesis in the mangrove species
Rhizophora mangle and Avicennio nitida. Bot. Mar. 23:289-296.
Moezel, van der, P.G., G.V.N. Pearce-Pinto and D.T. Bell, 1991. Screening for salt and waterlogging
tolerance in Eucalyptus and Melaleuca species. For. Beol. Manage. 40:27-37.
Montgomery, J. and M.T. Price, 1979. Release of trace metals by sewage sludge and the subsequent
uptake by members of a turtle grass mangrove ecosystem. Environ. Sci. Technol. 13:546-549.
Moon, GJ., B.F. Oough, C.A. Peterson and W.G. Allaway, 1986. Apoplastic and symplastic pathways
in Avicennio marina (Forssk.) Vierh. roots revealed by fluorescent tracer dyes. Aust. J. PI.
Physiol. 13:637-648.
Moore, H.E., 1973. Palms in the tropical forest ecosystems of Africa and South America. In: Meggers,
B.1., E.S. Ayensu and W.O. Duckworth (Eds.), Tropical forest ecosystems in Africa and South
America: a comparative review. Smithsonian Institution Press, Washington, pp. 63-88.
Moore, R.T., P.C. Miller, D. Albright and L.L. Tieszen, 1972. Comparative gas exchange
characteristics of three mangrove species during the winter. Photosynthetica 6:387-393.
Moore, R.T., P.C. Miller, J. Ehleringer and W. Lawrence, 1973. Seasonal trends in gas exchange
characteristics of three mangrove species. Photosynthetica 7:387-394.
Moriya, H., A. Komiyama, S. Prawiroatmodjo and K. Ogino, 1988. Forest as an ecosystem, its structure
and function. 4. Specific characteristics of leaf dynamics. In: Ogino, K. and M. Chihara (Eds.),
Biological system of mangroves - a report of east Indonesian mangrove expedition, 1986.
Ehime University, Ehime, pp. 123-136.
Morrisey, DJ., G.A. Skilleter, 1.1. Ellis, B.R. Bums, C.E. Kemp and K. Burt, in press. Differences in
fauna and sediment among mangrove stands of different ages. Est. Coastal Shelf Sci.
Morton, J.F., 1965. Can the red mangrove provide food, feed and fertilizer? Beon. Bot. 19:113-123.
Morton, R.M., 1989. Hydrology and fish fauna of canal developments in an intensively modified
Australian estuary. Est. Coastal Shelf Sci. 28:43-58. .
Morton, R.M., 1992. Fish assemblages in residential canal developments near the mouth of a subtropical
Queensland estuary. Aust. J. Mar. Freshw. Res. 43:1359-1371.
Mueller-Dombois, D. and F.R. Fosberg, 1998. Vegetation of the tropical Pacific islands. Springer
Verlag, New York.
Mulholland, D.A. and D.A.H. Taylor, 1992. Limonoids from Australian members of the Meliaceae.
Phytochem. 31 :4163-4166.
Mulik, N.G. and LJ. Bhosale, 1989. Flowering phenology of the mangroves from the west coast of
Maharashtra. J. Bombay Nat. Hist. Soc. 86:355-359.
Mullan, D.P., 1931a. Observations on the water storing devices in the leaves of some Indian halophytes.
J.lndian Bot. Soc. 10:126-133.
Mullan, D.P., 1931b. On the occurrence of glandular hairs (salt glands) on the leaves of some Indian
halophytes. J. Indian Bot. Soc. 10:184-189.
Muller, J., 1964. A palynological contribution to the history of the mangrove vegetation in Borneo. In:
Cranwell, L.M. (Ed.), Ancient Pacific Floras. University of Hawaii Press, Honolulu, pp. 33-42.
Muller, J., 1969. A palynological study of the genus SonneraJia (Sonneratiaceae). Pollen et Spores
11 :223-298.
Muller, J. and C. Caratini, 1977. Pollen of Rhizophora (Rhizophoraceae) as a guide fossil. Pollen et
Spores 19:361-390.
Murphy, D.H., 1990. The natural history of insect herbivory on mangrove trees in and near Singapore.
Raffles Bull. Zool. 38:119-204.
Murphy, J.M., 1899. Alligator shooting in Florida. Reprinted in: Oppel, F. and T. Meisel (Eds.), 1987,
Tales of old Florida, Castle, New Jersey, pp. 353-359.
Murray, D.R., 1995. Plant responses to carbon dioxide. Amer. J. Bot. 82:690-697.
Myers, J.G., 1935. Zonation along river courses. J. Ecol. 23:356-360.
Naidoo, G., 1980. Mangrove soils of the Beachwood Area, Durban. J. Sth. Afr. Bot. 46:293-304.
Naidoo, G., 1983. Effects of flooding on leaf water potential and stomatal resistance in Bruguiera
gymnorrhiZ/.l (L.) Lam. New Phytol. 93:369-376
Naidoo, G., 1985. Effects of waterlogging and salinity on plant-water relations and on the accumulation
of solutes in three mangrove species. Aquat. Bot. 22: 133-143.
Naidoo, G., 1989. Seasonal plant water relations in a South African mangrove swamp. Aquat. Bot.
33:87-100.
Naidoo, G. and Dol. von Willert, 1995. Diurnal gas exchange characteristics and water use efficiency
of three salt-secreting mangroves at low and high salinities. Hydrobiologia 295: 13-22.
332 Mangrove Ecology, Silviculture and Conservation

Naidoo. G. and D.J. von Willert, 1999. Gas exchange characteristics of the tropical mangrove.
Pelliciera rhizDphoreae. Mangroves and Salt Marshes 3:147-153.
Naidoo. G. H. Rogalla and D.J. von Willert, 1997. Gas exchange responses of the mangrove. Avicennia
marina to waterlogged and drained conditions. Hydrobiologia 352:39-47.
Naidoo. G.• H. Rogalla and D.J. von Willert. 1998. Field measurements of gas exchange in Avicennia
marina and Bruguiera gymnorrhizo. Mangroves and Salt Marshes 2:99-107.
Naskar. K. and R. Mandai. 1999. Ecology and biodiversity of Indian mangroves. Part I - Global status.
Part 2 - Morpho-anatomy of mangroves. Daya Publishing House. Delhi.
Naurois. R. de and F. Roux. 1965. Les mangroves d'Avicennia les plus septentrionales de la cate
occidentales d'Afrique. Bull. de I'I.F.A.N. 27:842-854.
Nedwell, D.B.• 1975. Inorganic nitrogen metabolism in an eutrophicated tropical estuary Water Res.
9:221-231.
Neilson. M.J. and G.N. Richards. 1989. Chemical composition of degrading mangrove leaf litter and
changes produced after consumption by mangrove crab Neosannatium smithi (Crustacea:
Decapoda: Sesannidae). J. Chem. Ecol. 15:1267-1284. ,
Neilson. MJ .• R.L. Giddins and G.N. Richards. 1986. Effect of tannins on the palatability of mangrove
leaves to the tropical sesarminid crab Neosarmatium smithi. Mar. Ecol. Prog. Ser. 34: 185-186.
Nickerson. OJ.• 1999. Trade-offs of mangrove area development in the Philippines. Ecol. Econ.
28:279-298.
Nickerson. N.H. and F.R. Thibodeau. 1985. Association between pore water sulfide concentrations and
the distribution of mangroves. Biogeochem. I: 183-192.
Nicole, M.• W.M. Egnankou and M. Schmidt, 1994. A preliminary inventory of coastal wetlands of
C6te d·Ivoire. IUCN. Gland.
Nicou. R.• 1956. Presence du Laguncularia racemosa dans la mangrove du pseudo-delta du Senegal.
Notes Africaines 71:67-68.
Nielsen. M.G .• 1997. Nesting biology of the mangrove. mud-nesting ant Polyrhachis sokolova Forel
(Hymenoptera: Fonnicidae) in northern Australia. Insects Sociaux 44:15-21.
Nishihira, M. and M. Urasaki. 1976. Production. settlement and mortality of seedlings of a mangrove.
Ktmdelia candel (L.) Druce in Okinawa. Abstracts of symposia and contributed papers.
International Symposium on the Ecology and Management of Some Tropical Shallow Water
Communities. Jakarta. July 1976.
Noakes. D.S.P.. 1952. A working plan for the Matang Mangrove Forest Reserve. Perak. Forest
Department. Perak, Federation of Malaya.
Noakes. D.S.P., 1955. Methods of increasing growth and obtaining natural regeneration of the
mangrove type in Malaya. Malay. For. 8:23-30.
Noske. RA .• 1993. Bruguiera haines;;: another bird-pollinated mangrove? Biotropica 25:481-483.
Nurkin. B.• 1994. Degradation of mangrove in South Sulawesi. Indonesia. Hydrobiologia 285:271-276.
Nye. L.B .• 1990. Trace metal accumulation under differing sediment conditions in the mangrove,
Rhiwphora mangle L.• in Key Largo. Florida. UnpUblished Thesis. University of Miami.
Miami.
Odum. E.P.• 1973. A description and value assessment of South Atlantic and Gulf Coast marshes and
estuaries. Bur. Fish. and Wildl. Publ.. Atlanta. Georp.
Odum. W.E.• 1970. Insidious alteration of the estuarine enV1fOnment. Trans. Amer. Fish. Soc. 99:836-
847.
Odum. W.E .• 1971. Pathways of energy flow in South Florida estuary. Sea Grant Tech. Bull. Miami
Univ.7:1-162.
Odum. W.E.• 1982. Environmental degradation and the tyranny of small decisions. BioScience 32:728-
729.
Odum. W.E. and E.J. Heald. 1972. Trophic analysis of an estuarine mangrove community. Bull. Mar.
Sci. 22:671-738.
Odum. W.E. and E.J. Heald. 1975a. Mangrove forests and aquatic productivity. In: Hasler. A.D. (Ed.).
Coupling of land and water systems. Ecological Studies 10. Springer Verlag. New York. pp.
129-136.
Odum. W.E. and E.J. Heald. 1975b. The detritus-ba.IIed food web of an estuarine mangrove community.
In: Cronin. L.E. (Ed.). Estuarine research. Vol. I. Academic Press, New York. pp. 265-286.
Odum. W.E .• C.C. Mcivor and T.S. Smith, 1982. The ecology of the mangroves of South Florida: a
community profile. US Fish and Wildlife Service. Office of Biological Services. Washington.
DC. FWSIOBS - 81124.
Ohnishi. T. and A. Komiyama. 1998. Shoot and root formation on cut pieces of viviparous seedlings of
a mangrove. Kandelia ctmdel (L.) druce. For. Ecol. Manage. 102:173-178.
Oliver. J. 1982. The geographic and environmental aspects of mangrove communities: climate. In:
Clough. B.F. (Ed.). Mangrove ecosystems in Australia - structure. function and management
ANU Press. Canberra, pp. 19-30.
Ollivier-Pierre. M.-F.• C. Gruas-Cavagnetto. E. Roche and M. Schuler. 1987. Elements de la flore de
type tropical et variations climatiques au paleag~ne dans quelques ba.c;sins d'Europe nord-
occidentale. McSm. Trav. E.P.H.E .• lnst. Montpellier 17:173-205.
References 333

Olmsted, I. And M.G. Juarez, 19%. Distribution and conservation of epiphytes on the Yucatan
Peninsula. Selbyana 17:58-70.
O'Neill, J., 1997. Managing without prices: the montary valuation of biodiversity. Ambio 26:546-550.
Ong, J.E., 1982. Mangroves and aquaculture in Malaysia. Ambio 11:252-257.
Ong, J.E., 1993. Mangroves - a carbon source and sink. Chemosphere 27:1097-1107.
Ong, J.E., 1995. The ecology of mangrove conservation & management. Hydrobiologia 295:343-351.
Ong. J.E., W.K. Gong and C.H. Wong 1981, Ecological monitoring of the Sungai Merbok estuarine
mangrove ecosystem. Universiti Sains Malaysia, Penang.
Ong, J.E., W.K. Gong and C.H. Wong, 1985. Seven years of productivity studies in a Malaysian
managed mangrove forest then what? In: Beardsley, K.N., J.D.S. Davie and C.D. Woodroffe
(Eds.), Coasts and Tidal Wetlands of the Australian Monsoon Region. NARU Mangrove
Monographs 1:213-223.
Ong, J.E., W.K. Gong, C.H. Wong and G. Dhanarajan, 1981. Productivity of a managed mangrove
forest in West Malaysia. In: Noor, Y.M. (Ed.), Trends in applied biology in southeast Asia.
University Sains Press, Penang, pp. 274-284.
Onuf, C.P., J.M. Teal and I. Valiela, 1977. Interactions of nutrients, plant growth and herbivory in a
mangrove ecosystem. Ecology 58:514-526.
Ormond, R.F.G., A.R.G. Price and A.R. Dawson-Shepherd, 1988. The distribution and character of
mangroves in the Red Sea, Arabian Gulf and Southern Arabia. Proceedings of the
UNDPIUNESCO Regional Mangrove Project, Colombo, 11-14 Nov. 1986, pp. 125-130.
Orozco, A., F. Rada, A. Azocar and G. Goldstein, 1990. How does a mistletoe affect the water,
nitrogen and carbon balance of two mangrove ecosystem species? Plant, Cell and Environ.
19:941-948.
Osborne, K. and TJ. Smith, 1990. Differential predation on mangrove propagules in open and closed
canopy forest habitats. Vegetatio 89: 1-6.
Oudot, J. and E. Dutrieux, 1989. Hydrocarbon weathering and biodegradation in a tropical estuarine
ecosystem. Mar. Environ. Res. 27:195-213.
Ozaki, K., S. Taka.~hima and O. Suko, 2000. Ant predation suppresses populations of the scale insect
Aulacaspis marina in natural mangrove forests. Biotropica 32:764-768.
Padron, C.M., 1996. Mangrove ecosystem restoration in Cuba: a case study in Havanna Province. In:
Field, C. (Ed.), Restoration of mangrove ecosystems. ISMFJITTO, Okinawa, pp. 160-169.
Paez, H.S., 1994. Los manglares de Colombia. In: Suman, D.O. (Ed.), EI Ecosistema de Manglar en
America Latina y la Cuenca del Caribe: su Manejo y Conservaci6n. The Tinker Foundation,
New York, pp. 21-33.
Paijmans, K. and B. Rollet, 1977. The mangroves of Galley Reach, Papua New Guinea. For. Ecol.
Manage. 1:\19-140.
Pailles, C., D.M. McConchie, A.V. Arakel and P. Saenger, 1993. The distribution of phosphate in
sediments of the Johnstone River catchment-estuary system, North Queensland, Australia.
Sedimentary Geol. 85:253-269.
Palis, H.G., 1996. Abortion rates of selected mangrove species under various reproductive
phenological stages. In: Khemnark, C. (Ed.), Proceedings of Ecotone IV: ecology and
management of mangrove restoration and regeneration in east and southeast Asia. Surat Thani,
Thailand, pp. 284-289.
Pandit, S. and B.C. Choudhury, 2001. Factors affecting pollinator visitation and reproductive success in
Sonneratia caseolaris and Aegiceras comiculatum in a mangrove forest in India. J. Trop. Ecol.
17:431-447.
Pannier, F., 1962. Estudio fisiologico sobre la viviparia de Rhizophora mangle L. Acta Cientifica
Venezolana (Botanical Series) 13:184-197.
Pannier, F., 1979. Mangroves impacted by human-induced disturbances: a case study of the Orinoco
Delta mangrove ecosystem. Environmental Management 3:205-216.
Pannier, F. and R. F. Pannier, 1975. Physiology of vivipary in Rhizophora mangle. In: Walsh, G.E., S.C.
Snedaker and H.1. Teas (Eds.), Proceedings of the International Symposium on Biology and
Management of Mangroves. Vol. 2, University of Florida, Gainesville, pp. 632-642.
Panshin, A.J., 1932. An anatomical study of the woods of the Philippine mangrove swamps. Philippine 1.
Sci. 48:143-208.
Paradis, G., 1979. Observations sur LaguncuLaria racemosa Gaertn. et Dalbergia ecastaphyllum (L.)
Taub. dans la successions secondaires de la mangrove du Benin. Bull. de I'I.F.A.N. 41:92-102.
Paradis, G., 1989. Observations sur la degradation d'origine anthropique des mangroves de Fresco et de
Grand Bassam (Cote d'lvoire). Candollea 44:453-483.
Parani, M., C.S. Rao, N. Mathan, C.S. Anuratha, K.K. Narayanan and A.K. Parida, 1997a. Molecular
phylogeny of mangroves. III. Parentage analysis of a Rhizophora hybrid using random
amplified polymorphic DNA (RAPD) and restriction fragment length polymorphism (RFLP)
markers. Aquat. Bot. 58:165-172.
Parani, M., M. Lakshmi, S. Elango, N. Ram, C.S. Anuratha and A. Parida, I 997b. Molecular phylogeny
of mangroves. II. Intra- and inter-specific variation in Avicennia revealed by RAPD and RFLP
markers. Genome 40:487-495.
334 Mangrove Ecology, Silviculture and Conservation

Parani, M., M. Lakshmi, P. SenthiIkumar, N. Ram and A.K. Parida, 1998. Molecular phylogeny of
mangroves. V. Analysis of genome relationship in mangrove species using RAPD and RFLP
markers. Theor. Appl. Genet. 97:617-625.
Parkes, A., 1997. Environmental change and the impact of Polynesian colonization: sedimentary
records from Central Polynesia. In: Kirch, P.V. and T.L. Hunt (Eds.), Historical ecology in the
Pacific Islands - prehistoric environmental and landscape change. Yale University Press, New
Haven, pp. 166-199.
Parkinson, R.W., R.D. DeLaune and 1R. White, 1994. Holocene sea-level rise and the fate of
mangrove forests within the wider Caribbean region. J. Coastal Res. 10:1077-1086.
Pasqualini, V., J. litis, N. Dessay, M. Lointier, O. Guelorget and L. Polidori, 1999. Mangrove mapping
in north-western Madagascar using SPOT-XS and SlR-C radar data. Hydrobiologia 413:127-
133.
Passioura, J.B., M.C. Ball and J.H. Knight, 1992. Mangroves may salinize the soil and in so doing limit
their transpiration rate. Functional Ecology 6:476-481.
Patterson, S., K.L. McKee and I.A. Mendelssohn, 1997. Effects of tidal inundation and predation on
Avicennia germinans seedling establishment and survival in a sub-tropical mangal/salt marsh
community. Mangroves and Salt Marshes 1:103-111.
Patterson-Zucca, C., 1982. The effects of road construction on a mangrove ecosystem. Trop. Ecol.
23: 105-124.
Pauly, D. and J. Ingles, 1999. The relationship between shrimp yields and intertidal vegetation
(mangrove) areas: a reassessment. In: Yanez-Arancibia, A. and A.L. Lara-DomInguez (Eds.),
Ecosistemas de mangIar en Am6rica tropical. Instituto de Ecologia, Xalapa, pp. 311-316.
Pegg, K.G. and J.L. Alcorn, 1982. Phytophthora operculata sp. nov., a new marine fungus. Mycotaxon
16:99-102.
Pegg, K.G. and I.M. Foresberg, 1981. Phytophthora in Queensland mangroves. Wetlands 1:2-3.
Pegg, K.G., N.C. Gillespie and L.I. Foresberg, 1980. Phytophthora sp. associated with mangrove death
in Central Coastal Queensland. Aust. Plant Path. 9:6-7.
Pelegri, S.P. and R.R. Twilley, 1998. Heterotrophic nitrogen fixation (acetylene reduction) during leaf-
litter decomposition of two mangrove species from south Florida, USA. Mar. BioI. 131 :53-61.
Pellegrin, F., 1952. Les Rhizophorac6es de l'Afrique 6quatoriale fran~aise. Notulae Systematicae
14:292-300.
Percival, M., 1974. Floral ecology of coastal scrub in south-east Jamaica. Biotropica 6:104-129.
Percival, M. and 1S. Womersley, 1975. Floristics and ecology of the mangrove vegetation of Papua
New Guinea. Bot. Bull. No.8, Dept. Forestry, Divn. Bot. Lae, Papua New Guinea.
Pernetta, J.C., 1993. Mangrove forests, climate change and sea level rise: hydrological influences on
community structure and survival. with examples from the Indo-West Pacific. IUCN. Gland.
Perrings. c., 2000. Sustainability indicators for fisheries in integrated coastal area management. Mar.
Freshw. Res. 51 :513-522.
Pezeshki, S.R .• R.D. Delaune and W.H. Patrick, 1990. Differential response of selected mangroves to
soil flooding and salinity: gas exchange and biomass partitioning. Can. J. For. Res. 20:869-874.
Phillipps. C., 1984. Current status of mangrove exploitation. management and conservation in Sabah. In:
Soepadmo, E .• A.N. Rao and OJ. McIntosh (Eds.), Proceedings of the Asian Symposium
Mangrove Environments - Research and Management. UNESCO. pp. 809-820.
Phillips. F.H. and AJ. Watson, 1959. Pulping studies on New Guinea woods. I. Sulphate. sulphite. and
neutral sulphite semichemical pulps from Excoecaria agallocho L. and Camptostemon schultzii.
CSIRO Division of Forest Products Technological Paper No.8. 19pp.
Pielou, E.C., 1977. Mathematical ecology. John Wiley & Sons. New York.
Pillai. G., 1985. The mana. or mangrove-lobster. Domodomo 3:2-20.
Pinto. L .• 1987. Environmental factors influencing the occurrence of juvenile fish in the mangroves of
Pagbilae, Philippines. Hydrobiologia 150:283-301.
Piyakamchana. T., 1981. Severe defoliation of Avicennia alba BI. by larvae of Cleora injectaria
Walker. J. Sci. Soc. Thailand 7:33-36.
Plaziat. J.-C .• 1995. Modem and fossil mangroves and mangals: their climatic and biogeographic
variability. In: Bosence. D.W.J. and P.A. Allison (Eds.). Marine palaeoenvironmental analysis
from fossils. Geol. Soc. Spec. Publ. No. 83. pp. 73-96.
Plaziat. J.-c. and C. Cavagnetto, 1996. Taphonomic and biogeographic processes controlling the
mangrove trees and mollusc associations of the Pyrenean Paleocene and Eocene.
Comunicaci6n de la" Reuni6n de Tafonomia y fosilizaci6n. pp. 331-336.
Plaziat, J.-c.. C. Cavagnetto. J.-C. Koeniguer and F. Baltzer, 2001. History and biogeography of the
mangrove ecosystem. based on a critical reassessment of the paleontological record. Wetlands
Ecology and Management 9:161-179.
Pole, M.S. and M.K. Macphail 1996. Eocene Nypa from Regatta Point, Tasmania. Rev. Palaeobot.
Palynol. 92:55-67.
References 335

Pollard, D.A., 1973. Estuaries: Development and 'progress' versus commonsense. Fisherman 4:28-32.
Pollard, D.A., 1981. Estuaries are valuable contributors to fisheries production. Aust. Fish. 40:7-9.
Poovachiranon, S. and P. Tantichodok, 1991. The role of sesarmid crabs in the mineralization of leaf
litter of Rhizophora apicuLata in a mangrove, southern Thailand. Phuket Mar. BioI. Center Res.
56:63-74.
Poovachiranon, S., K. Boto and N. Duke, 1986. Food preference studies and ingestion rate
measurements of the mangrove amphipod ParhyaLe hawaiensis (Dana). 1. Exp. Mar. BioI.
Ecol. 98:129-140.
Pool, DJ., S.C. Snedaker and A.E. Lugo, 1977. Structure of mangrove forests in Florida, Puerto Rico,
Mexico and Costa Rica. Biottopica 9:195-212.
Popp, M., 1984. Chemical composition of Australian mangroves. I. Inorganic ions and organic acids. Z.
Pflanzenphysiol. 113:395-409.
Popp, M., 1995. Salt resistance in herbaceous halophytes and mangroves. Progress in Botany 56:416-
429.
Popp, M., F. Larher and P. Weigel, 1984. Chemical composition of Australian mangroves. III. Free
amino acids, total methylated onium compounds and total nitrogen. Z. Pflanzenphysiol. 114: 15-
25.
Popp, M., J. Polania and M. Weiper, 1993. Physiological adaptations to different salinity levels in
mangrove. In: H. Lieth and A. Al Masoom (Eds.), Towards the rational use of high salinity
tolerant plants. Vol. I., Kluwer Academic Publishers, Utrecht, pp. 217-224.
Por, F.D., 1984. The ecosystem of the mangal: general considerations. In: Por, F.D. and I. Dor (Eds.),
Hydrobiology of the mangal. Dr. W. Junk, The Hague, pp. 1-14.
Por, F.D., I. Dor and A. Amir, 1977. The mangal of Sinai: limits of an ecosystem. Helgol. wiss.
Meeresunters. 30:411-424.
Potts, M., 1979. Nittogen-fixation (acetylene reduction) associated with communities of heterocystous
and non-heterocystous blue-green algae in the mangrove forests of Sinai. Oecologia 39:359-
374.
Prahl, H. von, 1986. Notas sobre la historia natural del mangle pinuelo PeLLiciera rhizophorae
(Theaceae) en el Pacifico Colombiano. Actual. BioI. 15:117-122,
Prakasah, N. and A.L. Lim, 1995. The systematic position of AegiaLitis, the enigmatic mangrove.
Wallaceana 7:11-15.
Prawiroatmodjo, S., 1988. Forest as an ecosystem, its structure and function. 5. Tidal effect on seed
dispersal. In: Ogino, K. and M. Chihara (Eds.), Biological system of mangroves - a report of
east Indonesian mangrove expedition, 1986. Ehime University, Ehime, pp. 137-140.
Primack, RB. and P.B. Tomlinson, 1978. Sugar secretions in buds of Rhizophora attractive to birds.
Biotropica 10:74-75.
Primack, RB. and P.B. Tomlinson, 1980. Variation in ttopical forest breeding systems. Biotropica
12:229-231.
Primack, RB., N.C. Duke and P.B. Tomlinson, 1981. Floral morphology in relation to pollination
ecology in five Queensland coastal plants. Austrobaileya 1:346-355.
Primavera, J.H., 1995. Mangroves and brackish water pond culture in the Philippines. Hydrobiologia
295:303-309.
Primavera, J.H., 1998. Mangroves as nurseries: shrimp popUlations in mangrove and non-mangrove
habitats. Est. Coastal Shelf Sci. 46:457-464.
Proches, S., 2001. Back to the sea: secondary marine organisms from a biogeographical perspective.
BioI. 1. Linn. Soc. 74:197-203.
Puente, M.E., G. Holguin, B.R. Glick and Y. Ba.~han, 1999. Root-surface colonization of black
mangrove seedlings by AzospiriLlum haLopraeferens and AzospiriLium brasilense in seawater.
FEMS Microbiol. Ecol. 29:283-292.
Pulver, T. R, 1976. Transplanting techniques for sapling mangrove trees, R. mangle, L racemosa, and
A. gemlinans, in Florida. Flo. Mar. Res. Publ. 22:1-14.
Purser, B.H., 1982. Sedimentation recente de la Mer Rouge et du Golfe Arabo-Persique et dynamique
de la Plaque Arabique. Geochromique 2:11-14.
Putz, F.E. and H.T. Chan, 1986. Tree growth, dynamics and productivity in a mature mangrove forest in
Malaysia. For. Ecol. Manage. 17:211-230.
Pynaert, L., 1933. La mangrove congolaise. Bull. Agro. Congo BeIge. 23:184-207.
Qureshi, T. M. D., 1990. Experimental plantation for rehabilitation of mangrove forest in Pakistan.
ISME Mangrove Ecosystems: Occasional Papers No. 4:1-37.
Qureshi, M.T., 1996. Restoration of mangroves in Pakistan. In: Field, C. (Ed.), Restoration of mangrove
ecosystems. ISMElITTO, Okinawa, pp. 126-142.
Rabinowitz, D. 1975. Planting experiments in mangrove swamps of Panama. In: Walsh, G.E., S.C.
Snedaker and HJ. Tea.~ (Eds.), Proceedings of the International Symposium on Biology and
Management of Mangroves, Vol. 1. University of Florida, Gainesville, pp. 385-393.
Rabinowitz, D., 1977. Effects of a mangrove borer, PoeciLips rhizophorae, on propagu\es of
Rhizophora harrisonii in Panama. Flo. Entomol. 60:129-134.
Rabinowitz, D., 1978a. Dispersal properties of mangrove propagules. Biotropica 10:47-57.
336 Mangrove Ecology, Silviculture and Conservation

Rabinowitz, D., 1978b. Mortality and initial propagule size in mangrove seedlings in Panama. J. Ecol.
66:45-51.
Rabinowitz, D., 1978c. Early growth of mangrove seedlings in Panama, and an hypothesis concerning
the relationships of dispersal and zonation. J. Biogeogr. 5: 113-133.
Rada, F., G. Goldstein, A. Orozco, M. Montilla, O. Zabala and A. Azocar, 1989. Osmotic and turgor
relations of three mangrove ecosystem species. Aust. J. PI. Physiol. 16:477-486.
Rafii, A.Z., R.S. Dodd and M.-T. Fauvel, 1999. A case of natural selection in Atlantic-East-Pacific
Rhizophora. Hydrobiologia 413:1-9.
Rafii, Z.A., R.S. Dodd and F. Fromard, 1996. Biogeographic variation in foliar waxes of mangrove
species. Biochemical and Systematic Ecology 24:341-345.
Rains, D.W. and E. Epstein, 1967. Preferential absorption of potassium by leaf tissue of the mangrove
Avicennia marina: an aspect of halophytic competence in coping with salt. Aust. J. BioI. Sci.
20: 847-857.
Rajendraprasad, M., P.N. Krishnan and P. Pushpangadan, 1998. The life form spectrum of sacred
groves - a functional tool to analyse the vegetation. Trop. Ecol. 39:211-217.
Ramassamy, V. and B. Kannabiran, 1996. Leaf epidermis and taxonomy in Rhizophoraceae. Indian
Forester 122:1049-1061.
Ramirez-Garcia, P., J. L6pez-Blanco and D. Ocana, 1998. Mangrove vegetation assessment in the
Santiago River mouth, Mexico, by means of supervised classification using Landsat TM
imagery. For. Ecol Manage. 105:217-229.
Ramsey, E.W. and J.R. Jensen, 1996. Remote sensing of mangrove wetlands: relating canopy spectra to
site-specific data. Photogrammetric Engineering and Remote Sensing 62:939-948.
Ramsey, E.W., G.A. Nelson and S.K. Sapkota, 1998. Classifying coastal resources by integrating
optical and radar imagery and colour infrared photography. Mangroves and Salt Marshes
2:109-119.
Rao, A.N. and H. Tan, 1984. Leaf structure and its ecological significance in certain mangrove plants.
In: Soepadmo, E., A.N. Rao and DJ. Mcintosh (Eds.), Proceedings of the Asian Symposium
Mangrove Environments - Research and Management, UNESCO, pp. 183-194.
Rao, C.S., P. Eganathan, A. Anand, L. Rangan and P. Balakrishna, P., I 998a. Application of
biotechnology and classical breeding methods in the genetic enhancement of mangroves.
Abstracts of the International Symposium on Mangrove Ecology & Biology, Kuwait, April 25-
27, 1998, p. 13.
Rao, C.S., P. Eganathan, A. Anand, P. Balakrishna and T.P. Reddy, I998b. Protocol for in vitro
propagation of Excoecaria agallocha L., a medicinally important mangrove species. Plant Cell
Reports 17:861-865.
Rao, D.W. and M. Sharma, 1968. The terminal sclereids and tracheids of Bruguiera gymnorhiza Blume
and the cauline sclereids of Ceriops roxburghiana Am. Proc. Nat. Inst. Sci. India. (Part B.
BioI. Sci.) 34: 267-275.
Rasolofo, V.M., 1993. Les mangroves de Madagascar. In: Diop, E.S. (Ed.), Conservation et utilisation
rationelle des forets de mangroves de l'Amerique Latine and de I'Afrique. Vol. II. Version
fran~aise du Rapport sur l'Afrique. ITTOIISME, Okinawa, pp. 248-265.
Rasowo, J., 1992. Mariculture development in Kenya: alternatives to siting ponds in the mangrove
ecosystem. Hydrobiologia 247:209-214.
Raymond, A. and T.L. Phillips, 1983. Evidence for an Upper Carboniferous mangrove community. In:
Teas, HJ. (Ed.), Biology and ecology of mangroves. Tasks for vegetation science, Vol. 8. Dr
W. Junk, The Hague, pp. 19-30.
Reinders-Gouwentak, C.A., 1953. Sonneratiaceae and other mangrove-swamp families, anatomical
structure and water relations. Flora Malesiana 4:513-515.
Richards, P.W., 1964. The tropical rain forest. Cambridge Univ. Press, Cambridge.
Ricklefs, R.E. and R.E. Latham, 1993. Global patterns of diversity in mangrove floras. In: Ricklefs, R.E.
and D. Schluter (Eds.), Species diversity in ecological communities. Historical and
geographical perspectives. University of Chicago Press, Chicago, pp. 215-33.
Rickson, F.R., 1979. Absorption of animal tissue breakdown product.~ into a plant stem - the feeding of
a plant by ants. Amer. J. Bot. 66:87-90.
Rico-Gray, V., 1993. Origen y rutas de dispersi6n de los mangles: una revisi6n con enfasis en las
especies de America. Acta Botanica Mexicana 25:1-13.
Rico-Gray, V. and L. da S.L. Sternberg, 1991. Carbon isotopic evidence for seasonal change in feeding
habits of Camponotus planatus Roger (Formicidae) in Yucatan, Mexico. Biotropica 23:93-95.
Ridd, P.V., 1996. Flow through animal burrows in mangrove creeks. Est. Coastal Shelf Sci. 43:617-625.
Ridd, P.V. and R. Sam, 1996. Profiling groundwater salt concentrations in mangrove swamps and
tropical salt flats. Est. Coastal Shelf Sci. 43:627-635.
Riley, R.W. and c.P.S. Kent, 1999. Riley encased methodology: principles and processes of mangrove
habitat creation and restoration. Mangroves and Salt Marshes 3:207-213.
Rintz, R.E., 1980. The peninsula Malayan species of Dischidia (Asclepiadaceae). Blumea 26:81-126.
Rivera-Monroy, V.H. and R.R. Twilley, 1996. The relative role of denitrification and immobilization in
the fate of inorganic nitrogen in mangrove sediments. Limnol. Oceanogr. 41:284-296.
References 337

Rivera-Monroy. V.H .• R.R. Twilley. R.G. Boustany. 1.W. Day. F. Vera-Herrera and M. del Carmen
Ramirez. 1995. Direct denitrification in mangrove sediments in Terminos Lagoon. Mexico.
Mar. Ecol. Progr. Ser. 126:97-109.
Riviere. C .• 1909. Les vegetaux tanniteres dans Ie nord d·Afrique. Journ. d·Agr. trop. 94:101-103.
Robertson. A.I.. 1986. Leaf-burying crabs: their influence on energy flow and export from mixed
mangrove forests (Rhizophora spp.) in northeastern Australia. 1. Exp. Mar. BioI. Ecol. 102:237-
248.
Robertson. A.1. and P.A. Daniel. 1989a. The influence of crabs on litter processing in high intertidal
mangrove forests in tropical Australia. Oecologia 78: 191-198.
Robertson. A.1. and N.C. Duke. I 987a. Insect herbivory on mangrove leaves in North Queensland.
Aust. l Ecol. 12:1-7.
Robertson. A.1. and N.C. Duke. 1987b. Mangroves as nursery sites: comparisons of the abundance and
species composition of fish and crustaceans in mangroves and other nearshore habitat~ in
tropical Australia. Mar. BioI. %:193-205.
Robertson. A.1. and N.C. Duke. 199Oa. Mangrove fish-communities in tropical Queensland. Australia:
spatial and temporal patterns in densities. biomass and community structure. Mar. BioI.
104:369-379.
Robertson. A.I. and N.C. Duke. 1990b. Recruitment. growth and residence time of fishes in a tropical
Australian mangrove system. Est. Coa.~t Shelf Sci. 31 :723-743.
Robertson. A.I. and MJ. Phillips. 1995. Mangroves as filters of shrimp pond effluent: predictions and
biogeochemical research needs. Hydrobiologia 295:311-321.
Robertson. A.I.. P.A. Daniel and P. Dixon. 1991. Mangrove forest structure and productivity in the Ay
River estuary. Papua New Guinea. Mar. BioI. 111:147-155.
Robertson. A.I.. R. Giddins and TJ. Smith. 1990. Seed predation by insects in tropical mangrove forests:
extent and effects on seed viability and the growth of seedlings. Oecologia 83:213-219.
Rodelli. M.R .• J.N. Gearing. P.N. Gearing. N. Marshall and A. Sasekumar. 1984. Stable isotope ratio as
a tracer of mangrove carbon in Malaysian ecosystems. Oecologia 61 :326-333.
Rodriguez. G .• 1987. Structure and production in neotropical mangroves. Trends Ecol. Evol. 2:264-267.
Roessler. M.A .• 1971. Environmental changes associated with a Florida power plant. Mar. Poll. Bull.
2:87-90.
Rosevear. D.R .• 1947. Mangrove swamps. Farm and Forest 8:23-30.
Ross. M.S .• J.F. Meeder. J.P. Sah, P.L. Ruiz and GJ. Telesnicki. 2000. The southea.~t saline Everglades
revisited: 50 years of coastal vegetation change. J. Veg. Sci. 11:101-112.
Ross. M.S .• P.L. Ruiz. G.J. Telesnicki and J.F. Meeder. 2001. Estimating above-ground bioma.~s and
production in mangrove communities of Biscayne Park. Florida (U.S.A.). Wetlands Ecology
and Management 9:27-37.
Rossi. G .• 1981. Le Quaternaire littoral du Kenya. Zeitschrift flir Geomorphologie 25:33-53.
Rossignol-Strick. M. and D. Duzer. 1979. West African vegetation and climate since 22.500 BP from
deep-sea core palynology. Pollen et Spores 21:105-134.
Roth. I.. 1%5. Histogenese der Lentizellen am Hypocotyl von Rhizophora mangle L. Ost. Bot. Zeit.
112: 640-653.
Roth. I.• 1992. Leaf structure: coastal vegetation and mangroves of Venezuela. Gebriider Borntraeger.
Berlin.
Roth. L.C .• 1992. Hurricanes and mangrove regeneration: effects of hurricane Joan. October 1988. on
the vegetation of Isla del Venado. Bluefields. Nicaragua. Biotropica 24:375-384.
Roth. L.C .• 1997. Implications of periodic hurricane disturbance for the sustainable management of
Caribbean mangroves. In: Kjerfve. B.• L.D. Lacerda and E.H.S. Diop (Eds.). Mangrove
ecosystem studies in Latin America and Africa. UNESCO. Paris. pp. 18-34.
Roth, L.C. and A. Grijalva, 1991. New record of the mangrove, Pelliciera rhizophorae, on the
Caribbean coast of Nicaragua. Rhodora 93:183-186.
Roy, P.S., B.G. Thom and L.D. Wright, 1980. Holocene sequences on an embayed high-energy coast:
an evolutionary model. Sedimentary Geol. 26:1-19.
Rubin, lA., C. Gordon and J.K. Amatekpor. 1998. Causes and consequences of mangrove
deforestation in the Volta estuary, Ghana: some recommendations for ecosystem rehabilitation.
Mar. Poll. Bull. 37:441-449.
Ruitenbeek, HJ., 1994. Modelling economy-ecology linkages in mangroves: economic evidence for
promoting conservation in Bintuni Bay. Indonesia. Ecol. Econ. 10:233-247.
Ruiz. G.M .• A.H. Hines and M.H. Posey. 1993. Shallow water as a refuge habitat for fish and
crustaceans in nonvegetated estuaries - an example from Chesapeake Bay. Mar. Ecol. Progr.
Ser.99:1-16.
Rull. V., T. Vegas-Vilamibia and N. Espinoza de Pernla, 1999. Palynological record of an Early-Mid
Holocene mangrove in eastern Venezuela. Implications for sealevel rise and disturbance
history. 1. Coa.~tal Res. 15:496-504.
Rumbold, D.G. and S.c. Snedaker, 1993. Do mangroves float? J. Trop. Ecol. 10:282-284.
Ruwa, R.K., 1993. Zonation and distribution of creek and fringe mangroves in the semi-arid Kenyan
coast. In: Lieth, H. and A. AI Masoom (Eds.), Towards the rational use of high salinity tolerant
plants. Vol. I.. Kluwer Academic Publishers, Utrecht, pp. 97-105.
338 Mangrove Ecology, Silviculture and Conservation

Ryther, J.H., 1975. Mariculture: how much protein and for whom? Oceanus 18:10-22.
Saad, S., M.L. Husain, R. Yaacob and T. Asano, 1999. Sediment accretion and variability of
sedimentological characteristics of a tropical estuarine mangrove: Kemaman. Terengganu.
Malaysia. Mangroves and Salt Marshes 3:51-58.
Sadiq, M. and T.H. Zaidi, 1994. Sediment composition and metal concentrations in mangrove leaves
from the Saudi coast of the Arabian Gulf. Sci. Tot. Environ. 155:1-8.
Saenger, P., 1982. Morphological, anatomical and reproductive adaptations of Australian mangroves.
In: Clough, BF. (Ed.), Mangrove ecosystems in Australia. Australian National University
Press, Canberra, pp. 153-191.
Saenger, P., 1984. Mangroves: life in a harsh environment. In: Talbot, F. (Ed.), Reader's Digest book of
the Great Barrier Reef. Reader's Digest, Sydney, pp. 304-323.
Saenger, P., 1985a. An initial attempt to assess life strategies in Australian mangroves. In: Beardsley,
K.N., J.D.S. Davie and C.D. Woodroffe (Eds.), Coasts and tidal wetlands of the Australian
monsoon Region. NARU Mangrove Monographs 1:187-200.
Saenger, P., 1985b. Definition and perceptions of the mangrove resource: the biological view. In:
Davie, J.D.S., J.R. Hanley and B.C. Russell (Eds.), Coa~tal management in northern Australia.
NARU Mangrove Monographs 2:15-20.
Saenger, P., 1986. Bangladesh mangrove afforestation project. Report prepared for the Government of
Bangladesh. Shedden Pacific Pty. Limited, Melbourne.
Saenger, P., 1987. Mangrove use and conservation. In: Field, C.D. and A.I. Dartnall (Eds.), Mangrove
ecosystems of Asia and the Pacific: status, exploitation and management. Australian Institute
Marine Science, Townsville, pp. 97-103.
Saenger, P., 1988. Gladstone environmental survey - final report. Queensland Electricity Commission,
Brisbane.
Saenger, P., 1994a. Cleaning up tbe Arabian Gulf: aftermath of an oil spill. Search 25: 19-22.
Saenger, P., 1994b. Mangroves and saltmarshes. In: Hammond, L.S. and R.N. Synnot (Eds.), Marine
biology. Longman Cheshire, Melbourne, pp. 238-256.
Saenger, P., 1995. The status of Australian estuaries and enclosed marine waters. In: Zann, L. and P.
Kailola (Eds.), State of the marine environment report for Australia: technical annex I, The
marine environment. GBRMPNOcean Rescue 2000, pp. 53-73.
Saenger, P., 1996. Mangrove restoration in Australia: a case study of Brisbane International Airport. In:
Field, C. (Ed.), Restoration of mangrove ecosystems. ISMElITTO, Okinawa, pp. 36-51.
Saenger, P., 1998. Mangrove vegetation: an evolutionary perspective. Mar. Freshw. Res. 49:277-286.
Saenger, P. and M.F. Bellan, 1995. The mangrove vegetation of the Atlantic coast of Africa. Universite
de Toulouse Press, Toulouse.
Saenger, P. and K. Bilham, 1996. Sustainable management of mangrove ecosystems. In: Khemnark, C.
(Ed.), Proceedings of Ecotone IV: ecology and management of mangrove restoration and
regeneration in east and southeast Asia. Surat Thani, Thailand, pp. 171-176.
Saenger, P. and M.S. Hopkins, 1975. Observations on the mangroves of the south-eastern Gulf of
Carpentaria, Australia. In: Walsh, G.E., S.c. Snedaker and H.I. Teas (Eds.), Proceedings of
the International Symposium on Biology and Management of Mangroves, Vol. I. University of
Florida, Gainesville, pp. 126-136.
Saenger, P. and G. Luker, 1997. Some phytogeographical considerations of mangroves. In: Spalding,
M., F. Blasco and C. Field, (Eds.), World Mangrove Atlas. ISMFJITTO, Okinawa, pp. 27-39.
Saenger, P. and C.C. McIvor, 1975. Water quality and fish populations in a mangrove estuary modified
by residential canal developments In: Walsh, G.E., S.C. Snedaker and H.J. Teas (Eds.),
Proceedings of the International Symposium on Biology and Management of Mangroves, Vol.
2. University of Florida, Gainesville, pp. 753-65.
Saenger, P. and J. Moverley, 1985. Vegetative phenology of mangroves along the Queensland
coastline. Proc. Ecol. Soc. Aust. 13:257-265.
Saenger, P. and J. Robson, 1977. A structural analysis of mangrove communities on the central
Queensland coastline. Mar. Res. Indonesia 18:101-118.
Saenger, P. and N.A. Siddiqi, 1993. Land from the sea: the mangrove afforestation program of
Bangladesh. Ocean and Coastal Managt. 20:23-39.
Saenger, P. and S.C. Snedaker, 1993. Pantropical trends in mangrove above-ground biomass and annual
litter fall. Oecologia 96:293-299.
Saenger, P., E.J. HegerJ and I.D.S. Davie (Eds.), 1983. Global status of mangrove ecosystems. The
Environmentalist 3 (Suppl.): 1-88.
Saenger, P., D. McConchie and M. Clark, 1991. Mangrove forests as a buffer zone between
anthropogenically polluted areas and the sea. In: Arakel, A.V. (Ed.), Proceedings of 1990
Workshop on Coastal Zone Management, Yeppoon, Queensland, pp. 280-299.
Saenger, P., Y. Sankare and T. Perry, 1996. Review of selection criteria and ecological guidelines for
mangrove restoration studies. Large Marine Ecosystem Project for the Gulf of Guinea.
UNIDO, UNDP, NOAA and UNEP, Abidjan.
Saenger, P., M.M. Specht, R.L. Specht and V.I. Chapman, 1977. Mangal and coastal salt marsh
communities in Australasia. In: Chapman, V.J. (Ed.), Wet coastal ecosystems. Elsevier
Scientific Publ. Co., Amsterdam, pp. 293-345.
References 339

Saifullah, S.M., 1997. Management of the Indus delta mangroves. In: Haq, B.U. (Ed.), Coastal zone
management imperative for maritime developing nations. Kluwer Academic Publishers,
Dordrecht, pp. 333-346.
Saintilan, N., 1997a. Above- and below-ground biomasses of two species of mangrove on the
Hawkesbury River estuary. Mar. Freshw. Res. 48:147-152.
Saintilan, N., 1997b. Above- and below-ground biomass of mangroves in a sub-tropical estuary. Mar.
Freshw. Res. 48: 601-604.
Saintilan, N. and T.R. Hashimoto, 1999. Mangrove-saltmarsh dynamics on a bay-head delta in the
Hawkesbury River estuary, New South Wales, Australia. Hydrobiologia 413:95-102.
Saintilan, N. and K. Wilton, 200 1. Changes in the distribution of mangroves and saltmarshes in Jervis
Bay, Australia. Wetlands Ecology and Management 9:409-420.
Salard-Cheboldaeff, M., 1976. Presence de l'Oligocene dans Ie ba~sin sedimentaire c<~tier du
Cameroun. Revue Micropaleontologique 18:236-245.
Salt, D.E., M. Blaylock, P.BA Kumar, V. Dushenkov, B.D. Ensley, I. Chet and I. Raskin, 1995. Phyto-
remediation: a novel strategy for the removal of toxic metals from the environment using
plants. Biotechnology 13:468-474.
Santos, M.C.F.V., J.e. Zieman and R.R.H. Cohen, 1997. Interpreting the upper mid-littoral zonation
patterns of mangroves in Maranhiio (Brazil) in response to microtopography and hydrology.
In: Kjerfve, B., L.D. Lacerda and E.H.S. Diop (Eds.), Mangrove ecosystem studies in Latin
America and Africa. UNESCO, Paris, pp. 127-144.
Sasekumar, A., V.C. Chong, M.U. Leh and R. Da Cruz, 1992. Mangroves as a habitat for fish and
prawns. Hydrobiologia 247:195-207.
Sattar, M.A. and D.K. Bhattacharjee, 1983. Strength properties of keora (Sonneratia apetala). Timber
Physics Series, Bulletin 7, Forest Research Institute, Chittagong, pp. 1-6.
Sattar, M.A. and D.K. Bhattacharjee, 1987. Physical and mechanical properties of sundri (Heritiera
fomes) and baen (Avieennia alba). Timber Physics Series, Bulletin 10, Forest Research
Institute, Chittagong, pp. 1-8.
Satuwong, I., I. Ninomiya and K. Ogino, 1995. Callus and multiple shoot formation in Bruguiera
gymnorrhiza. Bull. Ehime Univ. Forestry 32:25-33.
Satyanarayana, B., A.V. Raman, F. Dehairs, P. Chandramohan and e. Kalavati, in press. Mangrove
floristic and zonation patterns of Coringa, Kakinada Bay, east coast of India. Wetlands
Ecology and Management.
Saur, E., I. Bonheme, P. Nygren and D. Imbert, 1998. Nodulation of Pteroearpus officinalis in the
swamp forest of Guadeloupe (Lesser Antilles). J. Trop. Ecol. 14:761-770.
Saur, E., D. Imbert, 1. Etienne and D. Mian, 1999. Insect herbivory on mangrove leaves in Guadeloupe:
effects on biomass and mineral content. Hydrobiologia 413:89-93.
Savolainen, V., M.F. Fay, D.C. Albach, A. Backlund, M. Van Der Bank, K.M. Cameron, S.A. Johnson,
M.D. L1edo, J.C. Pintaud, M. Powell, M.C. Sheahan, D.E. Soltis, P.S. Soltis, P. Weston, W.M.
Whitten, K.J. Wurdack, and M.W. Chase, 2000. Phylogeny of the eudicots: a nearly complete
familial analysis based on rbeL gene sequences. Kew Bull. 55:257-309.
Savory, H.1., 1953. A note on the ecology of Rhizophora in Nigeria. Kew Bull. 8:127-128.
Schaeffer-Novelli, Y., R.R. Adaime, Y.M. de Camargo and G. Cintron, 1985. Variability of the
mangrove ecosystem along the Brazilian coast. Estuaries 8:107-112.
Scherrer, P. and F. Blasco, 1989. Croissance des plantules de Rhizophora mangle dans un substrat
tourbeux de mangrove massivement contamine par du petrol brnt. Bull. Ecol. 20:203-210.
Scherrer, P. and G. Mille, 1989. Biodegradation of crude oil in an experimentally polluted peaty
mangrove soil. Mar. Poll. Bull. 20:430-432.
Schoener, T.W., 1988. Leaf damage in island buttonwood, Conacarpus ereetus: correlations with
pubescence, island area, isolation and the distribution of major carnivores. Oikos 53:253-266.
Scholander, P.F., 1968. How mangrove desalinate seawater. Physiol. Plant. 21:251-261.
Scholander, P.F., E.D. Bradstreet, H.T. Hammel and E.A. Hemmingsen, 1966. Sap concentrations in
halophytes and some other plants. Plant Physiol. 41 :529-532.
Scholander, P.P., H.T. Hammel, E. Hemmingsen and W. Garey, 1%2. Salt balance in mangroves. Plant
Physiol. 37:722-729.
Scholander, P.F., H.T. Hammel, E.D. Bradstreet and E.A. Hemmingsen, 1965. Sap pressure in vascular
plants. Science 148:339-346.
Scholander, P.F., L. Van Dam and S.l. Scholander, 1955. Gas exchange in the roots of mangroves.
Amer. J. Bot. 42:92-98.
Schwarzbach, A.E. and L.A. McDade, in press. Phylogenetic relationships of the mangrove family
Avicenniaceae ba~ed on chloroplast and nuclear ribosomal DNA sequences. Syst. Bot.
Schwarzbach, A.E. and R.E. Ricklefs, 2000. Systematic affinities of Rhizophoraceae and
Anisophyllaceae, and intergeneric relationships within Rhizophoraceae, based on chloroplast
DNA, nuclear ribosomal DNA, and morphology. Amer. J. Bot. 87:547-564.
Schwarzbach, A.E. and R.E. Ricklefs, 2001. The use of molecular data in mangrove plant research.
Wetlands Ecology and Management 9:195-201.
Scotland, R.W., J.A. Sweere, P.A. Reeves and R.G. Olmstead, 1995. Higher-level systematics of
Acanthaceae determined by chloroplast DNA sequences. Amer. J. Bot. 82:266-275.
340 Mangrove Ecology, Silviculture and Conservation

Semeniuk, V., 1983. Mangrove distribution in northwestern Australia in relationship to regional and
local freshwater seepage. Vegetatio 53: 11-31.
Semeniuk. V.• 1985a. Development of mangrove habitats along ria shorelines in north and north-
western tropical Australia. Vegetatio 60:3-23.
Semeniuk. V.• 1985b. Mangrove environmenl~ of Port Darwin. Northern Territory: the physical
framework and habitats. J. Royal Soc. WA 67:81-97.
Semeniuk. V.• 1994. Predicting the effect of sea-level rise on mangroves in northwestern Australia. J.
Coastal Res. 10: 1050-1076.
Semeniuk, V .• K.F. Kenneally and P.G. Wilson. 1978. Mangroves of Western Australia. Handbook No.
12. WA Naturalist's Club. Perth.
Semesi. A.K., 1998. Mangrove management and utilization in eastern Africa. Ambio 27:620-626.
Sengupta. A. and S. Chaudhuri. 1991. Ecology of heterotrophic dinitrogen fixation in the rhizosphere of
mangrove plant community at the Ganges River estuary in India. Oecologia 87:560-564.
Setoguchi. H.• K. Kosuge and H. Tobe. 1999. Molecular phylogeny of Rhizophoraceae based on rbcL
gene sequences. J. Plant Res. 112:443-455.
Sharp. W.R .• 1980. Overview of potential in agricultural application. In Vitro 16:221.
Sheridan. R.P.• 1991. Epicaulous nitrogen-fixing microepiphytes in a tropical mangal community.
Guadeloupe. French West Indies. Biotropica 23:530-541.
Sheridan. R.P.• 1992. Nitrogen fixation by epicaulous cyanobacteria in the Pointe de la Saline mangrove
community. Guadeloupe, French West Indies. Biotropica 24:571-574.
Sherman. R.E., T.J. Fahey and R.W. Howarth. 1998. Soil-plant interactions in a neotropical mangrove
forest: iron. phosphorus and sulfur dynamics. Oecologia 115:553-563.
Sherman. R.E .• T.J. Fahey and U. Battles. 2000. Small-scale disturbance and regeneration dynamics in
a neotropical mangrove forest. J. Ecol. 88:165-178.
Sherrod. C. L. and C. McMillan. 1985. The distribution history and ecology of mangrove vegetation
along the northern Gulf of Mexico coastal region. Contr. Mar. Sci. 28:129-140.
Sherrod. C.L., D.L. Hockaday and C. McMillan. 1986. Survival of red mangrove. Rhizoplwra mangle.
on the Gulf of Mecico coast of Texas. Contr. Mar. Sci. 29:27-36.
Shields. L.M., 1950. Leaf xeromorphy as related to physiological and structural influences. Bot. Rev.
16:399-447.
Shinaq. R. and K. Bandel, 1998. The flora of an estuarine channel margin in the Early Cretaceous of
Jordan. Freiberger Forschungsheft C 474:39-57.
Shokita, S .• 2000. The role of aquatic animals in a mangrove ecosystem. In: Asia-Pacific cooperation
on research for conservation of mangroves: proceedings of an international workshop.
Okinawa, Japan, 26-30 March. 2000. United Nations University. Tokyo. pp. 1-30.
Shriadah. M.MA, 1999. Heavy metals in mangrove sediments of the United Arab Emirates shoreline
(Arabian GuU). Water. Air and Soil Pollution 116:523-534.
Siddiqi. N.A.. 1987. Observation on initial spacing in keora (Sonneratia apetala) plantation along the
coastal belt of Bangladesh. Malay. For. 50:204-216.
Siddiqi. N.A .• 1988. Growth. natural thinning and wood production in a keora (Sonneratia apetala)
stand. Bano Biggyan Patrika 17:91- 93.
Siddiqi. N.A .• 1992. Regeneration status and influence of animals on regeneration in the Sundarbans
mangrove forest. Unpublished Thesis. University of Dhaka, Dhaka.
Siddiqi. N.A.. 1996. Influence of deer on the artificial regeneration in the mangrove of Sundarbans.
Bangladesh. Indian Forester 122:937-942.
Siddiqi. N.A .• 2001. Mangrove forestry in Bangladesh. University of Chittagong. Chittagong.
Siddiqi. N.A. and M.R. Islam, 1988. Studies on fruit size. seed production and viability of keora
(Sonneratia apetala). Bano Biggyan Patrika 17:15-19.
Siddiqi. N.A. and M.A.S. Khan. 1996. Planting techniques for mangroves on new accretions in the
coastal areas of Bangladesh. In: Field, C. (Ed.). Restoration of mangrove ecosystems.
ISMElITTO. Okinawa. pp. 143-159.
Siddiqi. N.A.. M.J. Alam and M.A. Habib. 1995. Natural regeneration of mangroves in the coastal
areas of Bangladesh with particular reference to Noakhali. Bangladesh J. Forest Sci. 24:62-69.
Siddiqi. NA. M.R. Islam and MAS. Khan. 1989. Effect of salinity on germination success in keora
(Sonneratia apetala Buch.-Ham.) seeds. Bano Biggyan Patrika 18:57-62.
Siddiqi. N.A., M. Shahidullah and MAH. Shahjalal. 1991a. Studies on seeds. germination success and
raised seedlings of Nypa jruticans. Indian Forester 117:553-558.
Siddiqi. NA. M. Shahidullah and M.A.H. Shajalall991b. Studies on the size. viability and germination
of seeds of sundri (Heritiera jomes Buch.-Ham.). Indian J. Forestry 14: 119-124.
Siddiqi. N.A., M.R. Islam. MAS. Khan and M. Shahidullah. 1993. Mangrove nurseries in Bangladesh.
ISME Mangrove ecosystems occasional papers No.1. pp. 1-14.
Sidhu. S.S .• 1975. Structure of epidermis and stomatal apparatus of some mangrove species. In: Walsh.
G.E .• S.c. Snedaker and H.J. Teas (Eds.). Proceedins of the International Symposium on
Biology and Management of Mangroves. Vol. I. University of Florida. Gainesville, pp. 569-
578.
References 341

Silva, C.A.R. and A.A. Mozeto. 1997. Release and retention of phosphorus in mangrove sediments:
Sepetiba Bay. Brazil. In: Kjerfve. B.. L.D. Lacerda and E.H.S. Diop (Eds.). Mangrove
ecosystem studies in Latin America and Africa. UNESCO. Paris. pp. 179-190.
Silva, C.A.R.. L.D. Lacerda and C.E. Rezende. 1990. Metals reservoir in a red mangrove forest.
Biotropica 22:339-345.
Silva. c.A.R.. A.A. Mozeto and A.R.C. Ovalle. 1998. Distribution and fluxes as macrodetritus of
phosphorus in red mangroves. Sepetiba Bay. Brazil. Mangroves and Salt Marshes 2:37-42.
Silva. C.A.R .• L.D. Lacerda. A.R. Ovalle and C.E. Rezende. 1998. The dynamics of heavy metals
through Iitterfall and decomposition in a red mangrove forest. Mangroves and Salt Marshes
2:149-157.
Simberloff. D.S. and E.O. Wilson. 1969. Experimental zoogeography of islands: the colonization of
empty islands. Ecology 50:278-296.
Simberloff. D.S. and E.O. Wilson. 1970. Experimental zoogeograpohy of islands: a two-year record of
colonization. Ecology 51 :934-937.
Simpson. J.H .• W.K. Gong and J.E. Ong. 1997. The determination of the net fluxes from a mangrove
estuary system. Estuaries 20: 103-109.
Sinclair. D.F .• 1985. A new test for species sequencing. Mar. Ecol. Progr. Ser. 3:283-286.
Siqueiros Beltrones. D.A. and E. Sanchez Castrej6n. 1999. Structure of benthic diatom assemblages
from a mangrove environment in a Mexican subtropical lagoon. Biotropica 31 :48-70.
Skelton. N.1. and W.G. Allaway. 1995. Thermo-osmotic gas supply not detected in Avicennia marina
seedlings. Hydrobiologia 195:1-4.
Slim. F.J .• P.M. Gwada. M. Kodjo and M.A. Hemminga. 1996. Biomass and litterfall of Ceriops tagal
and Rhizophora mucronata in the mangrove forest of Gazi Bay. Kenya. Mar. Freshw. Res.
47:999-1007.
Smillie. R.M .• 1984. Cold and heat tolerances of mangroves and seagrass species. Unpubl. Final Report
(MST Grant) No. 81/032IT.
Smillie. R.M. and S.E. Hetherington. 1983. Stress tolerance and stress-induced injury in crop plants
measured by chlorophyll fluorescence in vivo. Chilling. freezing. ice cover. heat and high
light. Plant Physiol. 72:1043-1050.
Smith. A .• 1998. Latitudinal gradients in mangrove litter fall and crab productivity and diversity.
Unpublished Thesis, Southern Cross University. Lismore.
Smith. A.C .• 1973. Angiosperm evolution and the relationship of tbe floras of Africa and America. In:
Meggers. B.1 .• E.S. Ayensu and W.O. Duckworth (Eds.). Tropical forest ecosystems in Africa
and South America: a comparative review, Smithsonian Institution Press. Washington. pp. 49-
61.
Smith. A.H. and F. Berkes. 1993. Community-based use of mangrove resources in St. Lucia. Intern. 1.
Environ. Studies 43:123-131.
Smith. C.J. and R.D. Delaune. 1984. Influence of the rhizophere of Spartina altemiflora Loisel. on the
nitrogen loss from a Louisiana Gulf coast salt marsh. Environ. Exptl. Bot. 24:91-93.
Smith. J.A.c.. M. PoPp. U. Liittge. W.J. Cram. M. Diaz. H. Griffiths. H.S.J. Lee. E. Medina. C. Schafer.
K.-H. Stimmel and B. Thonke. 1989. Ecophysiology of xerophytic and halophytic vegetation of
a coastal alluvial plain in northern Venezuela. VI. Water relations and gas exchange of
mangroves. New Phytol. 111:293-307.
Smith. S.M. and S.C. Snedaker. 1995. Salinity responses of two populations of viviparous Rhizophora
mangle L. seedlings. Biotropica 27:435-440.
Smith. S.M. and S.C. Snedaker. 2000. Hypocotyl function in seedling development of the red mangrove,
Rhizophora mangle L. Biotropica 32:677-685.
Smith. T.J .• 1987a. Seed predation in relation to tree dominance and distribution in mangrove forests.
Ecology 68:266-273.
Smith. T.J., 1987b. Effects of seed predators and light level on the distribution of Avicennia marina
(Forsk.) Vierh. in tropical. tidal forests. Est. Coastal Shelf Sci. 25:43-51.
Smith, T.J .• 1987c. Effect.~ of light and intertidal position on seedling survival and growth in tropical
tidal forests. J. Exp. Mar. BioI. Ecol. 110:133-146.
Smith. T.J .• 1988. Differential distribution between subspecies of the mangrove Ceriops tagal:
competitive interactions along a salinity gradient. Aquat. Bot. 32:79-89.
Smith, T.1., 1992. Forest structure. In: Robertson, A.I. and D.M. Alongi (Eds.). Tropical mangrove
ecosystems. American Geophysical Union. Washington. DC. pp. 101-136.
Smith. T.J. and N.C. Duke. 1987. Physical determinants of inter-estuary variation in mangrove species
richness around the tropical coastline of Australia. J. Biogeogr. 14:9-20.
Smith. T.1 .• K.G. Bolo. S.D. Frusher and RL. Giddins. 1991. Keystone species and mangrove forest
dynamics: the influence of burrowing by crabs on soil nutrient status and forest productivity.
Est. Coa.~t. Shelf Sci. 33:419-432.
Smith. T.1 .• H.T. Chan. c.c. McIvor and M.B. Robblee. 1989. Comparisons of seed predation in
tropical. tidal forests from three continents. Ecology 70: 146-151.
Smith, T.J., M.B. Robblee. H.R. Wanless and T.W. Doyle, 1994. Mangroves, hurricanes and lightning
strikes. BioSci. 44:256-262.
342 Mangrove Ecology. Silviculture and Conservation

Snedaker, S.c., 1982. Mangrove species zonation: why? In: Sen, D.N. and K.S. Rajpurohit (Eds.),
Contribution to the ecology of halophytes, Vol. 2, Dr W. lunk, The Hague, pp. 111-125.
Snedaker, S.C., 1995. Mangroves and climate change in the Florida and Caribbean region: scenarios
and hypotheses. Hydrobiologia 295:43-49.
Snedaker, S.c. and RJ. Araujo, 1998. Stomatal conductance and gas exchange in four species of
Caribbean mangroves exposed to ambient and increased CO2, Mar. Freshw. Res. 49:325-327.
Snedaker, S.c. and P.D. Biber, 1996. Restoration of mangroves in the United States of America: a case
study in Florida. In: Field, C. (Ed.), Restoration of mangrove ecosystems. ISMElITIO,
Okinawa, pp. 170-188.
Snedaker, S.C. and M.S. Brown, 1982. The mystery of Pteroearpetum rhizophorosus. Biotropica
14:157-158.
Snedaker, S.C. and E.J. Lahmann, 1988. Mangrove understorey absence: a consequence of evolution?
J. Trop. Ecol. 4:311-314.
Snedaker, S.c., J.A. Jimenez and M.S. Brown, 1981. Anomalous aerial roots in Avieennia germinans
(L.) L. in Florida and Costa Rica. Bull Mar. Sci. 31 :467-470.
Snedaker, S.C.,1.F. Meeder, M.S. Ross and R.G. Ford, 1994. Discussion of Ellison, Joanna C. and
Stoddart, David R., 1991. Mangrove ecosystem collapse during predicted sea-level rise:
Holocene analogues and implications. Journal of Coastal Research, 7(1), 151-165. J. Coastal
Res. 10:497-498.
Sobrado, M.A, 1999a. Drought effects on photosynthesis of the mangrove, Avieennia germinans. under
contrasting salinities. Trees 13: 125- 130.
Sobrado, M.A., 1999b. Leaf photosynthesis of the mangrove Avieennia germinans as affected by
NaCI. Photosynthetica 36:547-555.
Sobrado, M.A., 2000. Relation of water transport to leaf gas exchange properties in three mangrove
species. Trees 14:258-262.
Sobrado, M.A and M.C. Ball. 1999. Light use in relation to carbon gain in the mangrove, Avieennia
marina. under hypersaline conditions. Aust. 1. PI. Physiol. 26:245-251.
Soltis, D.E., P.S. Soltis, M.W. Chase. M.E. Mort. D.C. Albach, M. Zanis, V. Savolainen, W.H. Hahn,
S.B. Hoot, M.F. Fay, M. Axtell. S.M. Swensen, L.M. Prince, W.J. Kress, K.C. Nixon, and J.S.
Farris, 2000. Angiosperm phylogeny inferred from 18S rDNA, rbeL. and atpB sequences. Bot.
J. Linn. Soc. 133:381-461.
Soto, R., 1988. Geometry, biomass allocation, and leaf life-span of Avieennia germinans (L.)
(Avicenniaceae) along a salinity gradient in Salina~, Puntarenas, Costa Rica. Rev. BioI. Trop.
36:309-323.
Sowunmi, M.A., 1981. Late Quaternary environmental changes in Nigeria. Pollen et Spores 23:125-
148.
Sowunmi, M.A., 1986. Change of vegetation with time. In: Lawson, G.W. (ed.), Plant ecology in West
Africa. John Wiley & Sons, Chichester, pp. 273-307.
Spalding, M., F. Blasco and C. Field (Eds.), 1997. World mangrove atlas. ISMEIITIO, Okinawa.
Specht, R.L., 1970. Vegetation. In: Leeper. G.W. (Ed.), The Australian environment. 4th ed., CSIRO
and Melb. Univ. Press, Melbourne, pp. 44-67.
Specht. R.L .• 1981a. Ecophysiological principles determining the biogeography of major vegetation
formations in Australia. In: Keast. A. (Ed.). Ecological biogeography of Australia. Dr W. Junk.
The Hague. pp. 299-332.
Specht. R.L .• 1981b. Biogeography of halophytic angiosperms (salt-marsh, mangrove and seagrass). In:
Keast, A. (Ed.), Ecological biogeography of Australia. Dr W. Junk. The Hague. pp. 577-589.
Specht, R.L., 1981c. Growth indices - their role in understanding the growth, structure and distribution
of Australian vegetation. Oecologia 50:347-356.
Specht, R.L. and D.G. Morgan, 1981. The balance between foliage projective covers of overstorey
and understorey strata in Australian vegetation. Aust. J. Ecol. 6: 193-202.
Specht. R.L. and A. Specht, 1999. Structure, growth and biodiversity of Australian plant communities.
Oxford University Press, Oxford.
Specht, R.L., R.B. Salt and S.T. Reynolds. 1977. Vegetation in the vicinity of Weipa, North Queensland.
Proc. Roy. Soc. Qld 88:17-38.
Spenceley. A.P .• 1976. Unvegetated saline tidal flats in North Queensland. J. Trop. Geogr. 42:78-85.
Spenceley, AP., 1977. The role of pneumatophores in sedimentary processes. Mar. Geol. 24:31-37.
Spenceley. AP., 1982. Sedimentation patterns in a mangal on Magnetic Island near TownsvilIe.
Northern Queensland, Australia. Singapore J. Trop. Geogr. 3:100-107.
Spenceley. AP.• 1983. Aspects of the development of mangals in the Townsville region, North
Queensland, Australia. In: Teas. H.J. (Ed.). Tasks for vegetation science. Dr W. Junk, The
Hague. pp. 47-56.
Sperry. J.S., M.T. Tyree and l.R. Donnelly, 1988. Vulnerability of xylem to embolism in a mangrove vs
an inland species of Rhizophoraceae. Physiol. Plant. 74:276-283.
Srivastave. P.B.L., S.L. Guan and A. Muktar, 1988. Progress of crop in some Rhizophora stands before
first thinning in Matang Mangrove Reserve of Peninsular Malaysia. Pertanika II :365-374.
Srivastava. P.B.L., G.B. Keong and A. Muktar, 1987. Role of Aerostiehum species in natural
regeneration of Rhizophora species in Malaysia. Trop. Ecol. 28:274-288.
References 343

St. 10hn, H., 1960. Flora of Eniwetok Atoll. Pac. Sci. 14:313-336.
Stace, C.A., 1966. The use of epidennal characters in phylogenetic considerations. New Phytol.
65:304-318.
Stanley, 0.1. and A.K. Hait, 2000. Holocene depositional patterns, neotectonics and Sundarban
mangroves in the western Ganges-Brahmaputra delta. 1. Coastal Res. 16:26-39.
Staples, 0.1., I 980a. Ecology of juvenile and adolescent banana prawns Penaeus merguiensis (de
Man) in a mangrove estuary and adjacent offshore area of the Gulf of Carpentaria. I.
Immigration and settlement of post larvae. Aust. J. Mar. Freshw. Res. 31 :635-652.
Staples, 0.1., 198Ob. Ecology of juvenile and adolescent banana prawns Penaeus merguiensis in a
mangrove estuary and adjacent offshore area of the Gulf of Carpentaria. II. Emigration,
population structure and growth of juveniles. Aust. 1. Mar. Freshw. Res. 31 :653-665.
Steane, O.A., R.W. Scotland, 0.1. Mabberley and R.G. Olmstead, 1999. Molecular systematics of
Clerodendrum (Lamiaceae): ITS sequences and total evidence. Amer. J. Bot. 86:98-107.
Steenis, C.G.G.1. van, 1958. Ecology of mangroves. In: Steenis, C.G.G.1. van (Ed.), Flora Malesiana.
Series I. Spennatophyta. P. Noordhoff, Haarlem, pp. 431-441.
Steenis, C.G.G.1. van, 1962. The distribution of mangrove plant genera and il~ significance for
palaeogeography. Proc. Kon. Net. Amsterdam, Ser. C. 65:164-169.
Steenis, C.G.G.1. van, 1979. Plant geography of ea.~t Malesia. Bot. J. Linn. Soc. 79:97-178.
Steenis, C.G.G.1. van, 1984. Three more mangrove trees growing locally in nature in freshwater.
Blumea 29:395-397.
Steers, J.A., 1977. Physiography. In: Chapman, V.1. (Ed.), Ecosystems of the world. Vol. I. Wet coastal
ecosystems. Elsevier, Amsterdam, pp. 31-60.
Steinke, T.O., 1975. Some factors affecting disperal and establishment of propagules of Avicennia
marina (Forsk.) Vierh. In: Walsh, G.E., S.c. Snedaker and H.1. Teas (Eds.), Proceedings of
the International Symposium on Biology and Management of Mangroves. Vol. 2. University of
Florida, Gainesville, pp. 402-414.
Steinke, T.O., 1979. Apparent transpirational rhythm~ of Avicennia marina at Inhaca Island,
Mozambique. J. S. Afr. Bot. 45:133-138.
Steinke, T.O., 1988. Vegetative and floral phenology of three mangroves in Mgeni estuary. S. Afr. J.
Bot. 54:97-102.
Steinke, T.O., 1999. Mangroves in South African estuaries. In: Allanson, B. and O. Baird (Eds.),
Estuaries of South Africa. Cambridge University Press, Cambridge, pp. 119-140.
Steinke, T.O. and Y. Naidoo, 1991. Respiration and net photosynthesis of cotyledons during
establishment and early growth of propagules of the mangrove, Avicennia marina, at three
temperatures. S. Afr. J. Bot. 57:171-174.
Steinke, T.O. and C.1. Ward, 1989. Some effects of the cyclones Oomoina and Imboa on mangrove
communities in the St Lucia Estuary. S. Afr. J. Bot. 55:340-348.
Steinke, T.O., A. Rajh and A.1. Holland, 1993. The feeding behaviour of the red mangrove crab
Sesarma meinerti de Man, 1887 (Crustacea: Decapoda: Grapsidae) and its effect on the
degradation of mangrove leaf litter. S. Afr. J. Mar. Sci. 13:151-160.
Sternberg, L. da S.L., N. Ish-Shalom-Gordon, M. Ross amd J. O'Brien, 1991. Water relations of coastal
plant communities near ocean/freshwater boundary. Oecologia 88:305-310.
Stevens, G.N. and R.W. Rogers, 1979. The macrolichen flora from the mangroves of Moreton Bay.
Proc. Roy. Soc. Qld. 90:33-49.
Stevens, G.N., 1979. Distribution and related ecology of macrolichens on mangroves on the east
Australian Coast. Lichenologist II :293-305.
Stewart, B., 1995. Rainforest seed ecology. Unpublished Thesis, Southern Cross University, Lismore.
Stewart, G.R. and M. Popp, 1987. The ecophysiology of mangroves. In: Crawford, R.M.M. (Ed.), Plant
life in aquatic and amphibious habitats. Blackwell, Oxford, pp. 333-345.
Stieglitz, T. and P.V. Ridd, 2001. Trapping of mangrove propagules due to density-driven secondary
circulation in the Nonnanby River estuary, NE Australia. Mar. Ecol. Progr. Ser. 211:131-142.
Stocker, G.C., 1976. Report on cyclone damage to natural vegetation in the Darwin area after cyclone
Tracey, 25 Oecember 1974. Forestry and Timber Bureau, Leaftlet No. 127.
Stoddart, O.R., 1980. Mangroves as successional stages inner reefs of the northern Great Barrier Reef.
J. Biogeogr. 7:269-284.
Stoddart, O.R., G.W. Bryan and P.E. Gibbs, 1973. Inland mangroves and water chemistry, Barbuda,
West Indies. J. Nat. Hist. 7:33-46.
Stoner, A.W. and R.J. Zimmennan, 1988. Food pathways associated with penaeid shrimps in a
mangrove-fringed estuary. Fish. Bull. 86:543-552.
Stover, L.E. and P.R. Evans, 1973. Upper Cretaceous-Eocene spore-pollen zonation, offshore
Gippsland Basin, Australia. Geol. Soc. Aust. Spec. Publ. 17:113-126.
Stowe, K.A., 1995. Intracrown distribution of herbivore damage on Laguncularia racemosa in a tidal
influenced riparian habitat. Biotropica 27:509-512.
Suarez, N.• M.A. Sobrado and E. Medina. 1998. Salinity effects on the leaf water relations components
and ion accumulation patterns in Avicennia germinans (L.) seedlings. Oecologia 114:299-304.
344 Mangrove Ecology, Silviculture and Conservation

Sudhersan, C., AboEI-NiI, M. and AI-Ajeel, A., 1998. In vitro tissue culture and differentiation of
Avicennia marina. Abstracts of the International Symposium on Mangrove Ecology & Biology,
Kuwait, April 25-27, 1998, p. 20.
Sugihara, K., N. Hanagata, Z. Dubinsky and I. Karube, 2000. Molecular cloning of cDNA for 33 Kda
oxygen evolving enhancer protein increased by salt stress in the mangrove Bruguiera
gymnarrhiza. Plant Cell Physiol. 41:1279-1285.
Sukardjo, S., 1987. Natural regeneration status of commercial mangrove species (Rhizophora apiculata
and Bruguiera gymna"hiza) in the mangrove forest of Tanjung Bungin, Banyuasin District,
South Sumatra. For. Ecol. Manage. 20:233-252.
Sukardjo.. S. and I. Yamada, 1992. Biomass and productivity of a Rhizophora mucronata Lamarck
plantation in Tritih. Central Java. Indonesia. For. Ecol. Manage. 49:195-209.
Sukardjo. S.• M.H. Azkab and A.V. Toro, 1987. Phenological studies of mangroves associated with
coral reef ecosystem: vegetative phenology and growth of Rhizophora stylosa Griff. seedling
in the coral reef flat at Pari Island. Biotrop Spec. Publ. 29:39-55.
Sun. M.• K.C. Wong and S.Y. Lee. 1998. Reproductive biology and population genetic structure of
Kandelia candel (Rhizophoraceae), a viviparous mangrove species. Amer. J. Bot. 85:1631-
1637.
Sun, Q. and P. Lin, 1997. Wood structure of Aegiceras comiculatum and il~ ecological adaptations to
salinities. Hydrobiologia 352:61-66.
Sussex. I., 1975. Growth and metabolism of the embryo and attached seedling of the viviparous
mangrove Rhizophora mangle. Amer. J. Bot. 62:948-953.
Suzuki, E. and H. Tagawa, 1983. Biomass ofa mangrove forest and a sedge marsh in Ishigaki Island,
South Japan. Japan. J. Ecol. 33:231-234.
Suzuki, K.• 1979. Vegetation of the Ryukyu Islands. Bull. Inst. Env. Sci. Technol. Yokohama Natl.
University 5:87-150.
Suzuki, K. and Y. Mochida, 1982. Phytosociological studies on the mangrove forest in the River Riko,
East Kalimantan, Indonesia. Bull. Inst. Env. Sci. Technol. Yokohama Natl. University 8:295-
318.
Suzuki, K. and P. Saenger, 1996. A phytosociological study of mangrove vegetation in Australia with a
latitudinal comparison of East Asia. Mangrove Sci. 1:9-27.
Sweatman, J., 1846. The journal of John Sweatman - a nineteenth century surveying voyage in north
Australia and Torres Strait. Reprinted in Allen. J. and P. Corris (Eds.). 1977, Queensland
University Press, Brisbane.
Taira, K.• 2001. Later quaternary palaeocenography in eastern Asia. J. Coastal Res. 17:114-117.
Takemura, T., N. Hanagata, K. Sugihara, K. Baba, I. Karube and Z. Dubinsky. 2000. Physiological and
biochemical responses to saIt stress in the mangrove Bruguiera gymnorrhiza. Aquat. Bot.
68:15-28.
Takhtajan, A.L., 1980. Outline of the classification of the flowering plants (Magnoliophyta). Bot. Rev.
46:225-359.
Tam, N.F.Y. and Y.S. Wong, 1995. Spatial and temporal variations of heavy metal contamination in
sediments of a mangrove swamp in Hong Kong. Mar. Poll. Bull. 31:254-261.
Tam, N.F.Y. and Y.S. Wong. 1997. Accumulation and distribution of heavy metals in a simulated
mangrove system treated with sewage. Hydrobiologia 352:67-75.
Tam. N.F.Y. and Y.S. Wong. 1999. Mangrove soils in removing pollutanl~ from municipal wastewater
of different salinities. J. Environ. Qual. 28:556-564.
Tam, N.F.Y .• Y.S. Wong. C.Y. Lan and G.Z. Chen. 1995. Community structure and standing crop
biomass of a mangrove forest in the Futian Nature Reserve. Shenzhen. China. Hydrobiologia
295: 193-201.
Tam. N.F.Y .• W. Yuk-Shan. C.Y. Lu and R. Berry. 1997. Mapping and characterization of mangrove
plant communities in Hong Kong. Hydrobiologia 352:25-37.
Tamai. S. and P. lampa. 1988. Establishment and growth of mangrove seedlings in mangrove forests of
southern Thailand. Ecol. Res. 3:227-238.
Tanai. T.. 1972. Tertiary history of vegetation in Japan. In: Graham. A. (Ed.). Floristics and
palaeofloristics of Asia and eastern North America. Elsevier. Amsterdam. pp. 235-255.
Tanaka. S .• T. Shinjo and M. Hoshino. 1994. Nutritive value and in vitro digestibility of some mangrove
species. In: Proceedings of VII Pacific Science Inter-Congress Mangrove Session. ISMENII
Pacific Science Inter-Congress. Okinawa. pp. 97-100.
Tang. H.T., H.B.HJ. Abu Hassan and E.K. Cheah. 1984. Mangrove forests of Peninsular Malaysia - a
review of management and research objectives and priorities. In: Soepadmo. E .• A.N. Rao and
OJ. Mcintosh (Eds.). Proceedings of the Asian Symposium Mangrove Environments -
Research and Management. UNESCO. pp. 796-808.
Tastet. J.P.• 1981. Morphologie des littoraux sedimentaires liees aux variations du niveau de la mer:
exemple du golfe de Guinee. Oceanis 7:455-472.
Tayab. M.R .• 1998. Conservation of mangrove (Avicennia marina) in Ras Laffan Industrial City (a case
study). In: Regional Conference on the Marine Environment of the Gulf. Doha. Qatar. 12-15
December. 1998. Abstracts p. 148.
References 345

Taylor, B.W., 1959. The classification of lowland swamp communities in north-eastem Papua. Ecology
40:103-111.
Taylor, F.J., 1919. Rhizophora in the Society Islands. Pac. Sci. 33:113-116.
Taylor, W.R., 1950. Plants of Bikini and other Northern Marshall Islands. Univ. Mich. Stud. Sci. Ser.
18:1-227.
Teas, H. J., 1911. Ecology and restoration of mangrove shorelines in Florida. Environ. Con.'lCrv. 4:51-
58.
Teas, H.J., 1919. Silviculture with saline water. In: Hollaender, A. (Ed.), The biosa\ine concept.
Plenum, New York, pp. 117-161.
Teas, H.1., W. Jurgens and M.C. Kimball, 1915. Plantings of red mangroves (Rhizophora mangle) in
Charlotte and St. Lucie counties, Florida. In: Lewis, R.R. (Ed.), Proceedings of the second
annual conference on the restoration of coastal vegetationb in Florida. Hillsborough
Community College, Tampa, pp. 132-161.
Teas, H.1., E.O. Duerr and J.R. Wilcox, 1981. Effects of South Louisiana crude oil and dispersants on
Rhizophora mangroves. Mar. Poll. Bull. 18:122-124.
Terchunian, A., V. Klemas, A. Segovia, A. Alvarez, B. Vascones and L. Guerrero, 1986. Mangrove
mapping in Ecuador: the impact of shrimp pond construction. Environmental Management
10:345-350.
TeStrake, D., R.B. Lassiter, J.A. Lassiter and D.A. Breil, 1986. Bryophytes from mangroves and
adjacent shoreline plant communities of Tampa Bay, Florida. Fla. Sci. 49:31-39.
Thangham, T.S. and K. Kathiresan, 1994. Mosquito larvicidal activity of Rhizophora apiculata Blume.
Int. J. Pharmacognosy 32:33-36.
Thayer, G.W., D.R. Colby and W.F. HeUler, 1981. Utilization of the red mangrove prop root habitat by
fishes in south Florioda. Mar. Ecol. 35:25-38.
Thibodeau, F.R. and N.H. Nickerson, 1986. Differential oxidation of mangrove substrate by Avicennia
gemnnans and Rhizophora mangle. Amer. J. Bot. 13:512-516.
Tho, Y.P., 1982. Gap formation by the termite Microcerotermes dubius in lowland forest of Peninsular
Malaysia. Malay. For. 45:184-192.
Thom, B.G., 1961. Mangrove ecology and deltaic geomorphology, Tabasco, Mexico. J. Ecol. 55:301-
343.
Thom, B.G., 1915. Mangrove ecology from a geomorphic viewpoint. In: Walsh, G.E., S.C. Snedaker
and H.1. Teas (Eds.), Proceedings of the International Symposium on Biology and
Management of Mangroves, Vol. 2. University of Florida, Gainesville, pp.469-81.
Thom, B.G., 1982. Mangrove ecology - a geomorphological perspective. In: Oough, B.F. (Ed.),
Mangrove ecosystems in Australia. Australian National University Press, Canberra, pp. 3-17.
Thom, B.G., 1984. Coastal landforms and geomorphic processes. In: Snedaker, S.C. and J.G. Snedaker
(Eds.), The mangrove ecosystem: research methods. UNESCO, Paris, pp. 3-11.
Thom, B.G., L.D. Wright and J.M. Coleman, 1915. Mangrove ecology and deltaic-estuwarine
geomorphology, Cambridge Gulf-Ord River, Western Australia. J. Ecol. 63:203-232.
Thomas, C. and J.E. Ong, 1984. Effect of heavy metals zinc and lead on Rhizophora mucronata Lam
and Avicennia alba BI. seedlings. In: Soepadmo, E., A.N. Rao and OJ. McIntosh (Eds.),
Proceedings of the Asian Symposium Mangrove Environments - Research and Management,
UNESCO, pp. 568-574.
Thompson, J.N., 1981. Reversed animal-plant interactions: the evolution of insectivorous and ant-fed
plants. BioI. J. Linn. Soc. 16:141-155.
Thorhaug, A., D. Segar and M.A. Roessler, 1913. Impact of a power plant on a subtropical estuarine
environment. Mar. Poll. Bull. 4:166-169.
Thome, R.F., 1973. Floristic relationships between tropical Africa and tropical America. In: Meggers,
B.J., E.S. Ayensu and W.O. Duckworth (Eds.), Tropical forest ecosystems in Africa and South
America: a comparative review. Smithsonian Institution Press, Washington, pp. 21-41.
Thome, R.F., 1992. An updated phylogenetic classification of the flowering plant.~. Aliso 13:365-389.
Thurairaja, V., 1994. Coastal resource development options in the southeast Asia and the Pacific
regions: economic valuation methodologies and applications in mangrove development.
Maritime Studies 19:1-13.
Tissot. C., M.R. Djuwansah and C. Marius, 1988. Evolution de la mangrove en Guyane au cours de
I'Holocene: ~tude palynologique. Inst. fro Pondich~ry, Trav. Sec. Sci. Tech. 25:125-137.
Tobey, J., J. Clay and P. Vergne, 1998. The economic, environmental and social impacts of shrimp
farming in Latin America. Coastal Resources Center, University of Rhode Island.
Toledo, G., Y. Bashan and A. Soeldner 1995. Cyanobacteria and black mangroves in northwestern
Mexico: colonization, and diurnal and seasonal nitrogen fixation on aerial roots. Can. J.
Microbiol. 41 :999-1 011.
Toledo, G., A. Rojas and Y. Bashan, 2001. Monitoring of black mangrove restoration with nursery-
reared seedlings on an arid coastal lagoon. Hydrobiologia 444:101-109.
Tomlinson, P.B., 1971. The shoot apex and its dichotomous branching in the Nypa palm. Ann. Bot.
35:865-819.
Tomlinson, P.B., 1986. The Botany of Mangroves. Cambridge University Press, Cambridge.
346 Mangrove Ecology, Silviculture and Conservation

Tomlinson, P.B. and P.A. Cox, 2000. Systematic and functional anatomy of seedlings in mangrove
Rhizophoraceae: vivipary explained? Bot. 1. Linn. Soc. 134:215-231.
Tomlinson, P.B. and D.W. Wheat, 1979. Bijugate pyllotaxis in Rhizophora (Rhizophoraceae). Bot. 1.
Linn. Soc. 78:317-321.
Tomlinson, P.B., R.B. Primack and 1.S. Bunt, 1979. Preliminary observations on floral biology in
mangrove Rhizophoraceae. Biotropica 11:256-277.
Tomlinson, P.B., J.S. Bunt, R.B. Primack and N.C. Duke, 1978. Lumnitzera rosea (Combretaceae) - its
status and floral morphology. J. Arnold Arbor. 59:342-351.
Tovilla, H.C. and E.G. Da La Lanza, 2000. Ecologia, producci6n y aprovechamiento del mangle
Conocarpus erectus L., en Barra de Tecoanapa Guerrero, Mexico. Biotropica 31: 121-134.
Trochain, T.L. and L. Dulau, 1942. Quelques particularites d'Avicennia nitida (Verbenacees) de la
mangrove ouest-africaine. Bull. Soc. Hist. Nat. Toulouse 77:271-281.
Trott, L.A. and D.M. Alongi, 2000. The impact of shrimp effluent on water quality and phytoplankton
biomass in a tropical mangrove estuary. Mar. Poll. Bull. 40:947-951.
Tsuchi, R., 1992. Pacific Neogene climatic optimum and accelerated biotic evolution in time and space.
In: Tsuchi, R. and J. Ingle (Eds.), Pacific Neogene environments, evolution and events.
University of Tokyo Press, Tokyo, pp. 237-350.
Tuan, M.S., I. Ninomiya and K. Ogino, 1995. Salt uptake and excretion in the mangrove, Avicennia
marino (Forsk.) Vierh. Tropics 5:69-79.
Tiiffers, A.V., G. Naidoo and DJ. von Willert. 1999. The contribution of leaf angle to photoprotection
in the mangroves Avicennia marino (Forssk.) Vierh. and Bruguiera gymnorrhiza (L.) Lam.
under field conditions in South Africa. Flora 194:267-275.
Turner, I.M., W.K. Gong, J.E. Ong, J.S. Bujang and T. Kohyama 1995. The architecture and allometry
of mangrove saplings. Functional Ecology 9:205-212.
Turner, R.E. and R.R. Lewis. 1997. Hydrologic restoration of coastal wetlands. Wetlands Ecology and
Management 4:65-72.
Twilley, R.R., 1995. Properties of mangrove ecosystems related to the energy signature of coastal
environments. In: Hall, C.A.S. (Ed.), Maximum power: the ideas and applications of H.T.
Odum. University of Colorado Press. Boulder. pp. 43-62.
Twilley, R.R., R.H. Chen and T. Hargis, 1992. Carbon sinks in mangroves and their implications to
carbon budget of tropical coastal ecosystems. Water, Air and Soil Pollut. 64:265-288.
Twilley, R.R., A.E. Lugo and C. Patterson-Zucca, 1986. Litter production and turnover in basin
mangrove forests in south-west Florida. Ecology 67:670-683.
Twilley, R.R .• R.R. Gottfried, V.H. Rivera-Monroy, W. Zhang. M.M. Armijos and A. Bodero, 1998. An
approach and preliminary model of integrating ecological and economic constraints of
environmental quality in the Guayas River estuary. Ecuador. Environmental Science and
Policy 1:271-288.
Twilley, R.R., M. POlO, V.H. Garcia, V.H. Rivera-Monroy, R. Zambrano and A. Bodera, 1997. Litter
dynamics in riverine mangrove forests in the Guayas River estuary, Ecuador. Oecologia
Ill:I09-122.
Uhl, N.W., 1972. Inflorence and flower structure in Nypafruiticans (Palmae). Amer. 1. Bot. 59:729-
743.
Ukpong, I.E., 1989. An evaluation of the ecological status and environmental relations of mangrove
swamps in south-eastern Nigeria. Unpublished Thesis, University of Ibadan, lbadan.
Ukpong, I.E., 1992. The structure and soil relations of A vicennia mangrove swamps in southeastern
Nigeria. Trop. Ecol. 33:1-16.
Ukpong, I.E., 1994. Soil-vegetation interrelationships of mangrove swamps as revealed by multivariate
analyses. Geoderma 64:167-181.
Ukpong, I.E., 1995. Vegetation and soil acidity of a mangrove swamp in southeastern Nigeria. Soil Use
Managt.II:141-144.
Ukpong, I.E., 1997. Vegetation and its relation to soil nutrient and salinity in the Calabar mangrove
swamp, Nigeria. Mangroves and Salt Marshes 1:211-218.
Ukpong, I.E., 2000. Ecological classification of Nigerian mangroves using soil nutrient gradient
analysis. Wetlands Ecology and Management 8:263-272.
UNEP, 1994. Assessment and monitoring of climatic change impacts on mangrove ecosystems. UNEP
Regional Seas Reports and Studies No. 154, pp. 62.
UNESCO 1979. Map of the world distribution of arid regions: explanatory note. MAB Technical Notes
7, pp. 1-54.
Untawale, A.G., 1996. Restoration of mangroves along the central west coast of India. In: Field, C.
(Ed.), Restoration of mangrove ecosystems. ISMElITTO, Okinawa, pp. 111-125.
Uphof,l.C.T., 1941. Halophytes. Bot. Rev. 7:1-58.
Valiela, I., 1984. Marine ecological processes. Springer-Verlag, New York.
Valiela, I. and S. Vince, 1976. Green borders of the sea. Oceanus 19:10-17.
Van der Hammen, T., 1972. Historia de la vegetaci6n y el medio ambiente del norte Sudamericano.
Mem. Simp. I Congr. Latinoamerican6 de Botanica, Mexico. pp. 119-134.
Van der Valk, A.G. and P.M. Attiwill, 1984. Acetylene reduction in an Avicennia marina community in
Southern Australia. Aust. J. Bot. 32:157-164.
References 347

Van Hove, C., 1976. Bacterial leaf symbiosis and nitrogen fixation. In: Nutman, p.s. (Ed.), Symbiotic
nitrogen fixation in plants. Cambridge University Press, London, pp. 551-560.
Van Speybroeck, D., 1992. Regeneration strategy of mangroves along the Kenya coast: a first
approach. Hydrobiologia 247:243-251.
Van Steveninck, RF.M., W.D. Armstrong, P.D. Peters and T.A. Hall, 1976. Ultrastructural localization
of ions: Ill. Distribution of chloride in mesophyll cells of mangrove (Aegiceras comiculatum
Blanco). Aust. 1. PI. Physiol. 3:367-376.
Vance, DJ., M.D.E. Haywood and DJ. Staples, 1990. Use of a mangrove estuary as a nursery area by
postlarval and juvenile banana prawns, Penaeus merguiensis de Man, in northero Australia.
Estuar. Coastal Shelf Sci. 31 :689-701.
Vannini, M., A. Oluoch and K. Ruwa, 1997. Tree-climbing mangrove decapods of Kenyan mangroves.
In: Kjerfve, B., L.D. Lacerda and S. Diop (Eds.), Mangrove ecosystem studies in Latin
America and Africa. UNESCO, Paris, pp. 325-338.
Vasil'eva, V.A. and B.R. Vasil'ev, 1988. The morphology and anatomy of the leaves of some mangrove
species. Vestn. Leningr. Univ. BioI. 1:24-32.
Vazquez, P., G. Holguin, M.E. Puente, A. Lopez-Cortes and Y. Bashan, 2000. Phosphate-solubilizing
microorganisms associated with the rhizosphere of mangroves in a semiarid coastal lagoon.
BioI. Fert. Soils 30:460-468.
Vegas Vilamibia, T., 2000. Zonation pattern of an isolated mangrove community at Playa Medina,
Venezuela. Wetlands Ecology and Management 8:9-17.
Venkateswarlu, J. and RS.P. Rao, 1964. The wood anatomy and taxonomic position of Sonneratiaceae.
Current Science 33:6-9.
Venkateswarlu, J. and RS.P. Rao, 1975. A contribution to the embryology of the tribe Hippomaneae of
the Euphorbiaceae. J. Indian Bot. Soc. 54:98-103.
Vermeij, GJ. and R Dudley, 2000. Why are there so few evolutionary transitions between aquatic and
terrestrial ecosystems? BioI. J. Linn. Soc. 70:541-554.
Verstraete, I.M., 1989. Le niveau de la mer Ie long des cotes de l'Afrique de l'Ouest et 11 l'equateur.
Hausse probable du niveau marin 11 I'echelle seculaire. COMARAF Publ. Ser. documentaire
No.4, Dakar, 5-43.
Viboth, H. and D. Ashwell, 1996. Ecology and management of mangroves in Cambodia. In: Khemnark,
C. (Ed.), Proceedings of Ecotone IV: ecology and management of mangrove restoration and
regeneration in east and southeast Asia. Surat Thani, Thailand, pp. 37-47.
Vidy, G., 2000. Estuarine and mangrove systems and the nursery concept: which is which? The case of
the Sine Saloum system (Senegal). Wetlands Ecology and Management 8:37-51.
Villiers, J-F., 1973. Etude floristique et phytosociologique d'une mangrove atlantique sur substrat
rocheux du littoral gabonais. Ann. Fac. Sci. Cameroun 14:3-46.
Vliet, GJ.C.M., van. 1976. Wood anatomy of Rhizophoraceae. Leiden Bot. Ser. 3:20-75.
Vliet, GJ.C.M., van, 1979. Wood anatomy of the Combretaceae. Blumea 25:141-223.
Vousden, D.H.P. and A.RG. Price, 1985. Bridge over fragile waters. New Scientist 1451:33-35.
Wada, K. and D. Wowor, 1988. Biology and behavior of benthic fauna. crabs and molluscs. 4. P1ant-
animal relationship of mangrove. foraging on Sonneratia pneumatophore by ocypodid crabs.
In: Ogino. K. and M. Chihara (Eds.). Biological system of mangroves - a report of east
Indonesian mangrove expedition. 1986. Ehime University, Ehime. pp. 67-83.
Wada. K. and D. Wowor. 1990. Foraging on mangrove pneumatophores by ocypodid crabs. J. Exp.
Mar. BioI. Ecol. 134:89-100.
Wafar. S .• A.G. Untawale and M. Wafar, 1997. Litter fall and energy flux in a mangrove ecosystem.
Est. Coastal Shelf Sci. 44: 111-124.
Waisel. Y .• A. Eshel and M. Agami. 1986. Salt balance of leaves of the mangrove Avicennia marina.
Physiol. Plant. 67:67-72.
Walker. D. (Ed.), 1972. Bridge and barrier - the natural and cultural history of Torres Strait. ANU
Press. Canberra.
Walsh. G.E.• 1967. An ecological study of a Hawaiian mangrove swamp. In: Lauff, G.H. (Ed.).
Estuaries. Amer. Assoc. Adv. Sci. Publ. No. 83, pp. 420-421.
Walsh. G.E .• 1974. Mangroves: a review. In: Reimhold. RJ. and W.H. Queen (Eds.). Ecology of
halophytes. Academic Press. New York. pp.51-174.
Walsh. G.E .• 1977. Exploitation of mangal. In: Chapman, V.I. (Ed.), Wet coastal ecosystems. Elsevier
Scientific Publ. Co., Amsterdam. pp. 347-362.
Walsh. G.E .• K.A. Ainsworth and R. Rigby, 1979. Resistance ofred mangrove (Rhizophora mangle L.)
seedlings to lead. cadmium and mercury. Biotropica II :22-27.
Walters. B.B.. 1997. Human ecological questions for tropical restoration: experiences from planting
native upland trees and mangroves in the Philippines. For. Ecol. Manage. 99:275-290.
Walters, B.B .• 2000. Local mangrove planting in the Philippines: are fisherfolk and fishpond owners
effective restorationists? Restoration Ecol. 8:237-246.
Walter. H. and M. Steiner. 1936. Die Okologie der Ost-Afrikanischen Mangroven. Zeitschrift fUr
Botanik 30:65-193.
Wang. W.Q. and P. Lin. 1999. Transfer of salt and nutrients in Bruguiera gymnorrhiza leaves during
development and senescence. Mangroves and Salt Marshes 3:1-7.
348 Mangrove Ecology, Silviculture and Conservation

Wardrop, J.A., A.J. Butler and J.E. Johnson, 1987. A field study of the toxicity of two oils and a
dispersant to the mangrove Avicennia marina. Mar. BioI. 96:151-156.
Warren, J.H. and A.J. Underwood, 1986. Effects of burrowing crabs on the topography of mangrove
swamps in New South Wales, Australia. J. Exp. Mar. BioI. Ecol. 102:223-236.
Watson, J.G., 1928. Mangrove forests of the Malay Peninsula. Malayan For. Rec. 6:1-275.
Weber, H.E., J. Moravec and J.-P. Theurillat, 2000. International code of phytosociological
nomenclature. 3M Edition. J. Veg. Sci. 11:739-768.
Weinstock, J.A., 1993. Rhizoplwra mangrove agroforestry. Econ. Bot. 48:210-213.
Weiper, M., 1995. Physiologische and strukturelle Untersuchungen zur Salzregulation bei Mangroven.
Unpublished Thesis, Westflihlischen Wilhelms-Universitlit, Miinster.
Weiss, H., 1966. Aper~u preliminaire sur une mangrove naturelle a l'interieure des terres au sud de
Tulear. Recueil des Travaux de la Station marine d'Endoume, Faculte des sciences de
Marseille. Fasc. H. Sene Suppl. 5:175-178.
Wells, A.G., 1982. Mangrove vegetation of northern Australia. In: Clough, B.F. (Ed.), Mangrove
ecosystems in Australia - structure, function and management. ANU Press, Canberra, pp. 57-
58.
Wells, A.G., 1983. Distribution of mangrove species in Australia. In: Teas, H.J. (Ed.), Biology and
ecology of mangroves. Tasks for vegetation science, Vol. 8. Dr W. Junk, The Hague, pp. 57-
76.
Wells, J.T. and J.M. Coleman, 1981. Periodic mudflat progradation, northeastern coast of South
America: a hypothesis. 1. Sediment. Petrol. 51:1069-1075.
Werner, A. and R. Stelzer, 1990. Physiological responses of the mangrove Rhizophora mangle grown in
the absence and presence of NaO. Plant, Cell and Environ. 13:243-255.
Weste, G., D. Cahill and D.J. Stamps, 1982. Mangrove dieback in North Queensland, Australia. Trans.
Brit. Mycol. Soc. 79:165-167.
Wester, L., 1981. Introduction and spread of mangroves in the Hawaiian Islands. Assoc. Pac. Coast
Geogr. Yearbook 43:125-137.
Westoby, M., 1984. The self-thinning rule. Adv. Ecol. Res. 14:167-225.
White, J., 1981. The allometric interpretation of the self-thinning rule. J. Theoret. BioI. 89:475-500.
Whitmore, T.C., 1989. Canopy gaps and the two major groups of forest trees. Ecology 70:536-538.
Whitten, A.J. and S.J. Damanik, 1986. Mass defoliation of mangroves in Sumatra, Indonesia. Biotropica
18:176.
Wier, A.M., T.A. Tattar and E.J. Klekowski, 2000. Disease of red mangrove (Rhizophora mangle) in
southwest Puerto Rico caused by Cytospora rhizoplwrae. Biotropica 32:299-306.
Wightman, G.M., 1989. Mangroves of the Northern Territory. Northern Territory Botanical Bull. No.7,
pp 1-130.
Wilcox, B.H.R., 1985. Angiosperm flora of the mangrove ecosystem of the Niger Delta. In: Wilcox,
B.H.R. and C.P. Powell (Eds.), The mangrove ecosystem of the Niger delta. University of Port
Harcourt, Port Harcourt, pp. 34-44.
Wilkinson, H.P., 1981. The anatomy of the hypocotyls of Ceriops Amott (Rhizophoraceae), Recent and
fossil. Bot. J. Linn, Soc. 82:139-164.
Williams, W.T., J.S. Bunt and H.J. Clay, 1991. Yet another method of species-sequencing. Mar. Ecol.
Prog. Ser. 72:283-287.
Wilson, E.O. and B. Holldobler, 1980. Sex differences in cooperative silk-spinning by weaver ant
larvae. Proc. Natl. Acad. Sci. 77:2343-2347.
Wilson, E.O. and D.S. Simberloff, 1969. Experimental zoogeography of islands. Defaunation and
monitoring techniques. Ecology 50:267-278.
Wilson, J.B. and A.D.Q. Agnew, 1992. Positive-feedback switches in plant communities. Adv. Ecol.
Res. 23:263-336.
Windolf, J., 1989. Bryophytes in a sub-tropical mangrove community. Austrobaileya 3:\03-\07.
Wise, R.R. and A.M. Juncosa, 1989. Ultrastructure of the transfer tissues during viviparous seedling
development in Rhizoplwra mangle (Rhizophoraceae). Amer. J. Bot. 76:1286-1289.
Wium-Andersen, S., 198\. Seasonal growth of mangrove trees in southern Thailand. III. Phenology of
Rhizoplwra inucronata Lamk. and Scyphiplwra hydrophyllacea Gaertn. Aquat. Bot. 10:371-
376.
Wium-Andersen, S. and B. Christensen, 1978. Seasonal growth of mangrove trees in southern Thailand.
II. Phenology of Bruguiera cylindrica, Ceriops taga~ Lumnitzera littorea and A vicennia marina.
Aquat. Bot. 5:383-390.
Wokoma, S.A. and B.I. Ezenwa, 1982. Construction offish ponds in the mangrove swamps of the Niger
delta. Technical Paper No.9, Nigerian Institute for Oceanography and Marine Research.
Victoria Island, Lagos.
Wolanski, E. and B. Cassagne, 2000. Salinity intrusion and rice farming in the mangrove-fringed
Konkoure River delta, Guinea. Wetlands Ecology and Management 8:29-36.
Wolanski, E. and P. Collis, 1976. Aspects of aquatic ecology of the Hawkesbury esturary. I.
Hydrodynamical processes. Aust. J. Mar. Freshw. Res. 27:565-582.
Wolanski, E. and R. Gardiner, 198\. Flushing of salt from mangrove swamps. Aust. J. Mar. Freshw.
Res. 32:681-683.
References 349

Wolanski, E., Y. Mazda and P. Ridd, 1992. Mangrove hydrodynamics. In: Robertson, A.1. and D.M.
Alongi (Eds.), Tropical mangrove ecosystems. American Geophysical Union, Washington,
DC, pp. 43-62.
Wolanski, E., S. Spagnol and T. Ayukai, 1998. Field and model studies of the fate of particulate carbon
in mangrove-fringed Hinchinbrook Channel, Australia. Mangroves and Salt Marshes 2:205-
221.
Wolanski, E., S. Spagnol and E.B. Lim, 1997. The importance of mangrove flocs in sheltering seagrass
in turbid coastal waters. Mangroves and Salt Marshes 1: 187-191.
Womersley, J.S., 1975. Management of mangrove forests: utilisation versus conservation with special
reference to the forests of the Papuan Gulf. In: Walsh, G.E., S.c. Snedaker and H. J. Teas
(Eds.), Proceedings of the International Symposium on Biology and Management of
Mangroves, Vol. 2. University of Florida, Gainesville, pp. 732-741.
Wong, c.H. and J.E. Ong, 1984. Electron microscopy study on the salt glands of Acanthus ilicifolius L.
In: Soepadmo, E., A.N. Rao and D.J. Macintosh (Eds.), Proceedings of the Asian symposium
on mangrove environment - research and management. Unversity of Malaya, Kuala Lumpur,
pp. 172-182.
Wong, Y.S., C.Y. Lan, G.Z. Chen, S.H. Li, X.R. Chen, Z.P. Liu and N.F.Y. Tam, 1995. Effect of
wastewater discharge on nutrient contamination of mangrove soils and plant~. Hydrobiologia
295:243-254.
Wong, Y.S., N.F.Y. Tam and c.y. Lan, 1997a. Mangrove wetlands as wastewater treatment facility: a
field trial. Hydrobiologia 352:49-59.
Wong, Y.S., N.F.Y. Tam, G.Z. Chen and H. Ma, 1997b. Response of Aegiceras comiculatum to
synthetic sewage under simulated todal conditions. Hydrobiologia 352:89-96.
Wongkaew, W. and S. Techapinyawat. 1996. Growth inhibitory substances from the leaves of Nypa
palm. In: Khemnark, C. (Ed.), Proceedings of Ecotone IV: ecology and management of
mangrove restoration and regeneration in east and southeast Asia. Surat Thani, Thailand, pp.
254-268.
Woodroffe, C.D., 1981. Mangrove swamp stratigraphy and Holocene transgression, Grand Cayman
Island, West Indies. Mar. Geol. 41:271-294.
Woodroffe, C.D., 1983. Development of mangrove forests from a geological perspective. In: Teas,
H.J. (Ed.), Tasks for vegetation science. Vol. 8. Dr W. Junk. The Hague, pp. 1-17.
Woodroffe, C.D., 1985a. Studies of a mangrove basin, Tuff Crater, New Zealand. I. Mangrove
biomass and production of detritus. Est. Coastal Shelf Sci. 20:431-446.
Woodroffe, C.D., 1987. Pacific Island mangroves: distribution and environmental settings. Pac. Sci.
41:166-185.
Woodroffe, C.D., 1988. Relict mangrove stand on last interglacial terrace, Christmas Island, Indian
Ocean. J. Trop. Ecol. 4:1-7.
Woodroffe, C.D., 1990. The impact of sea level rise on mangrove shorelines. Progress in Physical
Geography 14:483-520.
Woodroffe, C.D., 1992. Mangrove sediments and geomorphology. In: Robertson, A.l. and D.M. Alongi
(Eds.), Tropical mangrove ecosystems. American Geophysical Union, Washington, DC, pp. 7-
41.
Woodroffe, C.D., 1993a. Late Quaternary evolution of coastal and lowland riverine plains of Southeast
Asia and northern Australia: an overview. Sedimentary Geol. 83:163-175.
Woodroffe, C.D., 1993b. Sea-level. Progress in Physical Geography 17:359-368.
Woodroffe, C.D. and 1. Grindrod, 1991. Mangrove biogeography: the role of Quaternary
environmental and sea-level change. J. Biogeogr. 18:479-492.
Woodroffe, C.D. and T.J. Moss, 1984. Litter Fall beneath Rhizophora stylosa Griff., Vaitupu, Tuvalu,
South Pacific. Aquat. Bot. 18:249-255.
Woodroffe, C.D., Thom, B.G., and ChappeU, 1., 1985. Development of widespread mangrove swamps
in mid-Holocene times in northern Australia. Nature 317:711-7\3.
Wright,l.J. and K. Cannon, 2001. Relationships between leaf lifespan and structural defences in a low-
nutrient, sclerophyll flora. Functional Ecology 15:351-359.
Wright, L.D., 1978. River deltas. In: Davis, R.A. (Ed.), Coastal sedimentary environments. Springer
Verlag, New York, pp. 5-68.
Wright, L.D., 1.M. Coleman and M.W. Erickson, 1974. Analysis of major systems and their deltas:
morphologic and process comparisons. Coastal Studies Institute, Louisiana State University,
Tech. Report 156.
Wylie, R.B., 1949. Differences in foliar organisation among leaves from four locations in crown of an
isolated tree (Acer plantanoides). Proc. Iowa Acad. Sci. 56:189-198.
Wylie, R.B., 1954. Leaf organization of some wood dicotyledons from New Zealand. Amer. J. Bot.
41:186-191.
Wyn Jones, R.B. and R. Storey, 1981. Betaines. In: Paleg, L.G. and D. Aspinall (Eds.). Physiology and
biochemistry of drought resistance in plant~. Academic Press, Sydney, pp. 171-204.
Yabuki, K., Y. Kitaya and J. Sugi, 1985. Studies on the function of the mangrove pneumatophore. In:
Sugi, J. (Ed.), Studies on the mangrove ecosystem. Nodai Research Institute, Tokyo University
of Agriculture, Tokyo, pp. 76-79.
350 Mangrove Ecology, Silviculture and Conservation

Yamashiro, M., 1961. Ecological study of Kandelia candel (L.) Druce, with special reference to the
structure and falling of the seedlings. Hikobia 2:209-214.
Yanez-Arancibia, A.Y., A.LL. Dominguez and J.W. Dya, 1993. Interaction between mangrove and
seagrass habitats mediated by estuarine nekton assemblages: coupling of primary and
secondary production. Hydrobiologia 254:1-12.
yanez-Espinosa, L., T. Terrazas and L. L6pez-Mata, 2001. Effects of flooding on wood and bark
anatomy of four species in a mangrove forest community. Trees 15:91-97.
Yipp, M.W., C.H. Hau and G. Walthew, 1995. Conservation evaluation of nine Hong Kong mangals.
Hydrobiologia 295:323-333.
Young, B.M. and L.E. Harvey, 1996. A spatial analysis of the relationship between mangrove
(Avicennia marina var. australasica) physiognomy and sediment accretion in the Hauraki
Plains, New Zealand. Est. Coastal Shelf Sci. 42:231-246.
Youssef, T., 1995. Ecophysiology of water logging in mangroves. Unpublished Thesis, Southern Cross
University, Lismore.
Youssef, T., 1997. Approaches in mangrove planting: some options for Darwin mangroves. In: Hanley,
J.R., G. Caswell, D. Megirian and H.K. Larson (Eds.), Proceedings of the Sixth International
Marine Biological Workshop - The marine flora and fauna of Darwin Harbour, Northern
Territory, Australia. Northern Territory Museum of Arts and Sciences, Darwin, pp. 321-333.
Youssef, T. and A. Ghanem, in press. Salt secretion and stomatal behaviour in Avicennia marina
seedlings fumigated with the volatile fraction of light Arabian crude oil. Environ. Pollution.
Youssef, T. and P. Saenger, 1996. Anatomical adaptive strategies and rhizosphere oxidation in
mangrove seedlings. Aust. J. Bot. 44:297-313.
Youssef, T. and P. Saenger, 1998a. Photosynthetic gas exchange and accumulation of phytotoxins in
mangrove seedlings in response to soil physico-chemical characteristics associated with
waterlogging. Tree Physiol. 18:317-324.
Youssef, T. and P. Saenger, I998b. Photosynthetic gas exchange and water use in tropical and
subtropical populations of the mangrove Aegiceras comiculatum. Mar. Freshw. Res. 49:329-
334.
Youssef. T. and P. Saenger. 1999. Mangrove zonation in Mobbs Bay - Australia. Est. Coastal Shelf Sci.
49:43-50.
Youssef, T., M. EI Amry and A. Youssef, 2000. Post-spill behavior in an oil contaminated mangrove
stand Avicennia marina (Forrsk.) Vierh. in UAE. Arabian Gulf J. Sci. Res. 18:102-109.
Zann, L.P., 1999. A new (old) approach to inshore resources management in Samoa. Ocean and
Coastal Managt. 42:569-590.
Zahran, M.A., 1977. Africa A. Wet formations of the African Red Sea coast. In: Chapman, V.I. (Ed.),
Wet coastal ecosystems. Elsevier Scientific Publ. Co., Amsterdam, pp. 215-231.
Zahran, M.A., 1982. Ecology of the halophytic vegetation of Egypt. In: Sen, D.N. and K.S. Rajpurohit
(Eds.), Tasks for vegetation science, Vol 2, Dr. W. Junk. The Hague, pp. 3-20.
Zhang. R.T., 1985. A new species of Acanthus Linn. in Fujian (Family Acanthaceae). Wuyi Science
Journal 5:237-239.
Zhang, Y. and K. Wang. 1994. Distribution of mangrove pollen in the sediments from East China Sea
and South China Sea and it~ palaeoenvironmental significance. Oceanologia et Limnologia
Sinica 25:23-28.
Zhang, Y., K. Wang, Z. Li and L. Lui, 1997. Studies on pollen morphology of Sonneratia genus in China
and its palaeoecological significance. Marine Geology and Quaternary Geology 17:47-52.
Zhang, Z., 1991. Pollen flora and palaeoclimate of the Chao-Shan Plain during the last 50,000 years.
Acta Micropalaeontologica Sinica 8:461-480.
Zimmermann, U., F. Meinzer and F.-H. Bentrup. 1995. How does water a~cend in tall trees and other
vascular plants? Ann. Bot. 76:545-551.
Zimmermann, U., JJ. Zhu, F.C. Meinzer, G. Goldstein, H. Schneider, G. Zimmermann. R. Benkert, F.
Thiirmer. P. Melcher, D. Webb and A. Haase, 1994. High molecular weight organic
compounds in the xylem sap of mangroves: implications for long-distance water transport. Bot.
Acta 107:218-229.
Zuberer. D.A. and W.S. Silver, 1975. Mangrove associated nitrogen fixation developmenl~. In: Walsh,
G.E .• s.c. Snedaker and H.J. Teas (Eds.), Proceedings of the International Symposium on
Biology and Management of Mangroves. Vol. 2. University of Florida, Gainesville, pp. 643-
653.
Zuberer. D.A. and W.S. Silver, 1978. Biological dinitrogen fixation (acetylene reduction) associated
with Florida mangroves. Appl. Environ. Microbiol. 35:567-575.
Index

Alligator Rivers, Northern


A Territory, 196, 201
allometric equations, 258-265
abscisic acid, 86 Ambon, 19
Abu Dhabi, 32 Amphibolus, 86
ACANTHACEAE, 12, 14 Amyelon, 43
Acanthus, 12, 14, 52, 113, Amyema, 148
174, 247, 299 Andaman Sea, 26
A. ilicifolius, 53, 164 Angola, 27, 217
accretion rates, 80, 123 anoxia, 8, 76, 81-82, 129, 135
acid sulfate soils, 125-127, antagonism, 150-151
130, 167, 290 APOCYNACEAE, 14
Acrostichum, 12, 18, 106, aquaculture ponds, 25, 289-291
121, 212, 253 Arabian Gulf, see also Persian
A. aureum, 62, 121, 301 Gulf, 25, 32, 80, 115, 149,
A. speciosum, 46, 78, 253 278, 284
Adansonia, 4, 145 Arabian peninsula, 32-33
Adenaria, 16 ARECACEAE, 14
Aechmea, 154 aridity, 6, 22, 31, 292
AEGIALITlDACEAE, 11, 18 Asia, 8, 14, 177, 194, 237
Aegialitis, 12, 18, 55, 108 Atalantia, 12, 19
A. annulata, 52, 54, 55 Aulacaspis, 99, 178
Aegiceras, 12, 50, 52, 55, Australia, 4, 14, 15, 17,37,
106, 153, 201, 245, 279 102, 112, 151, 170, 177,
A. comiculatum, 52, 53, 179, 188, 194, 239, 275,
55, 106 293
aerenchyma, 76-79 Avicennia, 4, 12,20,35,37,
aerial roots, 76-77 3& 5~ 98, 151, 19~ 21~
afforestation, 25, 230, 240 248, 250
nursery techniques, 250- A. bicolor, 65
252 A. integra, 12,20,37, 141
plantation performance, A. germinans, 7, 54, 64,
256-269 79, 175
planting season, 256 A. marina, 35, 39, 53, 54,
site-species matching, 78,104,115,149,161,
249-250 230, 248, 281
spacing, 257 A. schaueriana, 59, 261
Africa, 177, 231 AVICENNIACEAE, II
East, 3, 8, 27, 33, 151, axial convergence, 92-93,
194, 210, 215, 224, 181, 202
271, 290 Azospirillum, 153
West, 3, 7, 8, 22, 23, 27, Azteca, 177
28, 94, 160, 194, 195,
209, 278, 290 B
aggregations of birds, 175
Aglaia, 12, 16 Bacillus, 153
air-layering, 245 Bahamas, 173
albinism, 90 Bali, Indonesia, 99, 178
352 Mangrove Ecology, Silviculture and Conservation

Bangladesh. 20. 25. 116. 169, B. parviflora. 166, 211,


212. 215. 217. 224, 229. 254
237. 255. 256. 269. 290 B. sexangula. 4, 165, 211
Barringtonia. 12, 142, 151, Brunei, 217
218 brushpark fisheries. 213-214
basal area (BA), 118, 263-265 Buenaventura Bay, Colombia,
Batis, 113, 155 17
Bay of Bengal, 26, 116. 219, buoyancy, 91
237 Burma, see Myanmar
Belize, 97, 111, 170, 179, 192 buttress roots, 76
Benin, 28, 32, 79. 188, 209,
214, 215 C
Bermuda, 32
BIGNONIACEAE,14 Cadaba, 4
biodiversity. 223-224 CAESALPINIACEAE, 13, 15,
biogeography, 3, 12-14, 21- 152
22, 25-26 Cairns, Queensland, 299
Arabian Peninsular, 32-36 Calabar, Nigeria 23
Atlantic, 28-32 calorific value of leaves, 173
Australasian, 37-39 Cambodia, 213, 277
East African, 32-36 Cameroon, 23, 28, 46, 139,
East Pacific, 28-32 210, 217
historical, 43-47 Camptostemon. 12,64, 104,
Indian Ocean, 32-36 202, 211
North-east Asian, 41-43 C. schultzii. 211
West Pacific, 39-41 canopy gaps, 110-112
biomass, 75, 156, 180, 257, Cape York, Australia 134, 145
301 carbon isotopes,
above-ground (AGB), 75, ratios, 70, 1 10, 143, 180
156, 257-265 Carboniferous, 43
below-ground (BGB), 75 Caribbean, 4. 9, 17,27, 32,
bioremediation. 285-286 134, 140. 179, 192, 238.
biotechnology, 301 240, 273, 276, 293, 297
bioturbation of sediments, Casamance, Senegal 29
166-168 Cassipourea. 68, 86
boardwalks, 217, 298-299 Casuarina. 150. 152
BOMBACEAE. 12, 15 cations, 137
Bonny River, Nigeria, 23, Celestun, Mexico. 121. 222
218, 282 Centrasema. 151
Bora Bora. 25 Cerbera. 12, 211
Borneo, 15, 19, 20, 26 Ceriops. 13, 50. 87. 97, 112,
botanic gardens 141, 196, 211, 278
Bogor, Indonesia. 6 C. australis. 196, 211, 260
Singapore, 23 C. decandra. 211
Bradyrhizobium. 152 C. ragal. 155, 211
Brazil, 32,46. 136, 170, 171, charcoal. 212. 234, 278
181, 188, 197. 280, 291, China. 21. 41. 104, 213. 229.
297 238, 276, 301
BROMELIACEAE, 153 Chiromenthes. 169
Brownlowia. 12, 20, 113 Chokoria Sundarbans, 290
Bruguiera. 4, 9, 10, 12, 18, Christmas Island, 4, 26
104. 142. 196. 201. 231, Cienaga Grande de Santa Marta.
245, 278 Colombia 99
B. cylindrica. 21 I Cistanche. 149
B. exaristata. 10, 38, 106 Cladium. 294
B. hainesii. 166 classification, 183-193
B. gymnorhiza. 4, 9, 10, geomorphological, 187-
35, 77, 78, 166, 196, 190
247
Index 353

physiographic. 190-193 Dominican Republic, 112,


phytosociological. 183- 122, 198
184 dredging, 287
structural. 184-187
Clerodendrum. 13.56, 113,
151, 174, 218
E
C. inerme, 20, 21, 56. 218 EBENACEAE. 13, 15
climate change. see global economic models, 298
climate change ecosystem modelling. 298
Coccoloba, 148 Ecuador, 25, 169, 179, 192,
Colombia, 15. 16. 24. 32. 213, 217, 253, 273, 298
169. 187. 224. 238. 291 effects of CO2, 291-292
COMBRETACEAE. 13. 15 Egypt, 32
compartmentalization. 60-61 EI Salvador, 188, 213, 291
competition, 154-160 Endeavour River, Queensland,
Conocarpus, 13, 15. 50. 148. 224-225
166. 292 Enewetak Atoll. 24
C. erectus, 56. 60, 166 Enhalus, 92
Cook Islands, 40 Eocene, 17,23, 27, 30, 37,
Cooktown. Queensland. 224, 46, 152, 154
225 epiphytes, 153-154. 223
Comer Inlet. Victoria. 39 Equatorial Guinea, 28, 32. 217
Costa Rica, 17, 154, 171, 179, ethanol production, 83-84
189. 192 etymology, 1
Cote d'Ivoire. 1, 28, 32. 188. Eucalyptus, 8
209 Euphorbia, 151
Cretaceous, 43, 44, 291 EUPHORBIACEAE, 13. 15
Late. 8 Excoecaria, 11, 13,50,57,
Cross River, Nigeria 23 106, 142, 148, 160, 161,
cryptovivipary. 84-86 162, 165, 169, 211, 235
Cuba. 238. 264, 267, 273 E. agallocha, 15, 16, 148,
currents, 5. 6, 27, 28, 42 211, 235, 247, 251
cyclones. 116-120 E. ovalis, 15, 16, 37, 106,
Cynometra, 4, 13. 52, 152, 203
172, 245 extrusion, 279
C. iripa, 57, 152, 245 rates of, 54
C. ramijlora, 4, 152
cytoplasmic osmoregulators,
62-63
F
Cytospora, 150 FABACEAE, 13, 15, 152
Farasan Islands, 35, 36, 295
D Federated States of Micronesia,
268-269
Dalbergia, 13 felling cycles, 233, 235, 268-
Damman. Saudi Arabia. 35 269
Darwin, Australia 97, 117 Fiji, 25, 40, 278, 291
definition, 10 firewood, 25, 209, 277
Derris, 13. 151,218,255 fisheries, 213-215, 233, 278
Dichrostachys, 4 flood mitigation, 286
diffusion coefficient, 5 Florida, 102, 110, 118, 122.
Diospyros, 13 132, 143, 155. 170. 185,
Dischidia, 176 193, 223, 242
disturbance, 157-158 flowering phenology, 160-163
by hail, 121 Fly River, Papua New Guinea,
by lightning. 121-122 101. 141, 146, 180, 187
by roosting, 175 foreshore development, 286-
Djibouti, 27, 33, 216 291
Dolichandrone, 13, 211 fossil record. 44
354 Mangrove Ecology. Silviculture and Conservation

France. 47. 184 Hainan Island. China. 19.41,


Frankia, 152 42. 252. 301
French Guiana. 23, 139, 140, Hawaiian Islands. 40. 165
156, 291 heavy metals. 279-282
fruiting phenology. 87-89 Helice, 169
Fuding, China. 41 heritage. 224
Fujian (Fukien). China. 41, Heritiera, 4.9, 13.20.62. 64.
293 76, 211. 235. 251, 278
fungi, 149-150 H. fomes, 211. 235, 247
H. littoralis, 22. 36. 211,
G 235
Hibiscus, 2. 13. 36. 251. 299
Gabon, 28. 32, 139, 195.217 Hinchinbrook Island,
Galapagos Islands. Ecuador, 39 Queensland. 129. 189. 190
Gambia, 28. 140. 209, 273 Holocene, 32, 40, 102, 293
Ganges delta, Bangladesh. 187 Hong Kong, 41, 163, 169,
Gazi Bay. Kenya. 33. 197 273. 297
genetic diversity. 99-100, 163 Horn of Africa (northern
geological periods, Somalia). 22. 33. 115
Carboniferous. 43 Hoya.176
Cretaceous. 43. 44. 291 hurricanes, 87, 116-120
Eocene, 17,23.27. 30, hybrids, 26
37. 43, 154 Hydnophytum, 176
Holocene. 32. 40. 102, hydrocarbon contamination,
293 282-286
Jurassic. 43 Hypochrysops, 176
Oligocene. 23 hypodermis. 64-67
Miocene. 23. 27. 43
Palaeocene. 45 I
Pliocene. 39, 43
Tertiary, 39 India. 14, 15. 26, 33, 163.
Ghana, 23, 188, 241, 274 187. 210, 215. 229. 238,
Gladstone. Queensland. 104, 256. 273. 291. 296
149 indices of mangrove 'health'.
global climate change. 291- 269-270
294 Indonesia. 19, 25, 169, 175.
grazing, 168-175 210. 214. 229, 265. 273,
Great Barrier Reef, 38, 190 290
groundwater, 129. 222. 289 Intsia, 246
Guadeloupe. 118, 133. 170, Irian Jaya, Indonesia, 19
250 Iridomyrmex, 176
Guam, 41 Iriomote Island. Japan. 43,
Guanabara Bay. Brazil, 171 201
Guangdong, China. 25. 41. isotope analysis. 70, 110,
252 143, 180
Guangxi. China, 41
Guiana. see French Guiana
Guinea. 28, 140, 209. 290
J
Guinea-Bissau. 28, 32 Jamaica, 46
Gulf of Carpentaria. Australia, Japan. 41. 276. 287
5. 27, 38. 116. 186. 188 Jeddah. Saudi Arabia, 190
Gulf of Guinea. West Africa. Juilong estuary, China, 293
28, 93-94, 188 Junonia, 99
Gulf of Mexico, 102, 221 Jurassic, 43

H K
hail damage. 121 Kagoshima, Japan, 41
Index 355

Kalimantan, Indonesia, 210 Malaysia, 19, 20, 111, 139,


Kandelia, 13, 81, 100, 104, 185, 188, 210, 229, 256,
172, 246, 251 265, 290
K. candel, 41, 100, 164, malic acid, 83-84
246, 293 MALVACEAE, 13, 16
Kawhia Harbour, New Zealand, mangrove borer, 174
39 mangroves,
Kenya, 24, 290 epiphytes, 153-154, 176-
Khafji, Saudi Arabia, 34, 284, 177, 223
285 extent of, 8
Kimberley Coast, Western fossil, 44
Australia, 119, 167, 189 introduced, 23, 24
knee-roots, 76-77 landlocked, 4, 5
Koehneria, 16 loss of, 273
Konkoure River, Guinea, 290 on postage stamps, 299-
Kosrae, Micronesia, 10, 41, 300
111, 139 species distributions, 12-
Kuwait Bay, Kuwait, 34,247, 14
284 Manila, Philippines 19
Martinique, 291
L Matang Mangrove Forest,
Malaysia, 231-234, 254,
Laguna, Brazil, 32 268, 294, 295
Laguncularia, 13,55, 104, Mauritania, 7, 25, 28, 31,115
107, 159-160, 195, 211, Mauritia, 27
292 mean annual increment (MAl),
L. racemosa, 55, 56, 137, 265-268
159, 171, 195, 211, Mediterranean Sea, 44-45, 47
250, 260, 292 Melaleuca, 8, 9, 174
Lamu Archipelago, Kenya, 33, MELIACEAE, 13, 16
273 Merope, 19
latitudinal limits, 21, 22, 36 mesocosm, 298
leaf herbivory, 168-175, 180 Metapenaeus, 213
leaf lifespans, 51-52, 170, 174 Mexico, 119, 155, 175, 186,
LECYTHIDACEAE, 12, 16 189, 192, 213, 223, 265,
lenticels, 77 273
Liberia, 27, 28, 29 Microcoleus, 153
lightning, 121-122 micropropagation, 247-249
Limonellus, 19 Middle East, 215, 240
Limonium, 18, 86 Miocene, 23, 27, 30, 43
litterfall (ALF), 221, 262 Mo~iimedes, Angola 31
Lobito, Angola. 31 Molokai, Hawaii, 24, 237
LORANTHACEAE,147-148 monetary value, 225-228
Lumnitzera, 13, 104, 151, Monomorium, 178
160, 211 Mo'orea, French Polynesia, 24
L. littorea, 104, 211 Mora, 13, 205
L. racemosa, 35, 137, 160, Moreton Bay, Queensland, 278
211 mortality, 149, 157
LYTHRACEAE, 13, 16 Mozambique, 25, 33, 273, 291
mud-lobsters, 166-167
M municipal garbage dumps, 282
mutualism, 152-154, 176-179
Macau, 41 Myanmar (Burma), 19, 229,
Macrophthalmus, 178 273, 295
macropropagation, 243-247 mycorrhizae, 152
Madagascar, 4, 16, 25, 27, 33, Myrmecodia, 176
213, 278, 291 myrmecophytes, 176
MYRSINACEAE, 12, 17, 153
MYRTACEAE, 8, 13, 17
356 Mangrove Ecology, Silviculture and Conservation

oxidation-reduction potential
(Eh), 98, 126, 127, 250
N
Natal. South Africa. 33. 35. p
168
Neosamartium. 97. 169 Pacific Ocean, 39-41
New Caledonia, 14. 24. 40 Pakistan, 27, 33, 215, 238
New Zealand. 6. 39. 80. 161. Palaeobruguiera. 44, 46
242 Palaeocene, 45
N 2-fixation. 133. 153. 179 Palau. 19,41,299,300
Nicaragua. 24. 117. 209. 215. PALMACEAE, 14
224 Panama, 23, 24,44, 165, 177,
Niger delta. 23. 28. 29. 30. 46. 192, 199, 201, 224, 240,
189. 209. 215. 217 283
Nigeria, 23, 27, 28, 139, 198, Papua New Guinea, 4, 14, 15,
217, 273, 290 19.27, 37, 121, 139, 146,
Nouakchottian transgression, 177. 213, 288
4,28 Parasesarma. 169
nutrients, 131. 153, 157, 175 parasitism, 147-150
and herbivory, 134, 171- Paratrechina, 178
174 Parhyale, 179
foliar concentrations. 131, Paspalum, 209
279 Pavonia. 13, 16
limitation, 131, 134 Pehria. 16
recycling, 136-137 Pelliciera, 13. 17, 27. 70, 76,
retention, 220 85, 154, 178
soil concentrations, 132, f. rhizophorae. 17, 138,
279 154, 166
Nypa, 13,27, 146, 151, 161, PELLICIERACEAE, 11, 13, 17
165, 212, 235 Pemphis, 13, 16, 60, 64, 107,
N. fruticans. 151, 165, 107, 162, 247
235, 251 penaeid prawns, 182. 290
NYPACEAE. 11, 14 Penaeus, 182. 213, 290
Pennisetum, 151
o Perisesarma. 169
Persian Gulf (see also Arabian
Oahu, Hawaii, 24, 165 Gulf), 25, 32, 80, 115, 149,
ocean currents. 6, 27, 28, 39, 278. 284
43 pharmaceuticals, 218
Oecophylla. 177 Phaseolus. 151
Ogooue. Gabon, 29 Philippines, 15, 19, 20, 25,
Ohiwa Harbour, New Zealand. 213, 215, 225. 229. 241.
39 273
oil production. 217-218 Phoenix. 86
Okinawa Island. Japan. 77, photoinhibition. 106
287 photosynthesis, 105-110
Oligocene. 23 Phthirusa, 148
Orinoco River. Venezuela. Phyllobacterium. 153
157, 187-188, 207 Phytophthora
OROBRANCHACEAa149 (Halophytophthora). 116,
Orono Nigeria 23 149
Osbornia. 8, 13, 37, 57. 107. phytostabilization, 300
172. 211 phytotoxins, 7, 75. 112. 113,
O. octodonta. 17,52, 57, 150
59, 63, 107, 211 Pitcairnia, 154
osmoregulators, 62-63 plantation performance, 256-
osmotic potential, 50 269
out welling, 179-182 Pliocene, 39, 43
Index 357

PLUMBAGINACEAE, 12, 18 cost of, 255-256


pneumatophores, 76, 178, 284 remote sensing, 185
Poecilips, 171 Repulse Bay, Queensland, 10,
Pohnpei (Ponape), Federated 204
States of Micronesia, 41, reserves, 294
224 restoration, 241-243
poles, 209 species selection, 249-250
pollination, 114, 160-166 Rhizobium, 152
Pongamia, 15 Rhizophora, 4, 13, 18,43,
Port Curtis, Queensland, 97, 108, 122, 144, 179, 211,
99, 116, 121, 199, 204 232, 251, 260, 278, 291
Port Stanley, Vanuatu 41 R. apicuLata, 50, 71, 131,
Posidonia, 92 137, 174, 211, 232,
prawn ponds, 239, 289 263, 292
prawns, 182, 213, 221, 233, R. harrisonii, 138
278, 289 R. mangle, 71, 108, 131,
Premna, 13, 20 137, 138, 143, 150,
Princess Charlotte Bay, 153, 188, 211, 241,
Queensland, 130, 155 248, 260, 261, 267,
propagule, 84, 87-91, 199, 291
243, 246, 284 R. mucronata, 36, 65, 71,
albinism, 89-90 94, 131, 137, 211, 232
buoyancy, 92 R. racemosa, 67, 188, 241
dispersal, 6, 85-100, 199, R. stylosa, 76, 77, 90,
253-255 130, 138, 167, 173,
establishment, 6, 91-100, 211, 263, 292
in canopy gaps, 111-112 RHIZOPHORACEAE, 11, 12,
longevity, 94 13, 18, 67
production, 87-91 Rio Longa, Angola, 31
predation, 88, 200 Rio Muni, Equatorial Guinea,
survival, 254-256 29
Proserpine, Queensland, 10, Rio San Juan, Venezuela, 237
97, 99, 104, 158, 250 rotation times, 233, 234, 268-
PTERIDACEAE, 12, 18 269
Pterocarpus, 9, 139, 152 RUBIACEAE, 19. 153
public awareness, 298 Rufiji delta, Tanzania, 224,
Puerto Rico, 103, 150, 181, 226, 290
192, 271, 272, 291 RUTACEAE, 12, 19
pyrite, 125-127 Ruyuku Islands. Japan, 43

Q s
Qatar. 240 Sabah, 210, 234. 278
quantum efficiency. 105-110 salinity. 137-146, 180, 196.
Queensland. 39. 133, 142. 197, 199, 202, 215-250.
168. 192, 201, 278 270
Quintana Roo. Mexico. 175 and potassium deficiency.
Qurma Island, Saudi Arabia, 34. 144-145
284 effects on growth. 98. 143,
290. 301
R gradients, 140-143, 203-
205
Ranong Biosphere Reserve. Saloum River estuary, Senegal.
Thailand, 295 29, 145, 215
Raphia, 9 salt. 3, 49. 145, 215, 301
Red Sea, 32, 34, 35. 115, 190. diffusion coefficient, 5, 6
284 exclusion, 5, 50-52
regeneration. 253-255 extrusion, 53-57, 283
358 Mangrove Ecology, Silviculture and Conservation

flats, 140, 205 S. alba, 36, 119, 141, 205,


glands, 53-57, 283 211
osmoregulation, 62-63 S. apetala, 25, 211, 235,
production, 216, 291 237
storage, 57-58 S. lanceolata, 141, 142,
tolerance, 9, 49, 204-205 105
ultrafiltration, 50 S. caseolaris, 6, 164. 172,
Samoa, 40, 275 211
Sapium, 15 SONNERATIACEAE, 14, 19
Sarawak, 210, 234, 254, 278 Soumba River, Guinea, 186,
Sarcocornia, 150 290
Saudi Arabia, 34, 190, 284, South Africa, 17, 120, 133,
295 213, 214
Sceura, 115 South America, 3, 6, 39, 154,
SCROPHULARIACEAE, 14 201, 231
Scylla, 168, 278 Spartina, 200
Scyphiphora, 13, 19, 104, Sporobolus. 150
153, 211, 246 Sri Lanka, 26, 62,213, 214
S. hydrophyllacea, 19, St. Lucia, West Indies, 212,
153, 211 241, 296
seabirds, 134, 175 stand (stem) density, 118, 156-
sealevel, 116, 293 158, 185, 263, 264
wind effects on, 116 stem diameter (DB H), 231,
importance of, 5 253, 258-265
changes in, 293 STERCULIACEAE, 13, 20
Senegal River, 23, 30 stilt roots, 76, 103
Senegal, 5, 23, 27, 28, 140, stomata, 65
145, 209, 291 stress, 157
Sepetiba Bay, 171 structural classification, 184-
Sesarma, 168 187
sewage effluents, 278 Suaeda, 113, 150
shade-tolerance, 107, 113 succulence, 58-60
shikimic acid, 83-84 Sudan, 25, 35, 47
shoreline protection, 219, Sulawesi, 26, 129, 214
229, 237 sulfides, 7, 75, 82, 113-114,
Sierra Leone, 290 126-127, 167, 300
silviculture, 229, 252 Sumatra, 169, 254
nursery and planting Sundarbans, 19, 25, 198, 217,
techniques, 250-252 235-238, 253
plantation performance, Biosphere Reserve, 236
256-269 syntax a, 183-184
planting season, 256
spacing, 257 T
Sinai Peninsula, 33, 34
Sine Saloum delta, Senegal, Tabebuia, 14, 205
29, 145, 215 Tahiti, 24
Singapore, 113, 273 Taiwan, 41
Society Islands, 24, 40 Taklong Island, Philippines,
soil, 50
development, 122-128 Tamil Nadu, India, 111
drainage, 128-130 tanbark extraction, 25, 208
evaporation from, 140 tannins and phenolics, 168-
nutrients, 131-137 169, 171-172, 174-175
water content, 250 Tanzania, 25, 224, 275, 291
Solomon Islands, 19, 20, 41, Tasmania, 37
177 Tecticornia, 113
Sonneratia, 4, 14, 19,44,47, temperature, 10 1-105
50, 52, 57, 59, 85, 172,
205, 211
Index 359

importance of, 3, 22, 38,


41, 103, 104
leaf, 73-75
v
and flowering, 87, 161 values of mangroves, 207-209
and ocean currents, 42-43 ecological, 208, 219-225
and pericarp shedding, 92 economic. 208-219, 225-
and transpiration, 70 228
Terebralia, 32 instrumental, 208
Terminos Lagoon, Mexico, intrinsic, 208, 223-225
116 symbolic, 208
Tertiary, 39 Vanuatu, 41
Tethys Sea, 30, 44-47 Venezuela, 47, 60, 62, 157,
Thailand, 19, Ill, 168, 169, 197, 207, 217, 237, 269
222, 229, 240, 256, 265, VERBENACEAE, 12, 13, 20
288, 289, 295, 297 vicariants, 26
Thais, 179 Vietnam, 19, 117, 229, 238,
Thalassina, 167 269, 273
Thalassodendron, 86, 92 Virgin Islands, 238
THEACACEAE, 17 vivipary, 84-86
thermal wastes, 282 Volta River, Ghana, 23, 29,
Thespesia. 2, 14, 64, 71,93, 241
151, 166, 245
Threskiornis. 175
tidal asymmetry, 216
w
TILIACEAE, 12, 20 water-use efficiency, 68, 143,
Tillandsia. 154 291
timber production, 230, 231. waterlogging, 75-84, 204,
267-268 250, 289
Timor, 38 tolerance, 8, 9
tissue culture, 247-249 Weichselia, 43
Togo, 28, 29 West Alligator Rivers,
Tonga, 40 Northern Territory, 196
top-dying, 236 Western Australia, 189, 289
Townsville, Queensland, 140 Westernport Bay, Victoria, 39,
transpiration, 68-73 106, 133
and waterlogging, 72 Wildman River, Northern
transplanting, 244, 255 Territory, 117, 142
Tribuga Gulf, 23 wind effects, 114-120
Triglochin. 150 wood, 120, 211
Trinidad and Tobago, 215, anatomy, 67, 68, 120
217, 223, 291 density, 211, 261
Truk,41 production, 211. 267-268
Turks and Caicos Islands, 185 volume, 258-260, 267-268
Tuvalu, 41 Woodfordia. 16
typhoons, 116-120 Wouri Estuary, Cameroon, 23

u x
Uca.181 xeromorphy, 63-67, 107
ultraviolet radiation, 109 Xylocarpus, II, 14, 16, 57,
understorey, 113-114 107, 159, 173, 211
United Arab Emirates, 32, 240, X. granatum. 22, 36, 52,
283 57, 162, 166, 211, 248
urbanization, 287 X. moluccensis, 22, 36,
USA, 238, 256, 298 52, 57, 161, 162, 166,
211
360 Mangrove Ecology, Silviculture and Conservation

Zambesi, Mozambique, 33
y Zhejian, China, 41
zonation, 7. 194-205
Yap, 41 shoreline. 194-201
Yemen. 25, 35, 115 upriver, 201-204
Yucatan, Mexico, 129 zoning of mangroves. 297

z
Zaire. 29. 231, 241

You might also like