Personal tools
Home » Working Groups » Valuation of Coastal Habitats » Relevant papers » Various Mangroves-Related Papers » Biology of Mangroves and Magrove Ecosystems (Kathiresam and Bingham, 2001)
Navigation
Log in


Forgot your password?
 
Document Actions

Biology of Mangroves and Magrove Ecosystems (Kathiresam and Bingham, 2001)

BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                                    1




           Biology of Mangroves and Mangrove Ecosystems

          ADVANCES IN MARINE BIOLOGY VOL 40: 81-251 (2001)


                    K. Kathiresan1 and B.L. Bingham2

        1
       Centre of Advanced Study in Marine Biology, Annamalai University,
                Parangipettai 608 502, India
     2
      Huxley College of Environmental Studies, Western Washington University,
    Bellingham, WA 98225, USA e-mail bingham@cc.wwu.edu (correponding author)

1. Introduction .............................................................................................. 4
    1.1. Preface........................................................................................ 4
    1.2. Definition ................................................................................... 5
    1.3. Global distribution ..................................................................... 5
2. History and Evolution ............................................................................. 10
    2.1. Historical background ................................................................ 10
    2.2. Evolution .................................................................................... 11
3. Biology of mangroves
    3.1. Taxonomy and genetics.............................................................. 12
    3.2. Anatomy..................................................................................... 15
    3.3. Physiology ................................................................................. 18
     3.4. Biochemistry ............................................................................. 20
    3.5. Pollination biology..................................................................... 21
    3.6. Reproduction, dispersal and establishment ................................ 22
    3.7. Biomass and litter production .................................................... 24
4. Mangrove-associated flora
    4.1. Bacteria ...................................................................................... 27
    4.2. Fungi and fungus-like protists.................................................... 29
    4.3. Microalgae.................................................................................. 33
    4.4. Macroalgae................................................................................. 34
    4.5. Seagrasses .................................................................................. 36
    4.6. Saltmarsh and other flora ........................................................... 37
5. Mangrove-associated fauna
    5.1. Zooplankton ............................................................................... 38
    5.2. Sponges and Ascidians............................................................... 39
    5.3. Epibenthos, infauna, and meiofauna .......................................... 41
    5.4. Prawns, shrimp and other crustaceans ....................................... 43
    5.5. Crabs .......................................................................................... 45
    5.6. Insects......................................................................................... 49
    5.7. Mollusks..................................................................................... 50
    5.8. Fish............................................................................................. 52
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                                   2


   5.9. Amphibians and Reptiles ........................................................... 56
   5.10. Birds ......................................................................................... 56
   5.11. Mammals.................................................................................. 57
6. Responses of mangroves and mangrove ecosystems to stress ................ 58
   6.1. Responses to light ...................................................................... 58
   6.2. Responses to gases ..................................................................... 59
   6.3. Responses to wind...................................................................... 61
   6.4. Responses to coastal changes..................................................... 61
   6.5. Responses to tidal gradients and zonation ................................. 63
   6.6. Responses to soil conditions ...................................................... 64
   6.7. Responses to salinity.................................................................. 66
   6.8. Responses to metal pollution ..................................................... 67
   6.9. Responses to organic pollution .................................................. 69
   6.10. Responses to oil pollution ........................................................ 70
   6.11. Responses to pests.................................................................... 71
   6.12. Responses to anthropogenic stress ........................................... 73
   6.13. Responses to global changes.................................................... 75
7. Ecological role of mangrove ecosystems
   7.1. Litter decomposition and nutrient enrichment ........................... 76
   7.2. Food webs and energy fluxes..................................................... 78
8. Concluding remarks ................................................................................ 80
   Acknowledgements ........................................................................... 82
   References .........................................................................................

    Mangroves are woody plants that grow at the interface between land and sea in
tropical and sub-tropical latitudes where they exist in conditions of high salinity, extreme
tides, strong winds, high temperatures and muddy, anaerobic soils. There may be no other
group of plants with such highly developed morphological and physiological adaptations
to extreme conditions.
     Because of their environment, mangroves are necessarily tolerant of high salt
levels and have mechanisms to take up water despite strong osmotic potentials. Some also
take up salts, but excrete them through specialized glands in the leaves. Others transfer
salts into senescent leaves or store them in the bark or the wood. Still others simply
become increasingly conservative in their water use as water salinity increases.
Morphological specializations include profuse lateral roots that anchor the trees in the
loose sediments, exposed aerial roots for gas exchange and viviparous water-dispersed
propagules.
     Mangroves create unique ecological environments that host rich assemblages of
species. The muddy or sandy sediments of the mangal are home to a variety of epibenthic,
infaunal, and meiofaunal invertebrates. Channels within the mangal support communities
of phytoplankton, zooplankton, and fish. The mangal may play a special role as nursery
habitat for juveniles of fish whose adults occupy other habitats (e.g., coral reefs and
seagrass beds)
     Because they are surrounded by loose sediments, the submerged mangroves roots,
trunks, and branches are islands of habitat that may attract rich epifaunal communities
including bacteria, fungi, macroalgae, and invertebrates. The aerial roots, trunks, leaves
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        3


and branches host other groups of organisms. A number of crab species live among the
roots, on the trunks or even forage in the canopy. Insects, reptiles, amphibians, birds and
mammals thrive in the habitat and contribute to its unique character.
    Living at the interface between land and sea, mangroves are well adapted to deal
with natural stressors (e.g., temperature, salinity, anoxia, UV). However, because they
live close to their tolerance limits, they may be particularly sensitive to disturbances like
those created by human activities. Because of their proximity to population centers,
mangals have historically been favored sites for sewage disposal. Industrial effluents have
contributed to heavy metal contamination in the sediments. Oil from spills and from
petroleum production has flowed into many mangals. These insults have had significant
negative effects on the mangroves.
    Habitat destruction through human encroachment has been the primary cause of
mangrove loss. Diversion of freshwater for irrigation and land reclamation has destroyed
extensive mangrove forests. In the past several decades, numerous tracts of mangrove have
been converted for aquaculture, fundamentally altering the nature of the habitat.
Measurements reveal alarming levels of mangrove destruction. Some estimates put global
loss rates at one million ha y-1, with mangroves in some regions in danger of complete
collapse. Heavy historical exploitation of mangroves has left many remaining habitats
severely damaged.
     These impacts are likely to continue, and worsen, as human populations expand
further into the mangals. In regions where mangrove removal has produced significant
environmental problems, efforts are underway to launch mangrove agroforestry and
agriculture projects. Mangrove systems require intensive care to save threatened areas. So
far, conservation and management efforts lag behind the destruction; there is still much to
learn about proper management and sustainable harvesting of mangrove forests.
     Mangroves have enormous ecological value. They protect and stabilize coastlines,
enrich coastal waters, yield commercial forest products and support coastal fisheries.
Mangrove forests are among the world’s most productive ecosystems, producing organic
carbon well in excess of the ecosystem requirements and contributing significantly to the
global carbon cycle. Extracts from mangroves and mangrove-dependent species have
proven activity against human, animal and plant pathogens. Mangroves may be further
developed as sources of high-value commercial products and fishery resources and as sites
for a burgeoning ecotourism industry. Their unique features also make them ideal sites for
experimental studies of biodiversity and ecosystem function. Where degraded areas are
being revegetated, continued monitoring and thorough assessment must be done to help
understand the recovery process. This knowledge will help develop strategies to promote
better rehabilitation of degraded mangrove habitats the world over and ensure that these
unique ecosystems survive and flourish.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                             4


1. INTRODUCTION

1.1. Preface

    Mangrove forests are among the world’s most productive ecosystems. They enrich
coastal waters, yield commercial forest products,
protect coastlines, and support coastal fisheries
(Figures 1 and 2). However, mangroves exist under
conditions of high salinity, extreme tides, strong
winds, high temperatures and muddy, anaerobic soils.
There may be no other group of plants with such
highly developed morphological, biological,
ecological and physiological adaptations to extreme
conditions.
    Mangroves and mangrove ecosystems have
been studied extensively but remain poorly
understood. With continuing degradation and
destruction of mangroves, there is a critical need to
understand them better. Aspects of mangrove biology
have been treated in several recent reviews. Tomlinson
                     (1986) described
                     the basic botany    Figure 2. A) General view of coastal
                                edge of a mangrove forest. B) A black
                     of mangroves.
                                mangrove thicket, Avicennia, showing
                     Snedaker and      aerial roots (pneumatophores). C) Closer
                     Snedaker (1984)    view of the pneumatophores of
                     reviewed earlier
                     mangrove research and made recommendations for
                     further research. An overview of tropical mangrove
                     community ecology, based primarily on Australian
                     work, can be found in Robertson and Alongi (1992).
                     Li and Lee (1997) reviewed much of the Chinese
                     mangrove literature published between 1950 and
                     1995. Ellison and Farnsworth (2000) have recently
                     published a general review of mangrove ecology.
                         As researchers continue to discover important
                     facts about mangroves and the role they play in the
                     global ecosystem, the volume of published
Figure 1. A) The seward edge of a
mangrove forest, showing red      information has grown enormously and increasing
mangroves, Rhizophora. B) A young    numbers of workers are drawn to these unique
plant of Rhizophora, showing prop
                     environments. Thus, there is a need for periodic
roots carrying epifauna, including
                     reviews of the rapidly expanding literature. In this
barnacles and oysters. C) The
                     review, we emphasize work on mangrove ecosystems
propagules of Rhizophora, developed
from the fruit, before release (photos: completed between 1990 and 2000, though for space
A, A.J. Southward; B, K. Kathiresan;  reasons we can list only a fraction of the studies. Our
C, B.L. Bingham)
                     intent is to make information more readily available to
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                         5


researchers around the world in hopes of facilitating and stimulating further study of the
mangrove environment.

1.2. Definition

    Mangroves are woody plants that grow at the interface between land and sea in
tropical and sub-tropical latitudes (Figures 1 and 2). These plants, and the associated
microbes, fungi, plants, and animals, constitute the mangrove forest community or
mangal. The mangal and its associated abiotic factors constitute the mangrove ecosystem
(Figure 3). The term “mangrove” often refers to both the plants and the forest community.
To avoid confusion, Macnae (1968) proposed that “mangal” should refer to the forest
community while “mangroves” should refer to the individual plant species. Duke (1992)
defined a mangrove as, “…a tree, shrub, palm or ground fern, generally exceeding one half
metre in height, and which normally grows above mean sea level in the intertidal zone of
marine coastal environments, or estuarine margins.” This definition is acceptable except
that ground ferns should probably be considered
mangrove associates rather than true mangroves.
The term “mangrove” is also used as an
adjective, as in ”mangrove tree” or “mangrove
fauna.” Mangrove forests are sometimes called
“tidal forests”, “coastal woodlands”, or “oceanic
rain forests.”
    The word “mangrove” is usually
considered a compound of the Portuguese word
“mangue” and the English word “grove.” The
corresponding French words are “manglier” and
“paletuvier” (Macnae, 1968) while the Spanish
term is “manglar”. The Dutch use            Figure 3. Physical and biological
“vloedbosschen” for the mangrove community       components of mangrove ecosystems.
and “mangrove” for the individual trees. German
use follows the English. The word “mangro” is a common name for Rhizophora in
Surinam (Chapman, 1976). It is believed that all these words originated from the
Malaysian word, “manggi-manggi” meaning “above the soil.” This word is no longer used
in Malaysia, but is used in eastern Indonesia to refer to Avicennia species.

1.3. Global distribution

     Mangroves are distributed circumtropically, occurring in 112 countries and
territories. Global coverage has been variously estimated at 10 million hectares (Bunt,
1992), 14-15 million hectares (Schwamborn and Saint-Paul, 1996), and 24 million hectares
(Twilley et al., 1992). Spalding (1997) gave a recent estimate of over 18 million hectares,
with 41.4% in south and southeast Asia and an additional 23.5% in Indonesia (Figure 4).
Mangroves are largely restricted to latitudes between 30° north and 30° south. Northern
extensions of this limit occur in Japan (31°22’N) and Bermuda (32°20’N); southern
extensions are in New Zealand (38°03’S), Australia (38°45’S) and on the east coast of
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                                          6


South Africa (32°59’S; Spalding, 1997, Yang
                            South and
et al., 1997). Mangroves are not native to the    Southeast Asia


Hawaiian Islands, but since the early 1900’s, at   The Americas


least 6 species have been introduced there.       West Africa

     Mangrove distributions within their       Australasia
ranges are strongly affected by temperature
                           East Africa and
(Duke, 1992) and moisture (Saenger and        the Middle East


Snedaker, 1993). Large-scale currents may also             0   2         4         6    8
                                      Area covered by mangrove forests (million ha)
influence distributions by preventing             Figure 4. Global coverage of
propagules from reaching some areas (De Lange and De mangrove forests (modified from
Lange, 1994). Individual mangrove species differ in the    Spalding, 1997).
length of time their propagules remain viable, their
establishment success, their growth rate, and their tolerance limits. These factors, which
appear quite consistent around the world, interact to produce characteristic distributional
ranges for most species (Duke et al., 1998a; Table 1).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                                                                                               7


Table 1. Mangrove species, their taxonomic authorities, and global distributions.




                                                                              Malay Archipeligo




                                                                                                    Southwest Pacific
                                          Southeast USA




                                                                     Southeast Asia
                                                  Central/South




                                                                                                              West Pacific
                                                               South Asia




                                                                                        East Asia

                                                                                              Australia
                                                   America
                                                          Africa
Family       Species




                                                                 !       !         !       !
Avicenniaceae   Avicennia alba Blume
                                                                                                !
          Avicennia balanophora Stapf and Moldenke ex Molodenke
                                                    !
          Avicennia bicolor Standley
                                                                                                !
          Avicennia eucalyptifolia (Zipp. ex Miq.) Moldenke
                                             !      !
          Avicennia germinans (L.) Stearn
                                                                                  !
          Avicennia lanata Ridley
                                                            !     !       !         !       !      !        !
          Avicennia marina (Forsk.). Vierh.
                                                                         !         !       !      !
          Avicennia officinalis L.
                                                    !
          Avicennia schaueriana Stapf and Leechman ex Moldenke
                                                            !
          Avicennia africana Palisot de Beauvois

                                                                 !       !         !
Bignoniaceae    Dolichandrone spathacea (L. f.) K. Schumann

                                                                                  !       !
Bombacaceae    Camptostemon philippinensis (Vidal) Becc.
                                                                                  !
          Camptostemon schultzii Masters

                                                                 !       !         !             !
Caesalpiniaceae  Cynometra iripa Kostel
                                                                 !       !         !
          Cynometra ramiflora L.

                                             !      !
Combretaceae    Conocarpus erectus L.
                                             !      !             !       !
          Laguncularia racemosa (L.) Gaertn. f.
                                                                 !       !         !
          Lumnitzera littorea (Jack) Voigt.
                                                            !     !       !         !       !
          Lumnitzera racemosa Willd.
                                                                                          !
          Lumnitzera X rosea (Gaud.) Presl. (hybrid of
           L. racemosa and L. littorea)
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                                   8


                                           !  !  !  !  !  !
Euphorbiaceae  Excoecaria agallocha L.
                                           !  !  !
         Excoecaria indica (Willd.) Muell. - Arg.
                                                   !
         Excoecaria dallachyana (Baill.) Benth.

                                         !  !  !  !  !  !
Lythraceae    Pemphis acidula Forst.
                                         !
         Pemphis madagascariensis (Baker) Koehne

                                           !
Meliaceae    Aglaia cucullata (Pellegrin ) Roxb.
                                           !  !  !  !  !    !
         Xylocarpus granatum Koen.
                                           !  !  !    !
         Xylocarpus mekongensis Pierre
                                           !  !  !
         Xylocarpus moluccensis (Lamk.) Roem.

                                           !  !  !  !  !
Myrsinaceae   Aegiceras corniculatum (L.) Blanco
                                             !  !  !
         Aegiceras floridum Roemer and Schultes

                                               !  !
Myrtaceae    Osbornia octodonta F. Muell. loc. cit.

                                             !
Pellicieraceae  Pelliciera rhizophoreae Triana and Planchon

                                               !
Plumbaginaceae  Aegialitis annulata R. Brown
                                           !  !
         Aegialitis rotundifolia Roxburgh

                                           !  !  !  !
Rhizophoraceae  Bruguiera cylindrica (L.) Bl.
                                               !
         Bruguiera exaristata Ding Hou
                                         !  !  !  !  !  !  !  !
         Bruguiera gymnorrhiza (L.) Lamk.
                                             !  !
         Bruguiera hainesii C. G. Rogers
                                           !  !  !  !  !
         Bruguiera parviflora Wight and Arnold ex Griffith
                                         !  !  !  !  !
         Bruguiera sexangula (Lour.) Poir.
                                           !  !  !  !
         Ceriops decandra (Griff.) Ding Hou
                                       !  !  !  !  !  !  !  !  !
         Ceriops tagal (Perr.) C. B. Robinson
                                           !  !  !  !
         Kandelia candel (L.) Druce
                                           !  !  !  !  !    !
         Rhizophora apiculata Bl.
                                     !  !
         Rhizophora mangle L.
                                         !  !  !  !  !  !  !  !
         Rhizophora mucronata Poir.
         Rhizophora racemosa Meyer
                                                     !
         Rhizophora samoensis (Hochr.) Salvoza
                                           !  !  !  !  !  !
         Rhizophora stylosa Griff.
                                           !    !    !
         Rhizophora X lamarckii Montr. (hybrid of R. apiculata
          and R. stylosa)
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                              9


                                        !
         Rhizophora X annamalayana Kathir. (hybrid of
          R. apiculata and R. mucronata )
                                                !
         Rhizophora X selala (Salvoza) Tomlinson (hybrid of
          R. stylosa and R. samoensis)
                                    !  !
         Rhizophora x harrisonii Leechman (hybrid of R.mangle
          and R. stylosa )

                                        !  !  !  !
Rubiaceae    Scyphiphora hydrophyllacea Gaetn. f.

                                      !  !  !  !  !  !  !
Sonneratiaceae  Sonneratia alba J. Smith
                                        !
         Sonneratia apetala Buch.-Ham.
                                        !  !  !  !
         Sonneratia caseolaris (L.) Engler
                                        !  !  !
         Sonneratia griffithii Kurz
                                            !
         Sonneratia lanceolata Blume
                                          !  !
         Sonneratia ovata Backer
                                            !
         Sonneratia X gulngai Duke (hybrid of S. alba
          and S. caseolaris)

                                        !  !  !
Sterculiaceae  Heritiera fomes Buch.-Ham.
                                            !
         Heritiera globosa Kostermans
                                        !  !  !  !
         Heritiera littoralis Dryand. In Aiton
                                              10


    Mangroves have broader ranges along the warmer eastern coastlines of the
Americas and Africa than along the cooler western coastlines. Mangroves prefer a humid
climate and freshwater inflow that brings in abundant nutrients and silt. Mangroves grow
luxuriantly in alluvial soils (loose, fine-textured mud or silt, rich in humus). They are
abundant in broad, sheltered, low-lying coastal plains where topographic gradients are
small and tidal amplitudes are large. Repeatedly flooded but well-drained soils support
good mangrove growth and high species diversity (e.g., Azariah et al., 1992). Mangroves
do poorly in stagnant water (Gopal and Krishnamurthy, 1993).


2. HISTORY AND EVOLUTION


2.1. Historical background

    Mangroves have been known and studied since ancient times. Descriptions by
Nearchus (325 B.C.) and Theophrastus (305 B.C) of Rhizophora trees in the Red Sea and
the Persian Gulf are the earliest known records. Plutarch (70 A.D.) and Abou’l Abass
(1230) wrote about Rhizophora and its seedlings (Macnae, 1968; Chapman, 1976). The
bibliography of mangrove research compiled by Rollet (1981), however, shows only 14
references before 1600, 25 references from the 17th century, 48 references in the 18th
century, and 427 in the 19th century. In contrast, there were 4500 mangrove references
between 1900 and 1975 and approximately 3000 between 1978 and 1997, illustrating the
explosion of interest in mangroves.
    Mangroves have a long historical link with human culture and civilization. In the
Solomon Islands, the bodies of the dead are disposed of and special rites are performed in
the mangrove waters (Vannucci, 1997). In the third century, a Hindu temple to the
mangrove Excoecaria agallocha was erected in south India. Rock carvings show the plant
being worshipped anciently as a “sacred grove” and even today it is believed that a dip in
the holy pond of the temple cures leprosy. The city where this temple is found bears the
name of the mangrove. In Kenya, shrines built in the mangrove forests are worshipped by
the local people, who believe spirits of the shrine will bring death to those who cut the
surrounding trees.
    The Portuguese, probably the first Europeans to visit the mangrove forests of the
Indian Ocean (around the 14th century), learned the traditional Indian technique of rice-
fish-mangrove farming, as demonstrated by letters from the Viceroys to the King of
Portugal. Some six centuries ago, this Indian technology was also transferred by Jesuit and
Franciscan Fathers to the African countries of Angola and Mozambique (Vannucci, 1997).
In the 19th century, the British used the practical knowledge gained over centuries by the
Indians to manage mangroves at Sunderbans for commercial timber production (Vannucci,
1997). An unusually creative use of mangroves is described in a traditional story from
India about two countries at war. The larger country planned to invade their small
neighbors during the night. The smaller nation, which had mangrove forests on its
coastline, plotted to discourage their enemies by placing lighted lamps on the aerial roots
of mangroves. What appeared to be a large flotilla of ships discouraged the invaders and
ended the hostilities.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       11




2.2. Evolution
    The evolutionary history of mangroves remains problematic with a number of
competing theories. Mangroves evolved from terrestrial rather than marine plants.
Mangrove pollen fossils have been found below marine foraminiferan assemblages (i.e., in
the lower deposits of estuarine environments) suggesting the evolution of these plants from
a non-marine habitat to an estuarine habitat (Srivastava and Binda, 1991). In the distant
past, these land plants adapted to brackish water and became the “core” mangrove flora.
The diversity of mangroves is much higher in the Indo-West Pacific than in the Western
Atlantic and Caribbean. Two competing hypotheses have been presented to explain this
pattern. The center-of-origin hypothesis suggests that all mangrove taxa first appeared in
the Indo-West Pacific and subsequently dispersed to other regions. The vicariance
hypothesis, on the other hand, states that all mangroves originated around the Tethys Sea.
Continental drift then isolated the flora in different regions of the earth where
diversification created distinct faunas.
    Ellison et al. (1999) evaluated these two hypotheses using 1) a review of the
mangrove fossil record, 2) a comparison of modern and fossil distributions of mangroves
and mangrove-associated gastropods, 3) an analysis of species-area relationships of
mangroves and gastropods, 4) an analysis of nestedness patterns of individual plants and
gastropod communities, and 5) an analysis of nestedness patterns of individual plants and
individual gastropod species. The evidence from all 5 analyses supported the vicariance
hypothesis, suggesting a Tethyan origin of mangroves. This argues that the much higher
diversity of mangroves in the Indo-West Pacific relates to conditions there that favored
diversification. For example, the continual presence of extensive wet habitat may have
allowed more species to make the transition from terrestrial to brackish-water habitats. The
Atlantic, Caribbean or and East Pacific all saw periods of drying which could have
prevented such adaptation. Ricklefs and Latham (1993) suggest that limited dispersal,
combined with the closure of the Tethys connection to the Atlantic Ocean in the mid-
Tertiary, restricted most mangrove taxa to the Indo-Pacific.
    Studies of mangrove biochemistry and genetics should provide further evidence
concerning mangrove evolution and dispersal. For example, Dodd et al. (1998) found
significant genetic differentiation between mangroves in eastern and western Atlantic
provinces. Three species from western Africa showed significantly greater lipid diversity
and longer carbon chains than conspecifics from eastern South America, suggesting that
the western Atlantic mangroves show derived characteristics. The authors concluded that
this evidence suggests it is unlikely that Atlantic mangroves dispersed from the Tethys via
the Pacific.
    Mangroves are quite old, possibly arising just after the first angiosperms, around
114 million years ago (Duke, 1992). Avicennia and Rhizophora were probably the first
genera to evolve, appearing near the end of the Cretaceous period (Chapman, 1976). Pollen
records provide important information about subsequent radiation. Fossil pollen from
sediments in the Leizhou Peninsula, China suggest that mangroves expanded from south to
north, reaching their northern limit on the Changjiang Delta by the mid-Holocene (Y.
Zhang et al., 1997). A similar study of pollen from late Holocene samples in Bermuda
suggests that mangroves were established there in the last 3000 years, when sea level rise
decreased from 26 to 7 cm per century (J.C. Ellison, 1996).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       12


     A detailed study of pollen records from Mexico, the Antilles, Central America and
northern South America (Graham, 1995) show that neotropical environments were first
occupied by Acrostichum, Brevitricolpites variabilis, Nypa and Pelliceria in the early
Eocene, about 50 million years ago. Avicennia appeared in this region in the late Miocene
(about 10 million years ago). Six mangrove species and three associated genera were
present by the middle Pliocene (3.5 million years ago), and fifteen plant genera were
present by the Quaternary period. Twelve additional species were added during the
Cenozoic to produce the present-day assemblage of about 27 genera of mangroves and
associated plants (Rico-Gray, 1993; Graham, 1995).
     Continental drift produced massive mixing and dispersal of genes in geologically
recent times, greatly enhancing evolutionary processes. Though mangroves evolved in the
tropics, one species, Avicennia marina, is found in temperate latitudes, particularly in the
southern hemisphere (Saenger, 1998). This genus is of a western Gondwanan origin with
the subsequent radiation of several taxa facilitated by tectonic dispersal of southern
continental fragments (Duke, 1995). Mangrove fossils have clearly provided valuable
information about prehistorical mangrove evolution and dispersal. However, Burnham
(1990) cautions that reconstructions based on organic remains can differ substantially
depending on the mangrove parts studied (e.g., fruits and seeds vs. leaf litter).
     Mangrove ecosystems, in general, are dynamic, undergoing changes on time scales
of 10 - 104 y(Woodroffe, 1992). Indeed fossil mangroves are often found in regions where
   2

they no longer exist: in Texas, USA (Westgate and Gee, 1990; Westgate 1994), west
Africa (Marius and Lucas, 1991), Hungary (Nagy and Kokay, 1991), India (Bonde, 1991;
Barni and Chanda, 1992), the Chao-Shan Plain of China (Z. Zheng, 1991), and Western
Australia (Kendrick and Morse, 1990), for example.
     Historical changes in mangrove distributions can reveal details about paleoclimates
and sea-level changes (Somboon, 1990; Khandelwal and Gupta, 1993; Y. Zhang and
Wang, 1994; Plaziat, 1995; Saito et al., 1995; Lezine, 1996; W. Zhang and Huang 1996; Y.
Zhang et al., 1997). For example, in the equatorial Pacific Ocean, there are alternating reef
and mangrove fossils in upper Miocene and lower Pliocene deposits (Cronin et al., 1991).
Similarly, Holocene sediments from the Maya Wetland of Belize indicate that mangrove
peat filled the lagoon by 4800 y ago (Alcala-Herrera et al., 1994). These patterns may
reflect fluctuating sea levels or large-scale climatic shifts. In Poverty Bay, New Zealand,
the presence of Avicennia marina var. resinifera during the early to mid-Holocene suggests
that the area then had a frost-free climate (Mildenhall, 1994). The mangrove fossil record
is clearly an area where continued research has the potential for providing significant
information, not only about the history of these unique plants, but also about the recent
history of the earth.

3. BIOLOGY OF MANGROVES
3.1. Taxonomy and genetics

3.1.1. Taxonomy
    Tomlinson (1986) recognized three groups of mangroves: major mangrove species,
minor mangrove species and mangrove associates. The major species are the strict or true
mangroves, recognized by most or all of the following features: 1) they occur exclusively
in mangal, 2) they play a major role in the structure of the community and have the ability
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        13


to form pure stands, 3) they have morphological specializations - especially aerial roots
and specialized mechanisms of gas exchange, 4) they have physiological mechanisms for
salt exclusion and/or excretion, 5) they have viviparous reproduction, and 6) they are
taxonomically isolated from terrestrial relatives. The strict mangroves are separated from
their nearest relatives at least at the generic level, and often at the sub-family or family
level.
     The minor mangrove species are less conspicuous elements of the vegetation and
rarely form pure stands. According to Tomlinson (1986), the major mangroves include 34
species in 9 general and 5 families. The minor species contribute 20 additional species in
11 genera and 11 families for a total of 54 mangrove species in 20 genera and 16 families.
Duke (1992), on the other hand, identified 69 mangrove species belonging to 26 genera in
20 families. One family falls in the fern division (Polypodiophyta); the remainder are in the
Magnoliophyta (angiosperms). Families containing only mangroves are the
Aegialitidaceae, Avicenniaceae, Nypaceae and Pellicieraceae. Two orders (Myrtales and
Rhizophorales) contain 25% of all mangrove families. By reconciling common features
from Tomlinson (1986) and Duke (1992), we recognize 65 mangrove species in 22 genera
and 16 families (Table I).
     There are a number of problems with mangrove taxonomy (Duke, 1992) and many
of these are based on hybridization between described species. For instance, the systematic
distinction between Rhizophora mucronata in eastern Africa, R. stylosa in Australia, and
their putative hybrids is unclear. Rhizophora lamarckii, which occurs in New Caledonia,
Papua New Guinea and Queensland, Australia, is a sterile F1 hybrid between R. apiculata
and R. stylosa. Rhizophora x annamalayana, found in a south Indian mangrove forest, was
first identified as R. lamarckii but has since been reidentified as a new species hybrid
between R. mucronata and R. apiculata (Kathiresan, 1995a). Some hybrids, like
Rhizophora x harrissoni, can not be confirmed with wax chemistry (Dodd et al., 1995).
Molecular analyses may help eventually resolve the taxonomic problems. For example,
DNA sequence data from the chloroplast gene rbcL indicate that the Rhizophoraceae
belongs not to the Myrtales, but to a rosid clade that includes the families Euphorbiaceae,
Humiriaceae and Malphighiaceae (Conti et al., 1996).

3.1.2. Genetic variation
    There is significant inter- and intraspecific variability among mangroves. For
example, physiological differences have been identified between West African and
Western Atlantic Avicennia germinans (Saenger and Bellan, 1995) and distinct
chemotypes have been described for A. germinans and Rhizophora (Corredor et al., 1995;
Dodd et al., 1995; Rafii et al., 1996). Variability may result from genotypic differences or
from phenotypic responses to local environments. Mean leaf area of Rhizophora mangle in
Mexico, for example, is positively correlated with annual precipitation and negatively
correlated with latitude. This morphological response to local conditions may allow the
trees to maximize their photosynthetic efficiency (Rico-Gray and Palacios-Rios, 1996a).
Similarly, leaf area indices can be used to differentiate Rhizophora mangle from basin and
dwarf forest types in southeast Florida, USA (Araujo et al., 1997). In contrast, variation in
Rhizophora mangle flower morphology appears to have a genetic basis. Dominguez et al.
(1998) found significant differences between populations on the Pacific and Atlantic coasts
of Mexico, among populations on each coast, and within individual populations. They
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        14


hypothesized that frequent extinctions, followed by recolonization of a few individuals, has
produced genetic differentiation.
     Genetic variability has been clearly demonstrated through biochemical markers like
iridoid glycosides (Fauvel et al., 1995), foliar leaf waxes (Dodd et al., 1995, 1998; Rafii et
al., 1996), and isoenzymes (Duke, 1991). It is also evident in differences in length and
volume of chromosomes (Das et al., 1994). Lakshmi et al. (1997) measured intraspecific
genetic variability in Acanthus ilicifolius through DNA-based molecular markers that are
insensitive to environmental influences (i.e., random amplified polymorphic DNAs and
restriction fragment length polymorphisms). They found 48 genotypes in eight distinct
populations. There were no differences in chromosome number (2n = 48). Genetic
polymorphism is even higher in Excoecaria agallocha. The E. agallocha polymorphism is
independent of morphological and sexual differences (Parani et al., 1997).
     Changes in gene frequency, such as those produced by inbreeding, can lead to
genetic differentiation. Inbreeding may result if pollen are shed before the flower opens
(Lowenfeld and Klekowski, 1992). If inbreeding is prevalent, a mangrove forest may be a
virtually monospecific stand with little genetic diversity. Pollination by bees produces
geitonogamous selfing in Kandelia candel. However, there is little genetic differentiation
among 13 populations along the coastlines of Hong Kong, indicating that dispersion of
propagules is sufficient to maintain high levels of gene flow in this species (Sun et al.,
1998). In contrast, genetic differentiation, has led to subspeciation in Avicennia marina
(Duke, 1991, 1995). It has been assumed that Avicennia propagules commonly move long
distances. However, allozyme studies suggest that Avicennia species in the Indo-West
Pacific and eastern North America have limited gene flow. This may indicate that true
dispersal distances are much shorter than has been commonly believed (Duke et al.,
1998b).
     Gene mutations can also cause species divergence. One or 2 gene mutations are
needed for biochemical differences, 5-10 for physiological changes, >10 for morphological
variations and >100 for taxonomic changes (Saenger, 1998). A single recessive gene
causes albinism in Rhizophora seedlings. This albino mutation is in the nuclear genome but
has a profound effect on ultrastructure of the chloroplasts (Klekowski et al., 1994a).
Pigment fingerprint studies of chlorophyll-deficient mutants show that most albino
genotypes are deficient in chlorophylls, xanthophylls, and carotenes (Corredor et al.,
1995). Recent studies of post-zygotic mutations reveal that fewer than 0.1% of the
Rhizophora in Puerto Rico exhibit somatic mutations. These mutations are often manifest
in shoot apices as complete or partial periclinal chimeras (Klekowski et al., 1996). Rates of
both mutation and outcrossing vary among mangrove populations. For instance, the Puerto
Rican Rhizophora are more outcrossed and have lower mutation rates for chlorophyll-
deficiency than Florida Rhizophora.

3.13. Tissue Culture
    There have been few studies of tissue culture in mangroves. This is because
explants frequently turn brown or black shortly after isolation, with tissue death usually
following (Kathiresan, 1990, 1994). The high tannin and phenol content of mangroves may
be responsible for the browning problem (Kathiresan and Ravi, 1990; Ravi and Kathiresan,
1990). Antioxidants can prevent phenolic browning in explants collected during the
monsoon season (Kathiresan and Ravikumar, l997).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                         15


    Callus induction has been achieved in Sonneratia apetala and Xylocarpus
granatum by supplementing the medium with double strength vitamins (Kathiresan and
Ravikumar, 1997). Baba and Onizuka (1997) have improved techniques for callus
induction and initiation of redifferentiation in the callus of Bruguiera gymnorrhiza,
Kandelia candel, Pemphis acidula and Rhizophora stylosa. Adventitious roots were
produced in P. acidula, but neither adventitious buds nor roots could be induced in the
remaining species.
    Researchers are currently working to identify and micropropagate unique plant
genotypes for commercial purposes. Mangals may provide good raw material for such
work. For instance, in vitro multiplication of the salt-marsh Sesuvium portulacastrum,
associated with Indian mangroves, has been achieved by axillary bud culture (Kathiresan,
1994; Kathiresan et al., 1997). In vitro cell cultures of this plant synthesize antibacterial
substances in higher quantities than do the intact plants, demonstrating the potential of
these systems for production of valuable metabolites (Kathiresan and Ravikumar, 1997).
    Cell protoplast fusion techniques may allow us to transfer salinity tolerance from
mangrove plants to non-salt-tolerant species (Swaminathan, 1991). Methods for extracting
and preparing protoplasts from tissue cultures of Bruguiera gymnorrhiza have been
developed by Eguchi et al. (1995). Sasamota et al. (1997) have done similar work with the
cotyledons of Avicennia marina and A. lanata. Such creative tissue culture work may
allow researchers to better understand, and make use of, the unique characteristics of
mangroves.



3.2. Morphology and anatomy

3.2.1. Root anatomy
    Mangroves are highly adapted to the coastal environment, with exposed breathing
roots, extensive support roots and buttresses, salt-excreting leaves, and viviparous water-
dispersed propagules. These adaptations vary among taxa and with the physico-chemical
nature of the habitat (Duke, 1992). Perhaps the most remarkable adaptations of the
mangroves, however are the stilt roots of Rhizophora, the pneumatophores of Avicennia,
Sonneratia and Lumnitzera, the root knees of Bruguiera, Ceriops and Xylocarpus and the
buttress roots of Xylocarpus and Heritiera. The roots of many mangroves do not penetrate
far into the anaerobic substrata. Instead, the trees produce profuse lateral roots for support.
Their effectiveness is well illustrated by the tallest mangrove trees, found in Ecuador,
which attain heights of more than 60 m and may be 100 yold (Emilio, 1997).
    The specialized roots are important sites of gas exchange for mangroves living in
anaerobic substrata. The exposed surfaces may have numerous lenticels (loose, air-
breathing aggregations of cells; Tomlinson, 1986). Avicennia possesses lenticel-equipped
pneumatophores (upward directed roots) through which oxygen passively diffuses. The
lenticels may be closed, partially opened or fully opened, depending on environmental
conditions (Ish-Shalom-Gordon and Dubinsky, 1992). The spongy pneumatophores are
generally short (< 30 cm), but grow much larger and become more numerous in Avicennia
marina living in anaerobic and oil-polluted conditions. This phenotypic response
apparently increases surface area for gas exchange (Saifullah and Elahi, 1992). In
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        16


Sonneratia, the pneumatophores may be 3 m long and stout from heavy secondary
thickening (Tomlinson, 1986).
     Oxygen may also pass through non-lenticellular portions of the pneumatophores.
Horizontal structures (subrisules) may be important in air exchange, particularly in rapidly
growing pneumatophores where the newly formed tip lacks lenticels (Hovenden and
Allaway, 1994). Pneumatophores are normally unbranched. However, following the 1991
Gulf War, mangroves in the Arabian Gulf began developing branched pneumatophores and
adventitious roots (Boeer, 1993).
     The general structure of mangrove roots is similar to that of most other vascular
plants. They typically have a root cap, lateral roots arising endogenously, exarch
protoxylem, and alternating strands of primary phloem and xylem. Many also have an
enlarged polyarch stele with a wide parenchymatous medulla. Aerial roots are modified for
life above ground. Compared to the underground roots, they have an exaggerated zone of
elongation behind the apical meristem (Tomlinson, 1986). They also have significant
secondary thickening (similar to the stems). When the aerial roots reach the ground, they
shift to having a short elongation zone and little to no secondary growth. They also become
spongy to adapt to sub-soil existence. In Rhizophora, the roots become thinner and form
“capillary rootlets” with a simple diarch stele and a narrow cortex. Like aquatic plants, true
mangroves lack root hairs. Hence, the endodermis is an effective absorbing layer
(Tomlinson, 1986).

3.2.2. Wood anatomy
    Tomlinson (1986) has summarized the unique anatomical features of mangrove
woods. Growth rings are conspicuously anomalous (as in Avicennia; Das and Ghose, 1998)
or completely absent. Hence, aging trees is difficult. Duke and Pinzon (1992) suggest that
leaf scar nodal number is a better way to estimate the age of Rhizophora seedlings.
    Mangrove wood has special features that enable the trees to overcome the high
osmotic potential of seawater and the transpiration caused by high temperatures. There are
numerous narrow vessels running through the wood. These range in density from 32 •
mm-2 in Excoecaria to 270 • mm-2 in Aegiceras (Das and Ghose, 1998). The vessels help
create high tensions in the xylem since a slight decrease in vessel diameter produces a
disproportionally large increase in flow resistance (Scholander et al., 1964, 1965;
Tomlinson, 1986). The vessel elements, which form the vessels, normally have simple
perforation plates (Tomlinson, 1986). However, mangroves in the family Rhizophoraceae
(except Kandelia candel) have scalariform perforation plates.
    Water conduction through wood is strongly influenced by size and distribution of
the vessels. Water moves most quickly through ring-porous woods in which the largest
vessels are in the outermost growth layer. Conduction is much slower in diffuse-porous
woods where vessels are more uniform in size and distribution. The wood of most
mangroves is diffuse-porous but Aegialitis rotundifolia has ring-porous wood (Das and
Ghose, 1998).
    Wier et al. (1996) studied wound repair in Rhizophora mangle. A closing layer
isolates necrotic tissue within 17 d, and the wound is completely enclosed by periderm by
52 d. Isolation of the damage site and development of wound periderm may prevent spread
of pathogens to undamaged tissues.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        17


3.2.3. Leaf anatomy
     Mangrove leaves are almost leathery with obscure leaf veins (there are no vein
sheaths). The cuticle is thick and smooth with small hairs, giving the plant a glossy
appearance. The leaves are of moderate size and are arranged in a modified decussate
(bijugate) pattern with each pair at an angle less than 180° to the preceding pair. This
arrangement reduces self-shading and produces branch systems that fill space in the most
photosynthetically efficient way (Tomlinson, 1986). The leaves generally show
dorsiventral symmetry though isolateral leaves are also found in Kandelia candel,
Sonneratia apetala and Phoenix paludosa (Das et al., 1996).
     Six types of stomata are known from mangrove leaves. These differ in their
arrangement of guard cells and subsidiary cells. In most species, a horn or beak-like
cuticular outgrowth covers either the outer side of the stomatal pore or both the inner and
outer sides. These structures reduce stomatal transpiration (Das and Ghose, 1993), which is
important given the high solute concentration of the water and the “physiological drought”
the trees experience. Heritiera fomes has deeply sunken stomata covered by trichomes. The
leaves in this species also have a palisade-spongy ratio that is small compared to other
halophytes (Das et al., 1995).
     Mangrove leaves have specialized idioblast cells including tannin cells
(Rhizophoraceae), mucous cells (Rhizophora, Sonneratia), crystalliferous cells
(Rhizophoraceae), oil cells (Osbornia) and laticifers (Excoecaria; Tomlinson, 1986). In
general, the leaves lack bundle sheath fibres and bundle sheath extensions, but possess
enlarged tracheids terminating in vein endings. Branched sclereids are abundant and well
developed in Aegiceras, Rhizophora, Sonneratia and Aegialitis. The sclereids may give
mechanical support to leaves or discourage herbivores. Both sclereids and tracheids may
also be involved in water storage (Tomlinson, 1986). Water is also stored in colourless,
non-assimilatory water-storage tissue that is hypodermal in dorsiventral leaves, but is deep-
seated in the extensive mesophyll region of isolateral leaves. In some species, the thick
layer of non-assimilatory tissue occurs in front of the assimilatory cells. This back scatters
incoming light, creating a gradient that may help the plant capture weak light, increasing
photosynthetic efficiency (Koizumi et al., 1998).
     Yoshihira et al. (1992) studied the distribution of pigments in mangrove leaves.
They found that different species concentrated the pigments in different parts of the leaves.
In Aegiceras corniculatum, the highest concentration of carotenoids and chlorophylls was
in the light-harvesting complex. In Rhizophora apiculata, however, chlorophyll was
concentrated in the chloroplast reaction center. The chlorophyll-binding proteins (including
the functional cytochrome B 6/f complex and the protein kinases) were found in the
thylakoid membranes in Bruguiera gymnorrhiza and Kandelia candel

3.2.4. Seed and seedling anatomy
    Avicennia marina forms endosperm haustoria during early embryonic
histodifferentiation. Once the growth phase is initiated, subsequent embryonic
development is extra-ovular. The mature seed, therefore, is enclosed by a pericarp that
originates entirely from the ovary wall. From the end of histodifferentiation until the
mature seeds are abscised, cotyledon cells become highly vacuolated and contain large
amounts of soluble sugars, which constitute the major nutrient reserves of the mature seed
(Farrant et al., 1992).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                         18


    Incipient phellogen usually develops toward the radicle end of mangrove seedlings
and masks the chlorophyllous tissue. Tannin cells are present in the aerenchymatous tissue,
stone cells are present in the outer cortex, and trichosclereids appear in the cortex and
medulla. Since the epidermis lacks stomata, numerous lenticels facilitate gas exchange.
    In experiments with six mangrove species, Youssef and Saenger (1996)
demonstrated that the seedlings have special features that allow them to tolerate flooding
and facilitate rhizosphere oxidation. Lacunae in the ground tissue constrict air flow
passages, conserving oxygen and enabling the mangrove to maintain aerobic metabolism
during periods of flooding. Variations in this anatomical feature are responsible for species
differences in tolerance to flood stress.

3.3. Physiology

3.3.1. Salt regulation
     Mangroves are physiologically tolerant of high salt levels and have mechanisms to
obtain fresh water despite the strong osmotic potential of the sediments (Ball, 1996). They
avoid heavy salt loads through a combination of salt exclusion, salt excretion, and salt
accumulation. For example, Rhizophora, Bruguiera, and Ceriops all possess ultrafilters in
their root systems. These filters exclude salts while extracting water from the soil. Other
genera (e.g., Avicennia, Acanthus, Aegiceras) take some salt up, but excrete it through
specialized salt glands in the leaves (Dschida et al., 1992; Fitzgerald et al., 1992). The salt-
excreting species allow more salt into the xylem than do the non-excretors, but still
exclude about 90% of the salts (Scholander et al., 1962, Azocar et al., 1992). Salt excretion
is an active process, as evidenced by ATPase activity in the plasmalemma of the excretory
cells (Drennan et al., 1992). The process is probably regulated by leaf hypodermal cells,
which may store salt as well as water (Balsamo and Thomson, 1995).
     Species of Lumnitzera and Excoecaria accumulate salts in leaf vacuoles and
become succulent. Salt concentrations in the sap may also be reduced by transferring the
salts into senescent leaves or by storing them in the bark or the wood (Tomlinson, 1986).
As water salinity increases, some species simply become increasingly conservative in their
water use, thus achieving greater tolerance (Ball and Passioura, 1993). In south Florida,
Rhizophora mangle decreases its salt stress by using surface water as its sole water source.
In the wet season, the fine root biomass increases in response to decreased salinity of the
surface waters, directly enhancing the uptake of low-salinity water (Lin and Sternberg,
1994).
     Most mangrove species directly regulate salts. However, they may also accumulate
or synthesize other solutes to regulate and maintain osmotic balance (Werner and Stelzer,
1990; Popp et al., 1993). For example, Aegiceras corniculatum, Aegialitis annulata and
Laguncularia racemosa accumulate mannitol and proline (Polania, 1990). Avicennia
marina accumulates glycine betaine, asparagine and stachyose (Ashihara et al., 1997).
Sonneratia alba synthesizes purine nucleotides that help it adapt to salt loads of 100 mM
NaCl (Akatsu et al., 1996). To facilitate the flow of water from root to leaves, the water
potential at the leaves is held lower (-2.5 to -6 MPa) than in the roots (-2.5 MPa;
Scholander et al., 1964).
     Because mangrove roots exclude salts when they extract water from soil, soil salts
could become very concentrated, creating strong osmotic gradients (Passioura et al., 1992).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        19


However, viscous, polymeric substances in the sap limit flow rate and decrease
transpiration (Zimmermann et al., 1994). This, combined with high water-use efficiency,
slows the rate of water uptake and prevents salts from accumulating in the soil surrounding
the roots. This helps the mangroves conserve water and regulate internal salt
concentrations (Ball and Passioura, 1993; Ball, 1996). Low transpiration and slow water
uptake, however, are not characteristic of all mangrove species. Becker et al. (1997)
measured relatively high transpiration rates in both Avicennia alba and Rhizophora
apiculata.
    Transpiration rates vary with season, being higher in the dry season than in the wet
season in Bruguiera cylindrica (Herppich and Von Willert, 1995; Hirano et al., 1996). This
corresponds to changes in stomatal movement. The oscillatory behaviour of Avicennia
germinans stomata is affected by any factor that changes hydraulic flow through the plant.
This includes increases in vapour pressure deficit and osmotic potential of the substrata
(Naidoo and Von-Willert, 1994).
    Fukushima et al. (1997) studied the effects of salt on sugar catabolism in leaves and
roots of Avicennia marina. They showed that sugar catabolic pathways are different in
roots and leaves. Over 50% of the 14C-labeled sucrose the gave the plants was incorporated
into an unidentified sugar in the leaves. The remainder appeared in the roots as glucose,
fructose and sucrose. Neither pathway was significantly affected by salt levels.

3.3.2. Photosynthesis
     Mangroves show characteristic C3 photosynthesis. Basak et al. (1996) found
significant intra- and interspecific variation in photosynthetic activity of 14 mangrove
species, suggesting that the rates of photosynthesis may have an underlying genetic basis.
This possibility is supported by observations that the photosynthetic rate of Bruguiera is
under direct internal control and is not influenced by stomatal activity induced by changes
in salinity or light (Cheeseman et al., 1991; Cheeseman, 1994).
     In contrast, other researchers have shown that photosynthetic rates of some species
are strongly affected by environmental conditions. For example, low salinity conditions
reduce carbon losses in Avicennia germinans and Aegialitis annulata and lead to greater
CO2 assimilation (Naidoo and Von-Willert, 1995). Fluctuating soil salinities lead to
significantly lower intercellular CO2 concentration and reduced photosynthesis in scrub
forests of south Florida (Lin and Sternberg, 1992). The stunted mangroves in these habitats
have much lower canopies, more main stems and smaller leaves than mangroves in fringe
forests that experience less salinity variability. Steinke and Naidoo (1991) also
demonstrated experimentally that temperature affects the photosynthetic rate of Avicennia
marina. Temperature-induced changes in the relative rates of photosynthesis and
respiration, in turn, influence overall growth rates.
     Strong sunlight can also reduce mangrove photosynthesis through inhibition of
Photosystem II (Cheeseman et al., 1991). The photosynthetic rates of mangroves saturate
at relatively low light levels despite their presence in high sunlight tropical environments.
The fairly low photosynthetic efficiency may be related to the concentration of zeaxanthin
pigments in the leaves (Lovelock and Clough, 1992). To prevent damage to the
photosystems, the mangroves dissipate excess light energy via the xanthophyll cycle
(Gilmore and Bjorkman, 1994) and through the conversion of O2 to phenolics and
peroxidases (Cheeseman et al., 1997).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       20


    Kathiresan and Moorthy (1994a) and Kathiresan et al. (1996c) demonstrated that
application of aliphatic alcohols can have a major stimulatory effect on mangrove
photosynthesis. Treatment with triacontanol (a long-chain aliphatic alcohol) increased the
photosynthetic rate of Rhizophora apiculata by 225%. A similar treatment with methanol
(a short-chain aliphatic alcohol) increased photosynthesis in R. mucronata by 612%.


3.4. Biochemistry

     Mangroves are biochemically unique, producing a wide array of novel natural
products. Excoecaria agallocha, for example, exudes an acrid latex that is injurious to the
human eye, hence its designation as “the blinding tree”. The latex is toxic to a variety of
marine organisms (Kathiresan and Thangam, 1987; Kathiresan et al., 1990b) and has
sublethal effects on the rice-field crab Oziotephusa senex senex, in which exposure
decreases whole-animal oxygen consumption and inhibits the ATPase system in gill and
hepatopancreas tissues (R. Ramamurthi et al., 1991). Soil bacteria and yeasts degrade the
toxic latex, preventing its accumulation in the mangal (Reddy et al., 1991).
     Researchers have isolated a variety of other mangrove compounds including
taraxerol careaborin and taraxeryl cis-p-hydroxycinnamate from leaves of Rhizophora
apiculata (Kokpol et al., 1990); 2-nitro-4-(2’-nitroethenyl phenol) from leaves of
Sonneratia acida (Bose et al., 1992); alkanes (46.7-97.9% wax) and triterpenoids (53.3%
wax) from leaves of Rhizophora species (Dodd et al., 1995); and iridoid glycosides from
leaves of Avicennia officinalis and A. germinans (Fauvel et al., 1995; Sharma and Garg,
1996). C.K.Rao et al. (1991) found arsenic in mangroves from the Goa Coast.
     Mangroves are also rich in polyphenols and tannins (Kathiresan and Ravi, 1990;
Ravi and Kathiresan, 1990; Achmadi et al., 1994). The levels of these substances may vary
seasonally (Basak et al., 1998), but older data should be interpreted cautiously since
standard methods for measuring tannins are very inaccurate for mangrove leaves (Benner
et al., 1990a).
     Substances in mangroves have long been used in folk medicine to treat disease
(Bandaranayake, 1998). Extracts have proven activity against human, animal and plant
pathogenic viruses including human immuno-deficiency virus (Premanathan et al., 1996),
Semliki forest virus (Premanathan et al., 1995), Tobacco Mosaic virus (Padmakumar and
Ayyakannu, 1997), Vaccinia virus (Premanathan et al., 1994a), Encephalomyocarditis
virus (Premanathan et al., 1994b), New castle disease virus (Premanathan et al., 1993), and
Hepatitis-B viruses (Premanathan et al., 1992). A few mangrove species, particularly those
belonging to the family Rhizophoraceae, show particularly strong antiviral activity
(Premanathan et al., 1992; Kathiresan et al., 1995a). Purified active fractions like acid
polysaccharides (galactose, galactosamine, glucose and arabinose) show potent anti-HIV
activity (Premanathan et al., 1999).
     Other unique mangrove biochemicals have potential commercial applications
(reviewed by Kathiresan, 2000). For example, mangrove extracts kill larvae of the
mosquitoes Anopheles stephensi (Thangam and Kathiresan, 1988), Culex tritaeniorhynchus
(Thangam and Kathiresan, 1989), Aedes aegypti (Thangam and Kathiresan, 1991, 1992a,
1994), and Culex quinquefasciatus (Thangam and Kathiresan, 1997). A pyrethrin-like
compound in stilt roots of Rhizophora apiculata shows strong mosquito larvicidal activity
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                          21


(Thangam, 1990). Smoke from burned extracts repels and kills both Aedes aegypti
(Thangam et al., 1992) and Culex quinquefasciatus (Thangam and Kathiresan, 1992b) and
extracts applied directly to human skin repel adult Aedes aegypti (Thangam and
Kathiresan, 1993a).
    Phenols and flavonoids in mangrove leaves serve as UV-screening compounds.
Hence, mangroves tolerate solar-UV radiation and create a UV-free, under-canopy
environment (Moorthy, 1995). These substances also contribute to a black tea that can be
extracted from mangrove leaves (Kathiresan, 1995b). The “mangrove tea” is rich in
theaflavin, the substance responsible for the briskness and colour of tea. The tea, which
shows no mammalian toxicity, can be improved by UV irradiation (Kathiresan and
Pandian, 1991, 1993, Kathiresan, 1995b).
    Moorthy and Kathiresan, (1997a) proposed a physiological grouping of mangrove
species based on pigments, which may differ significantly among species (Basak et al.,
1996). Pigments concentrations may also vary with environmental conditions and season.
For example, Menon and Neelakantan (1992) found that total chlorophyll content was
positively related to light levels. Oswin and Kathiresan (1994) found that mangrove
chlorophyll and carotenoid levels, in general, are high during the summer but anthocyanin
levels are highest in the monsoon months. Flavonoids increase during the premonsoon
period.


3.5. Pollination biology

     Mangroves have both self-pollinating and cross-pollinating mechanisms that vary
with species. For example, Aegiceras corniculatum and Lumnitzera racemosa are self-
pollinated. Avicennia officinalis is self-fertile, but can also cross-fertilize (Aluri, 1990). In
Avicennia marina, protandry makes self-pollination of an individual flower unlikely.
However, some fruits are set even when flowers are experimentally bagged to prevent
cross-pollination (between 4 and 41% of cross-pollinated flowers set fruit). Fruit abortion
is significantly higher in self-fertilized treatments, indicating some inbreeding depression
(Clarke and Myerscough, 1991a). There is a similar distinct trend for self-incompatibility
in Rhizophora, Ceriops and Sonneratia. This pattern is less clear in Bruguiera and
Kandelia (Ananda Rao, 1998).
     Mangroves are pollinated by a diverse group of animals including bats, birds, and
insects. Pollen is deposited on the animals as they deeply probe the flowers looking for
nectar; they subsequently transfer the pollen grains to the stigma of another flower. The
identity of the pollinators differs from species to species. Lumnitzera littorea, for example,
is pollinated primarily by birds while L. racemosa and small-flowered Bruguiera species
are pollinated by insects (Tomlinson, 1986). Sunbirds visit and may pollinate Acanthus
ilicifolius (Aluri, 1990) and large-flowered Bruguiera hainesii (Noske, 1993, 1995). Birds
are particularly important pollinators in the dry season when absence of terrestrial plant
flowers causes them to turn to mangroves as a food source.
     Bats are the major pollinators for Sonneratia, which opens its flowers to expose the
powdery stamens in the late night/early morning hours. If there are no bats, hawk moths
become the primary nighttime pollinators (Hockey and de Baar, 1991). Two lycaenid
butterflies may be important in the pollination of mangroves in Brisbane, Australia where
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       22


their abundance is directly correlated with the abundance of mangrove flowers (Hill,
1992). Bees regularly visit and pollinate species of Avicennia, Acanthus, Excoecaria,
Rhizophora, Scyphipora, and Xylocarpus. Some wasps and flies are highly dependent on
mangroves for nesting and are particularly important pollinators of Ceriops decandra,
Kandelia candel and Lumnitzera racemosa (Tomlinson, 1986). Rhizophora species
produce prolific amounts of pollen and are mainly wind-pollinated, though the stigma has
no special modifications to capture the wind-borne pollen (Tomlinson, 1986).


3.6. Reproduction, dispersal and establishment
    Bhosale and Mulik (1991) described four methods of mangrove reproduction:
viviparity, cryptoviviparity, normal germination on soil, and vegetative propagation.
Vivipary, the precocious and continuous growth of offspring while still attached to the
maternal plant, is a unique adaptation to shallow marine habitats (Thomas and Paul, 1996).
True viviparous species remain attached to the maternal plant for a full year while
cryptoviviparous offspring are only attached for 1-2 months (Bhosale and Mulik, 1991).
S.M. Smith and Snedaker (1995a) suggest that viviparous reproductive patterns allow
seedlings to develop some salinity tolerance before being released from the parent tree.
Figure 2c illustrates propagules of Rhizophora still attached to the parent.
    The timing of mangrove reproduction depends on local environmental conditions
and may differ broadly over the range of a species. For example, Duke (1990) found that
flowering in Avicennia marina occurred 6 months earlier in Papua New Guinea than in
Southern Australia and New Zealand. The period from flowering to fruiting was 2-3
months in the northern tropical site but stretched to 10 months in the southern temperate
locations. Flowering appeared to be controlled by daylength while air temperature set the
period for fruit maturation.
    Phytohormones are important in development, growth, and dispersal of mangrove
seeds, which may undergo no maturation drying, and remain metabolically active
throughout development (Farrant et al.; 1992, 1993). Phytohormones, like cytokinin
(particularly zeatin riboside) accumulate in both axes and cotyledons during reserve
accumulation. The level of abscissic acid (ABA) in the embryo stays low during this
period, making them sensitive to desiccation (though their dehydration tolerance increases
with development; Farrant et al., 1993). ABA levels in the pericarp increase throughout
seed development; the ABA in the pericarp may prevent precocious germination.
Farnsworth and Farrant (1998) suggest that ABA concentrations represent a trade-off
between salinity adjustment by the parental plant and developmental demands of the
embryo. Other biochemicals may be compartmentalized in the seeds. Mature propagules of
Rhizophora species exhibit high chlorophyll levels in the hypocotyl and high polyphenol
content in the radicle regions (Kulkarni and Bhosale, 1991).
    S.M. Smith et al. (1995) investigated the role of hormones in controlling flotation
and the development of roots and shoots in Rhizophora mangle propagules. Application of
gibberellic acid (GA3) caused the propagules to float horizontally, but painting with
naphthalene acetic acid (NAA) produced vertically floating propagules. NAA promoted
root elongation while GA3 enhanced stem elongation and leaf expansion (S.M. Smith et
al., 1996). A variety of hormones and chemicals (e.g., NAA, IBA, IAA, GA3, phenolics,
methanol, boric acids, triacontanol) promote root growth in propagules of other
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                              23


Rhizophora and Avicennia species (Kathiresan and Thangam, 1990b; Kathiresan and
Moorthy, 1992, 1994a,b,c,d; Kathiresan et al., 1990a, 1994b, 1996b).
     Mangrove propagules have an obligate dispersal phase of several weeks before the
radicle extends for root development. If, however, the propagules do not contact the
sediment, they remain viable in seawater for several months (Clarke, 1993). Dispersal of
propagules depends on their buoyancy and longevity and on the activity of tides and
currents. The propagules of Kandelia candel are sensitive to light; high levels inhibit
rooting. Fan and Chen (1993) suggest that this is adaptive as it keeps the floating
propagules alive during potentially long dispersal periods. It is unclear, however, how
common it is for mangrove propagules to travel great distances. It has been experimentally
shown that most Avicennia marina propagules strand and establish close to their parents; it
is uncommon for them to move very far (Clarke and Myerscough, 1991b; Kathiresan and
Ramesh, 1991; Kathiresan, 1999). This conclusion is supported by the observation of
Saifullah et al. (1994) that dispersal only determines small-scale distributional patterns of
mangroves in Karachi, Pakistan. Larger-scale patterns are created by environmental
heterogeneity.
     Mangrove propagules may suffer high mortality during their dispersal. In field
studies, propagules of Ceriops tagal in northern Australia dispersed very short distances
(only 9% moved more than 3 m from the parent tree). Within that short distance, however,
a high percentage of them were damaged or eaten by predators (McGuinness, 1997a;
Figure 5). Farnsworth and Ellison (1997a) measured predation on mangrove propagules in
42 mangrove swamps in 16 countries and found rates ranging from 0 - 93% with a global
average of 28.3%. The major predators were grapsid crabs and insects in the Coeleoptera,
and Lepidoptera. In Kenya, grapsid crabs cleared nearly 100% of the seeds from landward
                                mangrove plantations (Dahdouh-Guebas
  100
                                et al., 1998). Such high levels of seed
                                predation undoubtedly have significant
   80
Percent of propagules




                                effects on population dynamics and stand
                   Not taken by predators
                                regeneration.
   60

                                     Mortality is not restricted to
   40
                                propagules. Mangroves are also
                                vulnerable during establishment and
   20
              Not taken or damaged
                                early growth. In Belize, mortality of R.
                 by predators
                                mangle and A. germinans is highest
   0
    0    20    40     60     80     100
                                during establishment. The mortality can
              Days
                                be attributed to (1) a failure to establish
                                before seed viability is lost, (2)
Figure 5. Loss of Ceriops tagal propagules to
predators in a northern Australia mangal. Propagules      predation, and (3) desiccation (Ellison
were marked and tethered then monitored for           and Farnsworth, 1993).
disappearance and damage. Crab predators removed or
                                     After establishment, survival is
damaged 83% of the propagules within the first 90
                                strongly influenced by physicochemical
days (after McGuinness 1997a).
                                stresses. For example, shading,
orientation of the seedling axis (e.g., upright vs. horizontal), soil fertility, and flooding can
all have significant impacts on survival (Hovendon et al., 1995; McKee, 1995a; Koch,
1997; McGuinness, 1997a). Post-establishment growth is also affected by a suite of
physical and chemical factors. Experimental work with Rhizophora species demonstrates
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                         24


that propagule length, planting depth, soil type, salinity, concentration of leachates, pH and
light intensity are important determinants of growth (Kathiresan and Thangam, 1989,
1990a; Kathiresan and Ramesh, 1991; Kathiresan and Moorthy, 1993; Kathiresan et al.,
1993; Kathiresan et al., 1995b, 1996a; Kathiresan, 1999). Seedling growth can be
artificially stimulated by application of triacantanol and methanol. Both of these substances
increase the photosynthetic rate of the seedlings, the in vivo nitrate reductase activity, the
growth of roots and shoots, the protein and energy contents of leaves and roots, the
chlorophyll and carotenoid content in leaves, and the amount of chlorophyll in
photosystems I and II and in the light harvesting complex of the chloroplasts (Moorthy and
Kathiresan, 1993; Kathiresan and Moorthy, 1994a; Kathiresan et al., 1996a).
     New mangrove growth comes primarily from seeds and density of newly
established individuals can be very high (seedling densities reach 27,750 individuals • ha-1
in the Sunderbans of Bangladesh; Siddiqi, 1997). Vegetative regrowth from stump sprouts
(“copicing”) also occurs in some species (e.g., Excoecaria, Avicennia, Laguncularia,
Sonneratia; Tomlinson, 1986). Recently an air-layering technique has been used to
successfully induce vegetative propagation in Avicennia alba, A. officinalis, Sonneratia
apetala, Xylocarpus granatum and Rhizophora mangle. The technique was not successful
for A. marina or Kandelia (Kathiresan and Ravikumar, 1995a; Calderon and Echeverri,
1997; Ananda Rao, 1998). External application of auxins can stimulate growth of newly
planted mangrove cuttings. The auxins produce metabolic changes during initiation and
development of roots, enhancing levels of reducing sugars and increasing the mobilization
of nitrogen to the rooting zone (Basak et al., 1995; Das et al., 1997).

3.7. Biomass and litter production

     Mangroves and mangrove habitats contribute significantly to the global carbon
cycle. Mangrove forest biomass may reach 700 t ha-1 (Clough, 1992, Table 2) and Twilley
et al. (1992) estimate the total global mangrove biomass to be approximately 8.7 gigatons
dry weight (i.e., 4.0 gigatons of carbon). Accurate biomass estimates require measuring
volumes of individual trees. Da Silva et al. (1993) have developed equations for making
such measurements on living mangroves.
     Mangroves generally grow better in wet equatorial climates than they do in
seasonally monsoonal or arid climates (Clough, 1992) and the amount of litter they
produce is negatively correlated with latitude. Estimates of the annual global litterfall from
mangroves range from 130 to 1870 g m-2. In general, the litterfall is heaviest 1) in dry
summer months when thinning of the canopy reduces transpiration, and 2) in the wet rainy
season when fresh water input increases the nutrient supply (Roy, 1997; Wafar et al.,
1997). However, individual species may differ in the conditions that produce heavy litter.
For instance, Australian Rhizophora stylosa and Avicennia marina show heaviest litterfall
in hot climates with short dry seasons, but Ceriops tagal litterfall is heaviest in hot climates
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                           25


Table 2. Mangrove standing biomass measurements.


Location          Species      Biomass    Amount (t •  Reference
                                ha-1)
                        measured


Cuba (North America)    R. mangle    Roots        31.3   Fiala and Hernandez, 1993
              A. germinans  Roots        24.4

French Guiana (S.      Mixed forest  Total       31 – 315   Fromard et al, 1998
America)

Mgeni Estuary (S. Africa)  Mixed forest  Above-ground     94.4   Steinke et al., 1995
              A. germinans  Below-ground     9.6

Sunderbans (India)     Avicennia sp.  Total        147.7   Choudhuri, 1991
(6 yr old trees)      B.                  11.2
              gymnorrhiza
              S. apetala              34.5
              C. tagal               4.8

Tritih, Java (Indonesia)  R. mucronata  Above-ground     93.7   Sukardjo  and   Yamada,
                                      1992

Matang mangal        Mixed forest  Total        202.4   Gong and Ong, 1990
(Malaysia)

Hainan Island, (China)   Mixed forest  Total       9.6-14.2   Liao et al., 1993
              S. caseolaris  Total        47.2   Liao et al., 1990

Near Brisbane (Australia)  A. marina    Above-ground   110-340   Mackey, 1993
                      Below-ground+   109-126
                      pneumatophores

Mary River (Australia)   A.       Above-ground/   40/50    Saintilan, 1997
              corniculatum  below ground
              A. marina    Above-ground/   150/80
                      below ground
              E. agallocha  Above-ground/   140/40
                      below ground
              R. stylosa   Above-ground/   70/100
                      below ground
              C. australis  Above-ground/   110/50
                      below ground


with a long dry winter (Bunt, 1995). In India, Avicennia marina litter production is high in
the post-monsoon period and low in the pre-monsoon season (Ghosh et al., 1990).
Deviations from these general patterns of litterfall may result from habitat-specific stresses
(e.g., aridity, poor soils; Saenger and Snedaker, 1993; Imbert and Ménard, 1997).
     A number of researchers have measured mangrove litterfall. Results show a broad
range of litter volumes with production varying significantly from habitat to habitat. The
production appears to depend largely on local conditions, species composition, and
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                              26


productivity of the individual mangal. Litter production has been variously measured at
0.011 t ha-1 y-1 in the mangroves of Kenya, 9.4 t ha-1 y-1 in Bermuda, and 23.69 t ha-1
y-1 in Australia (Table 3).
Table 3. Litter production in mangrove forests.


Location                 Species     Litter production   Reference
                             (t • ha-1 • yr-1 )


Guyana (South America)        A. germinans      17.71       Chale, 1996

Teacapan-Ague Brava Lagoon      Mixed forest      14.17       Flores-Verdugo et al, 1990
(Mexico)

Bermuda (North America)        Mixed forest      9.40       Ellison, 1997

Bonny estuary (Nigeria)        R. racemosa       8.46       Abbey-Kalio, 1992
                   A. africana       6.41
                   Laguncularia sp.    8.18

South Africa             Mixed forest      4.50       Steinke and Ward, 1990

Gazi Bay (Kenya)           R. mucronata      0.02       Slim et al, 1996
                   C. tagal        0.01

Andaman Islands (India)        Mixed forest     7.10 - 8.50    Mall et al, 1991
                   B. gymnorrhiza    5.11 –7.09     Dagar and Sharma, 1993
                   R. apiculata     8.08 – 10.30    Dagar and Sharma, 1991

Mandovi-Zuari Estuary (India)     R. apiculata      11.70       Wafer et al, 1997
                   R. mucronata      11.10
                   S. alba        17.00
                   A. officinalis     10.20

Fly River Estuary (New Guinea)    Mixed forest     8.00 – 14.00    Twilley et al, 1992

Matang mangal (Malaysia)       Mixed forest      3.90       Gong and Ong, 1990

Jervis Bay, NSW (Australia)      A. marina        3.10       Clarke, 1994
                   A. corniculatum     2.10

Embley River (Australia)       R. stylosa       12.23       Conacher et al., 1996
                   C. tagal        5.39
                   A. marina        6.28

Australia               A. marina       15.98       Bunt, 1995
                   R. stylosa       23.69
                   C. tagal        12.90


    Litter from the mangroves is composed of leaves, twigs, branches, and seeds. Seeds
alone accounted for 25% of the total litterfall for Avicennia germinans and Rhizophora
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      27


mangle in a mangrove habitat in Martinique (Imbert and Ménard, 1997). In a temperate
mangal, the reproductive material was approximately 9% of the total for Avicennia marina
and 32% of the total for Aegiceras corniculatum. Clarke (1994) suggested that such
relatively high reproductive output may contribute to the low productivity and stunting of
mangroves at high latitude.
     Accumulated mangrove litter may wash into rivers and streams when rain or tides
inundate the forest. Consequently, mangrove litter may decompose either in the source
forest or in the river, with nutrients being retained or exported (Conacher et al., 1996).
Whether the litter (and its nutrients) remain in the habitat or are exported by water flow
may depend largely on the local animal community. On the east coast of Queensland, the
litter accumulation in a Ceriops forest was 6 g m-2 (0.06 t ha-1) while in an Avicennia
forest, it was closer to 84 g m-2 (0.84 t ha-1; Robertson et al., 1992). This enormous
difference in accumulation was attributed to the feeding activities of crabs.

4. MANGROVE-ASSOCIATED FLORA
4.1. Bacteria

    Mangroves provide a unique ecological environment for diverse bacterial
communities. The bacteria fill a number of niches and are fundamental to the functioning
of these habitats. They are particularly important in controlling the chemical environment
of the mangal. For example, sulfate-reducing bacteria (e.g., Desulfovibrio,
Desulfotomaculu, Desulfosarcina, and Desulfococcus; Chandrika et al., 1990; Loka-
Bharathi et al., 1991) are the primary decomposers in anoxic mangrove sediments. These
bacteria largely control iron, phosphorus, and sulfur dynamics and contribute to soil and
vegetation patterns (Sherman et al., 1998). Methanogenic bacteria are seasonally abundant
in sediments where Avicennia species dominate (T. Ramamurthy et al., 1990; Mohanraju
and Natarajan, 1992). Subsurface bacterial communities (along with epibenthic
microalgae) may sequester nutrients and hold them within nutrient-limited mangrove muds
(Alongi et al., 1993; Rivera-Monroy and Twilley, 1996).
    Bacteria are critical to the cycling of nitrogen in mangrove environments. Marine
cyanobacteria are a particularly important component of the microbiota, constituting a
source of nitrogen in every mangrove system (Sheridan, 1991, 1992; Hussain and Khoja,
1993; Krishnamurthy et al., 1995a; Palaniselvam, 1998). N2-fixing cyanobacteria isolated
from Avicennia pneumatophores in the Beachwood Mangrove Reserve, South Africa
supply 24.3% of the annual nitrogen requirements of that swamp. The N2-fixation rates are
controlled by light and temperature and show seasonal trends (low in the winter and high
in the summer; Mann and Steinke, 1993). Fixation rates are higher when the cyanobacteria
are on the mangrove than when they are held on an artificial growth medium (Toledo et al.,
1995b).
    N2-fixing bacteria are efficient at using a variety of mangrove substrates despite
differences in carbon content and phenol concentrations (Pelegri and Twilley, 1998).
However, their abundance may be dependent on physical conditions and mangrove
community composition. N2-fixing Azotobacter, which show potential as biofertilizers, are
abundant in the mangrove habitats of Pichavaram, south India. Their abundance in the
mangal exceeds that in marine backwaters and estuarine systems (S. Ravikumar, 1995).
Sengupta and Choudhuri (1991) studied N2-fixing bacteria in a Ganges River mangrove
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       28


community. They found high numbers in the rhizospheres of plants in inundated areas but
plants on occasionally inundated ridges and in degraded areas had fewer rhizosphere
bacteria. Ogan (1990) found similar distinct differences in nodulation and nitrogenase
activity among sites and among species in a Nigerian mangal.
     Two halotolerant N2- fixing Rhizobium strains have been isolated from root nodules
of Derris scandens and Sesbania species growing in the mangrove swamps of Sunderbans
(Sengupta and Choudhuri, 1990). If the non-N2-fixing bacteria are removed from the
rhizosphere, N2-fixing activity drops, indicating that other rhizosphere bacteria contribute
to the fixation process (Holguin et al., 1992). The non-N2 fixer, Staphylococcus sp.,
isolated from mangrove roots, promotes N2-fixation by Azospirillum brasilense. This can
be achieved by growing the two species in mixed culture or simply by adding a cell-free
dialysate of the Staphylococcus sp. to the A. brasilense culture. Aspartic acid is the
compound responsible for the effect (Holguin and Bashan, 1996).
     In addition to processing nutrients, mangrove bacteria may also help process
industrial wastes. Iron-reducing bacteria are common in mangrove habitats in some mining
areas (Panchanadikar, 1993). Eighteen bacterial isolates that metabolize waste drilling fluid
have been collected from a mangrove swamp in Nigeria (Benka-Coker and Olumagin,
1995). Interestingly, four additional bacterial strains isolated from the same swamp depress
growth rates of Staphylococcus and Pseudomonas species and could, therefore, decrease
normal rates of organic decomposition (Benka-Coker and Olumagin, 1996).
     Bacteria play a number of other roles in the mangal. Some live symbiotically with
other organisms. For example, rod bacteria can be commonly found in the hindguts of
mangrove detritivores (Harris, 1993) and deeply branched sulfur-oxidizing bacteria occur
as endosymbionts within members of the bivalve family Lucinacea in sulfide-rich, muddy
mangrove areas. Bauer-Nebelsick et al. (1996) and Ott et al. (1998) have described sulfur-
oxidizing bacteria that live as obligate ectosymbionts on colonial sessile ciliates
(Zoothamnium niveum) in a Belizian mangal.
     Other mangrove bacteria are parasitic or pathogenic. Bdellovibrios capable of
parasitizing Vibrio spp. are common in an Australian mangrove habitat. Their abundance
there (36.6 ml-1) is much higher than in nearby Great Barrier Reef habitats (9.5 ml-1;
Sutton and Besant, 1994). Also in Australia, Bacillus thuringiensis, which shows
insecticidal activity against mosquito larvae of Anopheles maculatus, Aedes aegypti and
Culex quinquefasciatus, has been isolated from mangrove sediments (Lee et al., 1990a;
Lee and Seleena, 1990). Actinomycetes (fungi-like bacteria) that occur in many mangrove
habitats (Kala and Chandrika, 1993; Vikineswary et al., 1997) may show antifungal
activity (Vikineswary et al., 1997).
     Bacterial populations show distinct spatial distribution patterns. Many live
epiphytically on the surfaces of mangroves, but different species appear to prefer different
parts of the tree. Leaves of Avicennia marina and Sesuvium portulacastrum harbour large
numbers of Flavobacterium while roots and stems have large populations of Vibrio spp.
(Abhaykumar and Dube, 1991). In many species, the aerial roots, especially
pneumatophores, harbour particularly dense bacterial cyanopopulations that may show
sharp vertical zonation. Coccoid forms occur in the upper zone of the pneumatophores.
Filamentous non-heterocystous forms predominate in the middle zone, and filamentous
heterocystous forms are largely restricted to the lower zones (Toledo et al., 1995a;
Palaniselvam, 1998). In the forests of Aldabra Lagoon, heterocystous forms like
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                                29


Scytonema sp. also form conspicuous growths on pneumatophores, but non-heterocystous
species are restricted to the sediment surface (e.g., Alongi and Sasekumar, 1992).
     Cyanobacteria in the mangal colonize any submerged surface including sediments,
roots, aerial roots, branches and trunks (Sheridan, 1991). Microbial mats in mangrove tidal
channels often have an outer layer of cyanobacteria
and a reddish inner layer of anoxygenic            100

phototrophic bacteria (Lopez-Cortes, 1990). A          80
cyanobacterium (Calothrix viguieri) isolated from




                                   % hairiness
                                60
the surface of mangrove roots show a peculiar
morphological response to salinity variation. In low      40

salinity, it develops hairs (Figure 6). The hairs are      20
shed if salinity is increased . The hairs may be an
                                 0
adaptation to hydrolyze pulses of organic             0   6      12      18   24

phosphorus that occur in the habitat after heavy        200

rains (Mahasneh et al., 1990).




                                   Mean hair length (µm)
     Bacterial counts are generally higher on       150


attached mangrove vegetation than they are on         100
fresh leaf litter. This is probably because attached,
undamaged leaves leak amino acids and sugars but        50

do not release much tannin (Kathiresan and
Ravikumar, 1995b). Shome et al. (1995) isolated         0
                                  0   6      12      18   24
thirty-eight distinct bacteria from mangrove leaf           Time after transfer to fresh water (h)

litter and sediments in south Andaman and           Figure 6. Effects of salinity on hair
                                formation in the mangrove bacterium
characterized the bacterial community. The bacteria
were generally gram-positive (76.3%), motile (87%),
fermentative (6.9-82.1%), pigmented (31%), and antibiotic resistant (100% against
polymixin B and 50% against chloramphenicol). Photosynthetic bacteria, including purple
sulfur bacteria (Chromatium spp.) and purple non-sulfur bacteria (Rhodopseudomonas
spp.), have been isolated from mangroves in Pichavaram, south India (Vethanayagam,
1991; Vethanayagam and Krishnamurthy, 1995). Nine species of purple non-sulfur
bacteria have also been found in mangroves of Egypt (Shoreit et al., 1994). Growth of the
purple sulfur bacteria in these habitats is limited by low light and sulfide. In contrast, high
light and sulfide limit growth of green sulfur bacteria (Chandrika et al., 1990).


4.2. Fungi and fungus-like protists

    Mangals are home to a group of fungi called “manglicolous fungi.” These
organisms are vitally important to nutrient cycling in these habitats (Hyde and Lee, 1995;
Kohlmeyer et al., 1995). Kohlmeyer and Kohlmeyer (1979) were the first to review this
group. They recognized 43 species of higher fungi, including 23 Ascomycetes, 17
Deuteromycetes, and 3 Basidiomycetes. Hyde (1990a) listed 120 species from 29
mangrove forests around the world. These included 87 Ascomycetes, 31 Deuteromycetes,
and 2 Basidiomycetes.
    Work in individual habitats has revealed surprisingly diverse fungal communities
(e.g., Hyde, 1990b; Hyde, 1996). Chinnaraj (1993a) identified 63 species of higher fungi in
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       30


mangrove samples from Andaman and Nicobar Islands alone. Similar samples from
Lakshadweep Island yielded 32 species (Chinnaraj, 1992) and 39 species were found in
mangrove samples from the Maldives (Chinnaraj, 1993b). D.R. Ravikumar and Vittal
(1996) found 48 fungal species in decomposing Rhizophora debris in Pichavaram, south
India. On the Indian Ocean coast of South Africa, Steinke and Jones, (1993) identified 93
species of marine fungi, including 55 from mangrove wood (particularly Avicennia
marina). Table 4 lists some of the fungal species identified in these studies.
Table 4. Some fungal species isolated from mangrove habitats.


Species                         Author


Aigialus striatispora                  Hyde (1992c)
Aniptodera longispora                  Hyde (1990b)
Aniptodera salsuginosa                 Nakagiri and Ito (1994)
Calathella mangrovei                  Jones and Agerer (1992)
Cryptovalsa halosarceicola               Hyde (1993)
Eutypa bathurstensis                  Hyde and Rappaz (1993)
Falciformispora lignatilis               Hyde (1992d)
Halophytophthora kandeliae               Ho et al (1991)
Halophytophthora kandeliae               Newell and Fell (1992b)
Halophytophthora vesicula                Newell and Fell (1992b)
Halophytophthora. spinosa                Newell and Fell (1992b)
Halosarpheia minuta                   Leong et al (1991)
Hapsidascus hadrus                   Kohlmeyer and Kohlmeyer (1991)
Hypoxylon oceanicum                   Whalley et al (1994)
Julella avicenniae                   Hyde (1992a)
Khuskia oryzae                     Pal and Purkayastha (1992a)
Lophiostoma asiana                   Hyde (1995)
M. ramunculicola                    Hyde (1991b)
Massarina armatispora                  Hyde et al (1992)
Massarina velatospora                  Hyde (1991b)
Payosphaeria minuta                   Leong et al (1990)
Pedumispora                       Hyde and Jones (1992)
Phomopsis mangrovei                   Hyde (1991a)
Saccardoella                      Hyde (1992b)
Trematospaeria lineolatispora              Hyde (1992d)

    Surveys are revealing a number of range extensions, new species, and even new
genera. Collections of mangrove fungi in Macau and Hong Kong, for instance, yielded 45
species. Twenty-eight of these were new records for Macau and 21 were new records for
Hong Kong (Vrijmoed et al., 1994). These discoveries are leading to significant taxonomic
revision of these groups (Jones et al., 1994, 1996; Alias et al., 1996; Goh and Yipp 1996;
Ho and Hyde, 1996; Vrijmoed et al., 1996; Honda et al., 1998).
    The fungus-like thraustochytrids are important endobionts in dead or living plants
and in calcareous shells (S. Raghukumar, 1990). A number of these occur in mangrove
swamps where they help decompose mangrove leaf litter (S. Raghukumar et al., 1994;
Bremer, 1995). Both thraustochytrids and chytridiomycetes (including Schizochytrium,
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       31


Thraustochytrium and Ulkenia) have been isolated from Costa Rican mangrove swamps
(Ulken et al., 1990). Two thraustochytrids, Thraustochytrium striatum and Schizochytrium
mangrovei, have been isolated from an Indian mangal at Goa. Both produce amoebae-like
structures, move using pseudopodia, and phagocytose bacterial cells (S. Raghukumar,
1992).
     Marine oomycetes (fungus-like protists) also occur in mangrove communities. T.K.
Tan and Pek (1997) found five Halophytophthora species in Singapore mangroves. This is
the first time three of them have been seen in tropical mangroves and the first time one has
been reported outside Australia. Oomycetes in the genus Halophytophthora have special
importance in mangrove habitats (Newell and Fell, 1992a). They greatly facilitate the
decomposition of mangrove material. Newly fallen Rhizophora mangle leaves are quickly
infested with mycelial growths of Halophytophora vesicula and H. spinosa. Rapid lateral
extension of the mycelia within the leaves apparently follows establishment of a single
zoospore (Newell and Fell, 1995). In laboratory cultures, the established Holophytophora
are subsequently colonized by bacteria and labyrinthulas (Newell and Fell, 1994).
Holophytophora species are generally good competitors against true fungi but have
difficulty colonizing leaves that already have bacterial films (Newell and Fell, 1997).
     Newell and Fell (1996) speculate that Halophytophora completes its colonization
of submerged leaves, from attachment of zoospore cysts to release of new zoospores, in the
early stages of leaf decomposition, before there is substantial entry into the leaves
themselves. Mild drying, low salinity and low temperatures may enhance zoospore release.
The release rates are low in older, decaying leaves and high in newer, less-decayed leaves
(Newell and Fell, 1996). Leaño et al. (1998) showed that the zoospores of other mangrove
fungi are chemically attracted to plant material and extracts. This undoubtedly aids in the
colonization of new substrata.
     A few researchers have studied the physiology and biochemistry of manglicolous
fungi. Many of the species produce interesting compounds. For example, most of the soil
fungi produce lignocellulose-modifying exoenzymes like laccase (C. Raghukumar et al.,
1994). Preussia aurantiaca synthesizes two new depsidones (Auranticins A and B) that
display antimicrobial activity (Poch and Gloer, 1991). Cirrenalia pygmea produces
melanin pigments that appear to protect the hyphae from sudden changes in osmotic
pressure; when melanin synthesis in cultures is inhibited with tricyclazole, the fungus
becomes sensitive to osmotic shock (Ravishanker et al., 1995). High salinities also
increase the number and types of amino acids this species produces (Ravishankar et al.,
1996).
     Ascus and ascospore ultrastructure have been studied in the fungi Swampomyces
armeniacus and Marinosphaera mangrovei (Read et al., 1995) and in Dactylospora
haliotrepha (Au et al., 1996). Ascocarp formation has been tested in single and mixed
cultures of Aigialus parvus, Lignincola laevis and Verruculina enalia growing on the wood
of Avicennia alba, Bruguiera cylindrica and Rhizophora apiculata. Sporulation was
delayed and fewer ascocarps were formed in mixed cultures, suggesting competition
among the fungi (T.K. Tan et al., 1995).
     A number of fungal species live directly on living mangroves but, in general, they
are not well known. Sivakumar and Kathiresan (1990) isolated ten fungal species from leaf
surfaces of seven mangrove species. The dominant phylloplane fungi were Alternaria
alternata, Rhizopus nigricans, Aspergillus and Penicillium spp. Abundances of these fungi
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       32


were negatively correlated with tannin content of the leaves. The fungi appear to prefer
leaf litter (which contain more amino acids) to fresh leaves (which contain more tannins
and sugars; S. Ravikumar and Kathiresan, 1993). Other fungi are harmful to the living
mangroves. Two new parasitic species (Pestalotiopsis agallochae and Cladosporium
marinum) have been isolated from the leaves of Excoecaria agallocha and Avicennia
marina (Pal and Purkayastha, 1992b). Pathogenic fungi may have contribute to diebacks of
Rhizophora mangle stands in Costa Rica (Tattar et al., 1994).
     A number of fungal species colonize subsurface mangrove roots. Nair et al. (1991)
found 25 fungal species from 15 genera in the rhizosphere of Avicennia officinalis;
adjacent non-rhizosphere soil held only 16 species from 10 genera. Sengupta and
Choudhuri (1994) found Rhizoctonia and VA-mycorrhiza-like fungi in the mangrove
community at Sunderbans. When Cajanas seedlings in nutrient-poor conditions were
inoculated with the VA-mycorrhizal isolates, there was a significant increase in growth.
This was due, in part, to mobilization of insoluble phosphate by the fungus.
     Distributions of fungal species within the mangrove habitat may reflect physical
conditions and/or habitat preference. It may also reflect age of the stand. Working in Belize
(Central America), Kohlmeyer and Kohlmeyer (1993) found that fungal diversity depends
on age of the mangrove stand. They discovered 43 species in established Rhizophora
stands but only 7 in recently introduced Rhizophora. Some species may be quite specific in
their habitat preferences. For example, of the 48 fungal species Ravikumar and Vittal
(1996) found in a south Indian mangal, 44 were on prop root while seedling and wood
samples only held 18 and 16 species respectively. The fungal species appeared to partition
the mangrove habitat. Verruculina enalia was most abundant on prop roots and seedlings
while Lophiostoma mangrovei was most common on wood. Physical conditions, or genetic
differentiation created by isolation, may lead to differences in fungal morphology and
physiology. Pestalotiopsis versicolor strains, isolated from Ceriops decandra growing in
different regions of the Sunderbans, vary in mycelial mat texture, growth rate and
sporulation intensity (Bera and Purkayastha, 1992).
     Differences in physical requirements may lead to vertical zonation of the fungi.
Hyde (1990b) found 57 intertidal fungal species on Rhizophora apiculata at Brunei
mangal. Most of these occurred above the mean tidal level. A similar study of senescent
Acanthus ilicifolius at Mai Po, Hong Kong revealed that the apical portions of the trees are
colonized by typical terrestrial fungi but the basal portions are colonized by marine species
(Sadaba et al., 1995). The authors attributed this to the nature of the substratum and the
frequency of tidal inundation. Other fungal species live directly on the sediment surface
but are still entirely restricted to mangrove habitats (Soares et al., 1997).
     Wood degrading fungi are well-known in mangrove habitats. Thirty species of such
lignicolous fungi have been recorded in Malaysian mangals. The most abundant are
Halosarpheia marina, Lulworthia sp., Lignincola laevis, Halosarpheia retorquens, Eutypa
sp., Kallichroma tethys, Marinosphaera mangrovei, Phoma sp. and Julelia avicenniae.
Diversity and abundance are greatest on Avicennia wood (T.K. Tan and Leong, 1992; Alias
et al., 1995). Test panels of different woods placed in mangrove waters along the Goa
coast of India showed four common lignicolous fungi (Periconia prolifica, Lignincola
laevis, Aniptoder sp. and Lulworthia sp.). Panels treated with copper chrome arsenic were
more resistant to fungal infestation than those treated with chrome boric (Santhakumaran et
al., 1994).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        33


    Nakagiri and Ito (1994) found a new lignicolous fungus (Aniptodera salsuginosa)
with unique ascospore appendages and an unusual ascus apical apparatus on decomposing
mangrove wood. The ascospore appendages are functional only when they are submerged
in brackish water. The ascospores are discharged through a fissure in the ascus wall at the
margin of the apical disc; the ascus pore in this disc does not function in ascospore release.



4.3. Microalgae

    Phytoplankton and benthic microalgal communities make important contributions
to the functioning of mangrove environments. However, their contribution to total
estuarine production is relatively small in most regions of southeast Asia, Australia,
Central America and tropical South America. Robertson and Blaber (1992) suggested that
the contribution of plankton to total net production in mangrove habitats ranges from 20-
50%. Careful measurements are verifying that predication for large systems. Phytoplankton
are responsible for 20% of the total production in mangrove estuaries in the Fly River
Delta in Papua New Guinea (Robertson et al., 1991, 1992) and 20-22% of the total
production in the Pichavaram mangroves of south India (Kawabata et al., 1993).
    Phytoplankton contributions to productivity in localized mangrove areas may be
much smaller. Lee (1990) found that phytoplankton and benthic macroalgae together
contribute less than 10% of the net primary production in Hong Kong mangals and Boto
and Robertson (1990), using nitrogen measurements, estimated that benthic cyanobacteria,
microalgae and macroalgae together contribute only 6% of the gross primary production in
mangrove ecosystem of northeastern Australia. Robertson and Blaber (1992) state that
phytoplankton productivity is significantly lower in estuarine mangrove areas than it is in
lagoons or open embayments fringed by mangroves.
    High turbidity, large salinity fluctuations and a generally small ratio of open
waterway to mangrove forest area contribute to the low light levels and shading that limit
productivity of the microalgae, especially the benthic forms (Alongi, 1994; Harrison et al.,
1994). High summer temperatures may also limit production (Lee, 1990). Rates of primary
production, which are generally low in the dry season, increase on ebb tides and decrease
on flood tides (Kitheka, 1996).
    In the Fly River delta of Papua New Guinea, Robertson et al., (1992) measured
very low production rates of only 0.022 to 0.0693 g C m-3 d-1. However, localized
conditions may lead to much higher rates. For example, daily production in the coastal
lagoons of Mexico may reach 2.4 g C m-3 d-1 (Robertson et al., 1992). Increased
productivity may relate to elevated nutrient levels. Production in lagoons of the Ivory
Coast reach 5 g C m-3 d-1. However, the effect is largely a result of nitrogen and
phosphorus input from nearby human population centers.
    Naturally occurring substances may also regulate phytoplankton growth. Selvam et
al. (1992) found phytoplankton productivity to be four times higher in mangrove waters
than in adjacent marine waters in south India. Refractive materials like humic acid, which
are abundant in the mangroves, stimulate phytoplankton growth there (Schwamborn and
Saint-Paul, 1996). In the Celestun Lagoon (northern Yucatan Peninsula, Mexico) low
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        34


concentrations of natural phenolics stimulate phytoplankton growth, but higher wintertime
levels depress the growth rates (Herrera Silveira and Ramirez Ramirez, 1996).
    While microalgae may make only small contributions to total productivity in
estuarine mangrove systems, they may be critical to supporting higher trophic levels
(Robertson and Blaber, 1992). This may be particularly true because of the high nutritional
quality of phytoplankton relative to mangrove detritus. Phytoplankton biomass,
productivity, and size are closely tied to diversity and abundance of higher trophic levels.
Teixeira and Gaeta (1991) determined the composition of the phytoplankton community in
a Brazilian mangal. Nanoplankton (cells from 2 - 20 µm) constituted over 80% of the total
phytoplankton. Laboratory testing showed that the smaller cells were responsible for a
significant part of the total productivity. Picoplankton (cells < 2 µm) accounted for 3-29%
of the total 14C uptake. The effects of this skewed phytoplankton size distribution on the
zooplankton community composition has not been studied.
    Despite relatively low productivity, mangrove phytoplankton communities can be
quite diverse. However, composition and density of the plankton community are strongly
affected by local environmental conditions (Lee, 1990). For example, low phytoplankton
diversity in Rhizophora habitats is related to the release of tannins by roots, decomposing
wood, and leaves (Robertson and Blaber, 1992). Phytoplankton populations also respond to
temperature and salinity variation. Thus, communities may show marked seasonal
variation (Mani, 1994). Phytoplankton studies at West Bengal, India revealed 46 species of
Bacillariophyceae, Dinophyceae and Cyanophyceae (Santra et al., 1991). Coscinodiscus,
Rhizosolenia, Chaetoceros, Biddulphia, Pleurosigma, Ceratium and Protoperidinium were
the dominant genera, existing almost year round. At least 82 phytoplankton species (72%
diatoms,15% dinoflagellates) occur in the Pichavaram mangroves of south India (Kannan
and Vasantha, 1992). The diatoms Nitzschia closterium, Pleurosigma spp., Thalassionema
nitzschioides and Thalassiothrix frauenfeldii are most abundant. Thirty-one of those
species may form seasonal blooms (Mani, 1992). Chaghtai and Saifullah (1992) reported
such a bloom of the diatom Navicula in the Karachi mangroves of Pakistan.
    Dinoflagellate assemblages have been particularly well studied in Belizean
mangrove habitats where a diverse collection of benthic and epiphytic species exists
(Faust, 1993a,b,c,d; Faust and Balech, 1993). Many are new species (e.g., Prorocentrum
maculosum, P. foraminosum, P. formosum, Plagiodinium belizeanum, Sinophysis
microcephalus). Faust and Gulledge (1996) found many microalgal species associated with
floating mangrove detritus. Dinoflagellates constituted the greatest proportion (50-90%),
followed by diatoms (5-15%), cyanobacteria (3-25%) and dinoflagellate cysts (1-7%).
Ciliates and nematodes were the major dinoflagellate consumers in the detritus.

4.4. Macroalgae

    The macroalgal flora is rich in mangrove habitats where it contributes to production
while also providing habitat and food for a number of invertebrate and fish species. Red
algae, especially in the genera Bostrychia, Caloglossa and Catenella, are most commonly
associated with mangroves and may be quite abundant. For instance, the total annual
biomass of Bostrychia tenella in a south Nigerian estuary reaches 1.84 mg • cm-2, which is
38% of the total algal production there (Ewa-Oboho and Abby-Kalio, 1993). The biomass
of algae in the mangrove lagoons of Puerto Rico is similar to the total annual leaf litterfall
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        35


from the Rhizophora fringe, leading to an algal-dominated foodweb (Rodriguez and
Stoner, 1990).
    Algal diversity can also be quite high in mangrove environments. Recent surveys
have revealed diverse macroalgal communities in Papua New Guinea (25 species; King,
1990), the Nicobar Islands in the Andaman Sea (61 species; Jagtap, 1992), and the coast of
Mauritius (127 species; Jagtap, 1993). King and Puttock (1994) and King (1995) provide
exhaustive reviews of the very diverse Australian mangrove macroalgal flora. Algal
assemblages tend to be richest in shallow areas with a mixture of hard and soft substrates.
Lowest diversity occurs where there is low light, soupy muds, or homogeneous, large-grain
sands (as in the Netherlands Antilles, Kuenen and Debrot, 1995).
    Algal surveys have produced new records for a number of species including
Stictosiphonia kelanensis from Atlantic mangroves (Fujii et al., 1990); Bostrychia pinnata,
Bostrychia simpliciuscula, Caloglossa angustalata (Rhodophyta) and Boodleopsis
carolinensis (Chlorophyta) from Singapore (West, 1991a); Bostrychia pinnata, Caloglossa
ogasawaraensis, C. stipitata and Halochlorococcum operculatum from Peru (West,
1991b); Bostrychia pinnata and Caloglossa ogasawaraensis from the Atlantic coast, USA
(West and Zuccarello, 1995); Bostrychia calliptera from the Central Gulf of Mexico
(Collado-Vides and West, 1996) and C. ogasawaraensis, C stipatata, C. lepricuriiI, B.
moritziana, B. pinnata, B. radicans and Catenella caespitosa in Southern Mexico and
Guatemala (Pedroche et al., 1995).
    Recent work has investigated genetic differentiation of some of the widely
distributed red algae. Male Caloglossa ogasawaraensis from a Peruvian mangle readily
hybridize with female C. ogasawaraensis from Brazil, producing viable tetrasporophytes
(West, 1991b). Similar studies with Bostrychia radicans from the Pacific and Atlantic
coasts of North America have been done. Almost all isolates from the northern Pacific
coast of Mexico are compatible and produce cystocarps that release viable carpospores.
However, isolates from the Atlantic coast of the United States show greater incompatibility
(Zuccarello and West, 1995).
    Algal abundance and diversity are largely determined by the physico-chemical
characteristics of the mangal (Mazda et al., 1990a) and these may be extremely variable.
As with the mangroves themselves, the most successful macroalgae have special
adaptations that help them tolerate extreme conditions. Work on the physiology of algae
associated with mangroves includes a study of salinity and the polyol (D-dulcitol, D-
sorbitol) content of Bostrychia (West et al., 1992). The success of B. simpliciuscula in the
mangrove swamps of Singapore may be attributed to its physiological adaptations to
salinity extremes. The polyols serve as osmoprotectors that B. simpliciuscula synthesizes
and sequesters as salinity increases (Karsten et al., 1994, 1996). Floridoside compounds,
which may be essential for survival of the algae, also change with salinity (Karsten et al.,
1995). Caloglossa leprieurii, which is also common in mangrove environments, has a
novel metabolic pathway that may be a similar biochemical adaptation to environmental
extremes (Karsten et al., 1997). Salinity gradients create distinct ecotypes of Caloglossa
leprieurii in mangals along the Brisbane River, Australia (Mosisch, 1993).
    Salinity, temperature, desiccation, tidal inundation, wave action, wetting frequency
and light intensity are all environmental factors likely to produce patterns of horizontal and
vertical distribution seen in many mangrove algae (e.g., Phillips et al., 1994; Farnsworth
and Ellison, 1996b). In the Gazi Bay of Kenya, there is distinct macroalgal zonation. The
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        36


upper intertidal is covered by Boodleopsis pusilla while the mid-intertidal is dominated by
Halimeda opuntia, Gracilaria salicornia and G. corticata. The low water mark has
primarily Halimeda macroloba and Avrainvillea obscura (Coppejans et al., 1992). A
distinct zonation has also been described for algae growing on the pneumatophores of
Avicennia marina (Steinke and Naidoo, 1990). There are generally three zones: an upper
Rhizoclonium zone; a middle Bostrychia zone, and a lower Caloglossa zone (Phillips et al.,
1996).
    The composition of the mangrove algal community may depend largely on the
nature of the early colonizers. Eston et al. (1992) monitored colonization of artificial
substrata by mangrove macroalgae and found that Bostrychia radicans and several other
species settled early. There was no evidence that later species could displace the early
colonists. This macroalgal community showed no succession; the pioneer community was
also the final community. Established macroalgae can also affect distribution of the
mangroves directly. For example, in southeastern Australia, the alga Hormosira banksii
inhibits intertidal establishment of grey mangrove (Avicennia marina) seedlings (Clarke
and Myerscough, 1993).
     A number of algae from mangrove habitats have potential commercial value. For
example, the red alga Gracilaria changii from Malaysian mangrove habitats is an excellent
source of agar; the agar content is between 12 and 25% of its dry weight (Phang et al.,
1996). Monostroma oxyspermum, Catenella impudica and Caloglossa lepriurii are all
edible food resources. The latter two species are also potential sources of dyes. Caulerpa
sp. has yielded bioactive substances that may hold promise as pharmaceutical agents (e.g.,
Ananda Rao et al., 1998).

4.5. Seagrasses

    Seagrasses are closely associated with mangrove habitats in many parts of the
world. In the Andaman Sea, there are three mangrove-associated sea-grasses, Thalassia
hemprichii, Enhalus acoroides and Halophila ovalis (Poovachiranon and Chansang, 1994).
Intertidal mangrove areas in the Gazi Bay, Kenya are colonized by Thalassia hemprichii,
Halophila ovalis and Halodule wrightii (Coppejans et al., 1992) while Halophila baccarii
occurs on intertidal mudflats of Indian mangals (Jagtap, 1991).
    The seagrass biomass in mangrove areas may be quite high. In an Andaman Sea
mangal, Poovachiranon and Chansang (1994) measured seagrass biomass ranging from 55-
1941 g wet wt • m-2, corresponding to 32-297 g dry wt • m-2. As with the macroalgal
communities, seagrass diversity and abundance are largely regulated by a combination of
light level and substrate type. In the Spaanse waters of the Netherland Antilles, the richest
assemblages of seagrasses occur in shallow areas with high light and a mix of hard and soft
substrates. Diversity is much lower where light is low and the substrates are loose muds or
homogeneous, coarse-grained sands (Kuenen and Debrot, 1995).
    Seagrasses generally require high light levels to grow and survive. Planktonic
primary producers require only about 1% of the surface irradiance to maintain a net
positive carbon balance. In contrast, seagrasses may require 10-20% of the daily average
surface irradiance to survive (Fourqurean and Ziemean, 1991). Growth rate may decrease
naturally in the winter months as a result of low temperatures and shortened daylengths.
However, in recent years, there have been precipitous declines of seagrass beds in
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        37


mangrove environments. Seagrass mortality has often been linked to reduced water quality
and increased turbidity that decrease light penetration (Giesen et al., 1990; Larkum and
West, 1990). In turbid waters, flocs from the mangroves themselves contribute to shading
of the seagrass (Wolanski et al., 1997).
     Though seagrass beds often occur in close proximity to mangroves, the two habitats
may not be closely coupled. Tussenbroek (1995) found that seagrass growth, biomass and
primary production were all higher in the vicinity of mangrove discharges than they were
in other habitats. Respiratory CO2 derived from mangrove particulate organic matter
(POM) could be a carbon source for seagrass and could promote faster growth. Ebb flows
are generally stronger than flood flows in mangrove creeks, which should promote a net
export of nutrients and POM. In general, however, fluxes from mangrove forests seem to
have little effect on adjacent seagrass beds (Fleming et al., 1990). For example, Hemminga
et al. (1994) failed to detect any input of mangrove POM in a seagrass bed only 3 km
away. POM was exported from the mangrove forest, but deposition was rapid and little
material reached the seagrass bed. Similarly, in the Gazi Bay of Kenya, leaf production and
nitrogen:phosphorus ratios of Thalassodendron ciliatum were unrelated to the input of
mangrove carbon and 13C studies confirmed that the mangroves contribute little reduced
carbon to adjacent seagrass beds (Lin et al., 1991). Nor does it appear that dissolved
nutrients move from the mangal to nearby grassbeds. The few dissolved nutrients
generated by the mangroves are likely to be used for primary production within the
mangrove zone itself (Kitheka et al., 1996).
     Mangroves and seagrasses serve parallel functions in the habitats they share. Both
trap sediments and help capture chemical elements, including trace metals (Costa and
Davy, 1992; Lacerda, 1998). Both also help support fish population by serving as food for
fish, as critical habitat for fish, and as growth surfaces for epizonts that fish eat. A number
of fish species may use seagrass/mangrove habitat as a nursery area. In Guadeloupe,
French West Indies, fish diversity is higher in Thalassia testudinum beds near mangroves
than in the adjacent coral reefs (Baelde, 1990). Similarly, in Belize, Central America, fish
abundance and biomass were highest in a mangrove creek, followed by a seagrass bed and
the sand-rubble zone of an adjacent lagoon (Sedberry and Carter, 1993). Arancibia et al.
(1993) found more than 80 fish species using the mangrove/seagrass habitat; seven species
were found only in these areas.

4.6. Saltmarsh and other flora

     Saltmarsh plants replace mangroves at their northern limit on the Gulf and Atlantic
coasts of North America but the southern limit of the saltmarsh distribution may be set by
competition with the mangroves. For example, the common saltmarsh grass Spartina
cannot survive high salinities and fast sediment accretion. As a result, it grows poorly in
areas where mangroves thrive (Kangas and Lugo, 1990). This usually leads to its
replacement by mangroves, as in Paranagua Bay, Brazil (Lana et al., 1991).
     Though saltmarsh species are generally not common in mangrove habitats, a large
number of other non-mangrove plant species may be found coexisting with the mangroves.
A floristic survey of the tidal mangrove flora in the Sunderbans, India, documented 1175
angiosperm species in 680 genera and 154 families (Nasakar and Bakshi, 1993). Working
in the tropical mangrove forests of the Yucatan Peninsula, Olmsted and Gomez (1996)
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        38


found approximately 100 epiphytic species in the families Orchidaceae, Bromeliaceae,
Cactaceae, Araceae, Piperaceae and Polypodiaceae scattered through the canopy and on
trunks of mangrove trees. The orchid Brassavola nodosa is an epiphyte on red mangroves
(Rhizophora mangle) in Belize, Central America, where it grows anywhere from 1-300 cm
above the ground. The largest specimens occur high above the ground where plentiful light
enables them to flower continuously through the summer (Murren and Ellison, 1996).
Lichens may also be abundant on the bark of the mangroves in some habitats (J.C. Ellison,
1997).

5. MANGROVE-ASSOCIATED FAUNA
5.1. Zooplankton

    Diverse communities of zooplankton exist in mangrove habitats and abundances
can be extremely high, reaching 105 individuals m-3 with biomasses up to 623 mg m-3.
These numbers are significantly higher than what is often recorded in offshore waters
(reviewed by Robertson and Blaber, 1992) and the planktonic organisms may contribute to
regional food webs. Such high abundances, however, do not occur in all mangrove
environments. On the west coast of India, for example, Goswami (1992) found lower
zooplankton biomass in the mangroves than in contiguous estuarine and neritic habitats.
    Zooplankton in mangrove waters can be grouped into three size classes. The
smallest organisms are the microzooplankton (organisms between 20 and 199 µm). This
group includes tintinnids, radiolarians, foraminiferans, ciliates, rotifers, copepod nauplii,
barnacle nauplii, and mollusk veligers. Krishnamurthy et al. (1995b) found 81 such species
in the Pichavaram mangroves of south India. Tintinnids were the dominant
microzooplankters with 50 species and densities ranging from 60 to 44,990 individuals m-
3
 . The most important genera were Tintinnopsis and Favella (Godhantaraman, 1994;
Krishnamurthy et al., 1995b). They also found 40 rotifer species in 17 genera. Except for
rotifers, whose populations peaked in the premonsoon and monsoon months, the microzoo-
plankters were most abundant in the summer, corresponding with highest phytoplankton
abundance.
    Copepods are the most abundant group in the mangrove mesoplankton (organisms
between 200 µm and 2 mm). In the Pichavaram mangroves of south India, copepod
densities reach 80,740 individuals • m-3 (Godhantaraman, 1994); the genera Acartia and
Acrocalanus (Calanoida), Macrosetella and Euterpina (Harpacticoida) and Oithona
(Cyclopoida) are the most abundant. In Kenyan mangrove waters, copepods constitute
48.5-92.4% of the zooplankton. Zooplankton counts are high in the creek mouth compared
to the inner creek. Abundances peak around May when heavy rains increase nutrient input
(Osore, 1992). Species in the cyclopoid genus Oithona are particularly abundant in many
studies of mangrove plankton. Harpacticoids (e.g., Pseudodiaptomus spp.) and calanoids
(e.g., Acartia spp., Paracalanus spp. and Parvocalanus spp.) are also important (Ambler et
al., 1991). Barnacle nauplii occur in mangrove canals throughout Raby Bay, Australia, but
the copepod Acartia tranteri is found only in the innermost canals (King and Williamson,
1995).
    Dioithona oculata is a particularly interesting member of the copepod assemblage
in some mangrove habitats. Individuals congregate to form swarms in light shafts among
mangrove prop roots. The swarms maintain their position in currents up to 2 cm • sec-1
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       39


(Buskey et al., 1996). Buskey et al. (1995) showed that the swarms form in response to an
endogenous rhythm. They cannot, therefore, be induced to swarm in artificial light shafts
created at night.
     Copepods and other mesoplanktonic organisms are food for the macrozooplankton
(organisms larger than 2 mm). Jellyfish are the most important macrozooplanktonic
species. The medusa Tripedalia cystophora is attracted to light shafts where non-breeding
individuals actively feed on copepods (reproductive males and gravid females do not feed;
R.W. Stewart, 1996). Planula larvae of Cassiopea species show a strong preference for
mangrove substrata, specifically settling and undergoing metamorphosis on submerged,
deteriorating mangrove leaves (Hofmann et al., 1996). The larvae are apparently attracted
to a soluble protein (molecular weight > 5000 daltons) leaching from the mangrove leaves
(Fitt, 1991, Fleck and Fitt, 1999).
     Meroplankton (planktonic larval stages of benthic invertebrates) may constitute up
to 70% of the zooplankton and span a range the full range of zooplankton sizes.
Brachyuran zoeae can be especially abundant. For example, decapod larval densities
reached 1000 individuals • m-3 in a mangrove area of Costa Rica. These early larval stages
are exported from the mangrove areas on outgoing tides; incoming tides bring the older
stages back to the habitat (Dittel and Epifanio, 1990).
     Bingham (1992) studied larval recruitment of invertebrates (e.g., sponges, oysters,
barnacles, bryozoans, ascidians) living epifaunally on Rhizophora mangle prop roots in the
Indian River Lagoon, Florida (USA). The major factor controlling adult distributions was
transport and recruitment of planktonic larvae as influenced by water flow through the
habitat. Physical factors also contributed to community structure, but on much larger
scales. Farnsworth and Ellison (1996a) reached similar conclusions for R. mangle root
communities in Belize, Central America. To better understand larval recruitment processes
and their importance to the structure and dynamics of mangrove marine communities,
Wolanski and Sarsenski (1997) have developed computer models that simulate the
dispersal of fish and shrimp larvae through mangrove habitats.

5.2. Sponges and Ascidians

     Because they are often surrounded by muddy or sandy sediments, submerged
mangroves roots, trunks, and branches are islands of habitat that attract rich epifaunal
communities. The epifauna may include a diverse array of invertebrate groups including
sponges, hydroids, anemones, polychaetes, bivalves, barnacles, bryozoans, and ascidians.
Encrusting sponges and ascidians are particularly important in many environments and
may be specially adapted to life there. A number of ascidian and sponge species are largely
restricted to mangrove surfaces (Goodbody, 1993, 1994, 1996; Bingham and Young
1991a; de Weerdt et al., 1991) and epifaunal species that do occur in other habitats may
show distinctly different growth forms when they are attached to mangrove roots
(Swearingen and Pawlik, 1998).
     As with mangrove bacterial, fungal, and algal communities the invertebrate
epifauna can show distinct distributional patterns correlated with desiccation, wave action,
temperature and salinity. Rützler (1995) described vertical zonation of sponges on the prop
roots of Rhizophora mangle in Belize. Differential desiccation tolerance produced the
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                          40


zonation, with the most resistant species occurring higher on the roots. Farnsworth and
Ellison (1996a) found a particularly rich ascidian epifauna on mangroves in leeward areas
of another Belizian mangal.
     Epifaunal organisms may play important roles in the structure and function of the
mangal. Sponges, for example, may be food resources for other invertebrates and fish.
Many sponges have anti-predator defenses including siliceous or calcareous spicules and
noxious or toxic chemicals (McClintock et al., 1997). However, mangrove species are
generally not as well defended chemically as sponges from reef habitats (Pawlik et al.,
1995; Dunlap and Pawlik, 1996). Surprisingly, the palatable species also seem to lack any
particular structural or nutritional features that would discourage predators (Swearingen
and Pawlik, 1998). In light of this vulnerability, the mangrove habitat itself may, to some
extent, be a refuge for less protected species. Species here may also rely on faster growth
or greater reproductive output to compensate for predation losses (Chanas and Pawlik,
1995). In contrast to the sponges, some of the mangrove ascidians may have unusual
chemicals that are potent feeding deterrents (Vervoort et al., 1997).
     Mangrove sponges may also lack the allelochemicals that protect them from
overgrowth by other species in space-limited coral environments. Bingham and Young
(1991b) tested 8 sponges commonly found on submerged roots of Rhizophora mangle in
the Indian River, Florida, the Florida Keys, and Belize, Central America. None of the
sponges appeared to use allelochemicals to reduce settlement or survival of potential
competitors. In fact, several epifaunal invertebrate species recruited more heavily in the
presence of the sponges.
     Despite a seeming lower level of anti-predator and anti-competitor chemicals in
mangal than in coral reef communities, epifauna invertebrates in the habitat may still be
sources for interesting, and valuable, compounds. Ecteinascidia turbinata, for instance, is a
colonial ascidian that grows primarily on the submerged prop roots of Rhizophora mangle
in many areas of the Caribbean. It was recently discovered that E. turbinata produces
compounds (ecteinascidins, Figure 7) that show strong activity against a variety of
carcinomas, melanomas, and lymphomas (Rinehart et al., 1990; Wright et al., 1990, Sakai
et al., 1992). This discovery has led to large scale collection of this species for extraction,
isolation, purification and testing of the compounds. Depending on the method used, these
collections adversely affect the wild populations (in addition to damaging the mangrove
trees on which they grow; Pain, 1996). Long
distance dispersal of E. turbinata appears to
depend on rafting of adult colonies; larval
dispersal is highly localized. Collection
techniques that damage the mangroves or
remove large patches of the population,
therefore, could have severe consequences of this
species (Bingham and Young, 1991a).
     The close association of invertebrate
epifauna and mangroves may have led to
                            Figure 7. Bioactive compound (ecteinascidin)
mutualisms between them. For example,         extracted from the mangrove ascidian
sponges and ascidians may protect the         Ecteinascidia turbinata. The ecteinascidins
mangroves on which they grow. Ellison and       have shown strong in vivo activity against a
                            variety of cancer cells.
Farnsworth (1990, 1992) found that epifaunal
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                              41


sponges and ascidians decreased the amount of damage wood-boring isopods did to the
roots of Rhizophora mangle. Roots without the sponge/ascidian cover showed significantly
more damage and 55% lower growth. In estuarine regions where physical conditions
prevented establishment of epifaunal sponges and ascidians, nearly 100% of the R. mangle
roots were damaged by the isopods.
    The invertebrate/mangrove mutualism may also take the form of a symbiotic
nutrient exchange. Sponges attached to submerged roots of Rhizophora mangle induce the
roots to produce fine rootlets that penetrate and grow throughout the sponge tissue.
Measurements indicate that the roots obtain dissolved inorganic nitrogen from the sponges.
The sponges, in turn, obtain carbon from the roots. Ellison et al. (1996) experimentally
transplanted sponges to bare R. mangle roots in a Belizean mangrove habitat. Within 4
weeks, adventitious rootlets had appeared over the surface of the root. The sponges
attached to the roots grew 1.4 – 10 times faster than did control sponges attached to PVC
pipes in the same habitat. Miller-Way and Twilley (1999) suggest that nitrogen-fixing
bacteria living symbiotically with Ulosa
                            80
rutzleri and Lissodentoryx isodictyalis on            E. turbinata     H. magniconulosa
mangrove roots release significant amounts
                            60
of NO3 to surrounding waters.
                          Percent Cover
    The epifaunal communities on
                            40        L. isodictyalis
mangrove roots may show strong
fluctuations. In the Florida Keys, USA,        20
Rhizophora mangle root communities
change dramatically over short time           0
intervals (1-2 months, Figure 8). Physical
                                  Jul




                                        Jul




                                              Jul
                                     Jan




                                           Jan




                                                 Jan
disturbance from tidal flows, species-
specific predation and fragmentation of the Figure 8. Fluctuations in cover of two epifaunal sponges
dominant sponges produce the variability.    (Lissodendoryx isodictyalis and Haliclona magniconulosa)
                        and a colonial ascidian (Ecteinascidia turbinata) on
The perturbations prevent competitive
                        submerged Rhizophora mangle prop root (Florida Keys,
processes from producing the more stable    USA). Photographic measurements were made at 2-3 month
equilibrium assemblages seen in some      intervals for 32 months (after Bingham & Young, 1995).
other mangrove epifaunal communities
(Bingham and Young, 1995).

5.3. Epibenthos, infauna, and meiofauna
    The muddy or sandy sediments of the mangal may be home to a variety of
epibenthic, infaunal, and meiofaunal invertebrates. The composition and importance of
these communities varies enormously from habitat to habitat depending on the sediment
characteristics of the individual mangal.
    Mangrove sediments generally support higher densities of benthic organisms than
do adjacent non-vegetated sediments (Edgar, 1990, Sasekumar and Chong, 1998). Sheridan
(1997) identified over 300 benthic taxa in red mangrove (Rhizophora mangle), seagrass
and mud habitats in southern Florida. Densities, which ranged from 22,591 - 52,914
individuals • m-2 were always higher in the red mangrove peat than in the other habitats.
The fauna was composed primarily of annelids and tanaids with maximum densities of
31,388 and 35,127 individuals • m-2 respectively (Sheridan, 1997).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       42


     The epibenthos may include hydrozoans. For instance, the hydrozoan Vallentinia
gabriellae, which feeds on a variety of zooplankters, is common in some south Floridan
mangals (Rey et al., 1992). Calder (1991) found that hydroids in a Belizean mangal
respond to water flow. The hydroid fauna is richer and more diverse in areas exposed to
waves and tidal currents than in sheltered, still-water areas of the mangal. Polychaetes are
the dominant macrobenthos in mangrove flats at Inhaca Island, Mozambique where their
distributions are controlled by sediment grain size, salinity and ground water (Guerreiro et
al., 1996). Oligochates may also be abundant in shallow mangal waters. Diaz and Erseus
(1994) found one oligochaete family, the Limnodriloidinae, entirely restricted to mangrove
muds.
     The most successful benthic species in the mangal are those that can adapt to the
salinity and temperature stresses that are characteristic of these environments (Ferraris et
al., 1994). Extreme fluctuations in these physical features may prevent colonization by
benthic species. For example, Lana et al. (1997) found that benthic infaunal abundance
and diversity were significantly lower in mangrove sites than in more seaward zones of
Paranagua Bay, Brazil.
     Mangrove meiofaunal communities may also include annelids (especially
oligochaetes) and crustaceans. However, they are generally dominated by nematodes. As a
result, nematodes have been better studied than any other members of the mangrove
meiofauna (Olafsson, 1996). In the dry tropical mangroves of northeastern Queensland,
nematode abundances may reach 2117 individuals • cm-2 with seasonal fluctuations
contributing to variability in the community (Alongi, 1990a). The study of mangrove
meiofaunal communities has led to descriptions of several new nematode species. These
include Parapinnanema ritae, P. alii and P. rhipsoides from Guadeloupe (Gourbault and
Vincx, 1994); Chromaspirina okemwai, Pseudochromadora interdigitatum and
Eubostrichus africanus from Ceriops sediments along the Belgian coast of the North Sea
(Muthumbi et al., 1995); and Papillonema danieli and Papillonema clavatum from Ceriops
sediments of Kenya (Verschelde et al., 1995).
     The distribution of the nematode fauna has been intensively studied in a temperate
mangrove mudflat of southeastern Australia (Nicholas et al., 1991). Approximately 85% of
the nematodes occurred in the top layer of the soft mud, but 5-7 species penetrated the
deeper anoxic muds down to 10 cm. Abundances were affected by tidal zonation.
Nematode biomass was approximately 888 mg dry wt m-2 (≈ 383 mg C m-2) in the low
tide zone but was only 19 mg dry wt m-2 (≈ 8 mg C m-2) in the upper tide zone.
     Nematode populations may vary with food content, grain size and organic content
of the mangrove sediment (Hodda, 1990). The meiofaunal community is undoubtedly part
of the detrital food web. Tietjen and Alongi (1990) found a significant correlation between
biomass of Avicennia marina litter, bacterial abundance, and nematode abundance. The
relationship disappears as detritus ages. However, a direct role of nematodes in organic
matter cycling could not be demonstrated experimentally. Nor does the meiobenthic
community appear to have much direct predator/prey interactions with the epibenthos.
Schrijvers et al. (1995, 1997) showed this experimentally through exclusions of
meiobenthic species from Kenyan Ceriops tagal and Avicennia marina habitats. There is
still much to learn about the role of these less-easily studied members of the mangal
community.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       43




5.4. Prawns, shrimp and other crustaceans
5.4.1. Prawns and shrimp
     Mangrove habitats and prawn/shrimp populations are tightly linked in many
regions. Analyses of commercial prawn catches have repeatedly shown strong correlations
between abundance and biomass of prawns and extent of the surrounding mangrove areas
(Sasekumar et al., 1992; Kathiresan et al., 1994c; Vance et al., 1996b).
     Robertson and Blaber (1992) proposed three explanations for this relationship.
First, organic detritus exported directly from the mangroves provides food and habitat for
juvenile penaeids in offshore areas (Daniel and Robertson, 1990). Second, the waters in
the numerous channels and small creeks of the mangrove receive high levels of terrestrial
runoff, rich in nutrients. Export of these nutrients (controlled largely by groundwater
flows; Mazda et al., 1990b; Ovalle et al., 1990) contribute to productivity. This
productivity, in turn, may support offshore penaeid populations. Third, the mangrove
waterways directly serve as nursery grounds for juvenile penaeids that move offshore and
enter the commercial fishery as they mature. This hypothesis is strongly supported by
surveys of larval, postlarval and juvenile penaeids in nearshore habitats (Vance et al.,
1990, 1996b, 1997; Mohan et al., 1997; Primavera, 1998; Rajendran and Kathiresan,
1999a).
     Sheridan (1992) found low shrimp abundance among Rhizophora mangle prop
roots in Rookery Bay, Florida. Only 4% of the collected animals were in the roots,
compared with 74% in adjacent seagrass beds. This, however, seems to be unusual, and
numbers and biomass of prawns and shrimp are generally higher in mangrove areas than in
adjacent nearshore habitats (Chong et al., 1990; Sasekumar et al., 1992). Study of these
diverse shrimp communities is revealing new species (Miya, 1991; Bruce, 1991).
     In a six-year study, Vance et al. (1997) determined the primary factors controlling
juvenile prawn abundance in mangroves to be larval supply and postlarval settlement. The
young of many shrimp species appear to use the mangal. Juveniles of eight penaeid prawn
species (primarily Metapenaeus monoceras and Penaeus indicus) are common in the
Pichavaram mangroves. Catches of the juveniles in core mangrove areas are greater than in
open waters (Rajendran, 1997). In Oman, R. Mohan and Siddeek (1996) similarly found
abundant postlarval and juvenile shrimp in the detritus-rich, muddy substrates of a mangal
they studied. Distributions of the juveniles within the mangal are strongly influenced by
salinity; densities are highest at intermediate salinities (R, Mohan et al., 1995).
     As the shrimp grow, they may eventually leave the mangal. In the Matang
mangroves of Malaysia, Chong et al. (1994) measured prawn densities of 4092 individuals
• ha-1 in the mangal but only 2668 individuals ha-1 in the adjacent mudflats. However,
biomass was approximately the same in both areas, suggesting that larger individuals move
out of the mangal. Using size distribution data, Rajendran and Kathiresan (1999a),
concluded that postlarvae of prawns recruit into the Pichavaram mangroves in the
postmonsoon period; subadults then leave during the premonsoon and monsoon periods.
The annual offshore commercial catch of adult P. merguiensis is significantly correlated
with number of prawn emigrating from the estuary during the wet season (Vance et al.,
1997). There is also a strong positive relationship between rainfall and subsequent offshore
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       44


commercial catch of adult shrimp (Staples et al., 1995), probably due to flushing from the
mangrove habitat as a result of heavy rains.
    Although the mangal may be a sink for settlement and early growth of shrimp and
prawns, it may also be a source for larvae that are transported to other habitats. Mangrove
waters in the Klang Strait of Malaysia may collect 65 billion penaeid prawn larvae before
their annual transport and settlement in coastal nursery grounds. Tidal currents and lateral
trapping in mangrove-lined channels cause this aggregation (Chong et al., 1996).
    There may be a number of benefits for juvenile shrimp and prawns living in
mangrove habitats. The habitat is complex and provides a variety of niches within which
species can exist. For example, in the mangroves of Muthupet, India, Penaeus indicus, P.
merguiensis and Metapenaeus dobsoni show clear preference for detritus-rich muddy
substrates in which they feed. In contrast, P. monodon shows no such preference. Other
shrimp feed directly on the mangroves. Cholesterol extracted from Rhizophora leaves
promotes growth of juvenile Penaeus indicus and increases their conversion efficiency
(Ramesh and Kathiresan, 1992). However, not all mangrove products are beneficial.
Excoecaria agallocha latex is toxic to larvae of the freshwater prawn Macrobrachium
lamarrei (Krishnamoorthy et al., 1995) and to penaeid prawns (Kathiresan and Thangam,
1987).
    The mangrove forest, with its small creeks and channels, its hanging roots, and soft
substrates may also provide refuge from predators. Prawns in these habitats tend to be most
active near high tide and at night (Stoner, 1991; Vance, 1992; Vance and Staples, 1992;
Rajendran, 1997). This presumably allows them to forage when food is most accessible
and predation danger is lowest.
    Some mangrove shrimp may avoid predation by burrowing in the muddy
sediments. Primavera and Lebata (1995) found that Metapenaeus were particularly active
burrowers. Penaeus monodon is also a burrower, but burrowing activity is size dependent
and increases as the animals grow. Shrimp may also escape predators by migrating with
the tides. Vance et al. (1996a) observed that juvenile P. merguiensis are very mobile,
moving substantial distances into the mangrove forest at high tide. An extreme example of
shrimp migration is the semi-terrestrial Merguia oligodon, a species common in some
Kenyan mangroves. This species lives among the aerial roots of Rhizophora mucronata. It
is active at night, grazes on mangrove bark, and climbs mangrove roots and trunks up to 80
cm above the ground (Vannini and Oluoch, 1993). Vance et al. (1997) have used a stake-
netting method to study distribution and movements of prawn in intertidal mangrove
forests. This technique shows promise as a way to provide better information about the
shrimp and prawns and their roles within the mangal.

5.4.2. Other Crustaceans
    While shrimp and prawns do not generally harm the mangroves, and may actually
be beneficial (e.g., through bioturbation of muddy sediments), other crustaceans do
significant damage. For example, barnacles can grow abundantly on mangrove roots and
pneumatophores (Foster, 1982; Anderson et al., 1988; Bayliss, 1993; Ross and
Underwood, 1997). Balanus amphitrite and other fouling organisms, for instance, kill
42.5% of the mangrove seedlings in Goa, India (Santhakumaran and Sawant, 1994). Some
species of barnacles belonging to the genera Euraphia, Elminius and Hexaminius appear
to prefer mangroves over other substrates. The settling barnacle larvae may even show
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       45


strong preferences for mangrove species and discriminate among parts of the trees. The
barnacle densities are controlled by the physical environment of the mangal (primarily
desiccation and temperature). Populations are greater on seaward than on landward areas
of the forest. Densities are also greater on lower surfaces than on upper surfaces of trunks
and leaves (Ross and Underwood, 1997). In some mangrove areas barnacle numbers are
also greater in the mid-intertidal than in the upper or lower intertidal zones (Kathiresan et
al., 2000), and the species show marked zonation (Baylisss, 1993), with chthamalids
occuring above Balanus amphitrite. The recruitment of barnacles, and other sessile
invertebrates within the mangal is largely controlled by larval abundance, tidal currents,
duration of larval life and density of the adult populations (Bingham, 1992; Farnsworth and
Ellison, 1995; Young, 1995). It was at one time thought that the selection by barnacles of
their mangrove habitat was so extreme that one species of Hexaminius (H. folorium)
occurred only on twigs and leaves, while another (H. popeiana) was restricted to the bark
(Anderson et al., 1988; Ross and Underwood, 1997). Recent studies indicate that
phenotypic variation is responsible, and that the leaf-occurring form is comparable
morphologically and by DNA to the form on mangrove bark (Ross, 1996; Ross and
Pannacciulli, personal communication). This situation can be compared with the changes
in form and colour seen in the bivalve Enigmonia aenigmatica and the snail Littoraria
pallescens when inhabiting different parts of the mangrove, trees as noted on page 81.
     Burrowing isopods (e.g., Sphaeroma terebrans and S. peruvianum) also do
tremendous damage to mangroves in many regions of the Atlantic, the Caribbean, and the
eastern Pacific. Numerous juveniles and adults can be found living inside a single root or
stem. Their burrowing can significantly affect root growth and development (Ellison and
Farnsworth, 1990; Santhakumari, 1991).
     Other crustaceans use mangrove waters temporarily during certain phases of their
life history. One of the better known is the Caribbean spiny lobster (Panulirus argus),
juveniles of which use the mangroves as nursery habitat (Monterrosa, 1991). Like the
shrimp and prawns, however, the lobsters migrate out of the mangal as they grow. Adult
lobsters remain in the mangal only if their preferred habitat (under coral heads) is
unavailable (Acosta and Butler, 1997). Migration to other habitats may reflect a search for
better food resources.

5.5. Crabs
    Crabs are characteristic members of the invertebrate mangrove fauna and have
received much attention. Some indication of the diverse array of mangrove-associated
crabs can be found in annotated checklists from India (Sethuramalingam and Ajmal Khan,
1991), Malaysia and Singapore (C.G.S. Tan and Ng, 1994) and Brazil (Vergara-Filho et
al., 1997).
    Within the complex mangrove environment, crabs fill a variety of niches. For some
species, the relationship with mangroves is obligatory; they depend directly on the
mangroves for survival (Vergara-Filho et al., 1997). Others simply have ranges that
overlap the mangal. The mud crab Scylla serrata inhabits seagrass and algal beds in the
mangroves of Pichavaram, south India (Chandrasekaran and Natarajan, 1994). Floating
leaves in a Costa Rican mangal harbor a unique community dominated by Uca crabs
(77.8% of all organisms counted; Wehrtmann and Dittel, 1990). Perhaps one of the most
striking associations is seen with the hermit crab Clibanarius laevimanus. Individuals of
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        46


this species climb the mangrove roots and rest on them during the entire low water period,
forming dense clusters of up to 5,000 individuals (Gherardi et al., 1991; Gherardi and
Vannini, 1993). Gherardi et al. (1994) have studied size, sex and shell characteristics of
this unique mangal species.
     Mangrove crabs are morphologically, physiologically, and behaviorally
well-adapted to their environment. For example, the semaphore crab, Heloecius cordifor-
mis, is active at low tide when it is completely exposed to air. Its branchial chambers are
modified for respiration both in the air and under water (Maitland, 1990). A number of
other crab species (particularly in the Family Grapsidae) live directly on the mangrove
trees. Species in this group generally have a square, flattened carapace, a relative
shortening of the dactylus on the walking legs and a lengthening of the propodus (Vannini
et al., 1997a). These structural specializations appear to be adaptations for their tree-
dwelling existence.
     Crabs living in the mangal must adjust to significant temperature and salinity
fluctuations. Some, like the grapsid Metopograpsus messor, retreat to burrows where
temperatures are less variable and consistently lower than the sediment or air temperatures.
When it is out of the burrow, M. messor uses evaporative cooling to keep its body
temperature lower than the surrounding air (Eshky et al., 1995).
     Other crabs have adopted a nocturnal lifestyle, possibly to escape high temperatures
and/or predators (Micheli et al., 1991). The hermit crabs Coenobita rugosus and Coenobita
cavipes are active 24 hours a day but are most active when they are among the mangrove
roots. Barnes (1997) suggests that they do this because wind speeds (and desiccation
potential) are lower there. Desiccation can significantly affect ion balances and mangrove
crabs are physiologically adapted to resist major changes. In Ucides cordatus and Carcinus
maenas total Na+ efflux is markedly reduced during emersion. The reduction in ion and
water loss results from decreased urine output (Harris et al., 1993). When U. cordatus is
placed in low salinity water, active sodium uptake increases 4-5 fold (Harris and Santos,
1993).
     Crabs in mangrove habitats show distinct distributional patterns related to substrate
characteristics, salinity, degree of tidal inundation, and wave exposure. In the Indian
Sunderbans, these conditions produce a vertical zonation of crab species (Chakraborty and
Choudhury, 1992, Kathiresan et al., 2000). Machiwa and Hallberg (1995) also found a
horizontal zonation of crabs in East Africa,. The terrestrial edge of the mangal was
occupied by grapsids while mixed associations of ocypodids dominated the open areas of
sand and mud.
     Different crab species respond differently to disturbance and this affects species
distributions. In Kenya, Sesarma guttatum prefers shaded habitats and is most common in
regions with an established mangrove canopy. In contrast, Uca urvillei and
Microphthalmus depressus prefer clear-cut areas. In the more landward Avicennia zone,
the species composition of the crab community remains constant whether the vegetation is
intact or clear-cut (Ruwa, 1997).
     Mangrove crabs can be divided into distinct guilds based on their feeding mode.
Some species (e.g., Uca and Macrophthalmus spp.) are detritivores that extract their food
from the sediments while others (e.g., the portunid Scylla serrata) are opportunistic
scavengers (Micheli et al., 1991). There are also a number of active predators. The
swimming portunid Thalamita crenata lives on the extreme seaward fringe of the
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        47


mangrove swamp where it preys heavily on bivalves and slow-moving crustaceans
(Cannicci et al., 1996c). This species is active during high tides, but only when the water is
between 10 and 30 cm deep, suggesting that their foraging behavior is controlled by
hydrostatic pressure changes associated with tidal flux (Vezzosi et al., 1995).
    Epixanthus denatus, which is very abundant in mangrove creeks along the Kenyan
coast, is another active predator. It forms dens among the mangrove roots and feeds on
almost any slow-moving invertebrate (including other crabs) that comes within a 3-m
radius. Their intense predation may be responsible for the climbing behaviour of many
potential prey species (Cannicci et al., 1998). Crabs in the mangrove face significant
predation risks and may show specific anti-predator adaptations. Diaz et al. (1995) suggest
that postlarval and juvenile crabs may avoid predation by responding to specific light cues.
Aratus pisonii, which lives among roots and branches, is attracted to narrow dark
rectangles but avoids large dark rectangles. The authors speculate that the narrow
rectangles resemble roots that represent refuges while the larger rectangles indicate
predators. In contrast, Chlorodiella longimana, a subtidal species, moves toward all dark
rectangles regardless of their size.
    In addition to scavenger and predator guilds, there is a guild of herbivorous
mangrove crabs that feed directly on mangrove litter. In Ao Nam Bor, Thailand, up to 82%
of the diet of sesarmid crabs consists of mangrove material (Poovachiranon and
Tantichodok, 1991). A number of these herbivores show clear feeding preferences. For
example, Sesarma meinertii generally prefers Bruguiera gymnorrhiza to Avicennia marina
leaves (Micheli et al., 1991). However, Steinke et al. (1993a) showed that the age of the
litter was more important than its source in determining preference. The crabs chose
yellow B. gymnorrhiza and A. marina leaves over green leaves of either species. Sesarma
messa and S. smithii both prefer decaying leaves to those that are simply senescent,
irrespective of leaf species. Neosarmatium meinerti does not choose among mangrove
species but does, however, strongly prefer fresh leaves. Its heavy, non-selective feeding on
mangrove seedlings and propagules could make it a significant threat to afforestation
efforts (Dahdouh-Guebas et al., 1997).
    It is unclear what factors are responsible for these feeding preferences. Micheli
(1993a, b) found that preferences were not affected by tannins, water content, % organics,
C:N ratio, or leaf toughness. Many of the herbivorous crabs store the leaves in their
burrows for some time. However, the nutritional value of the leaves does not increase
during the time they are stored, indicating that the crabs are simply storing the leaves and
not gardening them to encourage bacterial or fungal growth (Micheli, 1993a, b). Given that
the mangrove leaves, in general, have low nutritional value and the crabs do not have a
mechanism to promote bacterial or fungal growth, it may be very important for them to get
the maximum food value out of the leaves they eat. This may contribute to their specific
preferences.
    Although mangrove leaves are not particularly nutritious, they do produce
sufficient energy to influence survival, growth and reproduction of the crabs and it seems
reasonable to assume that is a sufficient selective force to produce feeding preferences.
Survival of two mangrove crabs, Chiromanthes bidens and Parasesarma plicata, is
directly related to litter type. They do best when fed brown Avicennia marina leaves,
followed by brown Kandelia candel, yellow A. marina, and finally, yellow K. candel
(Kwok and Lee, 1995). Mangroves at different sites in Venezuela produce leaves with
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                              48


different nutritional value. The crabs are smaller in sites where the stunted trees produce
leaves of low nutritional value (Conde and Diaz, 1992; Conde et al., 1995).
                                      Some of the herbivorous
                                  crabs do not simply graze on
                                  fallen leaves. Some actively
                                  forage in the canopy of the tree.
                                  In the mangrove swamps of East
                                  Africa, Sesarma leptosoma is an
                                  active climber that can reach the
                                  tops of the tallest trees. It never
                                  descends into the water nor
                                  venture out on the mud at low
                                  tide. This behaviour provides
                                  protection from predators
 Figure 9. Average daily migration patterns of the crab
                                  (Cannicci et al., 1996a). It spends
 Sesarma leptosoma into and out of the mangrove canopy (after
 Vannini and Ruwa, 1994).                     most of the night among the
                                  mangrove roots but, in the
morning, moves up into the canopy to feed on fresh leaves. Increasing temperatures and
the danger of desiccation eventually drive the crabs down to the bases of the trees where
they spend the hottest hours of the day. In the evening, they return to the canopy for
another short feeding period (Figure 9; Vannini and Ruwa, 1994; Vannini et al., 1997b).
The crabs tend to return to the same feeding
spot each time they visit the canopy and even
follow the same path to get there (Figure 10).
Cannicci et al. (1996b) suggest that site fidelity
is important as it takes the crabs near leaf buds
where they can find water trapped among the
scales. Reduced light delays the migration. It is
unclear why migration of this intertidal animal
is regulated by light instead of tides (Vannini
et al., 1995).
     Feeding by crabs hastens composting of
mangrove material and contributes to cycling
of nutrients through the mangal (Lee, 1998). In
the Gazi Bay, Kenya, crabs (along with large
snails) process over 18% of the fallen litter
(Slim et al., 1997). The large mangrove
grapsid Sesarma meinertii consumes Avicennia
marina leaves at a rate of 0.78 g m-2 d-1,
accounting for 43.58% of the leaf-fall in a        Figure 10. Branch fidelity in the mangrove
warm temperate mangrove swamp in southern         crab Sesarma leptosoma. Arrows indicate
                             movement of three individuals. Individuals B
Africa (Emmerson and McGwynne, 1992).
                             and C always returned to the same branch in
     Digging by crabs, in conjunction with
                             the canopy and even followed the same path
other benthic fauna like nematodes,            to get there. Individual A returned to one of 2
polychaetes, and mudskippers (Kristensen et        branches (after Cannicci et al, 1996b).
al., 1995) can also have a profound effect on
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                                        49


nutrient cycling and the physical and chemical environment of the mangal (Lee, 1998).
Burrows enhance aeration, facilitate drainage of the soils, and promote nutrient exchange
between the sediments and the overlaying tidal waters (Ruwa, 1990). Crab burrows
generally have two or more openings and may form extensive labyrinths of interconnected
tunnels. Using dye injections and flow measurements, Ridd (1996) estimated that, in a 1
km2 area of a North Queensland mangal, 1,000 to 10,000 m3 of water move through crab
burrows on each tidal cycle. T.J. Smith et al. (1991) removed burrowing crabs from a
mangal and observed significant increases in soil sulfide and ammonium levels relative to
control sites. These chemical changes led to decreased mangrove growth and reproduction.



5.6. Insects
    Insects constitute a significant portion of the fauna in many mangrove
communities. They may be permanent residents of the mangal or only transient visitors. In
either case, they often play important roles in the ecology of the system and contribute to
the unique character of these habitats. Surveys of mangrove insects are revealing complex
assemblages of species filling a wide variety of niches. For example, Veenakumari et al.
(1997) found 276 insect species in the mangals of Andaman and Nicobar Islands of India;
197 of these were herbivores, 43 were parasites and 36 were predators. Similar levels of
diversity and abundance have been found in the insect fauna of Thailand’s Ranong
mangroves (Murphy, 1990a). Many of the insects reported in mangals are only temporary
visitors; their ranges included many other habitat types. As a result, they provide linkages
between the mangal and other environments (Ananda Rao et al., 1998).
    Terrestrial organisms living in mangrove environments are faced with harsh
conditions of strong sunlight, high temperatures and desiccation. Many of the insects (and
other terrestrial arthropods) avoid these conditions by emerging only at night, or by living
entirely within the plants. In the mangals of Belize, wood-boring moths and beetles
excavate tunnels through the mangroves. The
                               12
tunnels then become home to more than 70 other                           Leaves always exposed
species of ants, spiders, mites, moths, roaches,                          Leaves submerged at high tide


termites, and scorpions (Rützler and Feller, 1996;
                                       Percent damaged




                                8

Feller and Mathis, 1997). A number of organisms
(including isopods, amphipods, myriapods, and
                                4
spiders in addition to insects) escape high
temperatures and desiccation by living in the
intertidal portions of the mangal. During periods       0
                                 Total leaf Holes through Damage to Internal leaf   Damage to
of high tide, these organisms retreat to air-filled        damage   the leaf  leaf margin  damage   underside of
                                                           primary vein

cavities where they remain until they are again      Figure 11. Effect of tidal submergence on
exposed by the falling water (Murphy, 1990b).       Laguncularia racemosa leaf damage
                              (Guanacaste Province, Costa Rica).
    Herbivorous insects can cause significant
                              Submergence limits access by terrestrial
damage to the mangroves, attacking leaves and
                              herbivores and may also cause chemical and
boring through the wood. Seedlings may be         structural changes in the foliage. Standard
particularly vulnerable to attack and are strongly     errors are shown. Statistical analysis showed
affected by proximity to adult trees. In a Belizean    that exposed leaves had significantly more
                              damage in all cases (after Stowe, 1995).
mangal, seedlings growing under the intact adult
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       50


canopy suffered twice as much herbivore damage as seedlings in areas without an
established canopy (Farnsworth and Ellison, 1991). Immersion in seawater may help
protect the trees. Portions of the mangrove canopy that are submerged by tidal waters
suffer significantly less herbivore damage than those that remain exposed (Stowe, 1995;
Figure 11).
    Recent records of insects in mangrove include 28 species of dragonflies in India
(Mitra, 1992), a water strider, Mesovelia polhemusi in Belize (Spangler, 1990), an unusual
psyllid, Telmapsylla, in Florida and Costa Rica (Hodkinson, 1992), and termites,
Nasutitermes nigriceps, in Jamaica (Clarke and Garraway, 1994). However, some of the
more important and best studied mangrove insects are bees, ants, and mosquitoes. Honey
bees produce significant quantities of honey from the mangroves of India, Bangaladesh,
the Caribbean and southwest Florida. The honey is an important food resource for humans
in some regions (e.g., Padrón et al., 1993). In India, the dominant bee species (Apis
dorsata), may travel hundreds of miles to forage in the mangrove forests during periods of
peak blooming (March and July). It builds honeycombs on several mangrove species, but
prefers Excoecaria (Krishnamurthy, 1990). In contrast, the same bee species in southern
Vietnam forages on mangrove vegetation primarily during the rainy season and rarely
builds combs (Crane et al., 1993).
    Twenty-two ant species are known from Brazilian mangrove habitats. Camponotus
and Solenopsis are most common genera (Cortes-Lopes et al., 1996). Clay and Andersen
(1996) found 16 ant species in an Australian mangal. Two of these, both in the genus
Polyrhachis, are apparently restricted to this habitat. In northern Australia, Polyrhachis
sokolova nests directly in the soft mud of the mangal (Nielsen, 1997). Adams (1994)
studied niche partitioning in four ant species common in Panamanian mangroves. These
species partition the mangrove canopy in non-overlapping territories that are maintained
through a combination of pheromonal signals and tactile displays.
    Holes in the mangrove trees (particularly Avicennia species) and crab-burrows
provide ideal sites for mosquito breeding (Thangam, 1990). Mosquitoes are ubiquitous in
mangrove habitats and may act as vectors for diseases of vertebrates. Populations are often
dense and species diversity can be high (eighteen species occur in the Pichavaram
mangroves of south India alone; Thangam and Kathiresan, 1993b). Predation by fishes
may reduce successful mosquito oviposition. Hence, mosquito populations are lower in
sites with high fish densities (Ritchie and Laidlaw-Bell, 1994). Regardless of its
composition, any mangrove forest that is flooded by < 14% of the highest daily tide can
potentially produce the mosquito Aedes taeniorhynchus (Ritchie and Addison, 1992).
Addison et al. (1992) identified A. taeniorhynchus oviposition sites and quantified larval
production by locating and counting egg shells.
     Mangals tend to be reservoirs for a number of pathogenic viruses including
Dengue, Haemorrhage Fever, Bakau and Ketapang. Mosquitoes are the most common
vectors for these viruses. However, several families of other diptera associated with faecal
contamination in mangroves of Singapore and Malay also contribute to the spread of
human disease (Murphy, 1990c).

5.7. Mollusks
    Mollusks are found throughout most mangrove habitats. They live on and in the
muds, firmly attached to the roots, or forage in the canopy. They occupy a number of
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        51


niches and contribute to the ecology of the mangal in important ways. The nature of the
molluskan community is strongly influenced by physical conditions. For example, in the
mangroves of China, Jiang and Li (1995) found that density and biomass of the mollusks
(including 52 species) were consistently highest in the high tide zones and decreased with
depth. In addition, species abundance increased with salinity. Such a pattern is likely to be
found in other mangals. This sensitivity of mollusks to their physical/chemical
environment may make them good bioindicators. Skilleter (1996) has used the composition
of the molluskan assemblage to assess the health of urban mangrove forests.
    The molluskan fauna in mangrove habitats is composed primarily of bivalves and
snails and most study has focused on these groups (e.g., Balasubrahmanyan, 1994). Other
mollusk groups (e.g., nudibranchs, chitons, scaphopods) are less obvious and have been the
subject of only a few studies (e.g., Sigurdsson, 1991). Much of the work with bivalves and
snails has concerned individual species and their specific adaptation to the mangrove
environment. For example, Dious and Kasinathan (1994) studied the high desiccation,

salinity, and temperatures tolerances of two pulmonate snails, Cassidula nucleus and
Melampus ceylonicus, from a south Indian mangal. Special conditions in the mangal may
result in local adaptation. Crow (1996) compared movement of the snail Bembicium
auratum in mangrove habitats and on rocky shores. Movement patterns in the mangroves
were very different (despite similar distributions). This suggests that models developed in
rocky intertidal communities may not be directly applicable to mangrove communities.
     Color variability may represent a special adaptation to the mangrove environment
or possibly a reaction to the complex chemical defenses of the plants. For example, the
leaf-inhabiting mangrove snail, Littoraria pallescens, has distinct color morphologies that
are sometimes associated with other shell differences. The color variation may reflect
predation pressures (Cook, 1990; Cook and Kenyon, 1993).
     The unique tree-climbing bivalve, Enigmonia aenigmatica, which occurs mostly on
Avicennia and Sonneratia also shows color variation. The shells are normally red to deep
purple. However, the shells of individuals attached directly to the mangrove leaves are
golden yellow. (Sigurdsson and Sundari, 1990). This enigmatic bivalve is one of the gems
of the mangal, and belongs to the Anomiidae, of which most species stay cemented to the
substratum. It uses its highly mobile foot to reach the desired level in the mangroves and
fastens temporarily with transparent byssus threads (Yonge, 1957; Berry, 1975; 1976)
     Other bivalves are adapted to the chemical environment of the mangal. Two
corbiculids, Geloina erosa and G. expansa, from Iriomote Island, Japan, occasionally
secrete thin organic sheets on the inner shell. Formation of these sheets may be a response
to shell dissolution in the acidic mangal environments; the sheets occur only in specimens
that have suffered extensive shell damage (Isaji, 1993, 1995).
     Frenkiel et al.. (1996) reported the bivalve Lucina pectinata from muddy mangrove
sediments. Like most other lucinids inhabiting sulfidic sediments, including also seagrass
beds and salt marshes, this species carries endosymbiotic chemoautotrophic sulfur-
oxidising bacteria in the gill, and the blood is rich in haemoglobin (see Somero et al., 1989
and Fisher, 1990 and references therein). These bivalves get their organic matter from the
bacteria, but the symbiosis requires proximity to both sulfide and oxygen. It has been
suggested that in seagrass beds these bivalves might benefit from the proximity to plant
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                              52


roots carrying oxygen (Fisher & Hand, 1984); a similar relationship could be suggested for
mangrove roots.
    Some mollusks are critical to the basic ecology of some mangals. For example, the
mangrove snail Thais kiosquiformis plays a central role in maintaining the function and
productivity of mangroves in Costa Rica by “cleaning” their root systems of encrusting
barnacles (Koch and Wolff, 1996). Ellison and Farnsworth (1992) measured similar effects
in the mangals of Belize. Detritivorous snails (e.g., Terebralaia palustris in Gazi Bay,
Kenya) aid nutrient cycling in the mangal by processing mangrove litter (Slim et al.,
1997). Bivalves may contribute significantly to the organic biomass in the habitat and may
be a link between phytoplankton communities and higher trophic levels (e.g., Ingole et al.,
1994; Deekae and Idoniboye-Obu, 1995).
    Researchers have collected detailed information on mangrove oysters, largely
because they can be valuable food (e.g., Tack et al., 1992; Ruwa and Polk, 1994). Around
Tuticorin, India, mangroves provide ideal conditions for production of edible oysters
(Crassostrea madrasensis) and oyster beds are an important part of the habitat
(Rajapandian et al., 1990). Newkirk and Richards (1991) have found that exposing the spat
of Crassostrea rhizophorae to air increases growth and enhances the yield of marketable
oysters. This response may reflect an adaptation to the tidal regime of their mangrove
environment.
    Teredinids (shipworms) and pholads are specialized bivalves that burrow through
wood. Some species within these groups do extensive damage to mangroves by destroying
submerged roots and branches. Seven such species (Bankia campanellata, B. carinata,
Dicyathifer manni, Lyrodus pedicellatus and Teredo furcifera, Martesia striata and M.
nari) live in the Pichavaram mangroves of south India (Sivakumar and Kathiresan, 1996).
Morton (1991) recently discovered the first mangrove shipworms in Hong Kong (Lyrodus
singaporeana). A survey for teredinids in the mangroves of Sao Paulo, Brazil revealed four
species (Nausitora fusticula, Bankia fimbriatula, B. gouldi and B. rochi). Differences in
salinity tolerance affect distributions of these species within the mangal (Lopes and Narchi,
1993).


5.8. Fish
     Mangroves have a rich and diverse
assemblage of fish (Figure 12), some with
commercial value. Other fish species are
important links in the mangrove food web.
Still others are only temporary residents
that spend most of their life history
elsewhere. Whatever their role, all are
important to the character of the mangal.
     Extensive studies of the fish
                       Figure 12. The canopy of Rhizophora mangle provides
community have been made in
                       habitat for terrestrial birds and insects. The submerged
Alligator Creek, northeastern
                       prop roots provide solid surfaces for attachment of a
Queensland, Australia (Robertson and     variety of marine invertebrates. In addition, many fish
Duke, 1990a, b); in the Embley River     species use the habitat as a nursery area; the complex
estuary (Blaber et al., 1990 a, b; Salini  tangle of roots provides a refuge from predators (from
                       Rutzler & Feller, 1988).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                          53


et al., 1990; Brewer et al., 1991) and the Leanyer Swamp of the Northern Territory; and in
the Dampier region of Western Australia (Robertson and Blaber, 1992). The fish fauna is
generally very rich; 197 species occur in the mangroves of the Embley River alone. Such
high diversity is not restricted to Australian mangals. One hundred and seventeen fish
species, in 49 genera, have been recorded in the Matang mangrove waters of Malaysia
(Sasekumar et al., 1994; Yap et al., 1994) while Hong and San (1993) reported 260 fish
species in the mangroves of Vietnam.
     Abundances of the fish can also be very high. In Mexican mangroves, fish
biomasses up to 10 g m-2 have been recorded (Flores-Verdugo et al., 1990; Arancibia et
al., 1993). In Moreton Bay, Australia, the biomass reaches 20 g m-2 and ninety-six percent
of the biomass (46% percent of the species, 75% of the total fish) is from species important
in regional fisheries (Morton, 1990). Robertson and Blaber (1992) measured fish
biomasses up to 29 g m-2, with densities up to 161 individuals m-2.
     In a Queensland, Australia mangal, sampling suggests that fish regularly move
through the habitat with the tidal flows. Density and biomass at high tide were 3.5
individuals m-2 and 10.9 g m-2 respectively. On the ebb tide, the fish moved to small,
shallow creeks where density and biomass reached 31.3 individuals m-2 and 29.0 g m-2
(Robertson and Duke, 1990a). Fish distributions and abundances may also change on diel
or seasonal cycles (Chandrasekaran and Natarajan, 1993). In southwestern Puerto Rico,
fish present in the mangal during the day may completely disappear at night (Rooker and
Dennis, 1991). Accurately assessing populations of highly mobile species in such a
complex environment requires special sampling techniques (Lorenz et al., 1997).
     A comparison of catches in various habitats suggests that some species specifically
choose to reside in the mangal. For example, the number of fish species in the coastal
mangroves of Malaysia (119) exceeds that in all other habitats (inshore waters held 92
species, mudflats held 70 species, and near inshore waters held only 58 species; Chong et
al., 1990). A similar result has been found for mangrove habitats in Belize (Sedberry and
Carter, 1993). The relative importance of the mangal as habitat, however, may decrease if
nearby environments include coral reefs. Acosta (1997) found much higher fish diversity
on the reefs of La Parguera, Puerto Rico than in adjacent mangroves.
     Fish in mangrove habitats are important predators, consuming amphipods, isopods,
shrimp, nematodes, insects, gastropods, other fish, and algae (e.g., Erondu, 1990; Brewer
and Warburton, 1992; Williamson et al., 1994; Rooker, 1995; Columbini et al., 1995,
1996). In the Matang mangroves of Malaysia, a suite of fishes feed on shrimp. Croakers
(Family Sciaenidae) specialize on penaeid shrimp, consuming about 1.2 kg of shrimp ha-1
d-1 or about 17% of the total shrimp biomass (Yap et al., 1994). In the Philippines, shrimp
predation is significantly higher in bare sand areas than among mangrove pneumatophores
(Primavera, 1997).
     Feeding activities of mangrove fish can be strongly affected by local conditions. In
northeastern Florida Bay, salinity influences feeding behavior of mojarras and gold-spotted
killifish. In upstream areas with high salinity variation, the fish eat nutritionally poor algae;
in less variable downstream areas, they eat a much better diet of benthic invertebrates. Ley
et al. (1994) suggested that fish gut content measurements can be a tool to assess
environmental conditions and habitat quality. As such, it could be useful in comprehensive
monitoring and restoration programs.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        54


    Mangals may play a special role as nursery habitat for juvenile fish. The juvenile
stages of adults that occur in other habitats (e.g., coral reefs and seagrass beds) may
migrate to the refuge on the mangal (Pinto and Punchihewa, 1996). It is common to find
large numbers of larvae and juvenile fish in net samples from mangrove habitats (Dennis,
1992; Tzeng and Wang Yu, 1992; Alvarez-Léon, 1993; Matheson and Gillmore, 1995) and
densities of juvenile fish in mangrove habitats are often higher than in adjacent habitats
(Robertson and Blaber, 1992). Thollot (1992) found that samples from the mangroves of
Southwest Lagoon, New Caledonia held 262 fish species, including the young of 30% of
the reef species. Most fish collected in the Lagos Lagoon (Nwadukwe, 1995) and in the
mangrove waters of Martinique Island (Louis et al., 1995) were small and sexually
immature. In Belize, most of the fish collected from mangrove waters are juveniles of
species that live out on the marine reefs as adults (Sedberry and Carter, 1993). Despite this
linkage to coral reefs, mangroves also have their own unique fish assemblages. Gill net
sampling in a tropical mangrove creek in SW Madagascar produced 60 species of juvenile
fish. Only six of those occurred on an adjacent coral reef (Laroche et al., 1997).
   Robertson and Blaber (1992) present three explanations for the high density of
juvenile fish in mangrove waters. First, mangrove estuaries supply an enormous amount of
food appropriate for juvenile fish (Chong et al., 1990). Second, reduced visibility in the
turbid mangrove waters may reduce predation by large fish. Third, the structural
complexity of the mangroves provides excellent shelter and protection for the juveniles.
    There is correlative evidence for the third possibility. In the Solomon Islands,
mangrove estuaries clogged with woody debris harbour pomacentrids and some species of
Apogonidae and Gobiidae. These groups are largely absent from mangrove estuaries that
are clear of the woody material (Blaber and Milton, 1990). Daniel and Robertson (1990)
also found a highly significant relationship between amounts of mangrove detritus and fish
densities or biomass in mangrove creeks. In Sri Lanka, fishermen increase their catches in
lagoons by placing thickets of dead mangrove twigs on the lagoon floor. Netting around
these thickets produces much higher catches than in adjacent bare mud areas (Robertson
and Blaber, 1992). Experiments with this technique in south India show that Avicennia
debris works better than Rhizophora debris, producing much higher catches (Rajendran,
1997; Rajendran and Kathiresan, 1998). The stilt and prop roots of some mangroves
provide a complex environment that would seem to provide an ideal refuge. However,
Mullin (1995) found more fish species in the open waters adjacent to Rhizophora mangle
stands than among the prop roots themselves, but it was unclear why the fish avoided the
roots.
    While mangroves in general may serve as nursery habitat for many fish species,
individual mangals may not. For example, Ley et al. (1999) found that mangrove habitats
in northeastern Florida Bay did not function as nursery-grounds. The authors suggest that
this particular mangrove estuary may be atypical for two reasons: (1) it has no lunar tides
and lacks typical tidal circulation, (2) it has little submerged vegetation; and (3) it
experiences severe hypersaline conditions. Conditions in this particular environment may
be sufficiently stressful to prevent its use by juvenile fish.
    Fish living in the mangal must adjust to temporal and spatial variability in physical
and chemical conditions, and some species possess specific adaptations to deal with this.
For example, the widely distributed hermaphrodite killifish (Rivulus marmoratus) is well
adapted to mangrove microhabitats (Taylor et al., 1995). Its specializations include an
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                          55


ability to survive in moist detrital substrates during periods of low water or drought and
reproduction by internal self-fertilization (Davis et al., 1995).
     Fish without such specific adaptations may respond behaviorally to physical cues
that indicate physically or chemically stressful microhabitats. This can lead to distinct
distributional patterns (Blaber et al., 1994). For example, Heath et al. (1993) demonstrate
experimentally that thermal cues affect fish distributions within mangrove ponds.
Cyprinodon prefers higher temperatures than other fish species tolerate. Temperature
differences within the mangal could, therefore, spatially separate fish populations. Fish
may also prefer certain areas of the mangal based on the nature of the substrate. Kimani et
al. (1996) sampled fish populations in an estuarine mangrove bay in Kenya for 12 months.
Diversity, density, and biomass were all lower in a silty area than in three other mangrove
areas with established seagrass beds.
     Hypoxia, which affects plasma osmolality, plasma chloride ion concentration, and
hematocrit in fish (Peterson, 1990; Peterson and Gilmore, 1991) can also influence their
distributions. For example, juvenile snook (Centropomus undecimalis) move toward
oxygenated surface waters when deeper waters become anoxic (Peterson and Gilmore,
1991). Habitat degradation can lead to changes in fish distribution. In the Virgin Islands,
differences in fish abundance and diversity between degraded and natural mangrove sites
are directly related to water quality (Boulon, 1992). The simple installation of culverts
creates better water exchange in some mangals, promoting reestablishment of marsh
vegetation, and increased fish production (Lin and Beal, 1995).
     Mudskippers are a group of unusual amphibious fish (Family Gobiidae: Subfamily
Oxudercinae) that are characteristic residents of many mangals. A variety of anatomical,
physiological, and behavioral adaptations help them tolerate the stressful mangrove
environment (Chew and Ip, 1990; Colombini et al., 1995, 1996; Ikebe and Oishi, 1996,
1997; Ip et al., 1991; Ishimatsu et al., 1999; Ogasawara et al., 1991; reviewed by Clayton,
1993).
     Researchers have long believed that one of the primary adaptations of mudskippers
is an ability to endure extremely hypoxic conditions. However, recent evidence suggests
that this may not be true for all species; many mudskippers are less tolerant of hypoxia
than has been assumed. Takeda et al. (1999) demonstrated that Periophthalmodon
schlosseri could recover from post-exercise oxygen debt, but only in air. Furthermore,
laboratory experiments demonstrate that P. schlosseri may rarely use aquatic gill
ventilation at all, apparently prefering aerial respiration (Aguilar et al.., in press). A limited
capacity for aquatic oxygen uptake has also been proposed for mudskippers in the genus
Periophthalmus. Aguilar (2000) suggests that Periophthalmus spp. lack physiological
mechanisms to tolerate hypoxia and instead have a range of adaptive behaviors that allow
them to avoid low-oxygen conditions.
     A unique behavioral mechanism may enable some species to completely escape
hypoxia. A number of mudskippers dig extensive burrows into anoxic mangrove
sediments and might be expected to experience extremely hypoxic conditions, particularly
during periods of low tide. Ishimatsu et al. (1998) found that Periophthalmodon schlosseri,
for example, creates burrows as deep as 125 cm. However, the burrows have special
excavated chambers that hold pockets of air. The P. schlosseri fill the pockets by gulping
air at the surface, carrying it down to the chamber and releasing it there. The fish,
therefore, escape rather than tolerate hypoxia. Such burrowing and air trapping,
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       56


particularly where mudskipper populations are dense, may significantly affect oxygenation
and chemistry of the mangrove sediments (Kristensen et al., 1995).
5.9. Amphibians and Reptiles
    Reptiles, including crocodiles, alligators, lizards, snakes and turtles live in many
mangroves. About 35 reptile species are known from the Sunderbans of Bangladesh alone.
The most notable ones are saltwater crocodiles (Crocodylus porosus), monitor lizards
(Varanus bengalensis, V. salvator and V. flavascens), King cobras (Ophiophagus hannah),
Green Pit vipers (Vipera trimeresurus), Rock pythons (Python molorus), and the olive
Ridley turtles (Lepidochelys olivacea, Hussain and Acharya, 1994). In Liberia, Nile
crocodiles occur only in the brackish water of mangrove swamps and at river mouths
(Kofron, 1992). Indo-Pacific crocodiles Crocodylus porosus are abundant in the upper
mangrove sections of the Klias River, Sabah, Malaysia (Stuebing et al., 1994).
    The amphibian fauna of the Sunderbans mangal includes four genera of frogs
(Rana, Bufo, Microhyla and Rhacophorus). The ground frog Rana hexadactyla, the tree
frog Rhacophorus maculatus, and the toad Bufo melanostictus are particularly common
(Gopal and Krishnamurthy, 1993; Hussain and Acharya, 1994). The amphibian fauna has
not been well studied in most other mangals.
    Human activities that impact mangroves have cascading effects on the reptile and
amphibian fauna. There have been drastic declines in the population of crocodiles in the
mangals of Liberia (Kofron, 1992) and of both crocodiles and snakes in the mangroves of
Bangladesh (Hussain and Acharya, 1994). Habitat loss through human encroachment is a
primary cause of the decline. These impacts are likely to continue, and worsen, as human
populations expand further into the mangals.

5.10. Birds
    Mangroves provide important habitat for landbirds, shorebirds and waterfowl, and
they are home to a number of threatened species including spoonbills (Ajala ajala), large
snowy egrets (Cosmorodium albus), scarlet ibis (Eudocimus ruber), fish hawks (Pandion
haliaetus), royal terns (Sterna hirundo), West Indian whistling-ducks (Dendrocygna
arborea), and Storm's Storks (Danielsen et al., 1997; Panitz, 1997; Staus, 1998). The birds
in the mangal may be permanent residents that forage and nest in the mangroves and the
mangrove waters or they can be temporary visitors. Lefebvre et al. (1994) measured bird
abundances and grouped species according to their diet and the frequency with which they
use mangrove habitats. Distributions and abundances of the feeding guild were consistent
with the abundance and distribution of their invertebrate prey (Lefebvre and Poulin, 1997).
    Lefebvre et al. (1992) studied distributions of passerine birds in the mangroves of
Venezuela and found that they form highly stable territories. In Singapore, sand pipers,
plovers, herons and egrets all regularly use mangrove habitat (Murphy and Sigurdsson,
1990). About 315 species of birds are known from the Sunderbans of Bangladesh. The
most common ones are white-bellied sea eagles (Haliaetus leucogaster) and Pallas’s fish
eagles (Haliaetus leucorhyphus; Hussain and Acharya, 1994). Alves et al. (1997) counted
32 bird species (2 marine species, 18 terrestrial species, and 12 waterfowl) in the
mangroves of Jequiaman, Brazil.
    Migratory birds visiting the mangroves may fly long distances to find food and
nesting places there. This may be particularly true in the Neotropics (Parrish and Sherry,
1994; Confer and Holmes, 1995; Lefebvre and Poulin, 1996; Panitz, 1997). Seventy-seven
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                                             57


bird species have been recorded in the Pacific mangroves of Colombia. Forty-three percent
of these are permanent residents, 22% are regular visitors and 18% are temporary winter
residents (Naranjo, 1997). One migratory species, the black-crowned night heron
(Nycticorax nycticorax) is an important vector for disease. This mangrove-breeding bird is
the principal host for Japanese Encephalitis Virus, which it widely disseminates during its
migrations (Murphy and Sigurdsson, 1990).
    Some of the resident bird species are highly dependent on mangroves for their
survival. The yellow warbler (Dendroica petechia) and the mangrove vireo (Vireo pallens)
are nearly confined to mangroves (Parkes, 1990; Buden, 1992). The mangrove gerygone
spends 80% of its time on Avicennia marina (Noske, 1996) while A. germinans provides
important breeding habitat for Florida Prairie Warblers (Dendroica discolor paludicola)
and Cuban Yellow Warblers (D. petechia gundlachi; Prather and Cruz,1995).
    Because of this dependence, disturbances to the mangal may reverberate through
the bird populations. This may be particularly true where the bird species show strong site
fidelity (Warkentin and Hernandez, 1996). The habitat disturbances may be natural, such
as the frequent cyclonic storms that strongly affect myna populations in the Pichavaram
mangroves of south India (Nagarajan and Thiyagesan, 1995). More frequently, however,
they are caused by human activities.
    Mangrove forest destruction and fragmentation, usually due to development, reduce
effective habitat and threaten mangrove-dependent species. Bancroft et al. (1995) found
reduced populations of mangrove cuckoos
(Cocyzus minor) in fragmented mangrove
areas of Florida, USA (Figure 13).                    Original      1991

                          100
Similarly, the mangrove finch (Cactospiza                                        Northern Flicker
heliobates), once present in six mangrove      80
                                     1 km

areas on two of the Galapagos Islands, are
                            Occurrence (%)




                                                  Mangrove Cuckoo
                           60
now restricted to four mangrove pockets on                White-eyed Vireo

one island. Habitat destruction has                                         Yellow-billed Cuckoo
                           40


completely eliminated the finches from the      20

other island (Grant and Grant, 1997).
                           0
Ironically, one potential threat to these        0-1  2-5    5 - 10         10 - 20 > 20

populations is the birdwatchers who explore                  Forest size (ha)

hoping to see the birds in their natural       Figure 13. Abundance of mangrove-associated
environment (Klein et al., 1995; Ellison and     land birds as a function of forest size. Inset
                           shows changes in the mangrove forests of Upper
Farnsworth, 1996a).
                           Matecumbe Key (Florida Keys, Florida, USA)
    Protection of the mangrove-dependent
                           from an original 252.2 ha forest to 105 forest
birds will require effective management of the    fragments totalling only 61.4 ha. Records of 4
entire mangrove habitat. This may be complex     bird species show that abundance is strongly
and require evaluation of habitat needs on a     affected by forest size (after Bancroft et al, 1995).
species by species basis. For example, in
Florida Bay, bald eagles (the U.S. national bird) nest almost exclusively in mangrove trees
(Avicennia germinans and Rhizophora mangle). Many of the nest sites are in snags (dead
trees) suggesting that a comprehensive eagle management plan will require preservation of
both living and dead mangroves (Curnutt and Robertson, 1994).

5.11. Mammals
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                          58


    A variety of mammals make their homes in the mangal. Their ecology within the
mangal and their associations with the mangroves themselves have been little studied and
are poorly known. Some of the noteworthy species present include dolphins (Platenista
gangetica), mangrove monkeys (Macaca mulatta) and otters (Lutra perspicillata ) in India
(Gopal and Krishnamurthy, 1993); flying fox (Pteropus conspicillatus and Pteropus
alecto) in northern Australia (Richards, 1990; Loughland, 1998) capuchin (Cebus apella
apella) in Brazil (Fernandes, 1991). In southeastern Brazil, distributions of some cetacean
species can also be related to the distribution of mangroves (Martuscelli et al., 1996).
Small clawed otters (Lutrinae) shelter amongst Acrostichum ferns during the dry season in
the mangroves of Singapore and Malay (Sivasothi and Burhanuddin, 1994).
    Thirty-two mammal species once lived in the Sunderban mangals of Bangladesh.
Four of these, Javan rhinoceros (Rhinoceros sondaicus), wild buffalo (Bubalus bubalis),
swamp deer (Cervus duvauceli) and hog deer (Axis porcinus), have gone extinct since the
beginning of this century (Hussain and Acharya, 1994). Two additional species, the Royal
Bengal Tiger (Panthera tigris) and the Chital deer (Axis axis), are currently endangered.
Studies have shown that the chital deer browses directly on the mangroves. Its feeding
damages Avicennia officinalis, Xylocarpus mekongensis, Bruguiera sexangula, and
Aegiceras corniculatum but has no effect on Heritiera fomes, Excoecaria agallocha or
Ceriops decandra (Siddiqi and Hussain, 1994).
    Loss of mammalian species in the world's mangrove environments is probably the
result of habitat fragmentation. This is particularly true for some of the larger species that
have large home ranges. This may largely explain the loss of species from the Sunderbans.
Habitat loss, however, can also have a major effect on smaller species. Forys and
Humphrey (1996) studied distribution and movement of an endangered marsh rabbit
(Sylvilagus palustrishefneri) in the Lower Florida Keys, USA. They found that the rabbits
use mangrove tracts as dispersal corridors between marsh habitats. Preservation of this
species will require protection of both the marsh and the mangrove corridors. Careful study
will be required to implement effective conservation plans for the mammals faced with
shrinking, and fragmenting, mangrove habitat.


6. RESPONSES OF MANGROVES AND MANGROVE ECOSYSTEMS TO
  ENVIRONMENTAL STRESS

6.1. Responses to light
    Although mangroves occur in tropical habitats where they are exposed to high light
intensities, their photosynthetic rates tend to plateau at relatively low light levels. The trees
possess mechanisms to deal with the high sunlight (see section 3.3.2). For example,
Avicennia marina shows good resistance to high sunlight, hot and dry conditions and is
well adapted to arid zones (ElAmry, 1998). However, there is evidence that intense light
can damage the mangroves despite these adaptations. For example, Cheeseman et al.
(1991) demonstrated that rates of photosynthesis drop in mangroves exposed to excessive
sunlight. This could explain why Rhizophora seedlings establish and sprout most readily
under the shady canopy of larger trees (Kathiresan and Ramesh, 1991; Kathiresan, 1999).
Kathiresan and Moorthy (1993) also demonstrated that the seedlings grow faster in the
shade, use NO3 more efficiently, and show more efficient photosynthesis.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        59


    The negative effects of sunlight may, in some cases, be due to the high doses of
UV-B radiation that the mangroves receive (Moorthy, 1995). To date, however, there have
been few studies of UV-B and mangroves (Lovelock et al., 1992; S.M. Smith and
Snedaker, 1995b). Moorthy (1995) and Moorthy and Kathiresan (1997a) studied mangrove
responses to UV-B in the Pichavaram mangroves of south India (total sunlight and UV-B
intensities here may exceed 1300 W m-2 and 0.31 W m-2 respectively). Species in the
Rhizophoraceae showed greater UV-B tolerance than did Avicennia species or other
succulent plants.
    To better understand potential effects of global increases in UV-B, Moorthy and
Kathiresan (1997b) grew Rhizophora apiculata seedlings under the UV-B regimes
predicted for 10, 20, 30 and 40% stratospheric ozone depletions. Net photosynthetic rates
of seedlings increased 45% under the 10% UV-B treatment (stomatal conductance
increased 47%). Raising the UV intensity to the 40% level, however, produced a 59%
decrease in net photosynthesis and a 73% increase in intercellular CO2 concentrations
(Moorthy, 1995).
    Increasing UV-B exposure also produced biochemical changes. Phenol and
flavonoids levels increased with UV-B dose, but anthocyanin concentrations dropped.
Small UV-B doses enhanced amino acid and protein levels but the effect was reversed at
higher UV-B levels. The UV-B, in general, enhanced saturated fatty acids (maximum
increase of 88%) and reduced unsaturated fatty acids (maximum decrease of 26%). Any
UV-B exposure also inhibited the activity of nitrate reductase while simultaneously
enhancing total tissue nitrate (Moorthy and Kathiresan, 1998). Both growth and
biochemical responses indicate that the mangroves are stressed by the high intensity UV-B.
    While excessive sunlight can damage mangroves, shading can also have negative
effects. Mangrove seedlings under a closed canopy showed lower growth in south Florida
(Koch, 1997) and Belizean mangals (Ellison and Farnsworth, 1993). The appearance of
gaps in the canopy produced rapid growth of the previously shaded trees. In dense
mangrove forests, shaded saplings have lower shoot biomass than those exposed to the sun.
The saplings may compensate for this by increasing development of the pneumatophores
(Turner et al., 1995).
    Shade tolerance differs among mangrove species. Clarke and Allaway (1993) found
that shading had no effect on growth or survival of Avicennia marina. But, McKee (1995b)
noted that shading with brief periods of light exposure increased biomass and growth in
Avicennia germinans and Laguncularia racemosa. The same treatment had little effect on
Rhizophora mangle. Ellison and Farnsworth (1993) found that R. mangle seedlings
performed better overall in larger canopy gaps. In R. mangle, ontogenetic developments
produce changes in light adaptation. Seedlings are apparently adapted for the shaded
understory environment while mature trees do better in the sunlit canopy (Farnsworth and
Ellison, 1996b).

6.2. Responses to gases
    Because of the environment they live in, mangroves may experience episodic, or
chronic oxygen stress. The consequences of low oxygen vary among species and appear
related to the physiological and morphological adaptations of each (McKee et al., 1996).
For example, McKee (1993) found that flooding and anoxia reduced the total biomass of
Avicennia germinans seedlings 20-40% relative to drained controls. However, under
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                                60


identical conditions, the biomass of Rhizophora mangle seedlings increased 9-24%. The
differential tolerance of these species for flooding and low oxygen may partly result from
differences in root aeration. R. mangle maintains high oxygen concentrations in its roots
even under reducing soil conditions; A. germinans does not. Skelton and Allaway (1996)
showed that a congener (A. marina) also does not maintain gas pressures under low-
oxygen conditions. Pressures in the aerial roots drop during high tide, probably due to
removal of respiratory CO2 from gas spaces during flooding. As the waters recede on the
low tide, a rapid influx of air may take place.
      High methane levels can be associated with anoxia in mangrove environments.
The methane flux from the sediments is strongly influenced by freshwater loading and
nutrient input (Sotomayor et al., 1994; Giani et al., 1996). Fluxes may also vary along tidal
gradients being generally low on the landward fringe and high in the seaward transition
zone between Avicennia and Rhizophora communities (e.g., Ye et al., 1997). Mangrove
species with pneumatophores may be best equipped to deal with high methane loads. The
pneumatophores themselves may be conduits for release of methane gas (Sotomayor et al.,
1994). Pneumatophore-bearing species also release more methane through their leaves
than do those lacking pneumatophores (Lu et al., 1998).
      Predicted global changes in atmospheric carbon dioxide are likely to have strong
                                 effects on mangroves. Farnsworth et al.
                                 (1996) grew Rhizophora mangle under
   25                400
                Ambient CO2
                                 double ambient CO2 for 1 year. Growth
                Elevated CO2
   20
                              Total stem length (cm)
      Number of branches




                                 rate, net assimilation, and photosynthetic
                    300

   15
                                 rate all increased significantly. Seedlings in
                    200
                                 the enhanced CO2 treatment had greater
   10

                                 biomass, longer stems, more branching, and
                    100
   5
                                 more leaf area than control seedlings
   0                  0
                                 (Figure 14). They also became reproductive
    0 100 200 300 400         0  100 200 300 400

                                 after only 1 year (2 years sooner than under
  180
                   2800
                                 normal conditions). Ellison (1994) found
Total plant biomass (g)




                        Total leaf area (cm2)




                                 that, in addition to stimulating productivity,
  140
                   2200

                                 increased CO2 led to more efficient use of
  100
                   1600
                                 water as a result of reduced stomatal
                                 conductance (Ellison, 1994).
   60                1000

                                     The effects of increased CO2 may
   20                400
                                 vary with other physical and chemical
    280 320  360 400         280  320  360 400

                                  conditions (Ball and Munns, 1992). For
           Days elapsed since planting
                                  instance, Rhizophora apiculata and R.
Figure 14. Effects of increased CO2 on Rhizophora
mangle growth and morphology. Predicted changes          stylosa both benefit from increased CO2,
in the global climate could result in significant         but the stimulatory effect is much greater
increases in CO2 levels. Experimentally doubling the
                                  under low salinity conditions (Ball et al.,
ambient CO2 levels produced significant increases in
                                  1997). The effects of increased CO2 may
number of branches, total stem length, plant biomass,
                                  also differ among habitats and species.
and total leaf area (after Farnsworth et al, 1996).
                                  Snedaker and Araújo (1998), for example,
studied the effects of a 6 - 34 % increase in CO2 concentration on four mangrove species in
south Florida (Rhizophora mangle, Avicennia germinans, Laguncularia racemosa, and
Conocarpus erectus). Elevated CO2 reduced stomatal conductance and transpiration and
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        61


significantly increased instantaneous transpiration efficiency in all of these species.
However, it did not increase net primary productivity in any species and actually reduced
the productivity of Laguncularia racemosa.

6.3. Responses to wind
     Tropical storms (hurricanes and cyclones) cause enormous damage to mangrove
forests, particularly in the Caribbean and the Bay of Bengal. The immediate consequence
is loss of the mangrove trees themselves. In 1988, for example, a severe cyclone in
Bangladesh destroyed 8.5 million trees with a loss of 66.3 million m3 of commercial saw
timber (e.g. Mastaller, 1996). Large numbers of trees have similarly been defoliated and
destroyed by storms in South Florida, Guadeloupe, Nicaragua, Belize, and Puerto Rico.
     Damage to the mangroves may also have indirect consequences. For example,
Hurricanes Gilbert and Joan, which both hit the Caribbean in the last quarter of 1988,
caused mass mortality of invertebrates growing on the roots of Rhizophora mangle
(Orihuela et al., 1991). Hurricane Hugo, while passing through the coastal environment of
Guadeloupe, French West Indies, killed large numbers of fish, producing changes in the
fish community (Bouchon et al., 1991, Imbert et al., 1996).
     A far-reaching consequence of mangrove mortality can be serious erosion of the
coastal habitat. Hurricane Andrew made landfall in 1992 on the mangrove-fringed coasts
of south Florida. Uprooted mangrove vegetation left behind unprotected intertidal and
subtidal sediments that were subsequently eroded by currents and waves. (Davis, 1995;
Doyle et al., 1995; Swiadek, 1997). In Bangladesh, a mangrove afforestation project was
initiated in 1966. A primary goal of the project was to provide a mangrove buffer that
would protect the coast from frequent cyclone damage. By 1993, 0.12 million ha had been
afforested (Saenger and Siddiqi, 1993).
     Some effects of storm damage may not be seen until some time after the event.
Often, trees that are broken or severely damaged by the storm later die. T.J.Smith et al.
(1994) found that mangrove mortality in south Florida continued for many weeks after
hurricane Andrew. Many mature trees later died from injuries sustained in the storm.
Propagules and seedlings in the habitat were also killed, largely by sedimentation and high
porewater sulfides.
     Differential species mortality associated with a major storm can change community
structure. For example, smaller Rhizophora mangle were not greatly affected by Hurricane
Andrew whereas large Laguncularia racemosa were heavily damaged (McCoy et al.,
1996). These differences in survival have produced a shift in species distributions in some
areas (Baldwin et al., 1995). A similar shift can be seen with Avicennia germinans and
Rhizophora mangle. Because A. germinans, cannot tolerate long periods of pneumatophore
submergence, storms and hurricanes in Florida may promote replacement of A. germinans
by R. mangle (Rey et al., 1990a). Even where the regenerated forest is composed of the
same species present before the hurricane, different regeneration rates may shift the
proportions of those species (Roth, 1992, 1997).

6.4. Responses to coastal changes
    Mangroves are tightly bound to the coastal environments in which they occur. Not
only are they influenced by chemical and physical conditions in their environment, they
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       62


help create those conditions. As a result, perturbations to the system can have cascading
long-term effects.
     Many changes seen in coastal mangals can be attributed to changes in hydrology.
Some of these changes are favorable. For example, in Singapore, increases in the
frequency of tidal inundation has created new mangal (comprised of Avicennia and
Sonneratia alba) adjacent to established mangrove stands (Lee et al., 1996). More often,
however, hydrological changes result in destruction of mangroves. In Pichavaram, south
India, changes in topography and tidal flushing have caused large-scale degradation of
mangroves. The mangroves are healthy and diverse where the land is flat. If water flow is
reduced, flat areas become shallow basins. The poor flushing and resultant hypersalinity
stunt the mangroves or replace them with saltmarsh (Suaeda spp.) or barren soil devoid of
vegetation. The process can be reversed by simply increasing the free flow of tidal waters
(Selvam and Ravichandran, 1998).
     In Senegal, decreasing rainfall and increasing evaporation have markedly changed
mangrove populations. The changes have been accelerated by altered tidal conditions
resulting from the breaching of a protective sand dune (Diop et al., 1997). Riverine
mangroves affect the dynamics of tidal currents (Wolanski et al., 1992), producing
asymmetrical tidal currents that may be 50% stronger on the ebb than on the flood. Erosion
from these flows creates deep channels through the habitat (Medeiros and Kjerfve, 1993).
Deforestation changes the tidal asymmetry and leads to changes in channel structure
(Wolanski et al., 1992).
     Human attempts to modify the physical character of the mangal by erecting hard
structures or by dredging can also drastically alter the system. In Florida, a culvert
connecting a mangrove marsh to a tidal lagoon was closed in 1979. This led to
overflooding and hypersalinity ( > 100 ppt) that eliminated the marsh. The culvert was
reopened in 1982, but the mangroves did not recover (Rey et al., 1990a, b).
     Another potential consequence of flow modification is change in sedimentation
patterns (e.g., Q. Zhang et al., 1996). On spring flood tides under normal conditions, about
80% of the sediment transported into the Middle Creek (Cairns, Australia) is trapped by
the mangroves. This corresponds to 10-12 kg of sediment m-1 on each spring tide and
could produce sediment accretions of 0.1 cm y-1 (Furukawa et al., 1997). Such levels of
sediment trapping can produce major changes to the habitat. Chakraborti (1995) traced the
evolutionary history of coastal quaternary deposits on the Bengal Plain of India and found
that mangroves were the dominant geomorphic agents in the evolution of tidal shoals and
their eventual accretion to the mainland.
     Damage to the mangroves strongly affects sediment budgets and promotes coastal
erosion (Kamaludin and Woodroffe, 1993). The eroded sediments may then cause further
damage the mangroves. For example, Young and Harvey (1996) showed that sediment
accretion interferes with root aeration in Avicennia marina var. australasica. Similarly,
movements of sand in mangrove habitats on Portuguese Island, Mozambique have caused
high mortality of Ceriops tagal. This has changed the mangrove species composition and
had significant secondary effects; all crustaceans and mollusks have also disappeared from
the mangal (Hatton and Couto, 1992).
     Major changes in mangrove distribution and abundance in coastal regions could
result from habitat loss associated with rising sea level (Fujimoto and Miyagi, 1990;
Woodroffe, 1990; Ellison, 1993; Parkinson et al., 1994). The vulnerability of individual
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                            63


mangals can be evaluated through annual measurements of soil elevation deficit (elevation
change minus sea-level rise). Cahoon and Lynch (1997) suggest that this is a better
measure than the more commonly used accretion deficit (accretion minus sea-level rise).
Historical patterns of sea level rise in a mangrove environment can be evaluated through
measurements of the mangrove trees themselves since the δ18 oxygen fraction in the wood
is an effective seawater tracer (Ish-Shalom-Gordon et al., 1992).
     Ellison and Stoddart (1991) suggested that mangroves are stressed by sea level rises
between 9 and 12 cm • 100 y-1 and concluded that faster rates could seriously threaten
mangrove ecosystems. This view has been challenged by Snedaker et al. (1994) who cite
historical records showing mangrove expansion under relative sea level changes nearly
twice that great. X. Tan and Zhang (1997) conclude that, given current rate estimates, sea
level rises pose no significant threat to most mangrove forests in China. The effects of sea-
level rise on any mangrove habitat will be influenced by local wetland type, geomorphic
setting, and human activities in the wetland. There is a need for better models predicting
effects of sea level and other coastal changes on individual mangals (e.g., Bacon,1994).


6.5. Responses to tidal gradients and zonation

     Zonation can be a structural feature of
mangrove forests in some parts of the world
(T.J. Smith, 1992; Woodroffe, 1992).
However, unlike open coast habitats where
zonation patterns are distinct, mangrove
distributions are extremely variable and
extensive surveys may be necessary to fully
document patterns, particularly if diversity is
high (Bunt 1996). The “zones” may be
obscured by broad overlap in species
distributions (e.g., Bunt and Bunt, 1999; Bunt
and Stieglitz, 1999; Figure 15) or they may
simply be absent in some mangals (Ellison et
al., in press). Bunt (1999) has developed
methods specifically for evaluating and
describing mangrove zonation.
     Where zonation does occur,
contributing factors may include plant       Figure 15. Distribution of mangroves south of Townsville,
succession, geomorphology, physiological      North Queensland, Australia. Stippled areas of the map
adaptation, propagule size, seed predation     represent mangal. Details of transect A-B are shown in the
and interspecific interactions (e.g., Bunt et   lower panel. The dominant mangrove species (i.e.,
                          Avicennia marina, Rhizophora stylosa, Ceriops tagal,
al., 1991; Woodroffe, 1992; Ellison and
                          Bruguiera gymnorhiza, and Osbornia octodonta) are
Farnsworth, 1993; Schwamborn and Saint-      represented by different symbols. Note that the mangroves
Paul, 1996). The relative importance of      are closely associated with rivers and creeks, resulting in
these factors, however, depends on the       broadly overlapping species distributions (after Macnae,
individual habitat (e.g., McKee, 1995c) and    1967).
there is disagreement about the general
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        64


importance of some of them. For example, Robertson et al. (1991) argue that succession
plays a minor role in mangrove zonation and that simple erosion and sedimentation control
the distribution of mangroves along the seaward edge of the mangal. The complexity and
uniqueness of these communities may make it difficult even to define successional stages
(Roth, 1992). The term “old-growth”, for instance, cannot be applied easily to mangrove
forests (Lugo, 1997).
    One potential cause of mangrove zonation is the differential ability of propagules to
establish at different tidal heights. This is directly related to propagule size. It has been
suggested that small propagules drift further inland and establish better in shallow water
than do large propagules (thus producing a species zonation dependent on propagule size;
Kathiresan, 1999). The importance of this process to the creation of mangrove zones has
been clearly demonstrated for Avicennia bicolor and Rhizophora racemosa on the Pacific
coast of Costa Rica (Jiménez and Sauter, 1991). However, the more general importance of
this process has been contested (T.J.Smith, 1992).
    Interspecific differences in tolerance for physiological stress is perhaps the best
demonstrated cause of mangrove zonation. However, while physiological responses to
physicochemical conditions undoubtedly influence mangrove distributions in some
habitats, conclusions must be made cautiously since field measurements do not always
support laboratory conclusions (Schwamborn and Saint-Paul, 1996). Despite this
limitation, it is clear that mangrove species respond differently to different tidal regimes.
For example, in the Indian Sunderbans, a mangrove forest that experiences total diurnal
inundation is dominated by Avicennia marina and A. alba while Excoecaria agallocha,
Ceriops decandra and Acanthus ilicifolius dominate sites that are not completely inundated
(Saha and Choudhury, 1995). Nypa fruticans also seems to prefer sites with low level of
tidal inundation (Siddiqi, 1995). Kathiresan et al. (1996a) studied growth of Rhizophora
apiculata seedlings in different tidal zones of a south Indian estuary. Individuals in the
low intertidal grew 2.5 times faster and sprouted 4 times as many leaves as individuals in
the highest zone. Similar patterns of differential survival and growth have been seen in the
mangroves of Qatar (Abdel-Razik, 1991) and New Zealand (Osunkoya and Creese, 1997).
    Ellison and Farnsworth (1993) studied survival and growth of Rhizophora mangle
and Avicennia germinans seedlings at tidal heights corresponding to lowest low water
(LLW), mean water (MW), and highest high water (HHW). R. mangle seedlings survived
in the MW (69%) and LLW (56%) treatments; Avicennia survived at MW (47%), but not
at LLW. Neither species survived at HHW. Seedlings of both species suffered twice as
much insect damage in the MW treatment as in the LLW treatment. The combination of
insect herbivory and differential flood tolerance may create the zonation of these two
species in the Caribbean. Further experimental studies of this nature should help clarify the
causes and consequences of zonation in mangrove communities.

6.6. Responses to soil conditions
    Soil properties have a major impact on mangrove nutrition and growth. Some of the
most important characteristics are siltiness, electrical conductivity, pH, and cation
exchange capacity (Kusmana, 1990; V.B. Rao et al., 1992; Pezeshki et al., 1997). The
most important factor, however, appears to be nutrient concentrations. Mangals are finely
balanced, highly effective nutrient sinks with net imports of dissolved nitrogen,
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       65


phosphorus, and silicon. Nutrient fluxes in these environments are closely tied to plant
assimilation and microbial mineralization (Alongi, 1996; Middelburg et al., 1996).
    Nutrients availability may limit growth and production in many mangals. Varying
nutrient concentrations can also change competitive balances and affect species
distributions (Chen and Twilley, 1998). As a result, nutrient pulses can create immediate,
and impressive, changes in the vegetation. For example, on the southeast coast of India,
high nutrient concentrations and low salinity from monsoons produce rapid growth in the
mangroves. Seedlings grow 5X as much and produce 4X as many leaves in the post-
monsoon season as they do in the dry season (Kathiresan et al., 1996a).
    The limiting nutrients may vary with individual mangrove habitats. For example,
potassium levels may be important in some regions. Rhizophora apiculata seedlings do
significantly better in plantation sites with enriched potassium (Kathiresan et al., 1994a).
In general, however, mangroves in low-nutrient carbonate soils are limited by phosphorus.
What phosphorus is present may be bound with calcium, effectively holding it within the
sediments (Silva and Mozeto, 1997). In mesocosm and field experiments with Rhizophora
mangle seedlings, phosphorus enrichment produced nearly a 7 fold increase in stem
elongation rates and a 3 fold increase in leaf area. Nitrogen addition produced no such
response (Koch and Snedaker, 1997). Low phosphorus availability similarly limits growth
of dwarf R. mangle in a Belizean mangal (Feller, 1995) and promotes development of hard,
long-lived leaves called sclerophylls. The sclerophylls may be an adaptation to conserve
nutrients in these oligotrophic habitats (Feller, 1996). Mangroves may have other
mechanisms to hold valuable nutrients. For example, mature photosynthetically active
leaves have much higher nitrogen and potassium concentrations than senescent leaves.
This is apparently a consequence of nutrients being translocated out of the aging leaves and
into other plant parts before the leaves fall (Soto, 1992).
    Damage to a mangal may compromise its ability to retain nutrients. For examples,
at severely damaged mangrove sites in North Queensland, Kaly et al. (1997) measured a
significant loss of both nitrogen and phosphorus from the soils. This may have been related
to declines in the crab populations and a dramatic decrease in density of burrows. The
effects of disturbance will differ from habitat to habitat and will depend on the sediment
characteristics and flow regimes of each site. For example, Triwilaida and Intari (1990)
found no differences in soil nutrient concentrations between degraded and healthy
mangrove stands in the Pedada Strait, Riau.
    Sulfides are a characteristic feature of mangrove sediments that influences
mangrove distributions and is influenced by their presence. Tidal mixing, bioturbation, and
the mangroves themselves (Holmer et al., 1994) control the distributions and
concentrations of the sulfides. For example, reoxidation of sulfides is facilitated by roots;
soils are often less reduced near the aerial roots of some species. This leads to lower
sulfide levels. In a neotropical Florida mangal, the zone dominated by Rhizophora mangle
(with its numerous aerial prop roots) has moderately reducing soils with low sulfide levels.
In contrast, the Avicennia germinans zone has strongly reducing soils with high sulfides
(McKee, 1993). Surprisingly, this pattern is not repeated in a similar mangal in Brazil. In
that location, the R. mangle soils are highly reducing with high sulfide concentrations. The
sulfide content of the A. germinans soil is highly variable as the rhizosphere changes from
oxygenated to anoxic conditions (Lacerda et al., 1995). The Avicennia soils also contain
more exchangeable trace metals (Lacerda et al., 1993).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        66


    Reduction of sulfate to sulfide is generally slower in young forests, resulting in
higher nutrient levels and lower sulfide toxicity (Alongi et al., 1998). High sulfide levels
can damage mangrove seedlings, causing stomatal closure, reduced gas exchange, reduced
growth, and high mortality (Youssef and Saenger, 1998). Disturbance can increase rates of
sulfate reduction. Clearing of mangrove forests, or simple formation of canopy gaps can
change the physical and chemical characteristics of the underlying soils (Ewel et al.,
1998a), leading to anaerobiosis and increased sulfide in the sediments (Ibrahim, 1990).
Heavy organic input can also increase sulfide production. In Ghana, pyrite (ferric sulfide)
accumulates directly in the upper layer of the mangrove soils. The rate of accumulation is
directly related to vegetation thickness (Nonaka et al., 1994).
    Under normal conditions, sulfides combine with metals in the sediment and
precipitate out as metal sulfides. When the metals available for sulfide precipitation are
exhausted, H2S is formed (Kryger and Lee, 1995). Kryger and Lee (1996) found that the
H2S from anaerobic processes accumulates in cable roots of Avicennia species as the
sediments age. Concentrations of H2S in the roots may be 30-40 times higher than in the
surrounding sediments. The H2S accumulation can kill the mangroves if their
pneumatophores are covered by silt and cannot transport oxygen to the rhizospheres.
Because they have aerial roots, Rhizophora species can better survive on aged mangrove
soils high in H2S. They may, therefore, be a natural successor to the less-tolerant Avicennia
species.


6.7. Responses to salinity
    Salinity, as controlled by climate, hydrology, topography and tidal flooding, affects
the productivity and growth of mangrove forests (Sylla et al., 1996; Twilley and Chen,
1998). It can also strongly influence competitive interactions among species (Ukpong,
1995; Ukpong and Areola, 1995; Cardona and Botero, 1998). The distributions of plant
species within the mangal, in many cases, can be explained primarily by salinity gradients
(Ukpong, 1994; Ball, 1998).
    In general, mangrove vegetation is more luxuriant in lower salinities (Kathiresan et
al., 1996a). However, low salinity associated with long periods of flooding contributes to
mangrove degradation through reduced cell turgor and decreased respiration (Triwilaida
and Intari, 1990). On the Pacific coast of Central America, freshwater availability (largely
from rainfall and surface runoff) controls reproductive phenology, growth and mortality of
Avicennia biocolor (Jiménez, 1990).
    Even in mangals with strong riverine input, the combined effects of evaporation
and transpiration may remove much of the fresh water entering the system (Simpson et al.,
1997). The plants must, therefore, have some salinity tolerance. True mangroves (e.g.,
Avicennia spp. and Rhizophora spp.) tolerate higher salinity than do non-mangroves, but
tolerance also varies among the true mangroves. For example, Rhizophora mucronata
seedlings do better in salinities of 30 g l-1, but R. apiculata do better at 15 g l-1
(Kathiresan and Thangam, 1990a; Kathiresan et al., 1996b). Sonneratia alba grows in
waters between 5 and 50 % seawater, but S. lanceolata only tolerates salinities up to 5%
seawater (Ball and Pidsley, 1995). Mangrove seedlings require low salinity (S.M. Smith et
al., 1996), but their salt tolerance increases as they grow (Bhosale, 1994).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       67


     Short periods of high salinity may trigger events in the mangrove life history. For
instance, high salinity at the end of dry period, followed by an extended rainy period
controls establishment of Rhizophora seedlings (Rico-Gray and Palacios-Rios, 1996b).
Chronic high salinity, however, is always detrimental to the mangroves. Hypersalinity
stunts tree growth in A. marina stands (Selvam et al., 1991), reduces biomass in
hydroponically grown Bruguiera gymnorrhiza (Naidoo, 1990), and causes denaturing of
terminal buds in Rhizophora mangle seedlings (Koch and Snedaker, 1997). Saline
interstitial water reduces leaf area, increases leaf sap osmotic pressure, increases the leaf
area/weight ratio and decreases total N, K, and P (Medina et al., 1995). Simple salinity
fluctuations also have significant negative effects on photosynthesis and growth (Lin and
Sternberg, 1993). In Senegal, hypersalinity (from a decade of low rainfall and high
evaporation) has caused salt flats to grow into mangrove areas, completely destroying the
vegetation (Diop et al., 1997).
     Extremely high salt concentrations in the groundwater of tropical salt flats are
responsible for the complete absence of macrophytes (including mangroves). There are
often very sharp changes in groundwater salt concentrations at the interface between salt
flats and mangroves, suggesting that the mangroves are modifying the salinity of the
groundwater (Ridd and Sam, 1996). Mathematical models of groundwater flow show that
human activity hundreds of kilometers inland can destroy vast mangrove areas by changing
groundwater flow and modifying salinity levels (Tack and Polk, 1997).

6.8. Responses to metal pollution
    Because of their proximity to population centers and industrialized regions,
mangrove habitats have often received inputs of heavy metals and the sediments may show
significant metal contamination (Mackey et al., 1992; Larcerda et al., 1993; Rivail et al.,
1996; Lacerda, 1998; Tam and Yao, 1998). The mangroves themselves, however, generally
have low concentrations of heavy metals. Consequently, they are very poor indicators of
trace metal contamination. For example, in Sepetiba Bay, Rio de Janeiro, the sediments
contain 99% of the Mn and Cu and almost 100% of the Fe, Zn, Cr, Pb and Cd in the total
mangrove ecosystem. The tissues of Rhizophora mangle contain less than 1% of the total
of these metals (Silva et al., 1990). On the Saudi coast of the Arabian Gulf, there is no
correlation between the concentrations of metals in sediments and in the leaves of
mangroves living on the contaminated soil (Sadiq and Zaidi, 1994).
    The low level of metals in the mangroves themselves may be due to 1) low
bioavailability in the mangal sediments 2) exclusion of the metals by the mangroves or 3)
physiological adaptations that prevent metal accumulation in the plants. Mangrove roots
appear to be barriers that prevent metals from reaching the more sensitive parts of the plant
(Tam and Wong, 1997). Oxygen exuded by underground roots forms iron plaques that
adhere to the root surfaces and prevent trace metals from entering the root cells. Where the
metals do enter, there are apparent mechanisms to keep them from circulating freely
through the plant. Heavy metal concentrations in Rhizophora apiculata seedlings decrease
from root to stem and from stem to leaves (Moorthy and Kathiresan, 1998a).
    The chemical and physical environment of the mangal may efficiently trap trace
metals in non-bioavailable forms. For example, rapid precipitation of stable metal sulfides
under anoxic conditions decreases the bioavailability of trace metals in the mangrove
sediments (Di-Toro, 1990; Mackey and Mackay, 1996; Figure 16). All but the most mobile
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       68


                            elements (e.g., Mn and Zn) may be
                            in strongly bound fractions. Thus,
                            the mangal may help control trace
                            metal pollution in tropical coastal
                            areas (Lacerda, 1997).
                                Trace metals may also be
                            bound in organic complexes that
                            show low bioavailability (Clark et
                            al., 1998; Lacerda, 1998). For
                            example, Cr, which does not form
                            sulfide minerals, is immobilized in
                            refractory organic compounds in
                            mangrove sediments (Lacerda et
                            al., 1991). Mercury can be similarly
                            bound. However, it may be bound
                            as the highly toxic dimethyl-
                            mercury. Under oxic conditions,
                            dimethyl-mercury is volatile and
                            unstable; in reducing mangrove
                            sediments, however, it may persist
                            and accumulate (Quevauviller et
                            al., 1992).
                                 While mangrove sediments
                             generally have a high capacity for
 Figure 16. Summary of major metabolic processes
 involving metals in anoxic mangrove sediments (after  absorbing and holding trace metals,
                             heavy loads may exceed the
binding capacity of the sediment (Stigliani, 1995). Tam and Wong (1996a, b) irrigated
mangrove soil samples with metal-laden artificial wastewater. They found that the upper
centimeter of the soils bound Cu, Cd, Mn and Zn. However, there were also higher
concentrations of Mn, Zn and Cd in the water-soluble, exchangeable fraction of the treated
sediments than in the untreated, native sediments.
     Disturbances may also cause the mangrove soils to lose their metal-binding
capacity, resulting in mobilization of the metals. The mangal then shift from a heavy metal
sink to a heavy metal source (Lacerda, 1998). Disturbances may be in the form of
prolonged dry periods (Clark et al., 1997), changes in the frequency and duration of tidal
flooding (Chiu and Chou, 1991) or changes in salinity (Spratt and Hodson, 1994). Often,
these disruptions are associated with human activities (Lacerda, 1998).
     S. Zheng et al. (1997) suggest that mangrove afforestation projects should not be
done on Cu or Zn-polluted soils since seedlings secrete organic acids that may increase
solubility of the metals. Rhizophora apiculata seedlings planted in an area formerly used
for tin mining showed high mortality (approximately 47% in the first four years). The
mortality, however, was attributed to altered microtopography and soil particle distribution
rather than metal contamination (Komiyama et al., 1996).
     Metals in mangrove sediments do not appear to strongly affect bacterial
populations, even under heavy loads (Tam, 1998). However, if the metals are bioavailable,
they may accumulate in the macroinvertebrate fauna. In Yingluo Bay, He et al. (1996)
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        69


found high Zn and Cd levels in mollusks living in the mangal. Crustaceans had elevated Cu
levels and sipunculids concentrated Pb. Meyer et al. (1998) similarly found that mangrove
oysters (Crassostrea rhizophorae) in a northeastern Brazil mangal accumulated mercury
and are good biomarkers for mercury contamination. Bioaccumulation of such substances
can carry substantial human health risks. Mangrove sediment chemistry and the fate of
heavy metals are subjects that merits much more study.


6.9. Responses to organic pollution
    Three characteristics have long made mangrove habitats favored sites for sewage
dumping: (1) flow through the habitat disperses wastes from a point source over vast areas,
(2) the vegetation itself filters nutrients from the water, and (3) the mangrove soil, algae,
microbes, and physical processes absorb large amounts of the pollutants (Wong et al.,
1995, 1997b).
    Nutrients (primarily nitrogen and phosphorus) are often major components of the
pollution. Researchers have studied the ability of mangals to absorb nutrients and the
effects of the pollutants on the mangal community as a whole. In general, mangrove soils
efficiently trap wastewater-borne phosphorus, but are less effective at removing nitrogen
(Tam and Wong, 1995). Tam and Wong (1996a, b) experimentally tested the ability of
mangrove soils to absorb nutrients when treated with synthetic wastewater. The soils
retained both nitrogen and phosphorus. The bulk of these were trapped in the upper 1 cm
of the sediment where they could be processed by bacterial communities (Corredor and
Morell, 1994).
    Wong et al. (1995, 1997a) found that two full years of sewage discharge did not
adversely affect mangrove growth in the Funtian mangal of China. Nor did sewage affect
biomass, density, or community structure of the benthic macrofauna (Yu et al., 1997).
Furthermore, wastewater input did not seem to increase litter production or litter decay
rates (N.F.Y. Tam et al., 1998).
    While these studies suggest that mangroves are tolerant of organic pollution, results
should be viewed cautiously since they may not hold in other habitat. The effects of
sewage dumping will depend on the quantities of sewage, the duration of dumping, and the
unique characteristics of each mangal. Particularly important are the patterns of water flow
through the habitat since this will determine flushing rates and residence times of the
pollutants (Ridd et al., 1990; Uncles et al., 1990; Wolanski et al., 1990; Wattayakorn et al.,
1990).
    High levels of organic pollution can contribute to disease, death, and changes in
species compositions within the mangal (Tattar et al., 1994). Mandura (1997) found that
sewage discharge killed pneumatophores of Avicennia marina in the Red Sea. The loss of
the pneumatophores decreased surface area for respiration and nutrient uptake and retarded
the growth of the trees. The pollution can also have cascading effects on invertebrate
populations (e.g., Sanches and Camargo, 1995).
    Beyond simple nutrients, organic pollution in mangrove environments may include
other anthropogenic chemicals and debris. Mangrove sediments in Cienaga Grande de
Santa Marta and Chengue Bay (Colombian Caribbean) contain significant organochlorine
pesticide residues. The concentrations of some of these vary seasonally (Espinosa et al.,
1995). Large amounts of plastic and non-mangrove wood are present in the mangroves of
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                         70


Jamaica. The volume of these solid wastes correlates strongly with total rainfall in a nearby
metropolitan area (Green and Webber, 1996).

6.10. Responses to oil pollution
     Oil pollution from oil or gas exploration, petroleum production and accidental spills
severely damages mangrove ecosystems (Mastaller, 1996). Clean-up operations after such
calamities are costly and difficult (IUCN, 1993). Oiling of mangroves has a number of
significant consequences. One of the most immediate and obvious is defoliation of the
trees. The toxicity of the oil may depend on environmental conditions; oil has the greatest
effect on survival and growth of Rhizophora mangle when the trees are in hot, bright
outdoor conditions (Proffitt et al., 1995). Toxicity may also differ among mangrove
species. For example, along the coast of Sao Paulo, Brazil, an oil spill caused 25.9%
defoliation of Rhizophora mangle, 43.4% defoliation of Laguncularia racemosa, and
64.5% defoliation of Avicennia schaueriana (Lamparelli et al., 1997). Differential
mortality of the trees can potentially lead to long-term changes in the community structure.
     Oil in a mangrove habitat (whether from a spill or chronic input) can have other
less obvious effects on the mangroves. For example, sediments can have significant
hydrocarbon pollution long after a spill event, even when there is no evidence of petroleum
contamination on the trees or in water samples from surrounding water (Bernard et al.,
1995, 1996). Munoz et al. (1997) followed the breakdown of Arabian light crude oil in
mangrove peat for 8 full years. Sediments contaminated in the Galeta spill in Panama
continued to hold oil residue, including the full range of aromatic hydrocarbons, 5 years
after the spill (Burns et al., 1994). The authors suggest that it will take at least 20 years for
the toxicity to completely disappear.
     Grant et al. (1993) demonstrated that sediment oil can inhibit establishment and
decrease survival of mangrove seedlings for several years. This residual toxicity may
interfere with mangrove afforestation efforts (S. Zheng et al., 1997). Dutrieux et al. (1990)
planted Sonneratia caseolaris in soils that had been treated with oil. Many of the plants
were killed; the survivors were significantly stunted. The retained oil can also cause
mutation. Klekowski et al. (1994b, c) found a positive correlation between concentrations
of polycyclic aromatic hydrocarbons in mangal sediments, and the frequency of
Rhizophora mangle carrying chlorophyll-deficient mutations.
     The extent of mangrove damage from oil pollution will depend on the kind of oil,
and the magnitude and frequency of spilling. For example, fresh oil causes more leaf loss
in Avicennia seedlings than does aged oil (Martin et al., 1990; Grant et al., 1993). Boeer
(1996b) measured effects of a mineral oil spill in the Arabian Sea off Fujairah. The
mangroves were relatively unaffected and all signs of the spill were nearly gone only 7
months after the spill.
     Proffitt and Devlin (1998) tested the effects of sequential oilings on potted
Rhizophora mangle seedlings, first treating the seedlings with No. 6 fuel oil, followed, 34
months later, by crude oil. They found no evidence of cumulative or synergistic effects, but
this conclusion has been challenged because of unnatural laboratory conditions and low
statistical power (Ellison, 1999). Given the sensitivity of mangroves to soil conditions, it is
essential to study oil effects under conditions that reflect the natural environment as closely
as possible. For example, salinity should be held at field levels and realistic oil
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       71


concentrations should be used to model chronic exposure of plants in oil-contaminated soil
(Ellison, 1999).
    The most realistic measure of repeated oil exposure comes from field habitats
exposed to natural oiliness. Two large oil spills (the first in 1968 and the second in 1986)
have caused large-scale damage to mangrove forests in Panama. In addition to killing trees
outright, the oil retained in the sediments caused apparent sublethal effects (Duke et al.,
1997). The residual effects of oiling may make the mangroves more vulnerable to future
damage. More careful, long-term laboratory experiments under natural conditions are
necessary to understand the responses of mangroves to oil and the consequences of oiling.
    Oil contamination can damage animals living in the mangal, both in the sediments
and on submerged mangrove roots (e.g., Mackey and Hodgkinson, 1996). Five years after
the Galeta oil spill in Panama, there was a 60% decrease in the number of isopods on
submerged Rhizophora mangle prop roots and a 40-50% drop in the number of spiny
lobsters (Levings and Garrity, 1994; Levings et al., 1994). Oyster populations dropped
65% along mangrove channels and 99% in mangrove streams. The population decreases
are due, in part, to loss of root surface on which to attach (the surface area of submerged
roots decreased 38% in the channels and 74% in streams; Garrity et al., 1994).
    In addition to killing the mangrove fauna directly, oil can have indirect effects
resulting from habitat modification. Oil released during the 1991 Gulf War left a black tar
layer in the mangals along the Saudi Gulf. The tar layer created higher than normal
temperatures in the soil. The ecological consequences of the higher temperatures, and the
effect on epifauna and infauna, are not yet fully known (Boeer, 1996a).
    The general response of a mangrove forest to oiling can be divided into four
phases: 1) immediate effects, 2) structural damage, 3) stabilization and 4) recovery. The
third and fourth phases may take many years to occur, if they occur at all. In Brazil, a
mangrove area damaged by oil did not began to recover until approximately 10 years after
the event (Lamparelli et al., 1997). Assessing the effects of oil on mangrove environments
will require the development of creative methods for measuring impacts and accurate
modeling of the physical and chemical events associated with the spill (e.g., Jacobi and
Schaeffer Novelli, 1990; Lamparelli et al., 1997). These efforts, however, will only be
effective if they are supported by careful monitoring and long-term data sets.

6.11. Responses to pests
    A few of the many plants and animals that make their homes in the mangal are
serious pests that damage the mangroves, decreasing growth and productivity and, in
extreme cases, killing the trees. Some of the harmful species do not directly injure the
mangroves. Instead, they cause damage by competing for scarce resources. Allelopathic
interactions among mangrove species suggest that interspecific competition is a normal
process in the mangal. Toxic leachates from leaf litter of some mangroves (e.g. Lumnitzera
racemosa, Ceriops decandra and R. apiculata) inhibit the growth of roots and shoots of
Rhizophora apiculata and R. mucronata seedlings (Kathiresan and Thangam, 1989;
Kathiresan et al., 1993).
    In general, stressful osmotic conditions that lead to lignification and suberization
prevent the development of a luxuriant herbaceous undergrowth in mangrove forests so
there is not normally strong competition between mangrove and non-mangrove plants
(Schwamborn and Saint-Paul, 1996). However, damage to established stands can open
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       72


windows of opportunity for invasive species that may restructure the community (Kangas
and Lugo, 1990; Lugo, 1998). The mangrove fern Acrostichum, for example, is a weedy
pest that causes significant losses to mangrove forestry (Chan, 1996). The pest is currently
controlled by application of herbicides, but efforts to control it are being refocused on its
responses to shading and salinity (Medina et al., 1990).
     Mangroves themselves can become pests when they are introduced to new habitats.
At least 6 mangrove species have been introduced to the Hawaiian Islands since the early
1900’s. Rhizophora mangle has been a particularly successful transplant, but two other
species (Bruguiera gymnorrhiza and Conocarpus erectus) also have self-sustaining
populations. The mangroves were planted to help stabilize sediments in coastal mud flats.
As invaders, however, the mangroves have had negative effects. In particular, they
compete with native plants and modify habitats that are important to Hawaiian birds
(including endangered species). They also cause drainage problems in some areas (Allen,
1998).
     Other pest organisms damage the mangroves, not by competing with them, but
simply by living on their surfaces. For example, the spiders Tetragantha nitens and
Chiracanthium live on Rhizophora. They lay their eggs on the leaves, which induces leaf
rolling, chlorosis and wilting. Heavy infestations can kill the trees (Irianto and Suharti,
1994). The semi-parasitic mistletoe, Phthirusa maritima, has a more direct effect on the
trees. Infections in Conocarpus erectus and Coccoloba uvifera induce higher transpiration
rates, lower CO2 assimilation rates, and lower water-use efficiency (Orozco et al., 1990).
     By far the most extensive and serious damage to mangroves occurs through the
feeding activities of herbivorous animals. While most of the damage is done by animals
feeding in the canopy, several kinds of crustaceans and mollusks bore directly into
submerged mangrove wood and do significant damage. Spaeromatids are generally the
most common wood-borers (e.g., Sivakumar, 1992; Huang et al., 1996). Infestations of
these isopods are heavier in dead mangrove stumps than in live wood but the stumps and
woody debris provide a perennial source of larvae that also attack the living wood
(Sivakumar and Kathiresan, 1996). Distributions of these pests are controlled largely by
currents and tidal regimes.
     Of the animals feeding on the mangrove canopy, insects are undoubtedly the most
destructive. Murphy (1990d) described 102 insect herbivores that attack 21 mangrove
species in Singapore. Veenakumari et al. (1997) listed 197 species of herbivores on the
Andaman and Nicobar Islands. Some of the insect herbivores are serious crop pests that
simply use mangroves as alternative hosts. Others have apparent preferences for
mangroves (e.g., Mictis on Sonneratia, Glaucias on Lumnitzera, Calliphara on
Excoecaria, and Antestiopsis on Avicennia; Murphy, 1990d).
     Insect herbivores can completely defoliate mangrove stands. Rhizophora leaves that
have been attacked by scale insects (Aspidiotus destructor) first turn yellow at the site of
feeding, then brown and necrotic. In extreme cases, the leaves dry up, drop off, and the
entire seedling dies (Kathiresan, 1993). Periodic outbreak populations of the moth Achaea
serva defoliate large stands of Excoecaria agallocha (McKillup and McKillup, 1997). In
Singapore, feeding by Paralebeda and Selepa caterpillars can lead to total loss of shoots in
Excoecaria; Trabala krishna has the same effect on Sonneratia. Apical bud destruction
may reduce leaf production and change the architecture of the plant (Murphy, 1990d).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        73


     Summer feeding by the caterpillar Nephpterix syntaractis in Hong Kong
completely defoliates Avicennia marina, severely reducing the reproductive output of the
trees (Anderson and Lee, 1995). Kandelia candel in the same region may experience a
35% defoliation (Lee, 1991). In Belize, Central America, an outbreak of the lepidopteran
Phocides pigmalion on Rhizophora mangle increases leaf abscission rates and reduces
above-ground net primary production by 5-20%. The lost production normally would have
been exported to surrounding marine environments (Ellison and Farnsworth, 1996b). The
insect defoliator, Pteroma plagiophleps (Lepidoptera: Psychidae), has been newly recorded
on the Indian west coast (Santhakumaran et al., 1995).
     Insect herbivores may show preferences among mangrove hosts. In an Ecuadorian
mangal, the bagworm, Oiketicus kirbyi removed 80% of the foliage of Avicennia
germinans, 10% of the Conocarpus erectus and < 5% of the Laguncularia racemosa (Gara
et al., 1990). The susceptibility of mangrove species, and individual mangrove plants, may
relate to their physico-chemical characteristics. High leaf toughness, measured as the ratio
of protein to fibre, reduces palatability and digestibility (Choong et al., 1992). Tannins also
deter herbivores. Avicennia species, which have low tannin levels, suffer more herbivore
damage than do Rhizophora species, which have more tannins (Kathiresan, 1992).
     Feeding preferences of the insects may also be influenced by the health of the
mangrove. Nutrient enriched trees tend to suffer higher herbivory. Herbivory by
Ecdytolopha (an endophytic insect that feeds in apical buds) and Marmara (which mines
stems) on Rhizophora mangle increased significantly when the trees were treated with P
and NPK. Fertilization with N alone did not increase herbivory (Feller, 1995). Damage
from feeding herbivores may also invite further attack. Farnsworth and Ellison (1993)
made small holes in the leaves of R. mangle and found that the artificial damage increased
natural damage from herbivorous insects; in 50 days, the size of the holes had increased
45.1%.
     Some herbivores feed specifically on the reproductive tissues and seeds of
mangroves. Crabs are particularly important seed predators (Osborne and Smith, 1990;
Robertson, 1991; McGuinness, 1997b; Dahdouh-Guebas et al., 1998). However, insects
can also attack mangrove seeds. Insect borers appear to impair the growth of Avicennia
marina propagules, but do not kill them (Robertson et al., 1990). A mite (Afrocypholaelaps
africana) feeds on mangrove pollen. Unopened flower buds are mite-free, but newly
opened flowers are infested by all post-embryonic stages of the mite. Egg-bearing female
mites are dispersed among the mangroves by the honeybee Apis mellifer. The mite
population declines as the mangrove flowering season ends (Seeman and Walter, 1995). It
is not clear what affect the mites have on the mangrove population.

6.12. Responses to anthropogenic stress
    In recent years, anthropogenic pressures have significantly damaged the world’s
mangroves, with alarming levels of habitat loss. For example, Ramirez-Garcia et al. (1998)
estimate a 32% decrease in mangroves in the Santiago River of Mexico in the past 23
years. Aksornkoae (1993) and Raine (1994) report more than a 50% reduction in the
mangrove forests of Thailand. Mndeme (1995) reports that the mangrove resources in the
Mafia District of Tanzania are in danger of collapse. In the Florida Keys, USA, Strong and
Bancroft (1994) report that 15% of the original mangrove forests have been cleared for
development; mean forest size has decreased 41%. Approximately 45% of the mangroves
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                         74


in Indonesia have been heavily impacted by human activities (Choong et al., 1990). Some
estimates put global mangrove loss rates at one million ha y-1 (Mohamed, 1996). Such
levels of destruction and habitat fragmentation raise concerns about conservation of
biodiversity in the mangrove habitats and preservation of the mangals themselves.
     Ellison and Farnsworth (1996a) classified anthropogenic disturbances into four
types: extraction, pollution, reclamation, and changing climate. These disturbances are
listed in the order of their increasing spatial scale, their increasing temporal scales, and the
increasing time required for recovery. Research suggests that even relatively low impact
human activities can affect the mangrove environment. For instance, boardwalks placed in
the mangals around Sydney, Australia to provide access for educational and recreational
activities have modified sediment composition and changed benthic invertebrate
community structure (Kelaher et al., 1998a, 1998b). It may require fairly long periods for
the mangal to recover from even minor disturbances (Snedaker et al., 1992).
     Diversion of freshwater for irrigation and land reclamation has historically been a
major cause of wide-scale mangrove destruction (Conde and Alarcón, 1993, Twilley et al.,
1998). Throughout the world, mangroves and mangrove products have also been used for
timber, fuel, food, clothing, perfume, dyes, tannins, and medicine (Rasolofo, 1997;
reviewed by Bandaranayake, 1998). In the past several decades, extensive tracts of
mangrove have been converted for aquaculture. Shrimp ponds have become particularly
common in many former mangals (Twilley et al., 1993; Primavera 1995; de Graaf and
Xuan, 1998). Menasveta (1997) reports that nearly 55% of the mangroves in Thailand were
converted to shrimp ponds between 1961 and 1993. Pond culture now surpasses open
ocean fishing as the major source of shrimp there. Unfortunately, ponds in many regions
are unsustainable and up to 70% of them may be left idle after some period of production
(Stevenson, 1997). Because of changes in the sediments caused by pond construction, the
abandoned sites are difficult to revegetate with mangroves even after the shrimp farming
has ceased (de Graaf and Xuan, 1998).
     Intact mangals process heavy organic loads and could help oxidatively process
nutrients in shrimp pond effluents (Eguchi et al., 1997; Twilley et al., 1998). Robertson
and Phillips (1995) estimated that 2 to 22 hectares of mangrove forest could completely
filter excess nitrogen and phosphorus from a one-hectare shrimp pond. The effluents, in
turn, could promote growth of the mangroves. A 70% dilution of effluent from a semi-
intensive shrimp culturing pond in south India significantly increased growth of mangrove
seedlings (Rajendran and Kathiresan, 1996).
     In the Mekong Delta of Vietnam, living mangroves actually increase productivity
of shrimp aquaculture facilities. Binh et al. (1997) collected data suggesting that yields are
greater in shrimp ponds with 30-50% mangrove coverage. Farmers who integrate shrimp
and mangrove farming may, therefore, realize better economic returns (Hong and San,
1993). P.T. Smith (1996) found that sediments in the shrimp ponds are very similar to
those from nearby mangrove habitats, again suggesting that mangrove and shrimp
aquaculture should be compatible.
     Heavy historical exploitation of mangroves has left many habitats severely
damaged. The damage has consequences beyond loss of the trees themselves. For
example, because mangroves serve as nursery habitats for many crustaceans and fish,
damage can have a direct effect on fishery resources and the lives of those who depend on
them (John and Lawson, 1990; Twilley et al., 1991, 1998; Ruitenbeek, 1994; Fouda and
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       75


Al-Muharrami, 1995; Primavera, 1998). Recently, community-based approaches to
conservation and resource management have been launched with the participation of local
people (A.H. Smith and Berkes, 1993; Kairo, 1995; Semesi, 1998). Guidelines for
evaluation, restoration, and management of mangrove ecosystems are also being developed
(Field, 1996; Siddiqi and Khan, 1996; Ewel et al., 1998b; Gilbert and Janssen, 1998; Kaly
and Jones, 1998; Twilley et al., 1998).
    Efforts are being made to rebuild damaged mangrove ecosystems in many parts of
the world (Semesi, 1992; Chowdhury and Ahmed, 1994; Field, 1998). The programs are
called regeneration, reclamation, rehabilitation, or ecodevelopment. Finding adequate
supplies of viviparous seedlings for use in such afforestation projects is a challenge and
more effective methods are needed. Living seedlings can be cut and the cuttings induced to
produce roots and shoots. However, success of the cuttings depends on how they are done
(growth and survival depend on where the location and length of the cutting; Ohnishi and
Komiyama, 1998). In vitro micropropagation methods have been recently developed for
Excoecaria agallocha (C.S. Rao et al., 1998). These techniques hold promise for
mangrove regeneration.
    In South Sulawesi, Indonesia, where mangrove removal has produced significant
environmental problems, efforts are underway to launch mangrove agroforestry projects.
Planting of Rhizophora mucronata along the coast is mitigating coastal erosion and
preventing flooding (which otherwise damages aquaculture facilities). Controlled
harvesting of the mangroves produces income as the product is sold for fuel wood
(Weinstock, 1994). Efforts at mangrove agriculture are also underway in the Federated
States of Micronesia (Devoe and Cole, 1998). However, there is still much to learn about
proper management and sustainable harvesting of mangrove forests. Despite nearly 100
years of careful management, timber yields from the Matang Mangrove Forest Reserve in
Malaysia are declining significantly (Gong and Ong, 1995).

6.13. Responses to Global changes
     It is expected that increasing concentrations of atmospheric CO2 and other
"greenhouse gases" will bring changes in the global climate. It has been predicted that each
decade could bring a 0.3° rise in air temperature and a 6 cm rise in the global sea level
(Titus and Narayanan, 1996; Wilkinson, 1996; Gregory and Oerlemans, 1998). Because of
their location at the interface between land and sea, mangroves are likely to be one of the
first ecosystems to be affected by global changes. Most mangrove habitats will experience
increasing temperature, changing hydrologic regimes (e.g., changes in rainfall,
evapotranspiration, runoff and salinity), rising sea level and increasing tropical storm
magnitude and frequency (R.W. Stewart et al., 1990, Field, 1995; Michener et al., 1997).
Davis et al. (1994) have developed a framework for assessing risks to mangrove
ecosystems in the context of a changing global climate but the seriousness of the effects
will be strongly site-specific (Kjerfve and Macintosh, 1997).
     Small increases in air temperature may have little direct effect on the mangroves
(Field, 1995), but if temperatures exceed 35° C, root structures, seedling establishment and
photosynthesis will all be negatively affected. The broader effects of temperature increases
may be in modifying larger-scale distribution and community structure, increasing species
diversity in higher latitude mangals and promoting spread of mangroves into sub-tropical
saltmarsh environments (Ellison, 1994).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        76


     Because they are so specialized, and may live so close to their tolerance limits,
mangroves are particularly sensitive to minor variation in hydrological or tidal regimes
(Blasco et al., 1996). Reduced rainfall and runoff would produce higher salinity and
greater seawater-sulfate concentrations. Both would decrease mangrove production
(Snedaker, 1995). The most important effects, however, would come from rising sea
levels, but responses will vary among locations and will depend on the local rate of the rise
and the availability of sediment to support reestablishment of the mangroves (Pernetta
1993; Parkinson et al., 1994; Semeniuk, 1994; Woodroffe, 1995, 1999). For example, in
the Caribbean, mangrove seedlings are very sensitive to low sediment availability,
suggesting that mangroves will not survive on Caribbean coral islands if sea levels increase
as predicted (Ellison, 1996).
     Ellison and Farnsworth (1997) studied the response of Rhizophora mangle to
increased inundation, mimicking the sea-level changes expected in the Caribbean in the
next 50-100 years. After 2.5 years of higher water, plants would have significantly lower
rates of photosynthesis and growth, be shorter and narrower, have fewer branches and
leaves, and more acid-sulfide in their soils. The authors suggested that increased mangrove
growth rates predicted for increasing atmospheric CO2 may be offset by decreased growth
resulting from changes in tidal regimes.
     Sayed (1995) tested the effects of higher water levels on Avicennia marina by
flooding potted seedlings. The treatment resulted in stomatal closure, loss of chlorophyll
fluorescence, and a slight reduction of leaf water potential. Post-flooding recovery,
however, was rapid, suggesting that sea level rises could lead to colonization of supratidal
flats by this species (Sayed, 1995). As sea level rises, mangroves, in general, would tend to
shift landward. Human encroachment at the landward boundary, however, may make this
impossible. Consequently, the width of mangrove systems would be likely to decrease as
the sea-level rose (Kjerfve and Macintosh, 1997).
     The mangrove-associated fauna would be affected both directly by climatic
changes and indirectly by changes in the mangroves. Species that are tolerant of increasing
temperatures (e.g., fish, gastropods, mangrove crabs and other crustaceans) may adjust
rapidly to the changes. In contrast, soft-bodied animals and bivalve mollusks would be
very sensitive to higher temperatures. Desiccation that would accompany increasing
temperatures would harm many marine species associated with mangroves (Kjerfve and
Macintosh, 1997). For mangrove-dependent species, however, the most serious
consequences of a changing climate would likely be the loss of habitat as the global
mangrove forests declined.

7. ECOLOGICAL ROLES OF MANGROVE ECOSYSTEMS

7.1. Litter decomposition and nutrient enrichment
    Mangrove ecosystems produce large amounts of litter in the form of falling leaves,
branches and other debris. Decomposition of the litter contributes to the production of
dissolved organic matter (DOM) and the recycling of nutrients both in the mangal and in
adjacent habitats. The organic detritus and nutrients could potentially enrich the coastal sea
and, ultimately, support fishery resources. The contribution of the mangroves could be
particularly important in clear tropical waters where nutrient concentrations are normally
low.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        77


     The nutrient cycling begins when leaves fall from the mangroves and are subjected
to a combination of leaching and microbial degradation (Lee et al., 1990b; Chale, 1993).
Leaching alone removes a number of substances and can produce high levels of DOM
(Benner et al., 1990b). Potassium is the most thoroughly leached element with up to 95%
of the total potassium being removed in a very short time (Steinke et al., 1993b).
Carbohydrates also leach quickly during early decomposition. Tannins, in contrast, leach
very slowly and the high tannin contents may slow establishment of bacterial populations
in the initial period of decomposition. As the tannins are eventually leached, the bacterial
populations rapidly increase (Steinke et al., 1990; Rajendran, 1997; Rajendran and
Kathiresan, 1999b).
     Bacteria and fungi contribute to decomposition of the mangrove material and to the
transformation and cycling of nutrients. Fungi are the primary litter invaders, reaching their
peak in the early phases of decomposition (Rajendran, 1997). The phylloplane fungi do not
attack live leaves; they begin to break the leaf material down only after it has been
submerged. There are two major phases of fungal decomposition. Cellulase-producing
fungi first attack the leaves between 0 and 21 days after submergence; xylanase producers
are active between 28 and 60 days. Pectinase, amylase and protease producers are present
throughout decomposition (Singh and Steinke, 1992; Raghukumar et al., 1994a).
     Bacterial colonies appear shortly after the litter has been colonized by fungi. The
bacteria grow quickly and can reach very high densities. Zhuang and Lin (1993) measured
bacterial densities from 2 x 105 to 10 x 105 • g-1 on Kandelia candel leaves that had
decomposed for 2-4 weeks. This was about 100 times higher than densities of
actinomycetes and filamentous fungi. The N2-fixing azotobacters are one of the important
groups in the decomposing litter (Rajendran, 1997) and their activities may increase the
nitrogen content of the leaves 2 - 3 times (Wafar et al., 1997; Rajendran, 1997).
     Chale (1993) measured a similar rapid nitrogen increase in leaves after six weeks of
decomposition and suggested that the litter 1) provides a surface for microbial nitrogen
synthesis and 2) acts as a nitrogen reservoir. The C:N ratio of decomposing Avicennia
marina leaves drops dramatically from approximately 1432 to 28, due primarily to a large
increase in their nitrogen content (Mann and Steinke, 1992; Singh and Steinke, 1992). In
another study, N.F.Y. Tam et al. (1990) saw the C:N ratio in decomposing leaves increase
for one week, then decrease, and finally stabilize at approximately 74. They hypothesized
that the initial increase resulted from the conversion of particulate and soluble nitrogen in
the litter to proteins in bacteria and fungi.
     A number of factors can affect the rate of litter decomposition and, therefore, the
rates of nutrient cycling. For example, litter decomposition rates vary among mangrove
species. Avicennia leaves, because they are thinner and have fewer tannins, decompose
faster than those of other species (Sivakumar and Kathiresan, 1990; Steinke et al., 1990;
Kristensen et al., 1995). Avicennia leaves also sink and begin to decompose immediately
whereas the leaves of other species (e.g., Sonneratia and Rhizophora) may float for several
days (Wafar et al., 1997). Lu and Lin (1990) found that litter of Bruguiera sexangula
decomposes quickly. Aegiceras corniculatum, in contrast, decomposes slowly (Tam et al.,
1990).
     Decomposition is influenced by tidal height, rainfall and temperature. In
subtropical mangrove forests, mangrove debris decomposes substantially faster in the rainy
season (e.g., Woitchik et al., 1997). Mackey and Smail (1996) studied decomposition of
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                             78


Avicennia marina. They found significantly faster decomposition in lower intertidal zones
with greater inundation. They also found an exponential relationship between leaf
decomposition rate and latitude with leaves decomposing most quickly at low latitudes.
They attributed the pattern to temperature differences, and concluded that seasonality can
have important effects on organic cycling and nutrient export from mangrove systems.
     Breakdown and decomposition of mangrove litter is accelerated by the feeding
activities of invertebrates (Camilleri, 1992). The animals may process large volumes of
the litter, contributing significantly to nutrient dynamics. Litter turnover rates have been
estimated by measuring rates of leaf decomposition. However, estimates made this way are
generally 10-20 times lower than rates calculated from actual measurements of leaf fall and
litter standing crop. The difference in the estimates can be attributed to 1) tidal export and
2) the feeding activities of crabs. The
                         7
crab feeding may be the more
important of these in many regions.              A. marina
                         6
                                Control
For example, in the Ao Nam Bor
                        -1
                        Juvenile peneid shrimp • haul
                         5
mangrove forest in Thailand, crabs
process about 80% of the litter         4
deposited in the mid-intertidal zone
                         3
and nearly 100% of the leaves
deposited in the high intertidal         2

(Poovachiranon and Tantichodok,         1
1991). In field experiments, Twilley
et al. (1997) found that mangrove        0

crabs process the mangrove material           10    20   30    40    50 60   70

very quickly. They removed a full                   Days of decomposition

day’s accumulation of mangrove leaf       Figure 17. Number of penaeid shrimp associated with
litter in only 1 hr. Because the mangrove    decomposing leaves of Avicennia marina (in situ litterbag
                         experiment by Rajendran, 1997). Populations increase
material is quite refractory, it may need
                         dramatically, but only after several weeks of
to decompose for some time before it is     decomposition.
useful to other invertebrates. Wafar et al.
(1997) estimated that litter needs to
decompose for about two months before it can be used in most detritivores’ diets. In situ
observations verify that mangrove leaves attract shrimp, crabs, and fish (particularly
juveniles), but only after several weeks of decomposition (e.g., Rajendran, 1997;
Rajendran and Kathiresan, 1999a; Figure 17).

7.2. Food webs and energy fluxes

     Mangals contribute to complex food webs and important energy transfers.
However, it is not clear how, or whether, these processes affect the larger ecosystem.
While the living vegetation is a valuable food resource for insects, crustaceans, and some
vertebrates, most of the mangrove production is transferred to other trophic levels through
litterfall and detrital pathways (Figure 18). Mangrove forests produce organic carbon well
in excess of the ecosystem requirements. Duarte and Cebrian (1996) estimate that the
excess photosynthetic carbon approaches 40% of net primary production. While some of
this organic matter simply accumulates in the sediments, large amounts could potentially
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        79


                        be transported offshore (Alongi, 1990b;
                        Robertson et al., 1991, 1992; Lee, 1995). The
                        amount of material exported, however,
                        depends strongly on local conditions and
                        varies enormously among mangals (Twilley et
                        al., 1992).
                             Material exported from the mangroves
                        could potentially support offshore
                        communities (Marshall, 1994; Robertson and
                        Alongi, 1995; Van Tussenbroek, 1995), but
                        the connections between mangal and adjacent
                        habitats are complex, dynamic, and have been
                        difficult to demonstrate unequivocally (Alongi
                        et al., 1992; Twilley et al., 1992; Hemminga et
                        al., 1995; Alongi, 1998). For instance,
                         Jennerjahn and Ittekkot (1997) found that
                         organic matter in continental sediments in
 Figure 18. A stylised food web in a mangrove
                         eastern Brazil was very different from that in
 ecosystem. The food web may be highly
 localized without strong connections to other mangrove environments and concluded the
 habitats. The foundations of the web are    mangrove matter is largely retained and
 detritus, microbes, algae and seagrasses.
                         decomposed within the mangal itself. Studies
with stable isotopes also suggest that mangroves do not make major contribution to coastal
food webs (Primavera, 1996; Loneragan et al., 1997). In fact, the data suggest that carbon
may instead be flowing from oceanic systems into the mangrove habitat (Figure 19).
Oceanic carbon contributed up to 86% of the particulate organic carbon (POC) in water
samples from a Brazilian mangal (Rezende et al., 1990).
     It appears that mangroves, in general, make only a localized contribution to the
food web (Fleming et al., 1990; Mohammed and Johnstone, 1995). Sediment meiofauna,
for example, feed directly on mangrove detritus. The composition of the meiofaunal
community changes during the process of litter decay, suggesting that the community is
responding to chemical changes in the leaves (Gee and Somerfield, 1997). The meiofaunal
community, though large in some habitats, may largely be a trophic dead end that
contributes little to the larger food web (Schrijvers et al., 1998).
     The mangroves may have stronger trophic linkages with epibenthic invertebrates
and fish living in the mangal and in nearby habitats (e.g., seagrass beds). For example,
mangrove detritus contributes to the nutrition of juvenile Penaeus merguiensis living in
tidal creeks. The juveniles feed directly on mangrove detritus, on other small detritivorous
invertebrates, and on benthic microalgae growing in the mangal (Newell et al., 1995).
Shrimp in mangrove estuaries may also feed heavily on seagrass epiphytes (Loneragan et
al., 1997). Invertebrates may also feed on the variety of cyanobacteria and microalgae that
live on submerged portions of the mangroves and on leaf litter (e.g., Sheridan, 1991;
Farnsworth and Ellison, 1995; Pedroche et al., 1995).
     Pinto and Punchihewa (1996) found that syngnathid fish (pipefish) in the Negombo
Estuary of Sri Lanka fed primarily on mangrove litter. However, mangroves apparently
contribute little of the carbon assimilated by other fishes. This is true despite the movement
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                             80


of a number of fish species between
                             -14
mangrove habitats and nearby seagrass                              A
                                                S
beds (Marguillier et al., 1997).                                S
                                    S
                             -16      S
    Mangrove detritus is probably
more important as a substrate for                   SS
                                    S
                             -18      S      S
                                   SS            S
microbial activity and nutrient                    S      A
                                          A
                                    S
regeneration than it is as a direct food                      S
                                    S      SS
                             -20
source for detritivores. Wafar et al.                       SS
                       δC
                        13                  S

(1997) analyzed energy and nutrient      (o/oo)         PP
                             -22            S
fluxes between mangroves and                            A
                                                'D
                                          S
estuarine waters and concluded that                  P
                             -24                  M
                                   A
                                          M
mangroves contribute significantly to                             'D

the estuarine carbon budget. However,          -26            M
                                                P
they contribute little to nitrogen and                 M
                                    M      M      P
                                   MM            M
phosphorus budgets. It is not clear                  M
                             -28           MM
                                   M'D
                                         M
whether any of these substances are                  M
                                                M
exported from the mangal in sufficient                M
                             -30

quantities to make significant
                               Sibunag River, Sementa Besar Laguna Joyuda,
contributions to energy flow and the              Philippines & Buloh River, Puerto Rico
                                       Malaysia
ecology of the broader ecosystem
(Alongi et al., 1992; Alongi, 1998).       Figure 19. Ratios of stable carbon isotopes in
Mangrove sediments efficiently uptake, retain shrimp collected from mangrove habitats in the
and recycle nitrogen (Rivera-Monroy et al.,    Philippines (Primavera, 1996), Malaysia (Rodelli et
                         al., 1984), and Puerto Rico (Stoner and
1995). Resident bacteria and benthic algae
                         Zimmerman, 1988). Shrimp tissue δ13 values (S)
rapidly assimilate available ammonium and
                         are much closer to the δ13 val ues of plankton (P)
prevent its export (Kristensen et al., 1995;
                         and algae (A) than they are to those of mangrove
Middelburg et al., 1996). The mangrove      leaves (M) or detritus (D). This suggests that the
environment may, therefore, represent a      shrimp are deriving their carbon primarily from
nutrient and carbon sink rather than a source   algae and the plankton; the mangrove detrital
                         pathway contributes little to their nutrition (after
for adjacent habitats. Careful measurements
                         Primavera, 1996).
and creative experimentation will be
necessary to clarify the role these habitats
play in larger-scale food webs and energy fluxes.

8. CONCLUDING REMARKS
    Mangrove ecosystems are receiving increasing attention, but we still lack much
basic information about their structure and function. There are still fundamental gaps in our
knowledge of the reproductive biology of mangroves, and mangrove evolution is poorly
understood. We are still far from understanding energy flow and food web dynamics in
mangrove environments and how the mangroves connect with other ecosystems. There is a
great need to better understand the effects of environmental change and pollution on
mangrove flora and fauna. Animals that are highly dependent on mangroves need
additional study, particularly with respect to larval supply and recruitment. Such
ecobiological research can be linked to management of mangroves and associated fishery
resources (e.g., Bacon and Alleng, 1992; Hudson and Lester, 1994; Fouda and Al-
Muharrami, 1995).
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                            81


     Mangrove ecosystems are seriously threatened, mainly by human activities that
impact the habitat (Pons and Fiselier, 1991; Fouda and Al-Muharrami, 1995; Farnsworth
and Ellison, 1997a; Figure
20). The value of mangroves
has gone unrecognized for
many years (Farnsworth,
1998a) and the forests are
disappearing in many parts
of the world. The full extent
of the damage is not yet fully
known, but technological
advances (e.g., airborne
multispectral sensors and
satellite imagery) are
                   Figure 20. Factors impacting mangroves and ecosystems. These
allowing researchers to map
                   valuable systems are under pressure from a variety of physical,
and monitor mangrove
                   chemical, and biological processes. Many of the stresses on these
habitats (Ibrahim and Hashim;    environments result from human activities.
1990; Gang and Agatsiva,
1992; Lin et al., 1994; Aschbacker et al., 1995; Wei et al., 1995; Long and Skewes, 1996;
Green et al., 1997, 1998; N.F.Y. Tam et al., 1997; Blasco et al., 1998; De Jesus and Bina,
1998). The results of such studies are not encouraging; mangrove habitats continue to
shrink around the world.
     Even where efforts have been made to slow the destruction, remaining forests have
a number of problems. In some areas, the health and productivity of the forests have
declined significantly. In Indian mangrove ecosystems, 67% of the mangrove plants, 52%
of the macroalgae, 10% of the invertebrates and 4% of the vertebrates are endangered (e.g.,
Ananda Rao et al., 1998). Similar losses have occurred in the mangals of Singapore
(Turner et al., 1994) and are likely to be seen in other regions of the world. Mangrove
systems require intensive care to save threatened taxa from extinction. The causes of these
tragic losses differ from habitat to habitat but are generally tied directly or indirectly to
human activities. Individual study is required to determine the most effective remedial
measures. Where degraded areas are being revegetated, continued monitoring and thorough
assessment must be done to help us understand the recovery process (van Speybroeck,
1992). This knowledge will help us develop strategies to effectively rehabilitate degraded
mangrove habitats the world over.
     It has long been known that mangrove protect and stabilize coastlines. They are
more effective than concrete barriers in reducing erosion, trapping sediments, stabilizing
shorelines, and dissipating the energy of breaking waves (Pearce, 1996). We have learned
that they are critical nursery habitats for important marine species. Pioneering
investigations are now showing that mangroves and their associated fauna can be sources
of valuable products like black tea, mosquitocides, gallotannins, microbial fertilizers,
antiviral drugs, anti-tumor drugs and UV-screening compounds (Ravi and Kathiresan,
1990; Premanathan et al., 1992; Kathiresan and Pandian, 1991, 1993; Kathiresan, 1995b;
Kathiresan et al., 1995a; Ravikumar, 1995; Moorthy and Kathiresan, 1997b;
Bandaranayake, 1998; Palaniselvam, 1998; Kathiresan, 2000). Mangroves may be
developed as sources of high value commercial products and fishery resources and as sites
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                            82


for a burgeoning ecotourism industry
(Thorhaug, 1990; Ruitenbeek, 1994;
Barton, 1995). Their unique features may
also make them ideal sites for
experimental studies of biodiversity and
ecosystem function (Osborn and
Polsenberg, 1996; Farnsworth, 1998b;
Field et al., 1998). All this will require
that the resource is understood, carefully
managed, and protected (Farnsworth and
Ellison, 1997b; Ammour et al.., 1999;
Figure 21). Involvement of local
communities in conservation and
education in wise use of our precious
mangrove resources will ensure that
these unique ecosystems survive and
flourish.

                       Figure 21. Mangroves are highly dynamic and complex
                       systems that are still poorly understood. Continued
                       study, combined with concerted conservation efforts will
                       be necessary to preserve these fragile and unique
                       environments (from Rutzler & Feller, 1996).


ACKNOWLEDGEMENTS
We gratefully acknowledge the help of the editors of ‘Advances in Marine Biology’ in
bringing this review to publication. We thank the authorities of Annamalai University and
Western Washington University for providing facilities and resources to complete the
project. We thank Mrs. Sumathi Kathiresan, T. Smoyer, K. Short, R. Lopresti, D. Morgan,
S. Strom, B. Kjerfve and dedicated research students (Dr. N. Rajendran, Dr. V.
Palaniselvam, Ms. B. Kamakshi, Messrs. T. Ramanathan, M. Masilamani Selvam and K.
Sivakumar) for their help during the preparation of the manuscript. A.M. Ellison, S. Strom,
N.M. Aguilar and D. Morgan read and commented on portions of the manuscript. A.M.
Ellison kindly provided unpublished data and information. Two anonymous reviewers
provided constructive criticism
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      83




REFERENCES


Abbey-Kalio, N.J. (1992). A pilot study of mangrove litter production in the Bonny estuary
    of southern Nigeria. Discovery and Innovation 4 (3), 71-78.
Abdel-Razik, M. S. (1991). Population structure and ecological performance of the
    mangrove Avicennia marina (Forssk.) Vierh. on the Arabian Gulf coast of Qatar.
    Journal of Arid Environments 20 (3), 331-338.
Abhaykumar, V.K. and Dube, H.C. (1991). Epiphytic bacteria of mangrove plants
    Avicennia marina and Sesuvium portulacastrum. Indian Journal of Marine
    Sciences 20 (4), 275-276.
Achmadi, S., Syahbirin, G., Choong, E.T. and Hemingway, R.W. (1994). Catechin-3-O-
    rhamnoside chain extender units in polymeric procyanidins from mangrove bark.
    Phytochemistry 35 (1), 217-219.
Acosta, A. (1997). Use of multi-mesh gillnets and trammel nets to estimate fish species
    composition in coral reef and mangroves in the southwest coast of Puerto Rico.
    Caribbean Journal of Science 33 (1-2), 45-57.
Acosta, C.A. and Butler, M.J. IV (1997). Role of mangrove habitat as a nursery for
    juvenile spiny lobster, Panulirus argus, in Belize. Marine & Freshwater Research
    48 (8), 721-727.
Adams, E.S. (1994). Territory defense by the ant Azteca trigona: Maintenance of an
    arboreal ant mosaic. Oecologia 97 (2), 202-208.
Addison, D.S., Ritchie, S.C., Webber, L.A. and Van-Essen, F. (1992). Eggshells as an
    index of aedine mosquito production. 2: Relationship of Aedes taeniorhynchus
    eggshell density to larval production. Journal of American Mosquito Control
    Association 8 (1), 38-43.
Aguilar, N.M. 2000. Comparative physiology of air-breathing gobies. Ph.D. Thesis, Scripps Institution of
    Oceanography, University of California, San Diego. 211pp.
Aguilar, N.M., Ishimatsu, A., Ogawa, K. and Khoo, K.H. (in press). Aerial ventilatory
    responses of the mudskipper, Periophthalmodon schlosseri, to altered aerial and
    aquatic respiratory gas concentrations. Comparative Biochemistry and Physiology,
    Series A.
Akatsu, M., Hosoi, Y., Sasamoto, H. and Ashihara, H. (1996). Purine metabolism in cells
    of a mangrove plant, Sonneratia alba, in tissue culture. Journal of Plant Physiology
    149 (1-2), 133-137.
Aksornkoae, S. (1993). Ecology and Management of Mangrove. IUCN - The World
    Conservation Union. Bangkok, Thailand. 176 pp.
Alcala-Herrera, J.A., Jacob, J.S., Machain Castillo, M.L. and Neck, R.W. (1994). Holocene
    paleosalinity in a Maya wetland, Belize, inferred from the microfaunal assemblage.
    Quaternary Research 41 (1), 121-130.
Alias, S.A., Kuthubutheen, A.J. and Jones, E.B.G. (1995). Frequency of occurrence of
    fungi on wood in Malaysian mangroves. Hydrobiologia 295 (1-3), 97-106.
Alias, S.A., Hyde, K.D. and Jones, E.B.G. (1996). Pyrenographa zylographoides from
    Malaysian and Australian mangroves. Mycological Research 5, 580-582.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       84


Allen, J.A. (1998). Mangroves as alien species: the case of Hawaii. Global Ecology and
    Biogeography Letters 7 (1), 61-71.
Alongi, D.M. (1990a). Community dynamics of free-living nematodes in some tropical
    mangrove and sandflat habitats. Bulletin of Marine Science 46 (2), 358-373.
Alongi, D.M. (1990b). Effect of mangrove detrital outwelling on nutrient regeneration and
    oxygen fluxes in coastal sediments of the central Great Barrier Reef lagoon.
    Estuarine, Coastal and Shelf Science 31 (5), 581-598.
Alongi, D.M. (1994). Zonation and seasonality of benthic primary production and
    community respiration in tropical mangrove forests. Oecologia 98 (3-4), 320-327.
Alongi, D.M. (1996). The dynamics of benthic nutrient pools and fluxes in tropical
    mangrove forests. Journal of Marine Research 54, 123-148.
Alongi, D.M. (1998). Coastal Ecosystems Processes. CRC Press, Boca Raton, FL, USA.
Alongi, D.M. and Sasekumar, A. (1992). Benthic communities. In “Tropical Mangrove
    Ecosystem” (A.L. Robertson and A.M. Alongi, eds), pp. 137-172. American
    Geophysical Union, Washington DC., USA.
Alongi, D.M., Boto, K.G. and Robertson, A.I. (1992). Nitrogen and phosphorous cycles. In
    “Tropical Mangrove Ecosystems” (A.I. Robertson and D.M. Alongi, eds), pp. 251-
    292. American Geophysical Union, Washington, D.C.
Alongi, D.M., Christoffersen, P. and Tirendi, F. (1993). The influence of forest type on
    microbial-nutrient relationships in tropical mangrove sediments. Journal of
    Experimental Marine Biology and Ecology 171 (2), 201-223.
Alongi, D.M., Sasekumar, A., Tirendi, F. and Dixon, P. (1998). The influence of stand age
    on benthic decomposition and recycling of organic matter in managed mangrove
    forests of Malaysia. Journal of Experimental Marine Biology and Ecology 225 (2),
    197-218.
Aluri, R.J. (1990). Observations on the floral biology of certain mangroves. Proceedings of
    the Indian National Science Academy, Part B, Biological Sciences 56 (4), 367-374.
Alvarez-Léon, R. (1993). Mangrove ecosystems of Colombia. In “Conservation and
    Sustainable Utilization of Mangrove Forests in Latin America and Africa Regions”
    (L.D. Lacerda, ed), pp. 75-114. Society for Mangrove Ecosystems, Okinawa.
Alves, V.S., Soares, A.A. and Ribeiro, A.B. (1997). Birds of the Jequia mangrove system,
    Ilha do Governador, Baía de Guanabara, Rio de Janeiro, Brazil. In “Mangrove
    Ecosystem Studies in Latin America and Africa” (B. Kjerfve, L.D. Lacerda and S.
    Diop, eds), pp. 163-170. UNESCO, Paris.
Ambler, J.W., Ferrari, F.D. and Fornshell, J.A. (1991). Population structure and swarm
    formation of the cyclopoid copepod Dioithona oculata near mangrove cays.
    Journal of Plankton Research 13 (6), 1257-1272.
Ammour, T., Imbach, A., Suman, D. and Windevoxhel, N. (1999). Manejo productivo de
    manglares in América Central. CATIE, Turrialba, Costa Rica.
Ananda Rao, T. (1998). Flowering phenology and pollination of the eumangroves and their
    associates to plan regeneration and breeding programmes. Journal of Economic
    Taxonomy and Botany 22, 19-27.
Ananda Rao, T., Molur, S. and Walker, S. (1998). Report of the workshop on
    “Conservation Assessment and Management Plan for Mangroves of India” (21-25,
    July 1997). Zoo Outreach Organization, Coimbatore, India.106 pp.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                     85


Anderson, C. and Lee. S.Y. (1995). Defoliation of the mangrove Avicennia marina in
    Hong Kong cause and consequences. Biotropica 27 (2), 218-226.
Anderson, D.T., Anderson, J.T. and Egan, E.A. (1988). Balanoid barnacles of the genus
    Hexaminius (Archaeobalaninae: Elminiidae) from mangroves in New South Wales,
    including a description of a species. Records of the Australian Museum 40, 205-
    223.
Anwahi, A., Al-Zarouni, M.A.R, Al-Janahi, A. and Cherian, T. (1998). Feasibility studies
    on mangrove Avicennia marina cultivation below ground level along the bank of a
    dug-out pond. Marine and Freshwater Research 49, 359-361.
Arancibia, A.Y., Dominguez, A.L.L. and Day, J.W. Jr. (1993). Interaction between
    mangrove and seagrass habitats mediated by estuarine nekton assemblages:
    Coupling of primary and secondary production. Hydrobiologia 254, 1-12.
Araujo, R.J., Jaramillo, J.C. and Snedaker, S.C. (1997). LAI and leaf size differences in
    two red mangrove forest types in south Florida. Bulletin of Marine Science 60 (3),
    643-647.
Aschbacker, J., Ofren, R., Delsol, J.P., Suselo, T.B., Vibulsresth, S. and Charrupat, T.
    (1995). An integrated comparative approach to mangrove vegetation mapping using
    advanced remote sensing and GIS technologies: preliminary results. Hydrobiologia
    295 (1-3), 285-294.
Ashihara, H., Adachi, K., Otawa, M., Yasumoto, E., Fukushima, Y., Kato, M., Sano, H.,
    Sasamoto, H., and Baba, S. (1997). Compatible solutes and inorganic ions in the
    mangrove plant Avicennia marina and their effects on the activities of enzymes.
    Zeitschrift fuer Naturforschung 52 (7-8), 433-440.
Au, D.W.T., Vrijmoed, L.L.P. and Jones, E.B.G. (1996). Ultrastructure of asci and
    ascospores of the mangrove ascomycete Dactylospora haliotrepha. Mycoscience 37
    (2), 129-135.
Azariah, J., Azariah, H., Gunasekaran, S. and Selvam, V. (1992). Structure and species
    distribution in Coringa mangrove forest, Godavari Delta, Andhra Pradesh, India.
    Hydrobiologia 247, 11-16.
Azocar, A., Rada, F. and Orozco, A. (1992). Relaciones hidricas e intercambio de gases en
    dos especies de mangle, con mecanismos contrastantes de regulacion de la
    salinidad interna. Ectropicos 5 (2), 11-19.
Baba, S. and Onizuka, R.I. (1997). Callus induction of five mangrove tree species. In “
    Mangrove Ecosystem Studies in Latin America and Africa” (B. Kjerfve, L.D.
    Lacerda and S. Diop, eds), pp. 339-347. UNESCO, Paris.
Bacon, P.R. (1994). Template for evaluation of impacts of sea level rise on Caribbean
    coastal wetlands. Ecological Engineering 3 (2), 171-186.
Bacon, P.R. and Alleng, G.P. (1992). The management of insular Caribbean mangroves in
    relation to site location and community type. In: “The ecology of mangrove and
    related ecosystems” (Jaccarini, V. and E. Martens, eds), pp. 235-241. Kluwer
    Academic Publishers, Netherlands.
Baelde, P. (1990). Differences in the structures of fish assemblages in Thalassia
    testudinum beds in Guadeloupe, French West Indies, and their ecological
    significance. Marine Biology 105 (1), 163-173.
Balasubrahmanyan, K. (1994). Micro-invertebrate benthic fauna of Pichavaram
    mangroves. In “Conservation of mangrove forest genetic resources, A Training
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      86


    manual” (V.D. Sanjay and V. Balaji, eds), pp. 257 -259. M.S. Swaminatham
    Research Foundation, Madras.
Baldwin, A.H., Platt, W.J., Gathen, K.L., Lessmann, J.M. and Rauch, T.J. (1995).
    Hurricane damage and regeneration in fringe mangrove forests of southeast Florida,
    USA. Journal of Coastal Research 21, 169-183.
Ball, M.C. (1996). Comparative ecophysiology of mangrove forest and
    tropical lowland moist forest. In “Tropical forest plant
    ecophysiology” (S.S. Mulkey, R.L. Chazdon and A.O. Smith, eds), pp.461-469.
    Chapman and Hall, New York.
Ball, M.C. (1998). Mangrove species richness in relation to salinity and waterlogging: a
    case study along the Adelaide River floodplain, northern Australia. Global Ecology
    and Biogeography Letters 7 (1), 73-82.
Ball, M.C. and Munns, R. (1992). Plant responses to salinity under elevated atmospheric
    concentrations of CO2. Australian Journal of Botany 40, 515-525.
Ball, M.C. and Passioura, J.B. (1993). Carbon gain in relation to water
    use: photosynthesis in mangroves. In “Ecophysiology of
    Photosynthesis” (E.D. Sehulze and N.M. Caldwell, eds), pp. 247-257. Springer,
    Kiedelberg, Berlin.
Ball, M.C. and Pidsley, S.M. (1995). Growth responses to salinity in relation to
    distribution of two mangrove species, Sonneratia alba and S. lanceolata, in
    northern Australia. Functional Ecology 9 (1), 77-85.
Ball, M.C., Cochrane, M.J. and Rawson, H.M. (1997). Growth and water use of the
    mangroves Rhizophora apiculata and R. stylosa in response to salinity and
    humidity under ambient and elevated concentrations of atmospheric CO2. Plant,
    Cell & Environment 20 (9), 1158-1166.
Balsamo, R.A. and Thomson, W.W. (1995). Salt effects on membranes of the hypodermis
    and mesophyll cells of Avicennia germinans (Avicenniaceae): a freeze-fracture
    study. American Journal of Botany 82 (4), 435-440.
Bancroft, G.T., Strong, A.M. and Carrington, M. (1995). Deforestation and its effects on
    forest-nesting birds in the Florida Keys. Conservation Biology 9 (4), 835-844.
Bandaranayake, W.M. (1998). Traditional and medicinal uses of mangroves. Mangroves
    and Salt Marshes 2, 133-148.
Barni, N.C. and Chanda, S. (1992). Late-Quaternary pollen analysis in relation to
    paleoecology, biostratigraphy and dating of Calcutta peat. Proceedings of the
    Indian National Science Academy Part-B, Biological Sciences 58 (4), 191-200.
Barnes, D.K.A. (1997). Ecology of tropical hermit crabs at Quirimba Island, Mozambique:
    Distribution, abundance and activity. Marine Ecology Progress Series 154, 133-
    142.
Barton, D.N. (1995). Partial Economic Valuation of management alternatives for the
    Terraba-Sierpe wetlands, Costa Rica. Senter for Miljoe- og Ressursstudier
    Rapport 21, 1-30.
Basak, U.C., Das, A.B and Das, P. (1995). Metabolic changes during rooting in stem
    cuttings of five mangrove species. Plant Growth Regulation 17 (2), 141-148.
Basak, U.C., Das, A.B. and Das, P. (1996). Chlorophylls, carotenoids, proteins and
    secondary metabolites in leaves of 14 species of mangrove. Bulletin of Marine
    Science 58 (3), 654-659.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      87


Basak, U.C., Das, A.B and Das, P. (1998). Seasonal changes in organic constituents in
    leaves of nine mangrove species. Marine and Freshwater Research 49 (5), 369-
    372.
Bauer-Nebelsick, M., Bardelo, C.F. and Ott, J.A. (1996). Electron microscopic studies on
    Zoothamnium niveum (Hemprich & Ehrenberg, 1831) Ehrenberg 1838
    (Oligohymenophora, Peritrichida), a ciliate with ectosymbiotic, chemoautotrophic
    bacteria. Europe Journal of Protistology 32 (2), 207-215.
Bayliss, D.E. (1993). Spatial distribution of Balanus amphitrite and Elminius adelaidae on
    mangrove pneumatophores. Marine Biology 116 (2), 251-256.
Becker, P., Asmat, A., Mohamad, J., Moksin, M. and Tyree, M.T. (1997). Sap flow rates of
    mangrove trees are not usually low. Trees 11, 432-435.
Benka-Coker, M.O. and Olumagin, A. (1995). Waste drilling fluid utilizing
    microorganisms in a tropical mangrove swamp oilfield location. Bioresource
    Technology 53 (3), 211-215.
Benka-Coker, M.O. and Olumagin, A. (1996). Effects of waste-drilling fluid on bacterial
    isolates from a mangrove swamp oilfield location in the Niger Delta of Nigeria.
    Bioresource Technology 53 (3), 175-179.
Benner, R., Weliky, K. and Hedges, J.I. (1990a). Early diagenesis of mangrove leaves in a
    tropical estuary; molecular-level analyses of neutral sugars and lignin-derived
    phenols. Geochimica et Cosmochimica Acta 54 (7), 1991-2001.
Benner, R., Hatcher, P.G. and Hedges, J.I. (1990b). Early diagenesis of mangrove leaves in
    a tropical estuary; bulk chemical characterization using solid-state (super 13) C
    NMR and elemental analyses. Geochimica et Cosmochimica Acta 54 (7), 2003-
    2013.
Bera, S. and Purkayastha, R.P. (1992). Physiological studies on strains of Pestalotiopsis
    versicolor isolated from a mangrove plant. Journal of Mycopathological Research
    30 (2), 157-165.
Bernard, D., Jeremie, J.J. and Pascaline, H. (1995). First assessment of hydrocarbon
     pollution in a mangrove estuary. Marine Pollution Bulletin 30 (2), 146-150.
Bernard, D., Pascaline, H. and Jeremie, J.J. (1996). Distribution and origin of
     hydrocarbons in sediments from lagoons with fringing mangrove communities.
     Marine Pollution Bulletin 32 (10), 734-739.
Berry, A.J. (1975). Molluscs colonizing mangrove trees with observations on Enigmonia
     rosea (Anomiidae). Proceedings of the Malacological Society of London 41, 589-
     600.
Bhosale, L.J. (1994). Propagation techniques for regeneration of mangrove forests-A new
     asset. Journal of Non-timber Forest Products 1 (3-4), 119-122.
Bhosale, L.J. and Mulik, N.G. (1991). Strategies of seed germination in mangroves. In
    “Proceedings on International Seed Symposium” (N. S. David and S. Mohammed,
    eds), pp. 201-205. Jodhpur, India.
Bingham, B.L. (1992). Life histories in an epifaunal community: coupling of adult and
    larval processes. Ecology 73 (6), 2244-2259.
Bingham, B.L. and Young, C.M. (1991a). Larval behaviour of the ascidian Ecteinascidia
    turbinata Herdman; an in situ experimental study of the effects of swimming of
    dispersal. Journal of Experimental Marine Biology and Ecology 145 (2), 189-204.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      88


Bingham, B.L. and C.M. Young (1991b). Allelopathy and the influence of sponges on the
    recruitment of marine invertebrate larvae. Marine Biology 109, 19-26.
Bingham, B.L. and Young, C.M. (1995). Stochastic events and dynamics of a mangrove
    root epifaunal community. Marine Ecology 16 (2), 145-163.
Binh, C.T., Phillips, M.J. and Demaine, H. (1997). Integrated shrimp-mangrove farming
    systems in the Mekong Delta of Vietnam. Aquaculture Research 28 (8), 599-610.
Blaber, S.J.M. and Milton, D.A. (1990). Species composition, community structure and
    zoogeography of fishes of mangrove estuaries in the Solomon Islands. Marine
    Biology 105 (2), 259-267.
Blaber, S.J.M., Salini, J.P. and Brewer, D.T. (1990a). A check list of the fishes of
    Albatross Bay and the Embley estuary, north eastern Gulf of Carpentaria. CSIRO
    Marine Laboratories Report Series 210, 1 - 22.
Blaber, S.J.M., Brewer, D.T., Salini, J.P. and Kerr, J. (1990b). Biomasses catch rates and
    patterns of abundance of demersal fishes, with particular reference to penaeid
    prawn predators, in a tropical bay in the Gulf of Carpentaria, Australia. Marine
    Biology 107, 397 - 408.
Blaber, S.J.M., Brewer, D.T. and Salini, J.P. (1994). Comparisons of fish communities of
    tropical estuarine and inshore habitats in the Gulf of Carpentaria, northern
    Australia. In “Changes in fluxes in estuaries: Implications from science to
    management” (K.R. Dyer and R.J. Orth, eds), pp. 363-372. Fredensborg, Olsen,
    Denmark.
Blasco, F., Saenger, P. and Janodet, E. (1996). Mangroves as indicators of coastal change.
    Catena 27 (3-4) 167-178.
Blasco, F., Guaquelin, T., Rasolofoharinoro, M., Denis, J., Aizpuru, M. and Caldairou, V.
    (1998). Recent advances in mangrove studies using remote sensing data. Marine
    and Freshwater Research 49 (4), 287-296.
Boeer, 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. Marine Pollution Bulletin 27, 207-211.
Boeer, B. (1996a). Increased soil temperatures in salt marshes and mangroves after the
    1991 Gulf War Oil Spill. Fresenius Environmental Bulletin 5 (7-8), 442-447.
Boeer, B. (1996b). Impact of a major oil spill off Fujairah. Fresenius Environmental
    Bulletin 5 (1-2), 7-12.
Bonde, S.D. (1991). Significance of mangrove and other coastal plants from the Tertiary
    sediments of India. In “Proceedings of the Symposium on significance of
    Mangroves” (A.D. Agate, S.D. Bonde and K.P.N. Kumaran, eds), pp. 39-46.
    Maharastra Association for the Cultivation of Science Research Institute, Pune,
    India.
Bose, A.K., Urbanczyk-Lipkowska, Z., Subbaraju, G.V., Manhas, M.S. and Ganguly, S.N.
    (1992). An unusual secondary metabolite from an Indian mangrove plant,
    Sonneratia acida Linn. f. In “Oceanography of the Indian Ocean” (B.N. Desai, ed.),
    pp. 407-411. Oxford and IBH, New Delhi.
Boto, K.G. and Robertson, A.I. (1990). The relationship between nitrogen fixation and
    tidal exports of nitrogen in a tropical mangrove system. Estuarine, Coastal and
    Shelf Science 31 (5), 531-540.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       89


Bouchon, C., Bouchon-Navaro,Y., Imbert, D. and Louis, M. (1991). The effect of
    Hurricane Hugo on the coastal environment of Guadeloupe Island (FWI). Annals of
    the Institute of Oceanography, Paris 67 (1), 5-33.
Boulon, R.H. Jr. (1992). Use of mangrove prop root habitats by fish in the northern U.S.
    Virgin Islands. In “Proceedings of the Forty first Annual Gulf and Caribbean
    Fisheries Institute, Curacao” (M.H. Goodwin, S.M. Kau and G.T. Waugh, eds),
    Vol. 41, pp. 189-204. Fisheries Institute, St. Thomas, United States Virgin Islands.
Bremer, G.B. (1995). Lower marine fungi (Labyrinthulomycetes) and the decay of
    mangrove leaf litter. Hydrobiologia 295 (1-3), 89-95.
Brewer, D.T. and Warburton, K. (1992). Selection of prey from a seagrass/mangrove
    environment by golden lined whiting, Sillago analis (Whitley). Journal of Fish
    Biology 40 (2), 257-271.
Brewer, D.T., Blaber, S.J.M. and Salini, J.P. (1991). Predation on penaeid prawns by fishes
    in Albatross Bay, Gulf of Carpentaria. Marine Biology 109, 231 - 240.
Bruce, A.J. (1991). The "African" shrimp genus Potamalpheops in Australia, with the
    description of Potamalpheops hanleyi, new species (Decapoda: Alpheidae).
    Journal of Crustacean Biology 11 (4), 629-638.
Buden, D.W. (1992). The birds of Long Island, Bahamas. Wilson Bulletins 104 (2), 220-
    243.
Bunt, J. S. (1992). Introduction. In “Tropical mangrove ecosystem ” (A.I. Robertson and
    D.M. Alongi, eds), pp. 1-6. American Geophysical Union, Washington DC., USA.
Bunt, J.S. (1995). Continental scale pattern in mangrove litter fall. Hydrobiologia 295 (1-
    3), 135-140.
Bunt, J.S. (1996). Mangrove zonation: An examination of data from seventeen riverine
    estuaries in tropical Australia. Annals of Botany 78 (3), 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 Bunt, E.D. (1999). Complexity and variety of zonal pattern in the
    mangroves of the Hinchinbrook area, Northeastern Australia. Mangroves and Salt
    Marshes 3, 165-176.
Bunt, J.S. and Stieglitz, T. (1999). Indicators of mangrove zonality: the Normanby River,
    N.E. Australia. Mangroves and Salt Marshes, 3, 177-184.
Bunt, J.S., Williams, W.T., Hunter, J.F. and Clay, H.J. (1991). Mangrove sequencing:
    Analysis of zonation in a complete river system. Marine Ecology Progress Series
    72 (3), 289-294.
Burnham, R.J. (1990). Paleobotanical implications of drifted seeds and fruits from modern
    mangrove litter, Twin Cays, Belize Palaios 5 (4), 364-370.
Burns, K.A., Garrity, S.D., Jorissen, D., MacPherson, J., Stoelting, M., Tierney, J. and
    Yelle-Simmons, L. (1994). The Galeta oil spill; II, Unexpected persistence of oil
    trapped in mangrove sediments. Estuarine, Coastal and Shelf Science 38 (4), 349-
    364.
Buskey, E.J., Peterson, J.O. and Ambler, J.W. (1995). The role of photoreception in the
    swarming behavior of the copepod Dioithona oculata. Marine and Freshwater
    Behaviour Physiology 26, 273-285.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       90


Buskey, E.J., Peterson, J.O. and Ambler, J.W. (1996). The swarming behaviour of the
    copepod Dioithona oculata: In situ and laboratory studies. Limnology and
    Oceanography 41 (3), 513-521.
Cahoon, D.R. and Lynch, J.C. (1997). Vertical accretion and shallow subsidence in a
    mangrove forest of Southwestern Florida, USA Mangroves and Salt Marshes 1,
    173-186.
Calder, D.R. (1991). Vertical zonation of the hydroid Dynamena crisioides (Hydrozoa,
    Sertulariidae) in a mangrove ecosystem at Twin Cays, Belize. Canadian Journal of
    Zoology 69 (12), 2993-2999.
Calderon, D.G. and Echeverri, B.R. (1997). Obtaining of Rhizophora mangle seedlings by
    stimulation of adventitious roots using an air layering technique. In “ Mangrove
    Ecosystem Studies in Latin America and Africa” (B. Kjerfve, L.D. Lacerda and S.
    Diop, eds), pp. 98-107. UNESCO, Paris.
Camilleri, J.C. (1992). Leaf-litter processing by invertebrates in a mangrove forest in
    Queensland. Marine Biology 114 (1), 139-145.
Cannicci, S., Ritossa, S., Ruwa, R.K. and Vannini, M. (1996a). Tree fidelity and hole
    fidelity in the tree crab Sesarma leptosoma (Decapoda: Grapsidae). Journal of
    Experimental Marine Biology and Ecology 196 (1-2), 299-311.
Cannicci, S., Ruwa, R.K., Ritossa, S., and Vannini, M. (1996b). Branch-fidelity in the tree
    crab Sesarma leptosoma (Decapoda, Grapsidae). Journal of Zoology 238 (4), 795-
    801.
Cannicci, S., Dahdouh, G.F., Anyona, D. and Vannini, M. (1996c). Natural diet and
    feeding habits of Thalamita crenata (Decapoda: Portunidae). Journal of
    Crustacean Biology 16 (4), 678-683.
Cannicci, S., Ruwa, R.K., Giuggioli, M. and Vannini, M. (1998). Predatory activity and
    spatial strategies of Expixanthus dentatus (Decapoda: Ozidae), an ambush predator
    among the mangroves. Journal of Crustacean Biology 18 (1), 57-63.
Cardona, P. and Botero, L. (1998). Soil characteristics and vegetation structure in a heavily
    deteriorated mangrove forest in the Caribbean coast of Colombia. Biotropica 30
    (1), 24-34.
Chaghtai, F. and Saifullah, S.M. (1992). First recorded bloom of Navicula bory in a
    mangrove habitat of Karachi. Pakistan Journal of Marine Sciences 1 (2), 139-140.
Chakraborti, K. (1995). Generic and species diversity of animal vegetation dynamics of
    Sunderban mangroves, South Bengal laterite tracts of West Bengal and North
    Bengal forests - an ecological study. The Indian Forester 112 (5), 407-416.
Chakraborty, S.K. and Choudhury, A. (1992). Population ecology of fiddler crabs (Uca
    spp.) of the mangrove estuarine complex of Sunderbans, India. Tropical Ecology 33
    (1), 78-88.
Chale, F.M.M. (1993). Degradation of mangrove leaf litter under aerobic conditions.
    Hydrobiologia 257 (3), 177-183.
Chale, F.M.M. (1996). Litter production in an Avicennia germinans (L.) stern forest in
    Guyana, South America. Hydrobiologia 330 (1), 47-53.
Chan, H.T. (1996). Mangrove reforestation in peninsular Malasyia:
    a case study of Matang. In “Restoration of Mangrove Ecosystems”, (C.
    Field, ed.), International Tropical Timber Organization and International Society
    for Mangrove Ecosystems” pp. 64-76. Okinawa, Japan.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      91


Chanas, B. and Pawlik, J.R. (1995). Defenses of Caribbean sponges against predatory reef
    fish. 2. Spicules, tissue toughness and nutritional quality. Marine Ecology Progress
    Series 127 (1-3), 195-211.
Chandrika, V., Nair, P.V.R. and Khambhadkar, L.R. (1990). Distribution of phototrophic
    thionic bacteria in the anaerobic and micro-aerophilic strata of mangrove ecosystem
    of Cochin. Journal of Marine Biological Association of India 32 (1-2), 77-84.
Chandrasekaran, V.S. and Natarajan, R. (1993). Mullet seed resources of Pichavaram
    mangrove, southeast coast of India. Journal of Marine Biological Association of
    India 35 (1-2), 167-174.
Chandrasekaran, V.S. and Natarajan, R. (1994). Seasonal abundance and distribution of
    seeds of mud crab Scylla serrata in Pichavaram mangrove, southeast India. Journal
    of Aquaculture Tropica 9 (4), 343-350.
Chapman, V.J. (1976). Mangrove Vegetation. J. Cramer, Vaduz.
Cheeseman, J.M. (1994). Depressions of photosynthesis in mangrove
    canopies. In “Photoinhibition of Photosynthesis: From Molecular Mechanisms
    to the Field” (N.R. Baker and J.R. Bowyer, eds), pp. 377-389. Bios Scientific
    Publishers, Oxford
Cheeseman, J.M., Clough, B.F., Carter, D.R., Lovelock, C.E., Eong, O.J. and Sim, R.G.
    (1991). The analysis of photosynthetic performance in leaves under field
    conditions: A case study using Bruguiera mangroves. Photosynthesis Research 29
    (1), 11-22.
Cheeseman, J.M., Herendeen, L.B., Cheeseman, A.T. and Clough, B.F. (1997).
    Photosynthesis and photoprotection in mangroves under field conditions. Plant Cell
    and Environment 20 (5), 579-588.
Chen, R. and Twilley, R.R. (1998). A gap dynamic model of mangrove forest development
    along gradients of soil salinity and nutrient resources. Journal of Ecology 86 (1),
    37-51.
Chew, S.F. and Ip, Y.K. (1990). Differences in the responses of two mudskippers,
    Boleophthalmus boddaerti and Periophthalmus chrysospilos to changes in salinity.
    Journal of Experimental Zoology 256 (2), 227-231.
Chinnaraj, S. (1992). Higher marine fungi of Lakshadweep Islands and a note on Quintaria
    lignatilis. Cryptogamie Mycologie 13 (4), 313-319.
Chinnaraj, S. (1993a). Higher marine fungi from mangroves of Andaman and Nicobar
    Islands. Sydowia 45 (1), 109-115.
Chinnaraj, S. (1993b). Manglicolous fungi from atolls of Maldives, Indian Ocean. Indian
    Journal of Marine Sciences 22 (2), 141-142.
Chiu, C.Y. and Chou, C.H. (1991). The distribution and influence of heavy metals in
    mangrove forests of the Tamshui Estuary in Taiwan. Soil Science and Plant
    Nutrition 37, 659-669.
Chong, V.C., Sasekumar, A., Leh, M.U.C. and D' Cruz, R. (1990). The fish and prawn
    communities of a Malaysian coastal mangrove system, with comparisons to
    adjacent mud flats and inshore waters. Estuarine, Coastal and Shelf Science 31 (5),
    703-722.
Chong, V.C., Sasekumar, A. and Lim, K.H. (1994). Distribution and abundance of prawns
    in a Malaysian mangrove system. In “Proceedings of Third ASEAN-Australian
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       92


    Symposium on Living Coastal Resources” (S. Sudara, C.R. Wilkinson and I.M.
    Chou, eds), Vol. 2, pp. 437-444. Ihula long korn University, Bangkok, Thailand.
Chong, V.C., Sasekumar, A. and Wolanski, E. (1996). The role of mangroves in retaining
    penaeid prawn larvae in Klang Strait Malaysia. Mangroves and Salt Marshes 1 (1),
    11-22.
Choong, E.T., Wirakusumah, R.S. and Achmadi, S.S. (1990). Mangrove forest resources in
    Indonesia. Forest Ecology and Management 33/34, 45-57.
Choong, M.F., Lucas, P.W., Ong, J.S.Y., Pereira, B., Tan, H.T.W. and Turner, I.M. (1992).
    Leaf fracture toughness and sclerophylly: Their correlations and ecological
    implications. New Phytologist 121 (4), 597-610.
Choudhuri, P.K.R. (1991). Biomass production of mangrove plantation in Sunderbans,
    West Bengal (India) - a case study. The Indian Forester 117 (1), 3-12.
Chowdhury, R.A. and Ahmed, I. (1994). History of forest management. In
    “Mangroves of the Sundarbans, Volume 2: Bangladesh” (Z. Hussain and G.
    Acharya, eds), pp. 155-180. IUCN Wetlands Programme, Switzerland.
Clark, M.W., McConchie, D., Saenger, P. and Pillsworth, M. (1997). Hydrological controls
    on copper, cadmium, lead and zinc concentrations in an anthropogenically polluted
    mangrove ecosystem, Wynnum, Brisbane, Australia. Journal of Coastal Research
    13 (4), 1150-1158.
Clark, M.W., McConchie, D., Lewis, D.W. and Saenger, P. (1998). Redox stratification
    and heavy metal partitioning in Avicennia-dominated mangrove sediments; a
    geochemical model. Chemical Geology 149, (3-4) 147-171.
Clarke, P.A. and Garraway, E. (1994). Development of nests and composition of colonies
    of Nasutitermes nigriceps (Isoptera: Termitidae) in the mangroves of Jamaica.
    Florida Entomologist 77 (2), 272-280.
Clarke, P.J. (1993). Dispersal of grey mangrove (Avicennia marina) propagules in
    southeastern Australia. Aquatic Botany 45, 195-204.
Clarke, P.J. (1994). Baseline studies of temperate mangrove growth and reproduction:
    Demographic and litterfall measures of leafing and flowering. Australian Journal
    of Botany 42 (1), 37-48.
Clarke, P.J. and Allaway, W.G. (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, P.J. and Myerscough, P.J. (1991a). Floral biology and reproductive phenology of
    Avicennia marina in south-eastern Australia. Australian Journal of Botany 39, 283-
    293.
Clarke, P.J. and Myerscough, P.J. (1991b). Buoyancy of Avicennia marina (Forssk.) Vierh.
    and Rhizophora stylosa Griff. in relation to salinity. Australian Journal of Plant
    Physiology 11, 419-430.
Clarke, P.J. and Myerscough, P.J. (1993). The intertidal distribution of the grey mangrove
    (Avicennia marina) in southeastern Australia: The effects of physical conditions,
    interspecific competition and predation on propagule establishment and survival.
    Australian Journal of Ecology 18 (3), 307-315.
Clay, R.E. and Andersen, A.N. (1996). Ant fauna of a mangrove community in the
    Australian seasonal tropics, with particular reference to zonation. Australian
    Journal of Zoology 44 (5), 521-533.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      93


Clayton, D.A. (1993). Mudskippers. Oceanography and Marine Biology Annual Review
    31, 507-577.
Clough, B.F. (1992). Primary productivity and the growth of mangrove forests. In
    “Tropical mangrove ecosystems” (A.I. Robertson and D.M. Alongi, eds), pp. 225-
    250. American Geophysical Society, Washington DC., USA
Collado-Vides, L. and West, J.A. (1996). Bostrychia calliptera (Montagne) Montagne
    (Rhodomelaceae: Rhodophyta), a new record for the central Gulf of Mexico.
    Ciencias Marinas 22 (1), 47-55.
Colombini, I., Berti, R., Ercolini, A. and Nocita, A. (1995). Environmental factors
    influencing the zonation and activity patterns of a population of Periophthalmus
    sobrinus Eggert in a Kenyan mangrove. Journal of Experimental Marine Biology
    and Ecology 190 (1), 135-149.
Colombini, I., Berti, R., Nocita, A. and Chelazzi, L. (1996). Foraging strategy of the
    mudskipper Periophthalmus sobrinus Eggert in a Kenyan mangrove. Journal of
    Experimental Marine Biology and Ecology 197 (2), 219-235.
Conacher, C.A., O’Brien, C., Horrocks, J.L. and Kenyon, R.K. (1996). Litter production
    and accumulation in stressed mangrove communities in the Embley river estuary,
    North eastern Gulf of Carpentaria, Australia. Marine and Freshwater Resources 47,
    737-743.
Conde, J.E. and Alarcón, C. (1993). Mangroves of Venezuela. In “Conservation
    and Sustainable Utilization of Mangrove Forests in Latin America and Africa
    Regions” (L.D. Lacerda, ed), pp. 211-243. International Society for Mangrove
    Ecosystems, Okinawa.
Conde, J.E. and Diaz, H. (1992). Extension of the stunting range in ovigerous females of
    the mangrove crab Aratus pisonii (H. Milne Edwards, 1837) (Decapoda:
    Brachyura: Grapsidae). Crustaceana 62 (3), 319-322.
Conde, J.E., Alarcón, C., Flores, S. and Diaz, H. (1995). Nitrogen and tannins in mangrove
    leaves might explain interpopulation variations in the crab Aratus pisonii. Acta
    Cintifica de Venezuela 46, 303-304.
Confer, J.L. and Holmes, R.T. (1995). Neotropical migrants in undisturbed and human-
    altered forests of Jamaica. Wilson Bulletin 107 (4), 577-589.
Conti, E., Litt, A. and Sytsma, K.J. (1996). Circumscription of Myrtales and their
    relationships to other rosids: evidence from rbcL sequence data. American Journal
    of Botany 83, 221-233.
Cook, L.M. (1990). Systematic effects on morph frequency in the polymorphic mangrove
    snail Littoraria pallescens. Heredity 65 (3), 423-427.
Cook, L.M. and Kenyon, G. (1993). Shell strength of colour morphs of the mangrove snail
    Littoraria pallescens. Journal of Molluscan Studies 59 (1), 29-34.
Coppejans, E., Beeckman, H. and Wit, M. (1992). The seagrass and associated macroalgal
    vegetation of Gazi Bay, Kenya. Hydrobiologia 247 (1-3), 59-75.
Corredor, J.E. and Morell, J.M. (1994). Nitrate depuration of secondary sewage effluents
    in mangrove sediments. Estuaries 17 (1B), 295-300.
Corredor, J.E., Morell, J.M., Klekowski, E.J. and Lowenfeld, R. (1995). Mangrove
    genetics: III. Pigment fingerprints of chlorophyll-deficient mutants. International
    Journal of Plant Sciences 156 (1), 55-60.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       94


Cortes-Lopes, B., Aguiar dos Santos, R. and Dos Santos, R.A. (1996). Aspects of the
    ecology of ants (Hymenoptera: Formicidae) on the mangrove vegetation of Rio
    Ratones, Santa Catarina Island, SC, Brazil. Boletin de Entomologia Venezolana 11
    ( 2), 123-133.
Costa, C.S.B. and Davy, A.J. (1992). Coastal salt marsh communities of Latin America. In
    “Coastal Plant Communities of Latin America” (U. Seeliger, ed), pp. 179-199.
    Academic Press, San Diego, USA.
Crane, E., Luyen, V., Mulder, V. and Tran Cong, T. (1993). Traditional management
    system for Apis dorsata in submerged forests in southern Vietnam and central
    Kalimantan. Bee World 74 (1), 27-40.
Cronin, T.M., Bybell, L.M., Brouwers, E.M., Gibson, T.G., Margerumer, R. and Poore,
    R.Z. (1991). Neocene biostratigraphy and paleoenvironments of Enewetak Atoll,
    equatorial Pacific Ocean. Marine Micropaleontology 18 (1-2), 101-114.
Crow, T. (1996). Different effects of microhabitat fragmentation on patterns of dispersal of
    an intertidal gastropod in two habitats. Journal of Experimental Marine Biology
    and Ecology 206 (1-2), 83-107.
Curnutt, J.L. and Robertson, W.B., Jr. (1994). Bald eagle nest site characteristics in south
    Florida. Journal of Wildlife Management 58 (2), 218-221.
Dagar, J.C. and Sharma, A.K. (1991). Litterfall beneath Rhizophora apiculata mangrove
    forests of Andamans, India. Tropical Ecology 32 (2), 231-235.
Dagar, J.C. and Sharma, A.K. (1993). Litterfall beneath Bruguiera gymnorrhiza in
    mangrove forest of South Andamans, India. Indian Journal of Forestry 16, 157-
    161.
Dahdouh-Guebas, F., Verneirt, M., Tack, J.F. and Koedam, N. (1997). Food preferences of
    Neosarmatium meinerti de Man (Decapoda: Sesarminae) and its possible effect on
    the regeneration of mangroves. Hydrobiologia 347 (1-3), 83-89.
Dahdouh-Guebas, F., Verneirt, M., Tack, J.F., Van Speybroeck, D., and Koedam, N.
    (1998). Propagule predators in Kenyan mangroves and their possible effect on
    regeneration. Marine and Freshwater Research 49 (4), 345-350
Daniel, P.A. and Robertson, A.I. (1990). Epibenthos of mangrove waterways and open
    embayments: Community structure and the relationship between exported
    mangrove detritus and epifaunal standing stocks. Estuarine, Coastal and Shelf
    Science 31 (5), 599-619.
Danielsen, F., Kadarisman, R., Skov, H., Suwarman, U. and Verheugt, W.J.M. (1997). The
    Storm's stork Ciconia stormi in Indonesia: Breeding biology, population and
    conservation. Ibis 39 (1), 67-75.
Das, S. and Ghose, M. (1993). Morphology of stomata and leaf hairs of some halophytes
    from Sunderbans, West Bengal. Phytomorphology 43 (1-2), 59-70.
Das, S. and Ghose, M. (1998). Anatomy of the woods of some mangroves of Sunderbans,
    West Bengal (India). In “ International Symposium on Mangrove Ecology and
    Biology”, April 25-27, 1998, Kuwait. 10p. Abstracts.
Das, A.B., Basak, U.C. and Das, P. (1994). Karyotype diversity in three species of
    Heritiera, a common mangrove tree on the Orissa coast. Cytobios 80 (321), 71-78.
Das, P.K., Chakravarti, V., Dutta, A. and Maity, S. (1995). Leaf anatomy and chlorophyll
    estimates in some mangroves. The Indian Forester 121 (4), 289-294.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       95


Das, S., Ghose, M. (1996). Anatomy of leaves of some mangroves and their associates of
    Sunderbans, West Bengal. Phytomorphology 46 ( 2), 139-150.
Das, P., Basak, U.C. and Das, A.B. (1997). Metabolic changes during rooting in pre-
    girdled stem cuttings and air-layers of Heritiera. Botanical Bulletin of Academia
    Sinica (Taipei) 38 (2), 91-95.
Da Silva, J.A.A., De Melo, M.R.C.S. and Borders, B.E. (1993). A volume equation for
    mangrove trees in northeast Brazil. Forest Ecology and Management 58, 129-136.
Davis, R.A. Jr. (1995). Geologic impact of Hurricane Andrew on Everglades coast of
    southwest Florida. Environmental Geology 25 (3), 143-148.
Davis, W.P., Thornton, K.W. and Levinson, B. (1994). Framework for assessing effects of
    global climate change on mangrove ecosystems. Bulletin of Marine Science 54 (3),
    1045-1058.
Davis, W.P., Taylor, D.S. and Turner, B.J. (1995). Does the autecology of the mangrove
    rivulus fish (Rivulus marmoratus) reflect a paradigm for mangrove ecosystem
    sensitivity? Bulletin of Marine Science 57 (1), 208-214.
Deekae, S.N. and Idoniboye-Obu, T.I.E. (1995). Ecology and chemical composition of
    commercially important molluscs and crabs of the Niger Delta, Nigeria.
    Environmental Ecology 13 (1), 136-142.
de Graaf, G.J. and Xuan, T.T. (1998). Extensive shrimp farming, mangrove clearance and
    marine fisheries in the southern provinces of Vietnam. Mangroves and Salt
    Marshes 2, 159-166.
De Jesus, B.R. Jr. and Bina, R.T. (1998). Mangrove management in the Philippines using
    remote sensing. Tropical Coastal Area Management 4 (3), 8-11.
De Lange, W.P. and De Lange, P.J. (1994). An appraisal of factors controlling the
    latitudinal distribution of mangrove (Avicennia marina var. resinifera) in New
    Zealand. Journal of Coastal Research 10 (3), 539-548.
Dennis, G.D. (1992). Island mangrove habitats as spawning and nursery areas for
    commercially important fishes in the Caribbean. In “Proceedings of the Forty First
    Annual Gulf and Caribbean” (M.H. Goodwin, S.M. Kau and G.T. Waugh, eds),
    Vol. 41, pp. 205-225. Fisheries Institute, St. Thomas, United States Virgin Islands.
Devoe, N.N. and Cole, T.G. (1998). Growth and yield in mangrove forests of the Federated
    State of Micronesia. Forest Ecology and Management 103 (1), 33-48.
de Weerdt, W.H., Rützler, K. and Smith, K.P. (1991). The chalinidae (Porifera) of Twin
    Cays, Belize, and adjacent waters. Proceedings of the Biological Society of
    Washington 104, 189-205.
Diaz, H., Orihuela, B., and Forward, R.B., Jr. (1995). Visual orientation of postlarval and
    juvenile mangrove crabs. Journal of Crustacean Biology 15 (4), 671-678.
Diaz, R.J. and Erseus, C. (1994). Habitat preferences and species associations of shallow-
    water marine Tubificidae (Oligochaeta) from the barrier reef ecosystems of Belize,
    Central America. Hydrobiologia 278 (1-3), 93-105.
Diop, E.S., Soumare, A., Diallo, N. and Guisse, A. (1997). Recent changes of the
    mangroves of the Saloum river estuary, Senegal. Mangroves and Salt Marshes 1,
    163-172.
Dious, S.R.J. and Kasinathan, R. (1994). Tolerance limits of two pulmonate snails
    Cassidula nucleus and Melampus ceylonicus from Pitchavaram mangroves.
    Environment and Ecology 12 (4), 845 - 849.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       96


Di-Toro, D.M. (1990). Toxicity of the cadmium in sediments: the role of acid volatile
    sulfides. Environmental Toxicological Chemistry 9, 1487-1502.
Dittel, A.I. and Epifanio, C.E. (1990). Seasonal and tidal abundance of crab larvae in
    tropical mangrove systems, Gulf of Nicoya, Costa Rica. Marine Ecology Progress
    Series 65, 25-34.
Dodd, R.S., Fromard, F., Rafii, Z.A. and Blasco, F. (1995). Biodiversity among West
    African, Rhizophora: Foliar wax chemistry. Biochemical Systematics and Ecology
    23 (7-8), 859-868.
Dodd, R.S., Rafiik, Z.A., Fromard, F. and Blasco, F. (1998). Evolutionary diversity among
    Atlantic coast mangroves. Acta Oecologica 19 (3), 323-330.
Dominguez, C.A., Eguiarte, L.E., Nunez-Farfan, J. and Dirzo, R. (1998). Flower
    morphometry of Rhizophora mangle (Rhizophoraceae): Geographical variation in
    Mexican populations. American Journal of Botany 85 (5), 637-643.
Doyle, T.W., Smith, T.J., and Robblee, M.B. (1995). Wind damage effects of Hurricane
    Andrew on mangrove communities along the southwest coast of Florida. Journal of
    Coastal Research 21, 159-168.
Drennan, P.M., Berjak, P. and Pammenter, N.W. (1992). Ion gradients and adenosine
    triphosphatase localization in the salt glands of Avicennia marina (Forssk.) Vierh.
    South African Journal of Botany 58 (6), 486-490.
Dschida, W., Platt-Aloia, K. and Thomson, W. (1992). Epidermal peels of Avicennia
    germinans (L.) Stern: a useful system to study the function of salt glands. Annals of
    Botany 70 (6), 501-509.
Duarte, C.M. and Cebrian, J. (1996). The fate of marine autotrophic production. Limnology
    and Oceanography 41 (8), 1758-1766.
Duke, N.C. (1990). Phenological trends with latitude in the mangrove tree Avicennia
    marina. Journal of Ecology 78, 113-133.
Duke, N.C. (1991). A systematic revision of the mangrove genus Avicennia
    (Avicenniaceae) in Australia. Australian Journal of Botany 27, 657-678.
Duke, N.C. (1992). Mangrove floristics and biogeography. In “Tropical
    Mangrove Ecosystems” (A.I. Robertson and D.M. Alongi, eds), pp.63-100.
    American Geophysical Union, Washington DC., USA.
Duke, N.C. (1995). Genetic diversity, distributional barriers and rafting continents - more
    thoughts on the evolution of mangroves. Hydrobiologia 295 (1-3), 167-181.
Duke, N.C. and Pinzon, S.M. (1992). Ageing Rhizophora seedlings from leaf scar nodes: A
    technique for studying recruitment and growth in mangrove forests. Biotropica 24
    (29), 173-186.
Duke, N.C., Pinzon, M.Z.S. and Prada, T.M.C. (1997). Large-scale damage to mangrove
    forests following two large oil spills in Panama. Biotropica 29 (1), 2-14.
Duke, N.C., Ball, M.C. and Ellison, J.C. (1998a). Factors influencing biodiversity and
    distributional gradients in mangroves. Global Ecology and Biogeography Letters 7,
    27-47.
Duke, N.C., Benzie, J.A.H., Goodall, J.A. and Ballment, E.R. (1998b). Genetic structure
    and evolution of species in the Mangrove genus Avicennia (Avicenniaceae) in the
    Indo-West Pacific. Evolution 52 (6), 1612-26.
Dunlap, M. and Pawlik, J.R. (1996). Video-monitored predation by Caribbean reef fishes
    on an array of mangrove and reef sponges. Marine Biology 126 (1), 117-123.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       97


Durand, P., Gros, O., Frenkiel, L. and Prieur, D. (1996). Phylogenetic characterization of
    sulfur-oxidizing bacterial endosymbionts in three tropical Lucinidae by 16S rDNA
    sequence analysis. Molecular Marine Biology and Biotechnology 5 (1), 37-42.
Dutrieux, E., Martin, F. and Debry, A. (1990). Growth and mortality of Sonneratia
    caseolaris planted on an experimentally oil-polluted soil. Marine Pollution Bulletin
    21 (20), 62-68.
Edgar, G.J. (1990). The influence of plant structure on the species richness, biomass and
    secondary production of macrofaunal assemblages associated with Western
    Australian sea grass beds. Journal of Experimental Marine Biology and Ecology
    137, 215-240.
Eguchi, F., Takei, T., Iijima, T. and Higaki, M. (1995). Preparation of protoplasts from the
    mesophyll of Bruguiera gymnorrhiza (L.) Lamk. Journal of the Japan Wood
    Research Society 41 (10), 932-937.
Eguchi, M., Rungsupa, S., Kawai, A. and Menasveta, P. (1997). Dissolved oxygen
    consumption by bottom sediments of shrimp pond and mangrove forest in
    Thailand. Fisheries Science 63 (3), 480-481.
ElAmry, M. (1998). Population structure, demography and life tables of Avicennia marina
    (Forssk.) Vierch. at sites on the eastern and western coasts of the United Arab
    Emirates. Marine and Freshwater Research 49 (4) 303-308.
Ellison, A.M. (1999). Cumulative effects of oil spills on mangroves. Ecological
    Applications 9 (4), 1490-1492.
Ellison, A.M and Farnsworth, E.J. (1990). The ecology of Belizean mangrove-root fouling
    communities: I. Epibenthic fauna are barriers to isopod attack of red mangrove
    roots. Journal of Experimental Marine Biology and Ecology 142, 91-104.
Ellison, A.M. and Farnsworth, E.J. (1992). The ecology of Belizean mangrove root-fouling
    communities: patterns of epibiont distribution and abundance and effects on root
    growth. Hydrobiologia 247, 87-98.
Ellison, A.M. and Farnsworth, E.J. (1993). Seedling survivorship, growth and response to
    disturbance in Belizean mangal. American Journal of Botany 80 (10), 1137-1145.
Ellison, A.M. and Farnsworth, E.J. (1996a). Anthropogenic disturbance to Caribbean
    mangrove ecosystems: Past impacts, present trends and future predictions.
    Biotropica 28 (4A), 549-565.
Ellison, A.M. and Farnsworth, E.J. (1996b). Spatial and temporal variability in growth of
    Rhizophora mangle saplings on coral cays: Links with variation in insolation,
    herbivory, and local sedimentation rate. Journal of Ecology 84 (5), 717-731.
Ellison, A.M. and Farnsworth, E.J. (1997). Simulated sea-level change alters anatomy,
    physiology, growth, and reproduction of red mangrove (Rhizophora mangle L.).
    Oecologia 112 (4), 435-446.
Ellison, A.M. and Farnsworth, E.J. (2000). Mangrove communities. In “Marine
    Community Ecology” (M.D. Bertness, S.D. Gaines and M.E. Hay, eds.), Sinauer
    Associates, Sunderland, MA, USA. (in press).
Ellison, A.M., Farnsworth, E.J. and Twilley, R.R. (1996). Facultative mutualism between
    red mangroves and root fouling sponges in Belizean mangal. Ecology 77 (8), 2431-
    2444.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       98


Ellison, A.M., Farnsworth, E.J. and Mertkt, R.E. (1999). Origins of mangrove ecosystems
    and the mangrove biodiversity anomaly. Global Ecology and Biogeography 8, 95-
    115.
Ellison, A.M., Mukherjee, B.B. and Karim, A. (2000), Testing patterns of zonation in
    mangroves: scale-dependence and environmental correlates in the Sunderbans of
    Bangladesh. Journal of Ecology (in press).
Ellison, J.C. (1993). Mangrove retreat with rising sea-level, Bermuda. Estuarine, Coastal
    and Shelf Science 37 (1), 75-87.
Ellison, J.C. (1994). Climate change and sea level rise impacts on mangrove ecosystems.
    In “Impacts of Climate Change on Ecosystems and Species: Marine and Coastal
    Ecosystems” (J. Pernetta, R. Leemans, D. Elder and S. Humphrey, eds.), pp. 11-30.
    IUCN, Gland.
Ellison, J.C. (1996). Pollen evidence of late Holocene mangrove development in Bermuda.
    Global Ecology and Biogeography Letters 5 (6), 315-326.
Ellison, J.C. (1997). Mangrove community characteristics and litter production in
    Bermuda. In “Mangrove Ecosystem Studies in Latin America and Africa” (B.
    Kjerfve, L.D. Lacerda and S. Diop, eds), pp. 8-17. UNESCO, Paris.
Ellison, J.C. and Stoddart, D.R. (1991). Mangrove ecosystem collapse during predicted
    sea-level rise: Holocene analogues and implications. Journal of Coastal Research
    7, 151-165.
Emilio, O. (1997). Majagual: The tallest mangroves in the world. International News
    Letter of Coastal Management-Intercoast Network, Special edition 1, 1-17.
Emmerson, W.D. and McGwynne, L.E. (1992). Feeding and assimilation of mangrove
    leaves by the crab Sesarma meinertii de Man in relation to leaf litter production in
    Mgazana, a warm-temperature southern African mangroves swamp. Journal of
    Experimental Marine Biology and Ecology 157, 41-53.
Erondu, E.S. (1990). The diet of wild and pond-cultured catfish Chrysichthys
    nigrodigitatus Bagridae, in mangrove swamps of the Niger Delta, Nigeria. Journal
    of African Zoology 104 (5), 367-374.
Eshky, A.A., Atkinson, R.J.A. and Taylor, A.C. (1995). Physiological ecology of crabs
    from Saudi Arabian mangrove. Marine Ecology Progress Series 126 (1-3), 83-95.
Espinosa, L.F., Ramirez, G. and Campos, N.H. (1995). Analisis de residuos de
    organoclorados en los sedimentos de zonas de manglar en la Cienaga Grande de
    Santa Marta la Bahia de Chengue, Caribe Colombiano. Anales del Instituto de
    Investigaciones Marinas de Punta de Betin. 24, 79-94.
Eston, V.R., Braga, M.R.A., Cordeiro Marino, M., Fujii, M.T. and Yokoya, N.S. (1992).
    Macroalgal colonization patterns on artificial substrates inside southeastern
    Brazilian mangroves. Aquatic Botany 42 (4), 315-325.
Ewa-Oboho, I.O. and Abby-Kalio, N.J. (1993). Seasonal variation and community
    structure of epibenthic algae on the roots of the mangrove Rhizophora mangle at a
    shallow tropical estuary. Tropical Ecology 34 (2), 160-172.
Ewel, K.C., Zheng, S., Pinzón, A.S. and Bourgeois, J.A. (1998a). Environmental effects of
    canopy gap formation in high-rainfall mangrove forests. Biotropica 30 (4), 510-
    518.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      99


Ewel, K.C., Twilley, R.R. and Ong, J.E. (1998b). Different kinds of mangrove forests
    provide different goods and services. Global Ecology and Biogeography Letters 7
    (1), 83-94.
Fan, H.Q. and Chen. J. (1993). Impacts of light on rooting of mangrove Kandelia candel
    propagules. Journal of the Guangxi Academy of Sciences 9 (2), 73-76.
Farnsworth, E.J. (1998a). Values, uses and conservation of mangroves in India: lessons for
    mangroves of the world. Environmental Awareness 21, 11-19.
Farnsworth, E.J. (1998b). Issues of spatial, taxonomic and temporal scale in delineating
    links between mangrove diversity and ecosystem function. Global Ecology and
    Biogeography Letters 7 (1), 15-25.
Farnsworth, E.J. and Ellison, A.M. (1991). Patterns of herbivory in Belizean mangrove
    swamps. Biotropica 23 (4b), 555-567.
Farnsworth, E.J. and Ellison, A.M. (1993). Dynamics of herbivory in Belizean mangal.
    Journal of Tropical Ecology 9 (4), 435-453.
Farnsworth, E.J. and A.M. Ellison. (1995). Scale-dependent spatial and temporal
    variability in biogeography of mangrove-root epibiont communities. Ecological
    Monograph 66, 45-66.
Farnsworth, E.J. and Ellison, A.M. (1996a). Scale-dependent spatial and temporal
    variability in biogeography of mangrove-root epibiont communities. Ecological
    Monographs 66 (1), 45-66.
Farnsworth, E.J. and Ellison, A.M. (1996b). Sun-shade adaptability of the Red Mangrove,
    Rhizophora mangle (Rhizophoraceae): Changes through ontogeny at several levels
    of biological organization. American Journal of Botany 83 (9), 1131-1143.
Farnsworth, E.J. and Ellison, A.M. (1997a). Global patterns of pre-dispersal propagule
    predation in mangrove forests. Biotropica 29 (3), 318-330.
Farnsworth, E.J. and Ellison, A.M. (1997b). Global conservation ecology of mangrove
    ecosystems. Ambio 26 (6), 328-334.
Farnsworth, E.J. and Farrant, J.M. (1998). Reductions in abscissic acid are linked with
    viviparous reproduction in mangroves. American Journal of Botany 85 (6), 760-
    769.
Farnsworth, E.J., Ellison, A.M. and Gong, W.K. (1996). Elevated CO2 alters anatomy,
    physiology, growth, and reproduction of red mangrove (Rhizophora mangle L.).
    Oecologia 108 (4), 599-609.
Farrant, J.M., Pammenter, N.W. and Berjak, P. (1992). Development of the recalcitrant
    (Homoiohydrous) seeds of Avicennia marina: anatomical, ultra structural and
    biochemical events associated with development from histodifferentiation to
    maturation. Annals of Botany 70 (1), 75-86.
Farrant, J.M., Pammenter, N.W. and Berjak, P. (1993). Seed development in relation to
    desiccation tolerance: a comparison between desiccation-sensitive (recalcitrant)
    seeds of Avicennia marina and desiccation-tolerant types. Seed Science Research 3
    (1), 11-13.
Faust, M.A. (1993a). Three new benthic species of Prorocentrum (Dinophyceae) from
    Twin Cays, Belize: P. maculosum sp. nov., P. foraminosum sp. nov. and P.
    formosum sp. nov.. Phycology 32 (6), 410-418.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      100


Faust, M.A. (1993b). Surface morphology of the marine dinoflagellate Sinophysis
     microcephalus (Dinophyceae) from a mangrove island, Twin Cays, Belize. Journal
     of Phycology 29 (3), 355-363.
Faust, M.A. (1993c). Sexuality in a toxic dinoflagellate, Prorocentrum lima. In “Toxic
     phytoplankton bloom in the sea” (T.J. Smayda and Y. Shimizu, eds), Vol. 3, pp.
     121-126. Amsterdam, Elsevier, Netherlands.
Faust, M.A. (1993d). Alternate asexual reproduction of Prorocentrum lima in culture. In
     “Toxic phytoplankton blooms in the sea” (T.J. Smayda and Y. Shimizu, eds), Vol.
     3, pp. 115-120. Amsterdam, Elsevier, Netherlands.
Faust, M.A. and Balech, E. (1993). A further SEM study of marine benthic dinoflagellates
     from a mangrove island, Twin Cays, Belize, including Plagiodinium belizeanum
     gen. et sp. nov.. Journal of Phycology 29 (6), 826-832.
Faust, M.A. and Gulledge, R.A. (1996). Associations of microalgae and meiofauna in
     floating detritus at a mangrove island, Twin Cays, Belize. Journal of Experimental
     Marine Biology and Ecology 197 (2), 159-175.
Fauvel, M.T., Bousquet Melou, A., Moulis, C., Gleye, J. and Jensen, S.R. (1995). Iridoid
     glycosides from Avicennia germinans. Phytochemistry 38 (4), 893-894.
Feller, I.C. (1995). Effects of nutrient enrichment on growth and herbivory of dwarf red
     mangrove (Rhizophora mangle). Ecological Monographs 65, 477-506.
Feller, I.C. (1996). Effects of nutrient enrichment on leaf anatomy of dwarf Rhizophora
     mangle L. (red mangrove). Biotropica 28 (1), 13-22.
Feller, I.C. and Mathis, W.N. (1997). Primary herbivory by wood-boring insects along an
     architectural gradient of Rhizophora mangle. Biotropica 29, 440-451.
Fernandes, M. E. B. (1991). Tool use and predation of oysters (Crassostrea rhizophorae)
     by the tufted capuchin, Cebus apella apella, in brackish-water mangrove swamp.
     Primates 32 (4), 529-531.
Ferraris, J.D., Fauchald, K. and Kensley, B. (1994). Physiological responses to fluctuation
     in temperature or salinity in invertebrates: Adaptations of Alpheus viridari
     (Decapoda, Crustacea), Terebellides parva (Polychaeta) and Golfinigia cylindrata
     (Sipunculida) to the mangrove habitat. Marine Biology 120 (3), 397-406.
Fiala, K. and Hernandez, L. (1993). Root biomass of a mangrove forest in southwestern
     Cuba (Majana). Ekologia Bratislava 12 (1), 15-30.
Field, C. (1998). Rationales and practices of mangrove afforestation. Marine and
     Freshwater Research 49, 353-358.
Field, C.B., Osborn, J.G., Hoffman, L.L., Polsenberg, J.F., Ackerly, D.D., Berry, J.A.,
     Bjoerkman, O., Held, A., Matson, P.A. and Mooney, H.A. (1998). Mangrove
     biodiversity and ecosystem function. Global Ecology and Biogeography Letters 7
     (1), 3-14.
Field, C. D. 1995. Impact of expected climate change on mangroves. Hydrobiologia 295,
     75-81.
Field, C.D. (1996). Rationale for restoration of mangrove ecosystems. In “Restoration of
     mangrove ecosystems” (C.D. Field, ed.), pp. 233-250. International Society for
     Mangrove Ecosystems, Okinawa, Japan.
Fisher, C.R. (1990). Chemoautotrophic and methanotrophic symbiosis in marine
     invertebrates. CRC Critical Reviews in Aquatic Science 2, 399-436.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      101




Fisher, M.R. and Hand, S.C. (1984). Chemoautotrophic synbionts in the bivalve Lucina
floridana from seagrass beds. Biological Bulletin, Marine Biological laboratory, Woods
Hole 167, 445-459.
Fitt, W.K. (1991). Natural metamorphic cues of larvae of a tropical jellyfish. American
    Zoology 31 (5), 106A.
Fitzgerald, M.A., Orlovich, D.A. and Allaway, W.G. (1992). Evidence that abaxial leaf
    glands are the sites of salt secretion in leaves of the mangrove Avicennia marina
    (Forsk.) Vierh. New Phytologist 120, 1-7.
Fleck, J. and Fitt, W.K. (1999). Degrading mangrove leaves of Rhizophora mangle Linne
    provide a natural cue for settlement and metamorphosis of the upside-down
    jellyfish Cassiopea xamachana Bigelow. Journal of Experimental Marine Biology
    and Ecology 234 (1), 83-94.
Fleming, M., Lin, G. and Sternberg, L. (1990). Influence of mangrove detritus in an
    estuarine ecosystem. Bulletin of Marine Science 47, 663-669.
Flores-Verdugo, F.J., Conalez-Farias, F., Amezcua, F., Yanez-Arancibia, A., Ramirez-
    Flores, O. and Day, J.W. (1990). Mangrove ecology, aquatic primary productivity
    and fish community dynamics in the Teacapan-Ague Brava lagoon-estuarine
    system (Mexican Pacific). Estuaries 13, 219-230.
Forys, E.A. and Humphrey, S.R. (1996). Home range and movements of the lower keys
    marsh rabbits in a highly fragmented habitat. Journal of Mammalogy 77, 1042-
    1048.
Foster, B.A. (1982). Two new intertidal barnacles from eastern Australia. Proceedings of
        the Linnean Society of New South Wales 106, 21-32.
Fouda, M.M. and Al-Muharrami, M. (1995). An initial assessment of mangrove resources
    and human activities at Mahout Island, Arabian Sea, Oman. Hydrobiologia 295 (1-
    3), 353-362.
Fourqurean, J.W. and Ziemean, J.C. (1991). Photosynthesis, respiration and the whole
    plant carbon budget of Thalassia testudinum. Marine Ecology Progress Series 69,
    161-170.
Frenkiel, L., Gros, O. and Moueza, M. (1996). Gill structure in Lucina pectinata (Bivalvia:
    Lucinidae) with reference to hemoglobin in bivalves with symbiotic sulphur-
    oxidizing bacteria. Marine Biology 125 (3), 511-524.
Fromard, F., Puig, H., Mougin, E., Marty, G., Betoulle, J.L. and Cadamuro, L. (1998).
    Structure, above-ground biomass and dynamics of mangrove ecosystems: New data
    from French Guiana. Oecologia 115 (1-2), 39-53.
Fujii, M.T., Yokoya, N.S. and Cordeiro-Marino, M. (1990). Stictosiphonia kelanensis, new
    record (Grunow ex post) King and Puttock (Rhodomelaceae: Rhodophyta), from
    Atlantic mangroves. Hoehnea 17 (2), 93-98.
Fujimoto, K. and Miyagi, T. (1990). Late Holocene sea level fluctuations and mangrove
    forest formation on Ponape Island, Micronesia. Journal of Geography 99 (5), 507-
    514.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      102


Fukushima,Y., Sasamoto, H., Baba, S. and Ashihara, H. (1997). The effect of salt stress on
    the catabolism of sugars in leaves and roots of a mangrove plant, Avicennia marina.
    Verlag der zeitschrift fur Naturforschung, 187-192.
Furukawa, K., Wolanski, E. and Mueller, H. (1997). Currents and sediment transport in
    mangrove forests. Estuarine, Coastal and Shelf Science 44 (3), 301-310.
Gang, P.O. and Agatsiva, J.L. (1992). The current status of mangroves along the Kenyan
    coast: A case study of Mida Creek mangroves based on remote sensing. In “The
    ecology of mangrove and related ecosystems” (Jaccarini, V. and Martens, E. eds),
    pp. 29-36. Kluwer Academic Publishers, Netherlands.
Gara, R.I., Sarango, A. and Cannon, P.G. (1990). Defoliation of an Ecuadorian mangrove
    forest by the bagworm, Oiketicus kirbyi Guidling (Lepidoptera: Psychidae).
    Journal of Tropical Forest Science 3 (2), 181-186.
Garrity, S.D., Levings, S.C. and Burns, K.A. (1994). The Galeta oil spill: I. Long-term
    effects on the physical structure of the mangrove fringe. Estuarine, Coastal and
    Shelf Science 38 (4), 327-348.
Gee, J.M. and Somerfield, P.J. (1997). Do mangrove diversity and leaf litter decay promote
    meiofaunal diversity? Journal of Experimental Marine Biology and Ecology 218
    (1), 13-33.
Gherardi, F. and Vannini, M. (1993). Hermit crabs in a mangrove swamp: proximate and
    ultimate factors in Clibanarius laevimanus clustering. Journal of Experimental
    Marine Biology and Ecology 168, 167-187.
Gherardi, F., Micheli, F. and Vannini, M. (1991). Preliminary observations on the
    clustering behaviour of the tropical hermit crab, Clibanarius laevimanus. Ethology,
    Ecology and Evolution 1, 151-153.
Gherardi, F., Zatteri, F. and Vannini, M. (1994). Hermit crabs in a mangrove swamp: the
    structure of Clibanarius laevimanus clusters. Marine Biology 121 (1), 41-52.
Ghosh, P.B., Singh, B.N., Chakrabarty, C., Saha, A., Das, R.L. and Choudhury, A. (1990).
    Mangrove litter production in a tidal creek of Lothian Island of Sunderbans, India.
    Indian Journal of Marine Sciences 19 (4), 292-293.
Giani L., Bashan Y., Holguin G. and Strangmann, A. (1996). Characteristics and
    methanogenesis of the Balandra lagoon mangrove soils, Baja California Sur,
    Mexico. Geoderma 72 (1-2), 149-160.
Giesen, W.B.J.T., Van-Katwijk, M.M. and Hartog, C. (1990). Eelgrass condition and
    turbidity in the Dutch Wadden Sea. Aquatic Botany 37, 71-85.
Gilbert, A.J. and Janssen, R. (1998). Use of environmental functions to communicate the
    values of a mangrove ecosystem under different management regimes. Ecological
    Economics 25 (3), 323-346.
Gilmore, A.M. and Bjorkman, O. (1994). Adenine nucleotides and the xanthophyll cycle in
    leaves. I. Effects of CO2- and temperature-limited photosynthesis on adenylate
    energy charge and violaxanthin de-epoxidation. Planta 192 (4), 526-536.
Godhantaraman, N. (1994). Species composition and abundance of tintinnids and copepods
    in the Pichavaram mangroves (South India). Ciencias Marinas 20 (3), 371-391.
Goh, T.K. and Yipp, M.W. (1996). In vivo and in vitro studies of three new species of
    Trimmatostroma associated with sooty spots of the mangrove Aegiceras
    corniculatum in Hong Kong. Mycological Research 100 (12), 1489-1497.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      103


Gong, W.K. and Ong, J.E. (1990). Plant biomass and nutrient flux in a managed mangrove
    forest in Malaysia. Estuarine, Coastal and Shelf Science 31 (5), 519-530.
Gong, W.K. and Ong, J.E. (1995). The use of demographic studies in mangrove
    silviculture. Hydrobiologia 295, 255-261.
Goodbody, I. (1993). The ascidian fauna of a Jamaican lagoon: thirty years of change.
    Revista de Biologia Tropical 41 (suppl.1), 35-38.
Goodbody, I. (1994). The tropical western Atlantic Perophoridae (Ascidiacea): I. The
    genus Perophora. Bulletin of Marine Science 55 (1), 176-192.
Goodbody, I. (1996). Pycnoclavella belizeana, a new species of ascidian from the
    Caribbean. Bulletin of Marine Science 58 (2), 590-597.
Gopal, B. and Krishnamurthy, K. (1993). Wetlands of South Asia. In “Wetlands of the
    world” (D.F. Whigham, D. Dy Kyjova and S. Hejny, eds), pp. 345 - 414. Kluwer
    Academic Publishers, Netherlands.
Goswami, S.C. (1992). Zooplankton ecology of the mangrove habitats of Goa. In
    “Tropical Ecosystems: Ecology and Management” (K.P. Singh and J.S. Singh,
    eds), pp. 321-332. Wiley Eastern, Delhi, India.
Gourbault, N. and Vincx, M. (1994). New species of Parapinnanema (Nematoda:
    Chromadoridae) are described, with a discussion of the genus. Australian Journal
    of Marine and Freshwater Research 45 (2), 141-159.
Graham, A. (1995). Diversification of Gulf/Caribbean mangrove communities through
    Cenozoic time. Biotropica 27 (1), 20-27.
Grant, D.L., Clarke, P.J. and Allaway, W.G. (1993). The response of grey mangrove,
    Avicennia marina (Forssk.) Vierh. seedlings to spills of crude oil. Journal of
    Experimental Marine Biology and Ecology 71 (2), 273-295.
Grant, P.R. and Grant, B.R. (1997). The rarest of Darwin's finches. Conservation Biology
    11 (1), 119-126.
Green, S. and Webber, M. (1996). A survey of the solid waste pollution in Kingston
    Harbour mangroves, near Port Royal, Jamaica. Caribbean Marine Studies 5, 14-22.
Green, E.P., Mumby, P.J., Edwards, A.J., Clark, C.D. and Ellis, A.C. (1997). Estimating
    leaf area index of mangroves from satellite data. Aquatic Botany 58 (1), 11-19.
Green, E.P., Mumby, P.J., Edwards, A.J., Clark, C.D. and Ellis, A.C. (1998). The
    assessment of mangrove areas using high resolution multispectral airborne imagery.
    Journal of Coastal Research 14, 433-443.
Gregory, J.M. and Oerlemans, J. (1998). Simulated future sea-level rise due to glacier melt
    based on regionally and seasonally resolved temperature changes. Nature 391
    (6666), 474-476.
Guerreiro, J., Freitas, S., Pereira, P., Paula, J. and Macia, A. (1996). Sediment
    macrobenthos of mangrove flats at Inhaca Island, Mozambique. Cahiers de
    Biologie Marine 37, 309-327.
Harris, J.M. (1993). Widespread occurrence of extensive epimural rod bacteria in the
    hindguts of marine Thalassinidae and Brachyura (Crustacea: Decapoda). Marine
    Biology 116 (4), 615-629.
Harris, R.R. and Santos, M.C.F. (1993). Sodium uptake and transport (Na+ + K+ and
    ATPase changes following Na+ depletion and low salinity acclimation in the
    mangrove crab Ucides cordatus (L.) ). Comparative Biochemistry and Physiology
    105A (1), 35-42.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      104


Harris, R.R., Carmo, F. and Santos, M. (1993). Ion regulatory and urinary responses to
    emersion in the mangrove crab Ucides cordatus and the intertidal crab Carcinus
    maenas. Journal of Comparative Physiology 163B (1), 18-27.
Harrison, P.J., Snedaker, S.C., Ahmed, S.I. and Azam, F. (1994). Primary producers of the
    arid climate mangrove ecosystem of the Indus River Delta, Pakistan: An overview.
    Tropical Ecology 35 (2), 155-184.
Hatton, J.C. and Couto, A.L. (1992). The effect of coastline changes on mangrove
    community structure, Portuguese island, Mozambique. Hydrobiologia 247 (1-3),
    49-57.
He, B., Dai, P. and Fan, H. (1996). A study on the contents of the heavy metals in the
    sediments and macrobenthos of Yingluo mangrove swamps, Guangxi. Marine
    Environmental Science 15 (1), 35-41.
Heath, A.G., Turner, B.J. and Davis, W.P. (1993). Temperature preferences and tolerances
    of three fish species inhabiting hyperthermal ponds on mangrove islands.
    Hydrobiologia 259 (1), 47-55.
Hemminga, M.A., Slim, F.J., Kazunga, J., Ganssen, G.M., Nieuwenhuize, J. and Kruyt,
    N.M. (1994). Carbon outwelling from a mangrove forest with adjacent sea grass
    beds and coral reefs (Gazi Bay, Kenya). Marine Ecology Progress Series 106 (3),
    291-301.
Hemminga, M.A., Gwada, P., Slim, F.J., De Koeyer, P. and Kazungu, J. (1995). Leaf
    production and nutrient contents of the seagrass Thalassodendron ciliatum in the
    proximity of a mangrove forest (Gazi Bay, Kenya). Aquatic Botany 50 (2), 159-
    170.
Herppich, W.B. and Von Willert, D.J. (1995). Dynamic changes in leaf bulk water
    relations during stomatal oscillations in mangrove species. Continuous analysis
    using a dewpoint hygrometer. Physiologia Plantarum 94 (3), 479-485.
Herrera Silveira, J.A. and Ramirez Ramirez, J. (1996). Effects of natural phenolic material
    (tannin) on phytoplankton growth. Limnology and Oceanography 41 (5), 1018-
    1023.
Hill, C.J. (1992). Temporal changes in abundance of two lycaenid butterflies (Lycaenidae)
    in relation to adult food resources. Journal of the Lepidopterists' Society 46 (3),
    173-181.
Hirano,T., Monji, N., Hamotani, K., Jintana, V. and Yabuki, K. (1996). Transpirational
    characteristics of mangrove species in southern Thailand. Environmental Control in
    Biology 34 (4), 285-293.
Ho, H.H., Chang, H.S. and Hsieh, S.Y. (1991). Halophytophthora kandeliae, a new marine
    fungus from Taiwan. Mycologia 83 (4), 419-424.
Ho, W.H. and Hyde, K.D. (1996). Pterosporidium gen. nov. to accommodate two species
    of Anthostomella from mangrove leaves. Canadian Journal of Botany 74 (11),
    1826-1829.
Hockey, M.J. and de Baar, M. (1991). Some records of moths (Lepidoptera) from
    mangroves in southern Queensland. Australian Entomological Magazine 18 (2),
    57-60.
Hodda, M. (1990). Variation in estuarine littoral nematode populations over three spatial
    scales. Estuarine, Coastal and Shelf Science 30 (4), 325-340.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      105


Hodkinson, I.D. (1992). Telmapsylla gen. n., an unusual psyllid from black mangrove in
    Florida and Costa Rica (Insecta: Homoptera: Psylloidea). Zoologica Scripta 21 (3),
    307-309.
Hofmann, D.K., Fitt, W.K. and Fleck, J. (1996). Checkpoints in the life-cycle of Cassiopea
    spp.: Control of metagenesis and metamorphosis in a tropical jellyfish.
    International Journal of Developmental Biology 40 (1), 331-338.
Holguin, G. and Bashan, Y. (1996). Nitrogen-fixation by Azospirillum brasilense is
    promoted when co-cultured with a mangrove rhizosphere bacterium
    (Staphylococcus sp.). Soil Biology and Biochemistry 28 (12), 1651-1660.
Holguin, G., Guzman, A. and Bashan, Y. (1992). Two new nitrogen-fixing bacteria from
    the rhizosphere of mangrove trees: their isolation, identification and in vitro
    interaction with rhizosphere Staphylococcus sp.. FEMS-Microbiology and Ecology
    101 (3), 207-216.
Holmer, M., Kristensen, E., Banta, G., Hansen, K., Jensen, M.H. and Bussawarit, N.
    (1994). Biogeochemical cycling of sulfur and iron in sediments of a South-East
    Asian mangrove, Phuket Island, Thailand. Biogeochemistry 26 (3), 145-161.
Honda, D., Yokochi, T., Nakahara, T., Erata, M. and Higashihara, T. (1998).
    Schizochytrium limacinum sp. nov., a new thraustochytrid from a mangrove area in
    the west Pacific Ocean. Mycological Research 102 (4), 439-448.
Hong, P.N. and San, H.T. (1993). Mangroves of Vietnam. IUCN - The World
    Conservation Union. Bangkok, Thailand. 173 pp.
Hovenden, M.J. and Allaway, W.G. (1994). Horizontal structures on pneumatophores of
    Avicennia marina (Forsk.) Vierh.: A new site of oxygen conductance. Annals of
    Botany 73 (4), 377-383.
Hovenden, M.J., Curran, M., Cole, M.A., Goulter, P.F.E., Skelton, N.J. and Allaway, W.G.
    (1995). Ventilation and respiration in roots of one-year-old seedlings of grey
    mangrove Avicennia marina (Forsk.) Vierh. Hydrobiologia 295 (1-3), 23-29.
Huang, Q., Zhou, S. and Li, F. (1996). Ecological studies on mangrove boring animals
    Fujian. Journal of oceanography in Taiwan Strait 15 (3), 305-309.
Hudson, D.A. and Lester, R.J.G.. (1994). Parasites and symbionts of wild mud crabs Scylla
    serrata (Forskal) of potential significance in aquaculture. Aquaculture 120: 183-
    199.
Hussain, Z. and Acharya, G. (1994). Mangroves of the Sundarbans, Volume 2:
    Bangladesh. 257 pp. IUCN, Gland, Switzerland.
Hussain, M.I. and Khoja, T.M. (1993). Intertidal and subtidal blue-green algal mats of
    open and mangrove areas in the Farasab archipelago (Saudi Arabia) Red Sea.
    Botanica Marina 36, 377-388.
Hyde, K.D. (1990a). A comparison of the intertidal mycota of five mangrove tree species.
    Asian Marine Biology 7, 93-108.
Hyde, K.D. (1990b). A new marine ascomycete from Brunei. Aniptodera longispora sp.
    nov. from intertidal mangrove wood. Botanica Marina 33 (4), 335-338.
Hyde, K.D. (1991a). Phomopsis mangrovei, from intertidal prop roots of Rhizophora spp.
    Mycological Research 95 (9), 1149-1151.
Hyde, K.D. (1991b). Massarina velatospora and a new mangrove inhabiting species, M.
    ramunculicola sp. nov., Mycologia 83 (6), 839-845.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      106


Hyde, K.D. (1992a). Julella avicenniae (Borse) comb. nov. (Thelenellaceae) from
    intertidal mangrove wood and miscellaneous fungi from the NE coast of
    Queensland. Mycological Research 96 (11), 939-942.
Hyde, K.D. (1992b). The genus Saccardoella from intertidal mangrove wood. Mycologia
    84 (5), 803-810.
Hyde, K.D. (1992c). Aigialus striatispora sp. nov. from intertidal mangrove wood.
    Mycological Research 96 (12), 1044-1046.
Hyde, K.D. (1992d). Intertidal mangrove fungi from the west coast of Mexico, including
    one new genus and two new species. Mycological Research 96 (1), 25-30.
Hyde, K.D. (1993). Cryptovalsa halosarceicola sp. nov. an intertidal saprotroph of
    Halosarceia halocnemoides. Mycological Research 97 (7), 799-800.
Hyde, K.D. (1995). Lophiostoma asiana sp. nov. from Thailand mangroves. Mycotaxon
    55, 283-288.
Hyde, K.D. (1996). Biodiversity of microfungi in north Queensland. Australian Systematic
    Botany 9 (2), 261-271.
Hyde, K.D. and Jones, E.B.G. (1992). Intertidal mangrove fungi: Pedumispora gen. nov.
    (Diaporthales). Mycological Research 96 (1), 78-80.
Hyde, K.D. and Lee, S.Y. (1995). Ecology of mangrove fungi and their role in nutrient
    cycling: What gaps occur in our knowledge? Hydrobiologia 295, 107-118.
Hyde, K.D. and Rappaz, F. (1993). Eutypa bathurstensis sp. nov. from intertidal
    Avicennia. Mycological Research 97 (7), 861-864.
Hyde, K.D., Vrijmoed, L.L.P., Chinnaraj, S. and Jones, E.B.G. (1992). Massarina
    armatispora sp. nov., a new intertidal Ascomycete from mangroves. Botanica
    Marina 35 (4), 325-328.
Ibrahim, S. (1990). The effects of clear felling mangroves on sediment anaerobiosis.
    Journal of Tropical Forest Science 3 (1), 58-65.
Ibrahim, S. and Hashim, I. (1990). Classification of mangrove forest by using 1:40,000-
    scale aerial photographs. Forest Ecology and Management 33/34, 583-592.
Ikebe, Y. and Oishi, T. (1996). Correlation between environmental parameters and
    behaviour during high tides in Periophthalmus modestus. Journal of Fish Biology
    49 (1) 139-147.
Ikebe, Y. and Oishi, T. (1997). Relationships between environmental factors and diel and
    annual changes of the behaviors during low tides in Periophthalmus modestus,
    Zoological Science 14 (1), 49-55.
Ishimatsu, A., Hishida, Y., Takita, T., Kanda, T., Oikawa, S., Takeda, T. and Huat, K.K.
    (1998). Mudskippers store air in their burrows. Nature 391, 237-238.
Ishimatsu, A., Aguilar, N.M., Ogawa, K., Hishida, Y., Takeda, T., Oikawa, S., Kanda, T.
    and Huat, K.K. (1999). Arterial blood gas levels and cardiovascular function during
    varying environmental conditions in a mudskipper, Periophthalmodon schlosseri,
    Journal of Experimental Biology 202, 1753-1762.
Imbert, D., Labbé, P. and Rousteau, A. (1996). Hurricane damage and forest structure in
    Guadeloupe, French West Indies. Journal of Tropical Ecology 12, 663-680.
Imbert, D. and Ménard, S. (1997). Structure de la végétation et production primaire dans la
    mangrove de la Baie de Fort-de-France, Martinique (F.W.I.). Biotropica 29 (4),
    413-426.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       107


Ingole, B.S., Krishna Kumari, L., Ansari, Z.A. and Parulekar, A.H. (1994). New record of
     mangrove clam Geloina erosa (Solander, 1786) from the west coast of India.
     Journal of Bombay Natural History Society 91, 338-339.
Ip, Y.K., Chew, S.F. and Low, W.P. (1991). Effects of hypoxia on the mudskipper,
     Periophthalmus chrysospilos (Bleeker, 1853). Journal of Fisheries Biology 38 (4),
     621-623.
Irianto, R.S.B. and Suharti, M. (1994). Damage to Rhizophora sp. by spiders and
     possibilities for its control. Bulletin of Penelitian Hutan 559, 33-50.
Isaji, S. (1993). Formation of organic sheets in the inner shell layer of Geloina (Bivalvia:
     Corbiculidae): An adaptive response to shell dissolution. Veliger 36 (2), 166-173.
Isaji, S. (1995). Defensive strategies against shell dissolution in bivalves inhabiting acidic
     environments: The case of Geloina (Corbiculidae) in mangrove swamps. Veliger 38
     (3), 235-246.
Ish-Shalom-Gordon, N. and Dubinsky, Z. (1992). Ultrastructure of the pneumatophores of
     the mangrove Avicennia marina. South African Journal of Botany 58 (5), 358-362.
Ish-Shalom-Gordon, N., Lin, G. and Da Silveira, L. (1992). Isotopic fractionation during
     cellulose synthesis in two mangrove species: salinity effects. Phytochemistry 31
     (8), 2623-2626.
IUCN (1993). Oil gas exploration and production in mangrove areas. Guidelines for
     environmental protection. IUCN - The World Conservation Union, Gland,
     Switzerland, 47 pp.
Jacobi, C.M. and Schaeffer Novelli, Y. (1990). Oil spills in mangroves: a conceptual
     model based on long-term field observations. Ecological Modelling 52 (1-2), 53-
     59.
Jagtap, T.G. (1991). Distribution of seagrasses along the Indian coast. Aquatic Botany 40
     (4), 379-386.
Jagtap, T.G. (1992). Marine flora of Nicobar group of islands in Andaman Sea. Indian
     Journal of Marine Sciences 21 (1), 56-58.
Jagtap, T.G. (1993). Studies on littoral and sublittoral macrophytes around the Mauritius
     coast. Atoll Research Bulletin 382, 1-10.
Jennerjahn, C. and Ittekkot, V. (1997). Organic matter in sediments in the mangrove areas
     and adjacent continental margins of Brazil: 1. Amino acids and hexosamines.
     Oceanologica Acta 20 (2), 359-369.
Jiang, J.X. and Li, R.G. (1995). An ecological study on the Mollusca in mangrove areas in
     the estuary of the Jiulong River. Hydrobiologia 295 (1-3), 213-220.
Jiménez, J.A. (1990). The structure and function of dry weather mangroves on the Pacific
     coast of Central America, with emphasis on Avicennia bicolor forests. Estuaries 13
     (2), 182-192.
Jiménez, J.A. and Sauter, K. (1991). Structure and dynamics of mangrove forests along a
     flooding gradient. Estuaries 14 (1), 49-56.
John, D.M. and G.W. Lawson (1990). A review of mangrove and coastal ecosystems in
     West Africa and their possible relationships. Estuarine, Coastal and Shelf Science
     31, 505-518.
Jones, E.B.G. and Agerer, R. (1992). Calathella mangrovei sp. nov. and observations on
     the mangrove fungus Halocyphina villosa. Botanica Marina 35 (4), 259-265.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       108


Jones, E.B.G., Vrijmoed, L.L.P., Read, S.J. and Moss, S. T. (1994). Tirispora, a new
    ascomycetous genus in the Halosphaeriales. Canadian Journal of Botany 72 (9),
    1373-1378.
Jones, E.B.G., Hyde, K.D., Read, S.J., Moss, S. T. and Alias, S.A. (1996). Tirisporella
    gen. nov., an ascomycete from the mangrove palm Nypa fruticans. Canadian
    Journal of Botany 74 (9), 1487-1495.
Kairo, J.G. (1995). Community participatory forestry for rehabilitation of deforested
    mangrove areas of Gazi Bay. A. first approach. Final Technical Report,
    Biodiversity Support Program, USAID, Washington DC., 100 pp.
Kala, R.R. and Chandrika, V. (1993). Effect of different media for isolation, growth and
    maintenance of Actinomycetes from mangrove sediments. Indian Journal of Marine
    Sciences 22 (4), 297-299.
Kaly, U.L. and Jones, G.P. (1998). Mangrove restoration: a potential tool for coastal
    management in tropical developing countries. Ambio 27, 656-661.
Kaly, U.L., Eugelink, G. and Robertson, A.I. (1997). Soil conditions in damaged North
    Queensland mangroves. Estuaries 20 (2), 291-300.
Kamaludin, B.H. and Woodroffe, C.D. (1993). The changing mangrove shorelines in
    Kuala Kurau, Peninsular Malaysia. Sedimentary Geology 83 (3-4), 187-197.
Kangas, P.C. and Lugo, A.E. (1990). The distribution of mangroves and saltmarsh in
    Florida (USA). Tropical Ecology 31 (1), 32-39.
Kannan, L. and Vasantha, K. (1992). Microphytoplankton of the Pichavaram mangals,
    southeast coast of India: Species composition and population density.
    Hydrobiologia 247, 77-86.
Karsten, U., Koch, S., West, J.A. and Kirst, G.O. (1994). The intertidal red alga Bostrychia
    simpliciuscula Harvey ex J. Agardh from a mangrove swamp in Singapore:
    Acclimation to light and salinity. Aquatic Botany 48 (3-4), 313-323.
Karsten, U., Barrow, K.D., Mostaert, A.S. and King, R.J. (1995). The osmotic significance
    of the heteroside floridoside in the mangrove alga Catenella nipae (Rhodophyta:
    Gigartinales) in eastern Australia. Estuarine, Coastal and Shelf Science 40 (3), 239-
    247.
Karsten, U., Mostaert, A.S., King, R.J., Kamiya, M. and Hara, Y. (1996). Osmoprotectors
    in some species of Japanese mangrove macroalgae. Phycology Research 44 (2),
    109-112.
Karsten, U., Barrow, K.D., Nixdorf, O., West, J.A. and King, R.J. (1997). Characterization
    of mannitol metabolism in the mangrove red alga Caloglossa leprieurii (Montagne)
    J. Agardh. Planta 201 (2), 173-178.
Kathiresan, K. (1990). Prospects of tissue culture studies in mangroves. In “Advances in
    Forestry Research” (Ram Parkash, ed.), Vol. 6, pp. 143-151. International Book
    Distributors, Dehra dun, India.
Kathiresan, K. (1992). Foliovory in Pichavaram mangroves. Environment and Ecology 10
    (4), 988-989.
Kathiresan, K. (1993). Dangerous pests on nursery seedlings of Rhizophora. The Indian
    Forester 119, 1026.
Kathiresan, K. (1994). Propagation of mangroves: some considerations. In “Conservation
    of mangrove forest genetic resources” - A training manual (S.V. Deshmukh and V.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                     109


    Balaji, eds), pp. 303-306. ITTO-CRSARD Project, M.S. Swaminathan Research
    Foundation, Madras.
Kathiresan, K. (1995a). Rhizophora annamalayana: A new species of mangrove.
    Environment and Ecology 13 (1), 240-241.
Kathiresan, K. (1995b). Studies on tea from mangrove leaves. Environment and Ecology
    13 (2), 321-323.
Kathiresan, K. (l999). Impact of mangrove biodiversity on associated fishery resource and
    fishers’ income . Project Report, WWF, Washington, USA. 108 pp.
Kathiresan, K. (2000). A review of studies on Pichavaram mangrove, southeast India.
    Hydrobiologia 430 (1), 185-205.
Kathiresan, K. and Moorthy, P. (1992). Influence of boric acid on rooting of Rhizophora
    apiculata Blume hypocotyl. Environment and Ecology 10 (4), 992-993.
Kathiresan, K. and Moorthy, P. (1993). Influence of different irradiance on growth and
    photosynthetic characteristics in seedlings of Rhizophora species. Photosynthetica
    29, 143 - 146.
Kathiresan, K. and Moorthy, P. (1994a). Photosynthetic responses of Rhizophora apiculata
    Blume seedlings to a long-chain aliphatic alcohol. Aquatic Botany 47 (2), 191-193.
Kathiresan, K. and Moorthy, P. (1994b). Effect of NAA, IBA and Keradix on rooting
    potential of Rhizophora apiculata Blume hypocotyls. The Indian Forester 120,
    420-422.
Kathiresan, K. and Moorthy, P. (1994c). Chemical induced rooting in hypocotyls of
    Rhizophora mucronata. Indian Journal of Forestry 17 (4), 310-312.
Kathiresan, K. and Moorthy, P. (1994d). Hormone-induced physiological response of
    tropical mangrove species. Botanica Marina 37, 139-141.
Kathiresan, K. and Pandian, M. (1991). Effect of UV on quality of black tea from Ceriops
    decandra. Science and Culture 57 (3-4), 93-95.
Kathiresan, K. and Pandian, M. (1993). Effect of UV on black tea constituents of
    mangrove leaves. Science and Culture 59, 61-63.
Kathiresan, K. and Ramesh, M.X. (1991). Establishment of seedlings of a mangrove. The
    Indian Forester 17 (3), 93-95.
Kathiresan, K. and Ravi, V. (1990). Seasonal changes in tannin content of mangrove
    leaves. The Indian Forester 116 (5), 390-392.
Kathiresan, K. and Ravikumar, S. (1995a). Vegetative propagation through air-layering in
    two species of mangroves. Aquatic Botany 50 (1), 107-110.
Kathiresan, K. and Ravikumar, S. (1995b). Influence of tannins, sugars and amino acids on
    bacterial load of marine halophytes. Environment and Ecology 13 (1), 94-96.
Kathiresan, K. and Ravikumar, S. (1997). Studies on tissue culture aspects of marine
    halophytes. In “Biotechnological applications of plant tissue and cell culture”
    (G.A. Ravishankar and L.V. Venkataraman, eds), pp. 290-295. Oxford and IBH,
    Publishing Co., Pvt. Ltd. India.
Kathiresan, K. and Thangam, T.S. (1987). Biotoxicity of Excoecaria agallocha L. latex on
    marine organisms. Current Science 56 (7), 314- 315.
Kathiresan, K. and Thangam, T.S. (1989). Effect of leachates from mangrove leaf on
    rooting of Rhizophora seedlings. Geobios 16 (1), 27-29.
Kathiresan, K. and Thangam, T.S. (1990a). A note on the effects of salinity and pH on
    growth of Rhizophora seedlings. The Indian Forester 116 (3), 243-244.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                     110


Kathiresan, K. and Thangam, T.S. (1990b). Effect of phenolics on growth of viviparous
    seedlings of Rhizophora apiculata. Geobios 17, 142-143.
Kathiresan, K., Ravishankar, G.A. and Venkataraman, L.V. (1990a). Auxin-phenol
    induced rooting in a mangrove Rhizophora apiculata Blume. Current Science 59
    (8), 430-432.
Kathiresan, K., Thangam, T.S. and Bose, K.S. (1990b). Effect of lates of Excoecaria
    agallocha L. on marine productivity. In “Perspectives in Phycology” (V.N. Raja
    Rao, ed.), pp. 319-321. Today and Tomorrow’s Publishers, New Delhi.
Kathiresan, K., Moorthy, P. and Ravikumar, S. (1993). Rooting of Rhizophora mucronata
    as influenced by mangrove leaf leachates. Journal of Tree Sciences 12 (1), 23-26.
Kathiresan, K., Moorthy, P. and Rajendran, N. (1994a). Seedling performance of
    mangrove Rhizophora apiculata (Rhizophorales: Rhizophoraceae) in different
    environs. Indian Journal of Marine Sciences 23 (3), 168-169.
Kathiresan, K., Moorthy, P. and Rajendran, N. (1994b). Promotory effect of some
    chemicals on seedling growth of Rhizophora apiculata. Environment and Ecology
    11, 716-717.
Kathiresan, K., Ramesh. M.X. and Venkatesan, V. (1994c). Forest structure and prawn
    seeds in Pichavaram mangroves. Environment and Ecology 12, 465-468.
Kathiresan, K., Thangam, T.S. and Premanathan, M. (1995a). Mangrove halophytes:
    potential source of medicines. In “Biology of salt tolerant plants” (M.A. Khan and
    I.A. Ungar, eds), pp. 361-370. University of Karachi, Pakistan.
Kathiresan, K., Moorthy, P. and Ravikumar, S. (1995b). Studies on root growth in
    seedlings of a tropical mangrove tree species. International Tree Crops Journal 8
    (2-3), 183-188.
Kathiresan, K., Rajendran, N. and Thangadurai, G. (1996a). Growth of mangrove seedlings
    in intertidal area of Vellar estuary southeast coast of India. Indian Journal of
    Marine Sciences 25, 240-243.
Kathiresan, K., Moorthy, P. and Ravikumar, S. (1996b). A note on the influence of salinity
    and pH on rooting of Rhizophora mucronata Lamk. Seedlings. The Indian Forester
    122 (8), 763-764.
Kathiresan, K., Moorthy, P. and Rajendran, N. (1996c). Methanol induced physiological
    changes in mangroves. Bulletin of Marine Science 59 (2), 454-458.
Kathiresan, K., Ravishankar,G.A. and Venkataraman, L.V. (1997). In vitro multiplication
    of a coastal plant Sesuvium portulacastrum L. by auxillary buds. In
    “Biotechnological applications of plant tissue and cell culture” (G.A. Ravishankar
    and L.V. Venkataraman, eds), pp. 185-192. Oxford and IBH, Publishing Co., Pvt.
    Ltd. India.
Kawabata, Z., Magendran, A., Palanichamy, S., Venugopalan, V.K. and Tatsukawa, R.
    (1993). Phytoplankton biomass and productivity of different size fractions in the
    Vellar estuarine system, southeast coast of India. Indian Journal of Marine
    Sciences 22 (4), 294-296.
Kelaher, B.P., Chapman, M.G. and Underwood, A.J. (1998a). Changes in benthic
    assemblages near boardwalks in temperate urban mangrove forests. Journal of
    Experimental Marine Biology and Ecology 228 (2), 291-307.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       111


Kelaher, B.P., Underwood, A.J. and Chapman, M.G. (1998b). Effect of boardwalks on the
    semaphore crab Heloecius cordiformis in temperate urban mangrove forests.
    Journal of Experimental Marine Biology and Ecology 227 (2), 281-300.
Kendrick, G.W. and Morse, K. (1990). Evidence of recent mangrove decline from an
    archaeological site in Western Australia (Australia). Australian Journal of Ecology
    15 (3), 349-354.
Khandelwal, A. and Gupta, H.P. (1993). Palynological evidence of mangrove degradation
    during mid-late Holocene at Rambha, Chilka Lake, Orissa. Geophytology 23 (1),
    141-145.
Kimani, E.N., Mwatha, G.K., Wakwabi, E.O., Ntiba, J.M. and Okoth, B.K. (1996). Fishes
    of a shallow tropical mangrove estuary, Gazi, Kenya. Marine and Freshwater
    Research 47 (7), 857-868.
King, C.R. and Williamson, I. (1995). Zooplankton distribution in Raby Bay, south-east
    Queensland, Australia. Proceedings of the Royal Society of Queensland 105 (2),
    23-31.
King, R.J. (1990). Macroalgae associated with the mangrove vegetation of Papua New
    Guinea. Botanica Marina 33 (1), 55-62.
King, R.J. (1995). Mangrove macroalgae: A review of Australian studies. Proceedings of
    the Linnean Society of New South Wales 115, 151-161.
King, R.J. and Puttock, C.F. (1994). Macroalgae associated with mangroves in Australia:
    Rhodophyta. Botanica Marina 37 (3), 181-191.
Kitheka, J.U. (1996). Water circulation and coastal trapping of brackish water in a tropical
    mangrove dominated bay in Kenya. Limnology and Oceanography 41 (1), 169-176.
Kitheka, J.U., Ohowa, B.O., Mwashote, B.M., Shimbira, W.S., Mwaluma, J.M. and
    Kazungu, J.M. (1996). Water circulation dynamics, water column nutrients and
    plankton productivity in a well flushed tropical bay in Kenya. Sea Resources 35
    (4), 257-268.
Kjerfve, B. and Macintosh, D.J. (1997). Climage change impacts on mangrove ecosystems.
    In “ Mangrove Ecosystem Studies in Latin America and Africa” (B. Kjerfve, L.D.
    Lacerda and S. Diop, eds), pp. 1-7. UNESCO, Paris.
Klein, M.L., Humphrey, S.R. and Percival, H.F. (1995). Effects of ecotourism on
    distribution of waterbirds in a wildlife refuge. Conservation Biology 9, 1454-1465.
Klekowski, E.J., Lowenfeld, R. and Hepler, P.K. (1994a). Mangrove genetics: II.
    Outcrossing and lower spontaneous mutation rates in Puerto Rican Rhizophora.
    International Journal of Plant Sciences 155 (3), 373-381.
Klekowski, E.J., Corredor, J.D., Morelli, J.M. and Del Castillo, C. (1994b). Petroleum
    pollution and mutation in mangroves. Marine Pollution Bulletin 28 (3), 166-169.
Klekowski, E.J., Corredor, J.D., Lowenfeld, R., Klekowski, E.H. and Morelli, J.M.
    (1994c). Using mangroves to screen for mutagens in tropical marine environments.
    Marine Pollution Bulletin 28 (6), 346-350.
Klekowski, E.J., Lowenfeld, R. and Klekowski, E.H. (1996). Mangrove genetics. IV.
    Postzygotic mutations fixed as periclinal chimeras. International Journal of Plant
    Sciences 157 (4), 398-405.
Koch, M.S. (1997). Rhizophora mangle L. seedling development into the sapling stage
    across resource and stress gradients in subtropical Florida. Biotropica 29 (4), 427-
    439.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       112


Koch, M.S. and Snedaker, S.C. (1997). Factors influencing Rhizophora mangle L. seedling
    development in Everglades carbonate soils. Aquatic Botany 59 (1-2), 87-98
Koch, V. and Wolff, M. (1996).The mangrove snail Thais kiosquiformis Duclos: A case of
    life history adaptation to an extreme environment. Journal of Shellfish Research 15
    (2), 421-432.
Kofron, C.P. (1992). Status and habitats of the three African crocodiles in Liberia. Journal
    of Tropical Ecology 8 (3), 265-273.
Kohlmeyer, J. and Kohlmeyer, E. (1979). Marine Mycology. Academic Press, New York.
    690 pp.
Kohlmeyer, J. and Kohlmeyer, V.B. (1991). Hapsidascus hadrus, new genus - new species
    (Ascomycotina) from mangroves in the Caribbean. Systema Ascomycetum 10 (2),
    113-120.
Kohlmeyer, J., Bebout, B. and Volkmann-Kohlmeyer, B. (1995). Decomposition of
    mangrove wood by marine fungi and teredinids in Belize. Marine Ecology 16, 27-
    39.
Kohlmeyer, V.B. and Kohlmeyer, J. (1993). Biogeographic observations on Pacific marine
    fungi. Mycologia 85 (3), 337-346.
Koizumi, M., Takahashi, K., Mineuchi, K., Nakamura, T. and Kano, H. (1998). Light
    gradients and the transverse distribution of chlorophyll fluorescence in mangrove
    and Camellia leaves. Annals of Botany 81 (4), 527-533.
Kokpol, U., Chavasiri, W., Chittawong, V. and Miles, D.H. (1990). Taraxeryl cis-p-
    hydroxycinnamate, a novel taraxeryl from Rhizophora apiculata. Journal of
    Natural Products 53 (4), 953-955.
Komiyama, A., Tanuwong, S. and Higo, M. (1996). Microtopography, soil hardness and
    survival of mangrove (Rhizophora apiculata BL.) seedlings planted in an
    abandoned tin-mining area. Forest Ecology and Management 81, 243-248.
Krishnamoorthy, P., Maruthanayagam, C. and Subramanian, P. (1995). Toxic effect of
    mangrove plant (Excoecaria agallocha L.) latex on the larvae of fresh water prawn
    Macrobrachium lamarrei. Environment and Ecology 13 (3), 708-710.
Krishnamurthy, K. (1990). The apiary of the mangroves. In “Wetland Ecology and
    Management: Case Studies” (D.F. Whigham, D. Dykyjova and S. Hejny, eds), pp.
    135-140. Kluwer Academic Press, Netherlands.
Krishnamurthy, K., Kathiresan, K., Kannan, L. Godhantaraman, N. and Damodara Naidu,
    W. (1995a). Cyanobacteria. In “Plankton of Parangipettai (Poroto Novo), India”.
    Fascile No.2. New Series, 27 pp.Memoirs of the CAS in Marine Biology,
    Annamalai University, Parangipettai.
Krishnamurthy, K., Damodara Naidu, W., Godhantaraman, N., Kannan, L. and Kathiresan,
    K. (1995b). Microzooplankton with special reference to Tintinnida (Protozoa:
    Ciliata: Tintinnida). In “Plankton of Parangipettai (Poroto Novo), India”. Fascile
    No.1. New Series, 81 pp. Memoirs of the CAS in Marine Biology, Annamalai
    University, Parangipettai.
Kristensen, E., Holmer, M., Banta, G.T., Jensen, M.H. and Hansen, K. (1995). Carbon,
    nitrogen and sulphur cycling in sediments of the AO NAM BOR Mangrove forest,
    Phuket, Thailand: A review. Phuket Marine Biological Centre Research Bulletin
    60, 37-64.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                     113


Kryger, L. and Lee, S.K. (1995). Effects of soil ageing on the accumulation of hydrogen
    sulphide and metallic sulphides in mangrove areas in Singapore. Environment
    International 21 (1), 85-92.
Kryger, L. and Lee, S.K. (1996). Effects of mangrove soil ageing on the accumulation of
    hydrogen sulfide in roots of Avicennia spp. Biogeochemistry 35, 367-375.
Kuenen, M.M.C.E. and Debrot, A.O. (1995). A quantitative study of the seagrass and algal
    meadows of the Spaanse Water, Curacao, Netherlands Antilles. Aquatic Botany 51
    (3-4), 291-310.
Kulkarni, P.K. and Bhosale, L.J. (1991). Studies on eco-physiology of Rhizophora
    mucronata and Rhizophora apiculata. Geobios 18, 19-24.
Kusmana, C. (1990). Soil as a factor influencing the mangrove forest communities in
    Talidendang Besar, Riau. Biotropica 4, 9-18.
Kwok, P.W. and Lee, S.Y. (1995). The growth performances of two mangrove crabs,
    Chiromanthes bidens and Parasesarma plicata under different leaf litter diets.
    Hydrobiologia 295 (1-3), 141-148.
Lacerda, L.D. (1997). Trace metals in mangrove plants: why such low concentrations? In
    “Mangrove Ecosystem Studies in Latin America and Africa” (B. Kjerfve, L.D.
    Lacerda and S. Diop, ets), pp. 171-178. UNESCO, Paris.
Lacerda, L.D. (1998). Trace metals biogeochemistry and diffuse pollution in mangrove
    ecosystems. ISME Mangrove Ecosystems Occasional Papers 2, 1-61.
Lacerda, L.D., Aragon, G.T., Ovalle, A.R.C. and Rezende, C.E. (1991). Iron and
    chromium distribution and accumulation in a mangrove ecosystem. Water, Air and
    Soil Pollution 57-58, 513-520.
Lacerda, L.D., Carvalho, C.E.V., Tanizaki, K.F., Ovalle, A.R.C. and Rezende, C.E. (1993).
    The biogeochemistry and trace metals distribution of mangrove rhizospheres.
    Biotropica 25, 251-256.
Lacerda, L.D., Ittekkot, V. and Patchineelam, S.R. (1995). Biogeochemistry of mangrove
    soil organic matter: a comparison between Rhizophora and Avicennia soils in
    south-eastern Brazil. Estuarine, Coastal and Shelf Science 40, 713-720.
Lakshmi, M., Rajalakshmi, S., Parani, M., Anuratha, C.S. and Parida, A. (1997). Molecular
    phylogeny of mangroves. I. Use of molecular markers in assessing the intraspecific
    genetic variability in the mangrove species Acanthus ilicifolius Linn.
    (Acanthaceae). Theoretical and Applied Genetics 94, 1121-1127.
Lamparelli, C.C., Rodrigues, F.O. and de Moura, D.O. (1997). A long-term assessment of
    an oil spill in a mangrove forest in São Paulo, Brazil. In “Mangrove Ecosystem
    Studies in Latin America and Africa” (B. Kjerfve, L.D. Lacerda and S. Diop, eds),
    pp. 191-203. UNESCO, Paris.
Lana, P.C., Couto, E.C.G. and Almeida, M.V. (1997). Distribution and abundance of
    polychaetes in mangroves of a subtropical estuary. Bulletin of Marine Science 60
    (2), 616-617.
Lana, P.D.C., Guiss, C. and Disaro, S.T. (1991). Seasonal variation of biomass and
    production dynamics for aboveground and belowground components of a Spartina
    alterniflora marsh in the Euhaline sector of Paranagua Bay (SE Brazil). Estuarine,
    Coastal and Shelf Science 32 (3), 231-242.
Larkum, A.W.D. and West, R.J. (1990). Long-term changes of seagrass meadows in
    Botany Bay, Australia. Aquatic Botany 37 (1), 55-70.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       114


Laroche, J., Baran, E. and Rasoanandrasana, N. B. (1997). Temporal patterns in a fish
    assemblage of a semiarid mangrove zone in Madagascar. Journal of Fish Biology
    51 (1), 3-20.
Leaño, E.M., Vrijmoed, L.L., Jones, P. and Gareth, E.B. (1998). Zoospore chemotaxis of
    two mangrove strains of Halophytophthora vesicula from Mai Po, Hong Kong.
    Mycologia 90 (6), 1001-1008.
Lee, H.L. and Seleena, P. (1990). Effect of sodium chloride on the growth of several
    isolates of Bacillus thuringiensis serotype H-14. Tropical Biomedicine 7 (2), 207-
    208.
Lee, H.L., Seleena, P. and Winn, Z. (1990a). Bacillus thuringiensis serotype H-14 isolated
    from mangrove swamp soil in Malaysia. Mosquito Borne Diseases Bulletin 7 (4),
    134-135.
Lee, K.H., Moran, M.A., Benner, R. and R.E. Hodson. (1990). Influence of soluble
    components of red mangrove (Rhozophora mangle) leaves on microbial
    decomposition of structural (lignocellulosic) leaf components in seawater. Bulletin
    of Marine Science 46 (2), 374-386.
Lee, S.K., Tan, W.H. and Havanond, S. (1996). Regeneration and colonization of
    mangrove on clay-filled reclaimed land in Singapore. Hydrobiologia 319, 23-35.
Lee, S.Y. (1990). Primary productivity and particulate organic matter flow in an estuarine
    mangrove-wetland in Hong Kong. Marine Biology 106, 453-463.
Lee, S.Y. (1991). Herbivory as an ecological process in a Kandelia candel
    (Rhizophoraceae) mangal in Hong Kong. Journal of Tropical Ecology 7, 337-348.
Lee, S.Y. (1995). Mangrove outwelling: A review. Hydrobiologia 295 (1-3), 203-212.
Lee, S.Y. (1998). Ecological role of grapsid crabs in mangrove ecosystems: a review.
    Marine and Freshwater Research 49, 335-343.
Lefebvre, G. and Poulin, B. (1996). Seasonal abundance of migrant birds and food
    resources in Panamanian mangrove forests. Wilson Bulletin 108 (4), 748-759.
Lefebvre, G. and Poulin, B. (1997). Bird communities in Panamanian black mangroves:
    potential effects of physical and biotic factors. Journal of Tropical Ecology 13 (1),
    97-113.
Lefebvre, G., Poulin, B. and McNeil, R. (1992). Settlement period and function of long-
    term territory in tropical mangrove passerines. The Condor 94, 83-92.
Lefebvre, G., Poulin, B. and McNeil, R. (1994). Temporal dynamics of mangrove bird
    communities in Venezuela with special reference to migrant warblers. Auk 111 (2),
    405-415.
Leong, W.F., Tan, T.K., Hyde, K.D. and Jones, E.B.G. (1990). Payosphaeria minuta gen.
    et sp. nov., an ascomycete on mangrove wood. Botanica Marina 33 (6), 511-514.
Leong, W.F., Tan, T.K., Hyde, K.D. and Gareth Jones, E.B. (1991). Halosarpheia minuta
    sp. nov., an ascomycete from submerged mangrove wood. Canadian Journal of
    Botany 69 (4), 883-886.
Levings, S.C. and Garrity, S.D. (1994). Effects of oil spills on fringing red mangroves
    (Rhizophora mangle): Losses of mobile species associated with submerged prop
    roots. Bulletin of Marine Science 54 (3), 782-794.
Levings, S.C., Garrity, S.D. and Burns, K.A. (1994). The Galeta oil spill. 3. Chronic
    reoiling, long-term toxicity of hydrocarbon residues and effects on epibiota in the
    mangrove fringe. Estuarine, Coastal and Shelf Science 38 (4), 365-395.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       115


Ley, J.A., Montague, C.L. and Conacher, C.C. (1994). Food habits of mangrove fishes: a
    comparison along estuarine gradients in northeastern Florida Bay. Bulletin of
    Marine Science 54, 881-899.
Ley, J.A., McIvor, C.C. and Montague, C.L. (1999). Fishes in mangrove prop-root habitats
    of northeastern Florida Bay: Distinct assemblages across an estuarine gradiant.
    Estuarine, Coastal and Shelf Science 48, 701-723.
Lezine, A.M. (1996). La mangrove ouest africaine, signal des variations du niveau marin et
    des conditions regionales du climat au cours de la derniere deglaciation. Bulletin de
    la Socièté Gèologique de France 167 (6), 743-752.
Li, M.S. and Lee, S.Y. (1997). Mangroves of China: a brief review. Forest Ecology and
    Management 96, 241-259.
Liao, B., Zheng, D. and Zheng, S. (1990). Studies on the biomass of a Sonneratia
    caseolaris stand. Forest Research 3 (1), 47-54.
Liao, B., Zheng, D. and Zheng, S. (1993). Biomass and leaf area index of secondary shrub
    in mangroves of Qingland Harbour in Hainan Island. Forest Research 6 (6), 680-
    685.
Lin, G.H. and Sternberg, L.D.S.L. (1992). Differences in morphology, carbon isotope
    ratios, and photosynthesis between scrub and fringe mangroves in Florida, USA.
    Aquatic Botany 42 (4), 303-313.
Lin, G.H. and Sternberg, L.D.S.L. (1993). Effects of salinity fluctuation on photosynthetic
    gas exchange and plant growth of the red mangrove (Rhizophora mangle L.).
    Journal of Experimental Botany 44 (258), 9-16.
Lin, G.H. and Sternberg, L.D.S.L. (1994). Utilization of surface water by red mangrove
    (Rhizophora mangle L.): An isotopic study. Bulletin of Marine Science 54 (1), 94-
    102.
Lin, G.H., Guanghui Banks, T. and Sternberg, L.D.S.L. (1991). Variation in delta 13C
    values for the seagrass Thalassia testudinum and its relations to mangrove carbon.
    Aquatic Botany 40 (4), 333-341.
Lin, J. and Beal, J.L. (1995). Effects of mangrove marsh management on fish and decapod
    communities. Bulletin of Marine Science 57 (1), 193-201.
Lin, R., Lin, M., Teng, J. and Zhang, W. (1994). Remote sensing survey and mapping of
    mangroves in western Xiamen Harbour. Journal of Oceanography in Taiwan Strait
    13 (3), 297-302.
Loka-Bharathi, P.A., Oak, S. and Chandramohan, D. (1991). Sulfate reducing bacteria
    from mangrove swamps. 2. Their ecology and physiology. Oceanologica Acta 14
    (2), 163-171.
Loneragan, N.R., Bunn, S.E. and Kellaway, D.M. (1997). Are mangroves and seagrasses
    sources of organic carbon for penaeid prawns in a tropical Australian estuary? A
    multiple stable-isotope study. Marine Biology 130 (2), 289-300.
Long, B.G. and Skewes, T.D. (1996). A technique for mapping mangroves with LandsatTM
    satellite data and geographic information system. Estuarine, Coastal and Shelf
    Science 43 (3), 373-381.
Lopes, S.G.B.C. and Narchi, W. (1993). A survey and distribution for Teredinidae
    (Mollusca: Bivalvia) at mangrove regions in Praia Dura, Ubatuba, Sao Paulo,
    Brazil. Bolivian Institute of Oceanography Sao Paulo 41 (1-2), 29-38.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      116


Lopez-Cortes, A. (1990). Microbial mats in tidal channels at San Carlos, Baja California
    Sur, Mexico. Geomicrobiological Journal 8 (2), 69-85.
Lorenz, J.J., McIvor, C.C., Pwell, G.V.N. and Frederick, P.C. (1997). A drop net and
    removable walkway used to quantitatively sample fishes over wetland surfaces in
    the dwarf mangroves of the southern Everglades. Wetlands 17 (3), 346-359.
Loughland, R.A. (1998). Mangal roost selection by the flying-fox Pteropus alecto
    (Megachiroptera: Pteropodidae). Maine and Freshwater Research 49, 351–352.
Louis, M., Bouchon, C. and Bouchon Navaro, Y. (1995). Spatial and temporal variations
    of mangrove fish assemblages in Martinique (French West Indies) Hydrobiologia
    295 (1-3), 275-284.
Lovelock, C.E. and Clough, B.F. (1992). Influence of solar radiation and leaf angle on leaf
    xanthophyll concentrations in mangroves. Oecologia 91 (4), 518-525.
Lovelock, C.E., Clough, B.F. and Woodrow, I.E. (1992). Distribution and accumulation of
    ultraviolet-radiation-absorbing compounds in leaves of tropical mangroves. Planta
    188 (2), 143-154.
Lowenfeld, R. and Klekowski, E.J. (1992). Mangrove genetics 1. Mating system and
    mutation rates of Rhizophora mangle in Florida and San Salvador Island, Bahamas.
    International Journal of Plant Sciences 153, 394-399.
Lu, C.Y. and Lin, P. (1990). Studies on litter fall and decomposition of Bruguiera
    sexangular community in Hainan Island, China. Bulletin of Marine Science 47,
    139-148.
Lu, C.Y., Wong, Y.S., Tam, N.F.Y., Ye, Y., Cui, S.H. and Lin, P. (1998). Preliminary
    studies on methane fluxes in Hainan mangrove communities. Chinese Journal of
    Oceanology and Limnology 16 (1), 64-71.
Lugo, A.E. (1997). Old-growth mangrove forests in the United States. Conservation
    Biology 11 (1), 11.
Lugo, A.E. (1998). Mangrove forests: a tough system to invade but an easy one to
    rehabilitate. Marine Pollution Bulletin 37 (8-12), 427-430.
Machiwa, J.F. and Hallberg, R.O. (1995). Flora and crabs in a mangrove forest partly
    distorted by human activities, Zanzibar. Ambio 24 (7-8), 492-496.
Mackey, A.P. (1993). Biomass of the mangrove Avicennia marina (Forssk.) Vierh. near
    Brisbane, south-eastern Queensland. Australian Journal of Marine and Freshwater
    Research 44 (5), 721-725.
Mackey, A.P. and Hodgkinson, M. (1996). Assessment of the impact of naphthalene
    contamination on mangrove fauna using behavioral bioassays. Bulletin of
    Environmental Contamination and Toxicology 56, 279-286.
Mackey, A.P. and Mackay, S. (1996). Spatial distribution of acid-volatile sulphide
    concentration and metal bioavailability in mangrove sediments from the Brisbane
    River, Australia. Environmental Pollution 93 (2), 205-209.
Mackey, A.P. and Smail, G. (1996). The decomposition of mangrove litter in a subtropical
    mangrove forest. Hydrobiologia 332 (2), 93-98.
Mackey, A.P., Hodgkinson, M. and Nardella, R. (1992). Nutrient levels and heavy metals
    in mangrove sediments from the Brisbane River, Australia. Marine Pollution
    Bulletin 24 (8), 418-420.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      117


Macnae, W. (1967). Zonation within mangroves associated with estuaries in North
    Queensland. In “Estuaries” (G.H. Lauff, ed), pp. 432-441. American Association
    for the Advancement of Science, Washington, D.C.
Macnae, W. (1968). A general account of a fauna and flora of mangrove swamps and
    forest in the Indo-Pacific region. Advances in Marine Biology 6, 73-270.
Mahasneh, I.A., Grainger, S.L.J. and Whitton, B.A. (1990). Influence of salinity on hair
    formation and phosphatase activities of the blue-green alga (Cyanobacterium)
    Calothrix viguieri D253. British Phycological Journal 25 (1), 25-32.
Maitland, D.P. (1990). Carapace and branchial water circulation and water related
    behaviours in the semaphore crab Heloecius cordiformis (Decapoda: Brachyura:
    Ocypodidae). Marine Biology 105 (2), 275-286.
Mall, L.P., Singh, V.P. and Garge, A. (1991). Study of biomass, litter fall, litter
    decomposition and soil respiration in monogeneric mangrove and mixed mangrove
    forests of Andaman Islands (India). Tropical Ecology 32 (1), 144-152.
Mandura, A.S. (1997). A mangrove stand under sewage pollution stress: Red Sea.
    Mangroves and Salt Marshes 1, 255-262.
Mani, P. (1992). Natural phytoplankton communities in Pichavaram mangroves. Indian
    Journal of Marine Sciences 21 (4), 278-280.
Mani, P. (1994). Phytoplankton in Pichavaram mangroves, east coast of India. Indian
    Journal of Marine Sciences 23 (1), 22-26.
Mann, F.D. and Steinke, T.D. (1992). Biological nitrogen-fixation (acetylene reduction)
    associated with decomposing Avicennia marina leaves in the Beachwood
    Mangrove Nature Reserve. South African Journal of Botany 58 (6), 533-536.
Mann, F.D. and Steinke, T.D. (1993). Biological nitrogen- fixation (acetylene reduction)
    associated with blue-green algal (cyanobacterial) communities in the Beachwood
    Mangrove Nature Reserve II. Seasonal variation in acetylene reduction activity.
    South African Journal of Botany 59 (1), 1-8.
Marguillier, S., Van-der-Velde, G., Dehairs, F., Hemminga, M.A. and Rajagopal, S.
    (1997). Trophic relationship in an interlinked mangrove-seagrass ecosystem as
    traced by delta 13C and delta 15N. Marine Ecology Progress Series 151, 115-121.
Marius, C. and Lucas, J. (1991). Holocene mangrove swamps of West Africa:
    sedimentology and soils. Journal of African Earth Sciences 12 (1-2), 41-54.
Marshall, N. (1994). Mangrove conservation in relation to overall environmental
    considerations. Hydrobiologia 285 (1-3), 303-309.
Martin, F., Dutriex, E. and Debry, A. (1990). Natural recolonization of a chronically oil-
    polluted mangrove soil after a depollution process. Ocean and Shoreline
    Management 14 (3), 173-190.
Martuscelli, P., Olmos, F., Silva, R.S.E., Mazzarella, I.P., Pino, F.V. and Raduan, E.N.
    (1996). Cetaceans of Sao Paulo, southeastern Brazil. Mammalia 60 (1), 125-140.
Mastaller, M. (1996). Destruction of mangrove wetlands-causes and consequences.
    Natural Resources and Development 43 - 44, 37-57.
Matheson, R.E. Jr. and Gilmore, R.G. Jr. (1995). Mojarras (Pisces: Gerreidae) of the
    Indian River Lagoon. Bulletin of Marine Science 57 (1), 281-285.
Mazda, Y., Sato, Y., Sawamoto, S., Yokochi, H. and Wolanski, E. (1990a). Links between
    physical, chemical and biological processes in Bashita-minato, a mangrove swamp
    in Japan. Estuarine, Coastal and Shelf Science 31 (6), 817-833.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      118


Mazda, Y., Yokochi, H., Sato, Y., Wolanski, E. and Boto, K.G. (1990b). Groundwater
    flow in the Bashita-Minato mangrove area, and its influence on water and bottom
    mud properties. Estuarine, Coastal and Shelf Science 31 (5), 621-638.
McClintock. B., Swenson, D., Trapido-Rosenthal, H. and Banghart, L. (1997).
    Ichthyodeterrent properties of lipophilic extracts from Bermudian sponges. Journal
    of Chemical Ecology 23 (6), 1607-1620.
McCoy, E.D., Mushinsky, H.R., Johnson, D. and Meshaka, W.E. Jr. (1996). Mangrove
    damage caused by Hurricane Andrew on the southwestern coast of Florida. Bulletin
    of Marine Science 59 (1), 1-8.
McGuinness, K.A. (1997a). Dispersal, establishment and survival of Ceriops tagal
    propagules in a north Australian mangrove forest. Oecologia 109 (1), 80-87.
McGuinness, K.A. (1997b). Seed predation in a tropical mangrove forest: a test of the
    dominance-predation model in northern Australia. Journal of Tropical Ecology 13
    (2), 293-302.
McKee, K.L. (1993). Soil physico-chemical patterns and mangrove species distribution:
    Reciprocal effects?. Journal of Ecology 81, 477-487.
McKee, K.L. (1995a). Seedling recruitment patterns in a Belizean mangrove forest: Effects
    of establishment ability and physico-chemical factors. Oecologia 101 (4), 448-460.
McKee, K.L. (1995b). Interspecific variation in growth, biomass partitioning and defensive
    characteristics of neotropical mangrove seedlings: response to light and nutrient
    availability. American Journal of Botany 82, 299-307.
McKee, K.L. (1995c). Mangrove species distribution and propagule predation in Belize: an
    exception to the Dominance-Predation Hypothesis. Biotropica 27 (3), 334-345.
McKee, K.L., Topa, M.A., Rygiewicz, P.T. and Cumming, J.R. (1996). Growth and
    physiological responses of neotropical mangrove seedlings to root zone hypoxia.
    Dynamics of physiological processes in woody roots. Tree Physiology 16 (11-12),
    883-889.
McKillup, S.C. and McKillup, R.V. (1997). An outbreak of the moth Achaea serva (Fabr.)
    on the mangrove Excoecaria agallocha (L.). Pan-Pacific Entomology 73, 184-185.
Medeiros, C.Q. and Kjerfve, B. (1993). Hydrology of a tropical estuarine system:
    Itamaraca, Brazil. Estuarine, Coastal and Shelf Science 36, 495-515.
Medina, E., Cuevas, E., Popp, M. and Lugo, A.E. (1990). Soil salinity, sun exposure, and
    growth of Acrostichum aureum, the mangrove fern. Botanical Gazette Chicago 151
    (1), 41-49.
Medina, E., Lugo, A.E. and Novelo, A. (1995). Mineral content of foliar tissues of
    mangrove species in Laguna de Sontecomapan (Veracruz, Mexico) and its relation
    to salinity. Biotropica 27 (3), 317-323.
Menasveta, P. (1997). Mangrove destruction and shrimp culture systems. World
    Aquaculture 28 (4), 36-42.
Menon, G.G. and Neelakantan, B. (1992). Chlorophyll and light attenuation from the
    leaves of mangrove species of Kali estuary. Indian Journal of Marine Sciences 21
    (1), 13-16.
Meyer, U., Hagen, W. and Medeiros, C. (1998). Mercury in a northeastern Brazilian
    mangrove area, a case study: Potential of the mangrove oyster Crassostrea
    rhizophorae as bioindicator for mercury. Marine Biology 131 (1), 113-121.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                     119


Micheli, F. (1993a). Effect of mangrove litter species and availability on survival,
    moulting, and reproduction of the mangrove crab Sesarma messa. Journal of
    Experimental Marine Biology and Ecology 171 (2), 149-163.
Micheli, F. (1993b). Feeding ecology of mangrove crabs in North Eastern Australia:
    Mangrove litter consumption by Sesarma messa and Sesarma smithii. Journal of
    Experimental Marine Biology and Ecology 171 (2), 165-186.
Micheli, F., Gherardi, F. and Vannini, M. (1991). Feeding and burrowing ecology of two
    East African mangrove crabs. Marine Biology 111 (2), 247-254.
Michener, W.K., Blood, E.R., Bildstein, K.L., Brinson, M.M. and Gardner, L.R. (1997).
    Climate change, hurricanes and tropical storms and rising sea level in coastal
    wetlands. Ecological Applications 7, 770-801.
Middelburg, J.J., Nieuwenhuize, J., Slim, F.J. and Ohowa, B. (1996). Sediment
    biogeochemistry in an East African mangrove forest (Gazi Bay, Kenya).
    Biogeochemistry 34 (3) 133-155.
Mildenhall, D.C. (1994). Early to mid Holocene pollen samples containing mangrove
    pollen from Sponge Bay, East Coast, North Island, New Zealand. Journal of the
    Royal Society of New Zealand 24 (2), 219-230.
Miller-Way, T. and Twilley, R.R. (1999). Oxygen and nutrient metabolism of a Carribean
    mangrove prop root community. Gulf Research Reports 11, 74.
Mitra, T.R. (1992). Odonata of the mangrove tidal forest of West Bengal, India.
    Odonatologia 3 (9), 141-143.
Miya, Y. (1991). Two alpheid shrimps (Crustacea: Decapoda) in mangal flats of Singapore
    and Darwin, NT, Australia. Zoological Science 8 (6), 1197p.
Mndeme, Y.E.S. (1995). Mafia marine resources in peril. Naga 18 (2), 12-13.
Mohamed, A.D. (1996) Mangrove forests: Valuable resources under the threat of
    development. Ocean Yearbook 12, 247-269.
Mohammed, S.M. and Johnstone, R.W. (1995). Spatial and temporal variations in water
    column nutrient concentrations in a tidally dominated mangrove creek: Chwaka
    Bay, Zanzibar. Ambio 24, 482-486.
Mohan, P.C., Rao, R.G. and Dehairs, F. (1997). Role of Godavari mangroves (India) in the
    production and survival of prawn larvae. Hydrobiologia 358, 317-320.
Mohan, R. and Siddeek, M.S.M. (1996). Marine habitat preference, distribution and
    growth of postlarvae, juvenile and pre adult Indian white shrimp, Penaeus indicus
    H. Milne Edwards, in Ghubat Hasish Bay, Gulf of Masirah, Sultanate of Oman.
    Fisheries Management Ecology 3 (2), 165-174.
Mohan, R., Selvam, V. and Azariah, J. (1995). Temporal distribution and abundance of
    shrimp postlarvae and juveniles in the mangroves of Muthupet, Tamil Nadu, India.
    Hydrobiologia 295, 183-191.
Mohanraju, R. and Natarajan, R. (1992). Methanogenic bacteria in mangrove sediments.
    Hydrobiologia 247, 187-193.
Monterrosa, O.E. (1991). Postlarval recruitment of the spiny lobster, Panulirus argus
    (Latreille), in southwestern Puerto Rico. In “Proceedings of the Fortieth Annual
    Gulf and Caribbean Fisheries Institute, Curacao” (G.T. Waugh and M.H. Goodwin,
    eds), Vol. 40, pp. 434-451. Netherlands, Antilles.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                     120


Moorthy, P. (1995). Effects of UV-B radiation on mangrove environment: Physiological
   responses of Rhizophora apiculata Blume. Ph.D. thesis, Annamalai University,
   India. 130 pp.
Moorthy, P. and Kathiresan, K. (1993). Physiological responses of mangrove seedlings to
   triacontanol. Biologia Plantarum 35 (4), 577-581.
Moorthy, P. and Kathiresan, K. (1997a). Photosynthetic pigments in tropical mangroves:
   Impacts of seasonal flux on UV-B radiation and other environmental attributes.
   Botanica Marina 40, 341-349.
Moorthy, P. and Kathiresan, K. (1997b). Influence of UV-B radiation on photosynthetic
   and biochemical characteristics of a mangrove Rhizophora apiculata Blume.
   Photosynthetica 34 (3), 465-471.
Moorthy, P. and Kathiresan, K. (1998). UV-B induced alterations in composition of
   thylakoid membrane and amino acids in leaves of Rhizophora apiculata Blume.
   Photosynthetica 35, (in press).
Morton, B. (1976). The biology, ecology and functional aspects of the organs of feeding
   and digestion of the S.E. Asian mangrove bivalve, Enigmonia aenigmatica
   (Mollusca: Anomiacea). Journal of Zoology, London 179, 437-466.

Morton, R.M. (1990). Community structure, density and standing crop of fishes in a
    subtropical Australian mangrove area. Marine Biology 105 (3), 385-394.
Morton, B. (1991). Notes on the first mangrove shipworm, Lyrodus singaporeana,
    recorded from Hong Kong. Journal of Molluscan Studies 57 (2) 295-296.
Mosisch, T.D. (1993). Effects of salinity on the distribution of Caloglossa leprieurii
    (Rhodophyta) in the Brisbane River, Australia. Journal of Phycology 29 (2), 147-
    153.
Mullin, S.J. (1995). Estuarine fish populations among red mangrove prop roots of small
    overwash islands. Wetlands 15 (4), 324-329.
Munoz, D., Guiliano, M., Doumenq, P., Jacquot, F., Scherrer, P. and Mille, G. (1997).
    Long term evolution of petroleum biomarkers in mangrove soil (Guadeloupe).
    Marine Pollution Bulletin 34 (11), 868-874.
Murphy, D.H. (1990a). The recognition of some insects assocciated with mangroves in
    Thailand. In “The recognition of some insects associated with mangroves in
    Thailand”. Mangrove Ecosystem Occasional Paper 7, pp. 15-23. UNDP /
    UNESCO, New Delhi.
Murphy, D.H. (1990b). The air-breathing arthropods of the mangrove system. In “Essays
    in Zoology” (L.M. Chou and P.K.L. Ng, eds), pp. 169-176. Department of Zoology,
    National University of Singapore.
Murphy, D.H. (1990c). Insects and public health in the mangrove ecosystem. In “Essays in
    Zoology” (L.M. Chou and P.K.L. Ng, eds), pp. 423-452. Department of Zoology,
    National University of Singapore.
Murphy, D.H. (1990d). The natural history of insect herbivory on mangrove trees in and
    near Singapore. The Raffles Bulletin of Zoology 38 (2), 119-203.
Murphy, D.H. and Sigurdsson, J.B. (1990). Birds, mangroves and man: prospects and
    promise of the new Sungei Buloh Bird Reserve. In “Essays in Zoology” (L.M.
    Chou and P.K.L. Ng, eds), pp. 223-244. Department of Zoology, National
    University of Singapore.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      121


Murren, C.J. and Ellison, A.M. (1996). Effects of habitat, plant size, and floral display on
    male and female reproductive success of the Neotropical orchid Brassavola
    nodosa. Biotropica 28 (1), 30-41.
Muthumbi, A., Verschelde, D. and Vincx, M. (1995). New Desmodoridae (Nematoda:
    Desmodoroidea): three new species from Ceriops mangrove sediments (Kenya) and
    one related new species from the North Sea. Cahiers de Biologie Marine 36 (3),
    181-195.
Nagarajan, R. and Thiyagesan, K. (1995). Avian mortality caused by a cyclone at the
    Pichavaram Mangroves, southern India. Pavo 33 (1-2), 117-121.
Nagy, E. and Kokay, J. (1991). Middle Miocene mangrove vegetation in Hungary. Acta
    Geologica Hungarica 34 (1-2), 45-52.
Naidoo, G. (1990). Effects of nitrate, ammonium and salinity on growth of the mangrove
    Bruguiera gymnorrhiza (L.) Lam. Aquatic Botany 38 (2-3), 209-219.
Naidoo, G. and Von-Willert, D.J. (1994). Stomatal oscillations in the mangrove Avicennia
    germinans. Functional Ecology 8 (5), 651-657.
Naidoo, G. and Von-Willert, D.J. (1995). Diurnal gas exchange characteristics and water
    use efficiency of three salt-secreting mangroves at low and high salinities.
    Hydrobiologia 295 (1-3), 13-22.
Nair, L.N., Rao, V.P. and Chaudhuri, S. (1991). Microflora of Avicennia officinalis Linn.
    In “Proceedings of the symposium on significance of mangroves” (A.D. Agate,
    S.D. Bonde and K.P.N. Kumaran, eds), pp. 52-55. Maharastra Association for
    Cultivation of Science and Research Institute, Pune, India.
Nakagiri, A. and Ito, T. (1994). Aniptodera salsuginosa, a new mangrove-inhabiting
    ascomycete, with observations on the effect of salinity on ascospore appendage
    morphology. Mycological Research 98 (8), 931-936.
Naranjo, L.G. (1997). A note on the birds of the Colombian Pacific mangroves. In “
    Mangrove Ecosystem Studies in Latin America and Africa” (B. Kjerfve, L.D.
    Lacerda and S. Diop, eds), pp. 64-70. UNESCO, Paris.
Nasakar, K. and Bakshi, D.N.G. (1993). The biological spectrum of the floral elements of
    the 24-Parganas district in West Bengal. Bangladesh Journal of Botany 22 (1), 15-
    20.
Newell, S.Y. and Fell, J.W. (1992a). Ergosterol content of living and submerged, decaying
    leaves and twigs of red mangrove. Canadian Journal of Microbiology 38, 979-982.
Newell, S.Y. and Fell, J.W. (1992b). Distribution and experimental responses to substrate
    of marine oomycetes (Halophytophthora spp.) in mangrove ecosystems.
    Mycological Research 96 (10), 851-856.
Newell, S.Y. and Fell, J.W. (1994). Parallel testing of media for measuring frequencies of
    occurrence for Halophytophthora spp. (Oomycota) from decomposing mangrove
    leaves. Canadian Journal of Microbiology 40 (4), 250-256.
Newell, S.Y. and Fell, J.W. (1995). Do Halophytophthoras (marine Phythiaceae) rapidly
    occupy fallen leaves by intraleaf mycelial growth? Canadian Journal of Botany 73
    (5), 761-765.
Newell, S.Y. and Fell, J.W. (1996). Cues for zoospore release by marine oomycotes in
    naturally decaying submerged leaves. Mycologia 88 (6), 934-938.
Newell S.Y, and Fell J.W. (1997). Competition among mangrove oomycotes, and between
    oomycotes and other microbes. Aquatic Microbial Ecology 12, 21-28.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      122


Newell, R.I.E., Marshall, N., Sasekumar, A. and Chong, V.C. (1995). Relative importance
    of benthic microalgae, phytoplankton, and mangroves as sources of nutrition for
    penaeid prawns and other coastal invertebrates from Malaysia. Marine Biology 123
    (3), 595-606.
Newkirk, G.F. and Richards, K. (1991). Effect of aerial exposure of Crassostrea
    rhizophorae spat on growth and survival during grow out. Journal of Shellfish
    Research 10 (1), 276.
Nicholas, W.L., Elek, J.A., Stewart, A.C. and Marples, T.G. (1991). The nematode fauna
    of a temperate Australian mangrove mudflat: its population density, diversity and
    distribution. Hydrobiologia 209 (1), 13-28.
Nielsen, M.G. (1997). Nesting biology of the mangrove, mud-nesting ant Polyrhachis
    sokolova Forel (Hymenoptera: Formicidae) in northern Australia. Insectes Sociaux
    44 (1), 15-21.
Nonaka, M., Sallah, N. and Kamura, T. (1994). Potential sulfate acidity on Ghanaian
    mangrove soils: effect of mangrove vegetation. Soil Microorganisms 44, 69-76.
Noske, R.A. (1993). Bruguiera hainesii: Another bird-pollinated mangrove? Biotropica 25
    (4), 481-483.
Noske, R.A. (1995). The ecology of mangrove forest birds in Peninsular Malaysia.
    Biotropica 137 (2), 250-263.
Noske, R.A. (1996). Abundance, zonation and foraging ecology of birds in mangroves of
    Darwin Harbour, Northern Territory. Wildlife Research 23 (4), 443-474.
Nwadukwe, F.O. (1995). Species abundance and seasonal variations in catch from two
    mangrove habitats in the Lagos Lagoon. Environment and Ecology 13 (1), 121-128.
Ogan, M.T. (1990). The nodulation and nitrogenase activity of natural stands of mangrove
    legumes in a Nigerian swamp. Plant and Soil 123 (1) 125-129.
Ogasawara, T., Ip, Y.K., Hasegawa, S., Hagiwara, Y. and Hirano, T. (1991). Changes in
    prolactin cell activity in the mudskipper, Periophthalmus chrysospilos, in response
    to hypotonic environment. Zoological Science 8 (1), 89-95.
Ohnishi, T. and Komiyama, A. (1998). Shoot and root formation on cut pieces of
    viviparous seedlings of a mangrove, Kandelia candel (L.) Druce. Forest Ecology
    and Management 102 (2-3), 173-178.
Olafsson, E. (1996). Meiobenthos in mangrove areas in eastern Africa with emphasis on
    assemblage structure of free-living marine nematodes. Hydrobiologia 312, 47-57.
Olmsted, I. and Gomez, J.M. (1996). Distribution and conservation of epiphytes on the
    Yucatan Peninsula. Selbyana 17 (1), 58-70.
Orihuela, B., Diaz, H. and Conde, J.E. (1991). Mass mortality in a mangrove roots fouling
    community in a hypersaline tropical lagoon. Biotropica 23 (4b), 592-601.
Orozco, A., Rada, F., Azocar, A. and Goldstein, G. (1990). How does a mistletoe affect the
    water, nitrogen and carbon balance of two mangrove ecosystem species? Plant Cell
    and Environment 13 (9), 941-948.
Osborn, J.G. and Polsenberg, J.F. (1996). Meeting of the mangrovellers: The interface of
    biodiversity and ecosystem function. Trends in Ecology and Evolution 11, 354-356.
Osborne, K. and Smith, T.J. III. (1990). Differential predation on mangrove propagules in
    open and closed canopy forest habitats. Vegetation 89 (1), 1-6.
Osore, M.K.W. (1992). A note on the zooplankton distribution and diversity in a tropical
    mangrove creek system, Gazi, Kenya. Hydrobiologia 247 (1-3), 119-120.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      123


Osunkoya, O.O. and Creese, R.G. (1997). Population structure, spatial pattern and seedling
    establishment of the grey mangrove, Avicennia marina var. australasica, in New
    Zealand. Australian Journal of Botany 45 (4), 707-725.
Oswin, S.D. and Kathiresan, K. (1994). Pigments in mangrove species of Pichavaram.
    Indian Journal of Marine Sciences 23 (1), 64-66.
Ott, J.A., Bright, M. and Schiemer, F. (1998). The ecology of a novel symbiosis between a
    marine peritrich ciliate and chemoautotrophic bacteria. Marine Ecology 19 (3),
    229-243.
Ovalle, A.R.C., Rezende, C.E., Lacerda, L.D., Silva, C.A.R., Wolanski, E. and Boto, K.G.
    (1990). Factors affecting the hydrochemistry of a mangrove tidal creek, Sepetiba
    Bay, Brazil. Estuarine, Coastal and Shelf Science 31 (5), 639-650.
Padmakumar, K. and Ayyakannu, K. (1997). Antiviral activity of marine plants. Indian
    Journal of Virology 13 (1), 33-36.
Padrón, C.M., Llorente, S.O. and Menendez, L. (1993). Mangroves of Cuba. In
    “Conservation and Sustainable Utilization of Mangrove Forests in Latin America
    and Africa Regions, Part I - Latin America” (L.D. Lacerda, ed), pp. 147-154.
    International Society for Mangrove Ecosystems, Okinawa, Japan.
Pain, S. (1996). Hostages of the deep. New Scientist 151 (2047), 38-42.
Pal, A.K. and Purkayastha, R.P. (1992a). Foliar fungi of mangrove ecosystem of
    Sunderbans, eastern India. Journal of Mycopathological Research 30 (2), 167-171.
Pal, A.K. and Purkayastha, R.P. (1992b). New parasitic fungi from Indian mangrove.
    Journal of Mycopathological Research 30 (2), 173-176.
Palaniselvam, V. (1998). Epiphytic cyanobacteria of mangrove: Ecological, physiological
    & biochemical studies and their utility as biofertilizer and shrimp feed, Ph.D.
    thesis., Annamalai University, India. 141 pp.
Panchanadikar, V.V. (1993). Studies of iron bacteria from a mangrove ecosystem in Goa
    and Konkan. International Journal of Environmental Studies 45 (1), 17-21.
Panitz, C.M.N. (1997). Ecological description of the Itacorubi mangroves, Ilha Santa
    Catarina, Brazil. In “Mangrove Ecosystem Studies in Latin America and Africa”
    (B. Kjerfve, L.D. Lacerda and S. Diop, eds), pp. 204-223. UNESCO, Paris.
Parani, M., Rao, C.S., Mathan, N., Anuratha, C.S., Narayanan, K.K. and Parida, A. (1997).
    Molecular phylogeny of mangroves. III. Parentage analysis of a Rhizophora hybrid
    using random amplified polymorphic DNA and restriction fragment length
    polymorphism markers. Aquatic Botany 58, 165-172.
Parkes, K.C. (1990). A revision of the mangrove vireo (Vireo pallens) (Aves: Vireonidae).
    Annals of the Carnegie Museum 59 (1), 49-60.
Parkinson, R.W., Delaune, R.D. and White, J.R. (1994). Holocene sea-level rise and the
    fate of mangrove forests within the wider Caribbean region. Journal of Coastal
    Research 10, 1077-1086.
Parrish, J.D. and Sherry, T.W. (1994). Sexual habitat segregation by American redstarts
    wintering in Jamaica: Importance of resource seasonality. Auk 111 (1), 38-49.
Passioura, J.B., Ball, M.C. and Knight, J.H. (1992). Mangroves may salinize the soil and in
    so doing limit their transpiration rate. Functional Ecology 6 (4), 476-481.
Pawlik, J.R., Charnas, B., Toonen, R.J. and Fenical, W. (1995). Defenses of Caribbean
    sponges against predatory reef fish. 1. Chemical deterrency. Marine Ecology
    Progress Series 127 (1-3), 183-194.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      124


Pearce, F. (1996). Living sea walls keep floods at bay. New Scientist 150, 7.
Pedroche, F.F., West, J.A., Zuccarello, G.C., Senties, A.G. and Karsten, U. (1995). Marine
    red algae of the mangroves in southern Pacific Mexico and Pacific Guatemala.
    Botanica Marina 38 (2), 111-119.
Pelegri, S.P. and Twilley, R.R. (1998). Heterotrophic nitrogen fixation (acetylene
    reduction) during leaf-litter decomposition of two mangrove species from South
    Florida, USA. Marine Biology 131 (1), 53-61.
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.
Peterson, M.S. (1990). Hypoxia-induced physiological changes in two mangrove swamp
    fishes: Sheepshead minnow, Cyprinodon variegatus Lacepede and Sailfin molly,
    Poecilia latipinna (Lesueur). Comparative Biochemistry and Physiology 97A (1),
    17-21.
Peterson, M.S. and Gilmore, R.G. Jr. (1991). Eco-physiology of juvenile snook
    Centropomus undecimalis (Bloch): Life-history implications. Bulletin of Marine
    Science 48 (1), 46-57.
Pezeshki, S.R., Delaune, R.D. and Meeder, J.F. (1997). Carbon assimilation and biomass
    partitioning in Avicennia germinans and Rhizophora mangle seedlings in response
    to soil redox conditions. Environmental and Experimental Botany 37 (2-3), 161-
    171.
Phang, S.M., Shaharuddin, S., Noraishah, H. and Sasekumar, A. (1996). Studies on
    Gracilaria changii (Gracilariales: Rhodophyta) from Malaysian mangroves.
    Hydrobiologia 326-327, 347-352.
Phillips, A., Lambert, G., Granger, J.E. and Steinke, T.D. (1994). Horizontal zonation of
    epiphytic algae associated with Avicennia marina (Forssk.) Vierh. pneumatophores
    at Beachwood Mangroves Nature Reserve, Durban, South Africa. Botanica Marina
    37 (6), 567-576.
Phillips, A., Lambert, G., Granger, J.E. and Steinke, T.D. (1996). Vertical zonation of
    epiphytic algae associated with Avicennia marina (Forssk.) Vierh. pneumatophores
    at Beechwood Mangroves Nature Reserve, Durban, South Africa. Botanica Marina
    39 (2), 167-175.
Pinto, L. and Punchihewa, N.N. (1996). Utilization of mangroves and seagrasses by fishes
    in the Negombo Estuary, Sri Lanka. Marine Biology 126 (2), 333-345.
Plaziat, J.C. (1995). Modern and fossil mangroves and mangals: their climatic and
    biogeographic variability. Geological Society Special Publication 83, 73-96.
Poch, G.K. and Gloer, J.B. (1991). Auranticins A and B: two new depsidones from a
    mangrove isolate of the fungus Preussia aurantiaca. Journal of Natural Products
    Lloydia 54 (1), 213-217.
Polania, J. (1990). Physiological adaptations in some species of mangroves. Acta Biologica
    Columbiana 2 (6), 23-36.
Pons, L.J. and Fiselier, J.L. (1991). Sustainable development of mangroves. Landscape
    and Urban Ecology 20 (1-3), 103-109.
Poovachiranon, S. and Chansang, H. (1994). Community structure and biomass of seagrass
    beds in the Andaman Sea. 1. Mangrove - associated seagrass beds. Research
    Bulletin of Phuket Marine Biological Centre 59, 53-64.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       125


Poovachiranon, S. and Tantichodok, P. (1991). The role of sesarmid crabs in the
    mineralization of leaf litter of Rhizophora apiculata in a mangrove, Southern
    Thailand. Research Bulletin of Phuket Marine Biological Centre 56, 63-74.
Popp, M., Polania, J. and Weiper, M. (1993). Physiological adaptations to different salinity
    levels in mangrove. In “Towards the Rational Use of High Salinity Tolerant
    Plants”(H. Lieth and A.A. Masoom, eds), Vol. 1, pp. 217-224. Kluwer Academic
    Publishers, Amsterdam.
Prather, J.W. and Cruz, A. (1995). Breeding biology of Florida prairie warblers and Cuban
    yellow warblers. Wilson Bulletin 107 (3), 475-484.
Premanathan, M., Chandra, K., Bajpai, S.K. and Kathiresan, K. (1992). A survey of some
    Indian marine plants for antiviral activity. Botanica Marina 35 (4), 321-324.
Premanathan, M., Kathiresan, K., Chandra, K. and Bajpai, S.K. (1993). Antiviral activity
    of marine plants against New Castle disease virus. Tropical Biomedicine 10, 31-33.
Premanathan, M., Kathiresan, K. and Chandra, K. (1994a). In vitro anti-vaccinia virus
    activity of some marine plants. Indian Journal of Medical Research 99, 236-238.
Premanathan, M., Kathiresan, K. and Chandra, K. (1994b). Anti-viral activity of marine
    and coastal plants from India. International Journal of Pharmacognosy 32, 330-
    336.
Premanathan, M., Kathiresan, K. and Chandra, K. (1995). Antiviral evaluation of some
    marine plants against Semliki forest virus. International Journal of Pharmacognosy
    33 (1), 75-77.
Premanathan, M., Nakashima, H., Kathiresan, K., Rajendran, N. and Yamamoto, N.
    (1996). In vitro anti-human immuno deficiency virus activity of mangrove plants.
    Indian Journal of Medical Research 130, 276-279.
Premanathan, M., Kathiresan, K. and Nakashima, H. (1999). Mangrove halophytes : A
    source of antiviral substances. South Pacific Study 19 (1-2), 49-57.
Primavera, J.H. (1995). Mangroves and brackish water pond culture in the Philippines.
    Hydrobiologia 295, 303-309.
Primavera, J.H. (1996). Stable carbon and nitrogen isotope ratios of penaeid juveniles and
    primary producers in a riverine mangrove in Guimaras, Philippines. Bulletin of
    Marine Science 58, 675-683.
Primavera, J.H. (1997). Fish predation on mangrove-associated penaeids: The role of
    structures and substrate. Journal of Experimental Marine Biology and Ecology 215,
    205-216.
Primavera, J.H. (1998). Mangroves as nurseries: shrimp populations in mangrove and non-
    mangrove habitats. Estuarine, Coastal and Shelf Science 46 (3), 457-464.
Primavera, J.H. and Lebata, J. (1995). Diel activity patterns in Metapenaeus and Penaeus
    juveniles. Hydrobiologia 295 (1-3), 295-302.
Proffitt, C.E. and Devlin, D.J. (1998). Are there cumulative effects in red mangroves from
    oil spills during seedling and sapling stages? Ecological Applications 8 (1) 121-
    127.
Proffitt, C.E., Devlin, D.J. and Lindsey, M. (1995). Effects of oil on mangrove seedlings
    grown under different environmental conditions. Marine Pollution Bulletin 30 (12),
    788-793.
Quevauviller, P., Donnard, O.F.X., Wasserman, J.C., Martin, F.M. and Schneider, J.
    (1992). Occurrence of methylated tin and dimethyl mercury compounds in a
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                    126


    mangrove core from Sepetiba Bay, Brazil. Applied Organometallic Chemistry 6,
    221-228.
Rafii, Z.A., Dodd, R.S. and Fromard, F. (1996). Biogeographic variation in foliar waxes of
    mangrove species. Biochemical Systematics Ecology 24, 341-345.
Raghukumar, C., Raghukumar, S., Chinnaraj, A., Chandramohan, D., D'Souza, T.M. and
    Reddy, C.A. (1994). Laccase and other lignocellulose modifying enzymes of
    marine fungi isolated from the coast of India. Botanica Marina 37 (6), 515-523.
Raghukumar, S. (1990). Speculations on niches occupied by fungi in the sea with relation
    to bacteria. Proceedings of Indian Academy of Sciences: Plant Sciences 100 (2),
    129-138.
Raghukumar, S. (1992). Bacterivory: A novel dual role for thraustochytrids in the sea.
    Marine Biology 113 (1), 165-169.
Raghukumar, S., Sharma, S., Raghukumar, C., Sathe Pathak, V. and Chandramohan, D.
    (1994). Thraustochytrid and fungal component of marine detritus. 4. Laboratory
    studies on decomposition of leaves of the mangrove Rhizophora apiculata Blume.
    Journal of Experimental Biology and Ecology 183 (1), 113-131.
Raine, R.M. (1994). Current land use and changes in land use over time in the coastal zone
    of Chanthaburi Province, Thailand. Biological Conservation 67 (3), 201-204.
Rajapandian, M.E., Gopinathan, C.P., Rodrigo, J.X. and Gandhi, A.D. (1990).
    Environmental characteristics of edible oyster beds in and around Tuticorin.
    Journal of Marine Biological Association of India 32 (1-2), 90-96.
Rajendran, N. (1997). Studies on mangrove - associated prawn seed resources of the
    Pichavaram, Southeast coast of India. Ph.D. thesis. Annamalai University, India.
    131 pp.
Rajendran, N. and Kathiresan, K. (1996). Effect of effluent from a shrimp pond on shoot
    biomass of mangrove seedlings. Aquaculture Research 27 (10), 745-747.
Rajendran, N. and Kathiresan, K. (1998). “Mangrove vegetation trap” technique for
    improving fishery resources in coastal waters. Current Science 75, 429.
Rajendran, N. and Kathiresan, K. (1999a). Seasonal occurrence of juvenile prawn and environmental
    factors in a Rhizophora magal, southeast coast of India. Hydrobiologia 394, 193-200.
Rajendran, N. and Kathiresan, K. (1999b). Do decomposing leaves of mangroves attract fishes? Current
    Science. Bangalore 77 (7) 972-976.
Ramamurthi, R., Jayasundaramma, B., Lakshmi Rajyam, C., Prasad, D.V.L.N. and
    Varalakshmi, C. (1991). Studies on marine bioactive substances from the Bay of
    Bengal: bioactive substances from the latex of the mangrove plant Excoecaria
    agallocha L., effects on the oxidative metabilism of crabs. In “Bioactive
    compounds from marine organisms with emphasis on the Indian Ocean” (M.F.
    Thompson, R. Sarojini and R. Nagabhushanam, eds), pp. 105-109. Oxford and
    IBH, Publishing Co, Pvt. Ltd., New Delhi.
Ramamurthy, T., Raju, R.M. and Natarajan, R. (1990). Distribution and ecology of
    methanogenic bacteria in mangrove sediments of Pitchavaram, east coast of India.
    Indian Journal of Marine Sciences 19 (4), 269-273.
Ramesh, M.X. and Kathiresan, K. (1992). Mangrove cholesterol in the diet of penaeid
    prawn Penaeus indicus. Indian Journal of Marine Sciences 21 (2), 164-166.
Ramirez-Garcia, P., Lopez-Blanco, J. and Ocana, D. (1998). Mangrove vegetation
    assessment in the Santiago River mouth Mexico, by means of supervised
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                     127


    classification using LandsatTM imagery. Forest Ecology and Management 105 (1-
    3), 217-229.
Rao, C.K., Chinnaraj, A., Inamdar, S.N. and Untawale, A.G. (1991). Arsenic content in
    certain marine brown algae and mangroves from Goa coast. Indian Journal of
    Marine Sciences 20 (4), 283-285.
Rao, C.S., Eganathan, P., Anand, A., Balakrishna, P. and Reddy, T.P. (1998). Protocol for
    in vitro propagation of Excoecaria agallocha L., a medicinally important mangrove
    species. Plant Cell Reports 17 (11), 861-865.
Rao, V.B., Rao, G.M.N., Sarma, G.V.S. and Rao, B.K. (1992). Mangrove environment and
    its sediment characters in Godavari estuary, east coast of India. Indian Journal of
    Marine Sciences 21 (1), 64-66.
Rasolofo, M.V. (1997). Use of mangroves by traditional fishermen in Madagascar.
    Mangroves and Salt Marshes 1, 243-253.
Ravi, A.V. and Kathiresan, K. (1990). Seasonal variation in gallotannin from mangroves.
    Indian Journal of Marine Sciences 19 (3), 224-225.
Ravikumar, D.R. and Vittal, B.P.R. (1996). Fungal diversity on decomposing biomass of
    mangrove plant Rhizophora in Pichavaram estuary, east coast of India. Indian
    Journal of Marine Sciences 25 (2), 142-144.
Ravikumar, S. (1995). Nitrogen-fixing azotobacters from the mangrove habitat and their
    utility as bio-fertilizers. Ph.D. thesis., Annamalai University, India. 120 pp.
Ravikumar, S. and Kathiresan, K. (1993). Influence of tannins, amino acids and sugars on
    Fungi of marine halophytes. Mahasagar 26, 21-24.
Ravishankar, J.P., Muruganandam, V. and Suryanarayanan, T.S. (1995). Isolation and
    characterization of melanin from a marine fungus. Botanica Marina 38 (5), 413-
    416.
Ravishankar, J.P., Muruganandam, V. and Suryanarayanan, T.S. (1996). Effect of salinity
    on amino acid composition of the marine fungus Cirrenalia pygmea. Current
    Science 70 (12), 1086-1087.
Read, S.J., Jones, E.B.G., Moss, S.T. and Hyde, K.D. (1995). Ultrastructure of asci and
    ascospores of two mangrove fungi: Swampomyces armeniacus and Marinosphaera
    mangrovei. Mycological Research 99 (12), 1465-1471.
Reddy, T.K.K., Rajasekhar, A., Jayasunderamma, B. and Ramamurthi, R. (1991). Studies
    on marine bioactive substances from the Bay of Bengal: Bioactive substances from
    the latex of mangrove plant Excoecaria agallocha L. - antimicrobial activity and
    degradation. In “Bioactive compounds from marine organisms with emphasis on
    the Indian Ocean” (M.F. Thompson, R. Sarojini and R. Nagabhushanam, eds), pp.
    75-78. Oxford & IBH Publishers Co. Pvt. Ltd. New Delhi.
Rey, J.R., Shaffer, J., Tremain, D., Crossman, R.A. and Kain, T. (1990a). Effects of re-
    establishing tidal connections in two impounded subtropical marshes on fishes and
    physical conditions. Wetlands 10 (1), 27-45.
Rey, J.R., Crossman, R.A. and Kain, T.R. (1990b). Vegetation dynamics in impounded
    marshes along the Indian River Lagoon, Florida, USA. Environmental Management
    14 (3), 396-410.
Rey, J.R., Kain, T. and De-Freese, D.E. (1992). Observations on the feeding behaviour and
    local distribution of Vallentinia gabriellae (Hydrozoa: Olindiidae): A new record
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      128


    from mangrove wetlands of the Indian River Lagoon, Florida. Wetlands 12 (3),
    225-229.
Rezende, C.E., Lacerda, L.D., Ovalle, A.R.C., Silva, C.A.R. and Martinelli, L.A. (1990).
    Nature of POC transport in a mangrove ecosystem; a carbon stable isotopic study.
    Estuarine, Coastal and Shelf Science 30 (6), 641-645.
Richards, G.C. (1990). The spectacled flying-fox, Pteropus conspicillatus, (Chiroptera:
    Pteropodidae) in north Queensland (Australia): 1. Roost sites and distribution
    patterns. Australian Mammology 13 (1-2), 17-24.
Ricklefs, R.E. and R.E. Latham. (1993). Global patterns of species
    diversity in mangrove floras. In “Species diversity in ecological
    communities” (Ricklefs, R.E. and D. Schluter, eds), pp. 215-229. University of
    Chicago Press, Chicago.
Rico-Gray, V. (1993). Origen y rutas de dispersion de los mangles: Une revision con
    enfasis en las especies de America. Acta Botanica Mexicana 25, 1-13.
Rico-Gray, V. and Palacios-Rios, M. (1996a). Leaf area variation in Rhizophora mangle L.
    Rhizophoraceae) along a latitudinal gradient in Mexico. Global Ecology and
    Biogeography Letters 5 (1), 30-35.
Rico-Gray, V. and Palacios-Rios, M. (1996b). Salinity and water level as factors in the
    distribution of vegetation in the marshes of NW Campeche, Mexico. Acta Botanica
    Mexicana 34, 53-61.
Ridd, P.V. (1996). Flow through animal burrows in mangrove creeks. Estuarine, Coastal
    and Shelf Science 43 (5), 617-625.
Ridd, P.V. and Sam, R. (1996). Profiling groundwater salt concentrations in mangrove
    swamps and tropical salt flats. Estuarine, Coastal and Shelf Science 43 (5), 627-
    635.
Ridd, P.V., Wolanski, E. and Mazda, Y. (1990). Longitudinal diffusion in mangrove-
    fringed tidal creeks. Estuarine, Coastal and Shelf Science 31 (5), 541-554.
Rinehart, K.L., Holt, T.G., Fregeau, N.L., Stroh, J.G., Kiefer, P.A., Sun, F., Li, L.H. and
    Martin, D.G. (1990). Ecteinascidins 729, 743, 745, 759A, 759B, and 770: Potent
    antitumor agents from the Caribbean tunicate Ecteinascidia turbinata. Journal of
    Organic Chemistry 55, 4512-4515.
Ritchie, S.A. and Addison, D.S. (1992). Oviposition preferences of Aedes taeniorhynchus
    (Diptera: Culiocidae) in Florida mangrove forests. Environment and Entomology 21
    (4), 737-744.
Ritchie, S.A. and Laidlaw Bell, C. (1994). Do fish repel oviposition by Aedes
    taeniorhynchus? Journal of American Mosquito Control Association 10 (3), 380-
    384.
Rivail, D.M., Lamotte, M., Donard, O.F.X., Soriano-Sierra, E.J. and Robert, M. (1996).
    Metal contamination in surface sediments of mangroves, lagoons and Southern Bay
    in Florianopolis Island. Environmental Technology 17 (10), 1035-1046.
Rivera-Monroy, V.H. and Twilley, R.R. (1996). The relative role of denitrification and
    immobilization in the fate of inorganic nitrogen in mangrove sediments (Terminos
    Lagoon, Mexico). Limnology and Oceanography 41 (2), 284-296.
                                             129


Rivera-Monroy, V.H., Twilley, R.R., Boustany, R.G., Day, J.W. and Vera-Herrera, F.
    (1995). Direct denitrification in mangrove sediments in Terminos Lagoon, Mexico.
    Marine Ecology Progress Series 126, (1-3), 97-109.
Robertson, A.I. (1991). Plant-animal interactions and the structure and function of mangrove
    forest ecosystems. Australian Journal of Ecology 16, 433-443.
Robertson, A.I. and Alongi, D.M. (1992). Tropical Mangrove Ecosystems. 329 pp. American
    Geophysical Union, Washington DC., USA.
Robertson, A.I. and Alongi, D.M. (1995). Role of riverine mangrove forests in organic
    carbon export to the tropical coastal ocean; a preliminary mass balance for the Fly
    Delta (Papua New Guinea). Geo-Marine Letters 15 (3-4), 134-139.
Robertson, A.I. and Blaber, S.J.M. (1992). Plankton, epibenthos and fish communities. In
    “Tropical Mangrove Ecosystem” (A.I. Robertson and D.M. Alongi, eds), pp. 173-
    224. American Geophysical Union, Washington DC, USA.
Robertson, A.I. and Duke, N.C. (1990a). Mangrove fish-communities in tropical Queensland,
    Australia: Spatial and temporal patterns in densities, biomass and community
    structure. Marine Biology 104 (3), 369-379.
Robertson, A.I. and Duke, N.C. (1990b). Recruitment, growth and residence time of fishes in
    a tropical Australian mangrove system. Estuarine, Coastal and Shelf Science 31 (5),
    723-743.
Robertson, A.I. and Phillips, M.J. (1995). Mangroves as filters of shrimp pond effluent:
    Predictions and biogeochemical research needs. Hydrobiologia 295 (1-3), 311-
    321Robertson, A.I., Giddins, R. and Smith, T.J. (1990). Seed predation by insects in
    tropical mangrove forests: Extent and effects on seed viability and the growth of
    seedlings. Oecologia 83 (2), 213-219.
Robertson, A.I., Daniel, P.A. and Dixon, P. (1991). Mangrove forest structure and
    productivity in the Fly River estuary, Papua New Guinea. Marine Biology 111, 147-
    155.
Robertson, A.I., Alongi, D.M. and Boto, K.G. (1992). Concluding remarks: research and
    mangrove conservation. In “Tropical mangrove ecosystem” (A.I. Robertson and D.M.
    Alongi, eds), pp. 293-326. American Geophysical Union, Washington DC., USA.
Rodriguez, C. and Stoner, A.W. (1990). The epiphyte community of mangrove roots in a
    tropical estuary: distribution and biomass. Aquatic Botany 36 (2), 117-126.
Rollet, B. (1981). Bibliography on mangrove research 1600-1975. UNESCO, U.K. 479 pp.
Rooker, J.R. (1995). Feeding ecology of the schoolmaster snapper, Lutjanus apodus
    (Walbaum), from southwestern Puerto Rico. Bulletin of Marine Science 56 (3), 881-
    894.
Rooker, J.R. and Dennis, G.D. (1991). Diel, lunar and seasonal changes in a mangrove fish
    assemblage of southwestern Puerto Rico. Bulletin of Marine Science 49 (3), 684-698.
Ross, P.M. (1996). An examination of differences in morphology and reproduction of the
    barnacles Elminius covertus and Hexaminius spp. from mangrove forests in nthe
    Sydney region, New South Wales. Marine and Freshwater Research 47, 715-721
Ross, P.M. and Underwood, A.J. (1997). The distribution and abundance of barnacles in a
    mangrove forest. Australian Journal of Ecology 22 (1), 37-47.
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 (3), 375-384.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                         130


Roth, L.C. (1997). Implications of periodic hurricane disturbance for sustainable
    management of Caribbean mangroves. In “ Mangrove Ecosystem Studies in Latin
    America and Africa” (B. Kjerfve, L.D. Lacerda and S. Diop, eds), pp. 18-33.
    UNESCO, Paris.
Roy, S.D. (1997). Study of litterfall and its decomposition in a mangrove stand, South
    Andaman. Journal of the Andaman Science Association 13 (1-2), 119-121.
Ruitenbeek, H.J. (1994). Modelling economy ecology linkages in mangroves: economic
    evidence for promoting conservation in Bintuni Bay, Indonesia. Ecological
    Economics 10 (3), 233-247.
Rützler, K. (1995). Low-tide exposure of sponges in a Caribbean mangrove community.
    Marine Ecology 16 (2), 165-179.
Rützler, K. and Feller, I.C. (1996). Caribbean mangrove swamps. Scientific American 274,
    94-99.
Ruwa, R.K. (1990). The effects of habitat complexities created by mangroves on
    macrofaunal composition in brackish water intertidal zones at the Kenya coast.
    Discovery and Innovation 2, 49-55.
Ruwa, R.K. (1997). Zonation of crabs that burrow or bury in mangrove vegetation soils on
    the east coast of Kenya. In “Mangrove Ecosystem Studies in Latin America and
    Africa” (B. Kjerfve, L.D. Lacerda and S. Diop, eds), pp. 316-324. UNESCO, Paris.
Ruwa, R.K. and Polk, P. (1994). Patterns of spat settlement recorded for tropical oyster
    Crassostrea cucullata (Born 1778) and the barnacle, Balanus amphitrite (Darwin
    1854) in a mangrove creek. Tropical Zoology 7 (1), 121-130.
Sadaba, R.B., Vrijmoed, L.L.P., Jones, E.B.G. and Hodgkins, I.J. (1995). Observations on
    vertical distribution of fungi associated with standing senescent Acanthus ilicifolius
    stems at Mai Po Mangrove, Hong Kong. Hydrobiologia 295 (1-3), 119-126.
Sadiq, M. and Zaidi, T.H. (1994). Sediment composition and metal concentrations in
    mangrove leaves from the Saudi coast of the Arabian Gulf. Science of the Total
    Environment 155 (1), 1-8.
Saenger, P. (1998). Mangrove vegetation: An evolutionary perspective. Marine and
    Freshwater Research 49 (4), 277-286.
Saenger, P. and Bellan, M.F. (1995). The mangrove vegetation of the Atlantic coast of
    Africa: A review, 96 pp. Universite de Toulouse Press, Toulouse.
Saenger, P. and Siddiqi, N.A. (1993). Land from the sea: The mangrove afforestation
    program of Bangladesh. Ocean and Coastal Management 20 (1), 23-39.
Saenger, P. and Snedaker, S.C. (1993). Pantropical trends in mangrove above-ground
    biomass and annual litter fall. Oecologia 96, 293-299.
Saha, S. and Choudhury, A. (1995). Vegetation analysis of restored and natural mangrove
    forest in Sagar Island, Sunderbans, east coast of India. Indian Journal of Marine
    Sciences 24, 133-136.
Saifullah, S.M. and Elahi, E. (1992). Pneumatophore density and size in mangroves of
    Karachi, Pakistan. Pakistan Journal of Botany 24 (1), 5-10.
Saifullah, S.M., Shaukat, S.S. and Shams, S. (1994). Population structure and dispersion
    patterns in mangroves of Karachi, Pakistan. Aquatic Botany 47 (3-4), 329-340.
Saintilan, N. (1997). Above- and below-ground biomass of mangroves in a sub-tropical
    estuary. Marine and Freshwater Research 48 (7), 601-604.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        131


Saito, T., Yamanoi, T., Morohoshi, F. and Shibata, H. (1995). Discovery of mangrove plant
    pollen from the "Shukunohora Sandstone facies", Akeyo Formation, Mizunami Group
    (Miocene), Gifu Prefecture, Japan. Journal of the Geological Society of Japan 101
    (9), 747-749.
Sakai, R., Rinehart, K.L., Guan, Y. and Wang, A.H.J. (1992). Additional antitumor
    ecteinascidians from a Caribbean tunicate: crystal structures and activities in vivo.
    Proceedings of the National Academy of Science USA 89, 11456-11460.
Salini, J.P., Blaber, S.J.M. and Brewer, D.T. (1990). Diets of piscivorous fishes in a tropical
    Australian estuary with particular reference to predation on penaeid prawns. Marine
    Biology 105, 363 - 374.
Sanches, A.K. and Camargo, A.F.M. (1995). Effects of organic pollution in a mangrove of
    Cananeia island: Evidences from physical and chemical variables and composition of
    zooplankton. Naturalia 20, 125-133.
Santhakumari, V. (1991). Destruction of mangrove vegetation by Sphaeroma terebrans along
    Kerala coast. Fisheries Technology Society Fish Technology Kochi 28 (1), 29-32.
Santhakumaran, L.N. and Sawant, S.G. (1994). Observations on the damage caused by
    marine fouling organisms to mangrove saplings along Goa coast. Journal of the
    Timber Development Association of India 40 (1), 5-13.
Santhakumaran, L.N., Chinnaraj, A. and Sawant, S.G. (1994). Lignicolous marine fungi from
    panels of different timbers exposed along Goa coast (India). Journal of the Timber
    Development Association of India 40 (2), 9-19.
Santhakumaran, L.N., Remadevi, O.K. and Sivaramakrishnan, V.R. (1995). A new record of
    the insect defoliator, Pteroma plagiophleps Hamp. (Lepidoptera: Psychidae) from
    mangroves along the Goa coast (India). The Indian Forester 121 (2), 153-155.
Santra, S.C., Pal, U.C. and Choudhury, A. (1991). Marine phytoplankton of the mangrove
    delta region of West Bengal. Journal of Marine Biological Association of India 33 (1-
    2), 292-307.
Sasamoto, H., Wakita, Y. and Baba, S. (1997). Effect of high sorbitol concentration on
    protoplast isolation from cotyledons of mangroves, Avicennia marina and A. lanata.
    Plant Biotechnology 14 (2), 101-104.
Sasekumar, A. and Chong, V.C. (1998). Faunal diversity in Malaysian mangroves. Global
    Ecology and Biogeography Letters 7 (1), 57-60.
Sasekumar, A., Chong, V.C., Leh, M.U. and D' Cruz, R. (1992). Mangroves as a habitat for
    fish and prawns. Hydrobiologia 247 (1-3), 195-207.
Sasekumar, A., Chong, V.C., Lim, K.H. and Singh, H.R. (1994). The fish community of
    Matang mangrove waters, Malaysia. Hydrobiologia 2, 457-464.
Sayed, O.H. (1995). Effects of the expected sea level rise on Avicennia marina L: A case
    study in Qatar. Qatar University Science Journal 15 (1), 91-94.
Scholander, P.F., Hammel, H.T., Hemmingsen, E.A. and Cray, W. (1962). Salt balance in
    mangroves. Plant Physiology 37, 722-729.
Scholander, P.F., Hammel, H.T., Hemmingsen, E.A. and Bradstreet, E.D. (1964).
    Hydrostatic pressure and osmotic potential in leaves of mangroves and some other
    plants. Proceedings of National Academy of Sciences, USA 52, 119-125.
Scholander, P.F., Hammel, H.T., Bradstreet, E.D. and Hemmingsen, E.A. (1965). Sap
    pressure in vascular plants. Science 148, 339-346.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      132


Schrijvers, J., Okondo, J.P., Steyaert, M. and Vincx, M. (1995). The influence of epibenthos
    on the meiobenthos of a Ceriops tagal mangrove sediment at Gazi Bay, Kenya.
    Marine Ecology Progress Series 128, 247-259.
Schrijvers, J., Schallier, R., Silence, J., Okondo, J.P. and Vincx, M. (1997). Interaction
    between epibenthos and meiobenthos in a high intertidal Avicennia marina mangrove
    forest. Mangroves and Salt Marshes 1, 137-154.
Schrijvers, J., Camargo, M.G., Pratiwi, R. and Vincx, M. (1998). The infaunal macrobenthos
    under East African Ceriops tagal mangroves impacted by epibenthos. Journal of
    Experimental Marine Biology and Ecology 222 (1-2), 175-193.
Schwamborn, R. and Saint-Paul, U. (1996). Mangroves - Forgotten Forests? Natural
    Resources and Development 43-44, 13-36.
Sedberry, G.R. and Carter, J. (1993). The fish community of a shallow tropical lagoon in
    Belize, Central America. Estuaries 16 (2), 198-215.
Seeman, O.D. and Walter, D.E. (1995). Life history of Afrocypholaelaps africana (Evans)
    (Acari: Ameroseiidae), a mite inhabiting mangrove flowers and phoretic on
    honeybees. Journal of the Australian Entomological Society 34 (1), 45-50.
Selvam, V. and Ravichandran, K.K. (1998). Restoration of degraded mangrove wetlands: a
    case study of Pichavaram (India). In “International Symposium on Mangrove
    Ecology and Biology”, April 25-27, 1998, Kuwait. p. 16.
Selvam, V., Mohan, R., Ramasubramanian, R. and Azariah, J. (1991). Plant communities and
    soil properties of three mangrove stands of Madras Coast. Indian Journal of Marine
    Sciences 20, 67-69.
Selvam, V., Azariah, J. and Azariah, H. (1992). Diurnal variation in Physical - chemical
    properties and primary production in the interconnected marine, mangrove and
    freshwater biotopes of Kakinada coast, Andhra Pradesh, India. Hydrobiogia 247, 181-
    186.
Semeniuk, V. (1994). Predicting the effect of sea-level rise on mangroves in northwestern
    Australia. Journal of Coastal Research 10 (4), 1050-1076.
Semesi, A.K. (1992). Developing management plans for the mangrove forest reserves of
    mainland Tanzania. Hydrobiologia 247 (1-3), 1-10.
Semesi, A.K. (1998). Mangrove management and utilization in eastern Africa. Ambio 27,
    620-626.
Sengupta, A. and Choudhuri, S. (1990). Halotolerant Rhizobium strains from mangrove
    swamps of the Ganges River Delta. Indian Journal of Microbiology 30 (4), 483-484.
Sengupta, A. and Choudhuri, S. (1991). Ecology of heterotrophic dinitrogen fixation in the
    rhizosphere of mangrove plant community at the Ganges River estuary in India.
    Oecologia 87 (4), 560-564.
Sengupta, A. and Choudhuri, S. (1994). A typical root endophytic fungi of mangrove plant
    community of Sunderban and their possible significance as mycorrhiza. Journal of
    Mycopathological Research 32 (1), 29-39.
Sethuramalingam, S. and Ajmal Khan, S. (1991). Brachyuran crabs of Parangipettai coast,
    CAS in Marine Biology publication, Annamalai University, India. 92 pp.
Sharma, M. and Garg, H.S. (1996). Iridoid glycosides from Avicennia officinalis. Indian
    Journal of Chemistry 35 (5), 459-462.
Sheridan, R.P. (1991). Epicaulous, nitrogen-fixing microepiphytes in a tropical mangal
    community, Guadeloupe, French West Indies. Biotropica 23, 530-541.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        133


Sheridan, P.F. (1992). Comparative habitat utilization by estuarine macrofauna within the
    mangrove ecosystem of Rookery Bay, Florida. Bulletin of Marine Science 50 (1), 21-
    39.
Sheridan, P.F. (1997). Benthos of adjacent mangrove, seagrass and non-vegetated habitats in
    Rookery Bay, Florida. USA Estuarine, Coastal and Shelf Science 44, 455-469.
Sherman, R.E., Fahey, T.J. and Howarth, R.W. (1998). Soil-plant interactions in a
    neotropical mangrove forest: Iron, phosphorus and sulfur dynamics. Oecologia 115
    (4), 553-563.
Shome, R., Shome, B.R., Mandal, A.B. and Bandopadhyay, A.K. (1995). Bacterial flora in
    mangroves of Andaman. Part 1: Isolation, identification and antibiogram studies.
    Indian Journal of Marine Sciences 24, 97-98.
Shoreit, A.A.M., El Kady, J.A. and Sayed, W.F. (1994). Isolation and identification of purple
    nonsulfur bacter of mangal and non-mangal vegetation of Red Sea coast, Egypt.
    Limnologica 24 (2), 177-183.
Siddiqi, N.A. (1995). Site suitability for raising Nypa fruticans plantations in the Sunderban
    mangroves. Journal of Tropical Forest Science 7 (3), 405-411.
Siddiqi, N.A. (1997). Management of Resources in the Sunderbans Mangroves of
    Bangladesh. International News letter of coastal Management - Intercoast Network.
    Special edition 1, 22-23.
Siddiqi, N.A. and Hussain, K.Z. (1994). The impact of deer on natural regeneration in the
    Sunderbans mangrove forest of Bangladesh. Bangladesh Journal of Zoology 22 (2),
    223-234.
Siddiqi, N.A. and Khan, M.A.S. (1996). Planting techniques for mangroves on new
    accretions in the coastal areas of Bangladesh. In “Restoration of Mangrove
    Ecosystems” (C. Field, ed), pp. 143-159. International Tropical Timber Organization
    and International Society for Mangrove Ecosystems, Okinawa, Japan.
Sigurdsson, J.B. (1991). A nudibranch: Murphydoris singaporensis a new genus and species
    from Singapore mangroves (Gastropoda: Opistobranchia: Goniodoriidae). The Raffles
    Bulletin of Zoology 39 (1), 259-263.
Sigurdsson, J.B. and Sundari, G. (1990). Colour changes in the shell of the tree-climbing
    bivalve Enigmonia aenigmatica (Holten, 1802) (Anomidae). The Raffles Bulletin of
    Zoology 38 (2), 213-218.
Silva, C.A.R. and Mozeto, A.A. (1997). Release and retention of phosphorus in mangrove
    sediment: Sepetiba Bay, Brazil. In “Mangrove Ecosystem Studies in Latin America
    and Africa” (B. Kjerfve, L.D. Lacerda and S. Diop, eds), pp. 179-190. UNESCO,
    Paris.
Silva, C.A.R., Lacerda, L.D. and Rezende, C.E. (1990). Heavy metal reservoirs in a red
    mangrove forest. Biotropica 22, 339-345.
Simpson, J.H., Gong, W.K. and Ong, J.E. (1997). The determination of the net fluxes from a
    mangrove estuary system. Estuaries 20 (1), 103-109.
Singh, N. and Steinke, T.D. (1992). Colonization of decomposing leaves of Bruguiera
    gymnorrhiza (Rhizophoraceae) by fungi, and in vitro cellulolytic activity of the
    isolates. African Journal of Botany 58 (6), 525-529.
Sivakumar, A. (1992). Studies on wood biodeterioration of mangroves. Ph.D. thesis,
    Annamalai University, India. 55 pp.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      134


Sivakumar, A. and Kathiresan, K. (1990). Phylloplane fungi from mangroves. Indian Journal
    of Microbiology 30 (2), 229-231.
Sivakumar, A. and Kathiresan, K. (1996). Mangrove wood bored by molluscs, southeastern
    coast of India. Phuket Marine Biological Centre Special Publication 16, 211-214.
Sivasothi, N. and Burhanuddin H.M.N. (1994). A review of otters (Carnivora: Mustelidae:
    Lutrinae) in Malaysia and Singapore. Hydrobiologia 285, 151-170.
Skelton, N.J. and Allaway, W.G. (1996). Oxygen and pressure changes measured in situ
    during flooding in roots of the grey mangrove Avicennia marina (Forssk.) Vierh.
    Aquatic Botany 54 (2-3), 165-175.
Skilleter, G.A. (1996). Validation of rapid assessment of damage in urban mangrove forests
    and relationships with molluscan assemblages. Journal of the Marine Biological
    Association of the United Kingdom 76 (3), 701-716.
Slim, F.J., Gwada, P.M., Kodjo, M. and Hemminga, M.A. (1996). Biomass and litterfall of
    Ceriops tagal and Rhizophora mucronata in the mangrove forest of Gazi Bay, Kenya.
    Marine and Freshwater Research 47 (8), 999-1007.
Slim, F.J., Hemminga, M.A., Ochieng, C., Jannink, N.T., Cocheret De La Moriniere, E. and
    Van Der Velde, G. (1997). Leaf litter removal by the snail Terebralia palustris
    (Linnaeus) and sesarmid crabs in an East African mangrove forest (Gazi Bay, Kenya).
    Journal of Experimental Marine Biology and Ecology 215 (1), 35-48.
Smith, A.H. and Berkes, F. (1993). Community-based use of mangrove resources of St.
    Lucia. International Journal of Environmental Studies 43, 123-131.
Smith, P.T. (1996). Physical and chemical characteristics of sediments from prawn farms and
    mangrove habitats on the Clarence River, Australia. Aquaculture 146 (1-2), 47-83.
Smith, S.M. and Snedaker, S.C. (1995a). Salinity responses in two populations of viviparous
    Rhizophora mangle L. seedlings. Biotropica 27 (4), 435-440.
Smith, S.M. and Snedaker, S.C. (1995b). Developmental responses of established red
    mangrove, Rhizophora mangle L., seedlings to relative levels of photosynthetically
    active and ultraviolet radiation. Florida Scientist 58 (1), 55-60.
Smith, S.M., Snedaker, S.C. and Davenport, T.L. (1995). Observations on the effects of
    naphthalene acetic acid and gibberellic acid on the floating behaviour and early
    development of red mangrove (Rhizophora mangle L.) seedlings. Tropical Ecology
    36 (1), 129-135.
Smith, S.M., Yang, Y.Y., Kamiya, Y. and Snedaker, S.C. (1996). Effect of environment and
    gibberellins on the early growth and development of the red mangrove, Rhizophora
    mangle L.. Plant Growth Regulation 20 (3), 215-223.
Smith, T.J. III. (1992). Forest structure. In “Tropical mangrove ecosystems” (A.I. Robertson
    and D.M. Alongi, eds), pp. 101-136. American Geophysical Union, Washington DC.,
    USA.
Smith, T.J. III., Boto, K.G., Frusher, S.D. and Giddins, R.L. (1991). Keystone species and
    mangrove forest dynamics: The influence of burrowing by crabs on soil nutrient and
    forest productivity. Estuarine, Coastal and Shelf Science 33, 419-432.
Smith, T.J., III, Robblee, M.B., Wanless, H.R. and Doyle, T.W. (1994). Mangroves,
    hurricanes, and lightning strikes: Assessment of Hurricane Andrew suggests an
    interaction across two differing scales of disturbance. Bioscience 44 (4), 256-262.
Snedaker, S.C. (1995). Mangroves and climate change in the Florida and Caribbean region:
    Scenarios and hypotheses. Hydrobiologia 295 (1-3), 43-49.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       135


Snedaker, S.C. and Araújo, R.J. (1998). Stomatal conductance and gas exchange in four
    species of Caribbean mangroves exposed to ambient and increased CO2. Marine and
    Freshwater Research 49, 325–327.
Snedaker, S.C. and Snedaker J.G. (1984). The mangrove ecosystem: research methods. In
    “The mangrove ecosystem: research methods” (S.C. Snedaker and J.G. Snedaker,
    eds), pp. 251. UNESCO, Paris.
Snedaker, S.C., Brown, M.S., Lahmann, E.J. and Araujo, R.J. (1992). Recovery of mixed
    species mangrove forest in South Florida following canopy removal. Journal of
    Coastal Research 8 (4), 919-925.
Snedaker, S.C., Meeder, J.F., Ross, M.S. and Ford, R.G. (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. Journal of Coastal Research 10 (2), 497-498.
Soares, C.A.G., Maury, M., Pagnocca, F.C., Araujo, F.V., Mendonca-Hagler, L.C. and
    Hagler, A.N. (1997). Ascomycetous yeasts from tropical intertidal dark mud of
    southeast Brazilian estuaries. Journal of General and Applied Microbiology 43 (5),
    265-272.
Somboon, J.R.P. (1990). Palynological study of mangrove and marine sediments of the Gulf
    of Thailand. Journal of Southeast Asian Earth Science 4 (2), 85-97.
Somero, G.N., Childress, J.J. and Anderson, A.E. (1989). Transport, metabolism and
    detoxification of hydrogen sulfide in animals from sulfide-rich marine environments.
    CRC Critical Reviews in Aquatic Sciences 1, 591-614.
Soto, R. (1992). Nutrient concentration and retranslocation in coastal vegetation and
    mangroves from the Pacific coast of Costa Rica. Brenesia 37, 33-50.
Sotomayor, D., Corredor, J.E. and Morell, J.M. (1994). Methane flux from mangrove
    sediments along the southwestern Coast of Puerto Rico. Estuaries 17 (1B), 140-147.
Spalding, M. (1997). The global distribution and status of mangrove ecosystems.
    International News Letter of Coastal Management-Intercoast Network, Special
    edition 1, 20-21.
Spangler, P.J. (1990). A new species of halophilous water-strider, Mesovelia polhemusi, from
    Belize and a key and checklist of New World species of the genus (Heteroptera:
    Mesoveliidae). Proceedings of Biology Society Washington 103 (1), 86-94.
Spratt, H.G., Jr. and R.E. Hodson (1994). The effect of changing water chemistry on rates of
    manganese oxidation in surface sediments of a temperate saltmarsh and a tropical
    mangrove estuary. Estuarine, Coastal and Shelf Science 38 (2), 119-135.
Srivastava, S.K. and Binda, P.L. (1991). Depositional history of the Early Eocene Shumaysi
    Formation, Saudi Arabia. Palynology 15, 47-61.
Staples, D.J., Vance, D.J. and Loneragan, N.R. (1995). Penaeid prawn recruitment
    variability: effect of the environment. In “Proceedings of the workshop on spawning
    Stock-Recruitment Relationships (SRRs) in Australian crustacean fisheries” (A.J.
    Courtney and M.G. Cosgrove, eds), June 1, 1994, pp. 41-50. Department of Primary
    Industries, Brisbane, Queensland, Australia.
Staus, N.L. (1998). Habitat use and home range of West Indian whistling-ducks. Journal of
    Wildlife Management 62 (1), 171-178.
Steinke, T.D. and Jones, E.B.G. (1993). Marine and mangrove fungi from the Indian Ocean
    coast of South Africa. South African Journal of Botany 59 (4), 385-390.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       136


Steinke, T.D. and Naidoo, Y. (1990). Biomass of algae epiphytic on pneumatophores of the
    mangrove, Avicennia marina, in the St. Lucia estuary. South African Journal of
    Botany 56, 226-232.
Steinke, T.D. and Naidoo, Y. (1991). Respiration and net photosynthesis of cotyledons
    during establishment and early growth of propagules of the mangrove, Avicennia
    marina, at three temperatures. South African Journal of Botany 57 (3), 171-174.
Steinke, T.D. and Ward, C.J. (1990). Litter production by mangroves. III. Wavecrest
    (Transkei) with predictions for other Transkei estuaries. South African Journal of
    Botany 56, 514-519.
Steinke, T.D., Barnabas, A.D. and Somaru, R. (1990). Structural changes and associated
    microbial activity accompanying decomposition of mangrove leaves in Mgeni estuary
    (South Africa). South African Journal of Botany 56 (1), 39-48.
Steinke, T.D., Rajh, A. and Holland, A.J. (1993a). The feeding behaviour of the red
    mangrove crab Sesarma meinertii De Man, 1887 (Crustacea: Decapoda: Grapsidae)
    and its effect on the degradation of mangrove leaf litter. African Journal of Marine
    Science 13, 151-160.
Steinke, T.D., Holland, A.J. and Singh, Y. (1993b). Leaching losses during decomposition of
    mangrove leaf litter. South African Journal of Botany 59, 21-25.
Steinke, T.D., Ward, C.J. and Rajh, A. (1995). Forest structure and biomass of mangroves in
    the Mgeni estuary, South Africa. Hydrobiologia 295 (1-3), 159-166.
Stevenson, N.J., (1997). Disused shrimp ponds: Options for redevelopment of mangroves.
    Coastal Management 12 (4), 425-435.
Stewart, R.W., Kjerfve, B., Milliman, J. and Dwivedi, S.N. (1990). Relative sea-level
    change: A critical evaluation. UNESCO reports in Marine Science 54, 1-22.
Stewart, S.E. (1996). Field behavior of Tripedalia cystophora (class Cubozoa). In
    “Symposium on the sensory ecology and physiology of zooplankton” (D.K. Hartline,
    J. Purcell, P. Lenz and D. Macmillan, eds), 8-12, January 1995. Vol. 27 (2-3), pp.
    175-188. Hawaii.
Stigliani, W.M. (1995). Global perspectives and risk assessment. In “Biogeodynamics of
    Pollutants in Soils and Sediments” (W. Salomons and W.M. Stigliani, eds), pp. 331-
    344. Springer Verlag, Berlin.
Stoner, A.W. (1991). Diel variation in the catch of fishes and penaeid shrimps in a tropical
    estuary. Estuarine, Coastal and Shelf Science 33, 57-69.
Stowe, K.A. (1995). Intracrown distribution of herbivore damage on Laguncularia racemosa
    in a tidally influenced riparian habitat. Biotropica 27 (4), 509-512.
Strong, A.M. and Bancroft, G.T. (1994). Patterns of deforestation and fragmentation of
    mangrove and deciduous seasonal forests in the upper Florida Keys. Bulletin of
    Marine Science 54 (3), 795-804.
Stuebing, R.B., Ismail, G. and Ching, L.H. (1994). The distribution and abundance of the
    Indo-Pacific crocodile Crocodylus porosus Schneider in the Klias River, Sabah, east
    Malaysia. Biological Conservation 69 (1), 1-7.
Sukardjo, S. and Yamada, I. (1992). Biomass and productivity of a Rhizophora mucronata
    Lamk. plantation in Tritih, central Java, Indonesia. Forest Ecology and Management
    49 (3-4), 195-209.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       137


Sun, M., Wonga, K.C. and Lee, J.S.Y. (1998). Reproductive biology and population genetic
    structure of Kandelia candel (Rhizophoraceae), a viviparous mangrove species.
    American Journal of Botany 85, 1631-1637.
Sutton, D.C. and Besant, P.J. (1994). Ecology and characteristics of bdellovibrios from three
    tropical marine habitats. Marine Biology 119 (2), 313-320.
Swaminathan, M.S. (1991). Foreward. In “Proceedings of the project formulation Workshop
    for establishment a global network of mangrove genetic resource centres for
    adaptation to sea level rise”, January 15-19, 1991 (S.V. Deshmukh and M.
    Rajeshwari, eds), Vol. 2, pp. 1-3. CRSARD, Madras, India.
Swearingen, D.C., III and Pawlik, J.R. (1998). Variability in the chemical defense of the
    sponge Chondrilla nucula against predatory reef fishes. Marine Biology 131 (4), 619-
    627.
Swiadek, J.W. (1997). The impacts of Hurricane Andrew on mangrove coasts in southern
    Florida: A review. Journal of Coastal Research 13 (1), 242-245.
Sylla, M., Stein, A. and van Mensvoort, M.E.F. (1996). Spatial variability of soil actual and
    potential acidity in the mangrove agroecosystem of West Africa. Soil Science Society
    of America Journal 60, 219-229.
Tack, J.F. and Polk, P. (1997). Groundwater flow in coastal areas influences mangrove
    distribution. In “International Seminar on Mangroves”, held in the Department of
    Zoology, 25-27 March, 1997. Waltair, Andra Pradesh. 7p. Abstract.
Tack, J.F., Vanden-Berghe, E. and Polk, P. (1992). Ecomorphology of Crassostrea cucullata
    (Born 1778) (Ostreidae) in mangrove creek (Gazi, Kenya). Hydrobiologia 247 (1-3),
    109-117.
Takeda, T., Ishimatsu, A., Oikawa, S., Kanda, T., Hishida, Y. and Khoo, K.H. (1999).
    Mudskipper Periophthalmodon schlosseri can repay oxygen debts in air but not in
    water. Journal of Experimental Zoology 284, 265-270.
Tam, N.F.Y. (1998). Effects of wastewater discharge on microbial populations and enzyme
    activities in mangrove soils. Environmental Pollution 102 (2-3), 233-242.
Tam, N.F.Y. and Wong, Y.S. (1995). Mangrove soils as sinks for waste waters borne
    pollutants. Hydrobiologia 295 (1-3), 231-241.
Tam, N.F.Y. and Wong, Y.S. (1996a). Retention and distribution of heavy metals in
    mangrove soils receiving wastewater. Environmental Pollution 94 (3), 283-291.
Tam, N.F.Y. and Wong, Y.S. (1996b). Retention of wastewater-borne nitrogen and
    phosphorus in mangrove soils. Environmental Technology 17 (8) 851-859.
Tam, N.F.Y. and Wong, Y.S. (1997). Accumulation and distribution of heavy metals in a
    simulated mangrove system treated with sewage. Hydrobiologia 352 (1-3), 67-75.
Tam, N.F.Y. and Yao, M.W.Y., (1998). Normalization and heavy metal contamination in
    mangrove sediments. Science of the Total Environment 216 (1-2), 33-39.
Tam, N.F.Y., Vrijmoed, L.L.P. and Wong, Y.S. (1990). Nutrient dynamics associated with
    leaf decomposition in a small subtropical mangrove community in Hong Kong.
    Bulletin of Marine Science 47, 68-78.
Tam, N.F.Y., Yuk-Shan, W., Lu, C.Y. and Berry, R. (1997). Mapping and characterization of
    mangrove plant communities in Hong Kong. Hydrobiologia 352 (1-3), 25-37.
Tam, N.F.Y., Wong, Y.S., Lan, C.Y. and Wang, L.N. (1998). Litter production and
    decomposition in a subtropical mangrove swamp receiving wastewater. Journal of
    Experimental Marine Biology and Ecology 226 (1), 1-18.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       138


Tan, C.G.S. and Ng, P.K.L. (1994). An annotated checklist of mangrove brachyuran crabs
    from Malaysia and Singapore. Hydrobiologia 285 (1-3), 75-84.
Tan, T.K. and Leong, W.F. (1992). Lignicolous fungi of tropical mangrove wood.
    Mycological Research 96 (6), 413-414.
Tan, T.K. and Pek, C.L. (1997). Tropical mangrove leaf litter fungi in Singapore with an
    emphasis on Halophytophthora. Mycological Research 101 (2), 165-168.
Tan, T.K., Teng, C.L. and Jones, E.B.G. (1995). Substrate type and microbial interactions as
    factors affecting ascocarp formation by mangrove fungi. Hydrobiologia 295 (1-3),
    127-134.
Tan, X. and Zhang, Q. (1997). Mangrove beaches' accretion rate and effects of relative sea-
    level rise on mangroves in China. Marine Science Bulletin, Haiyang 16 (4), 29-35.
Tattar, T.A., Klekowski, E.J. and Stern, A.I. (1994). Dieback and mortality in red mangrove,
    Rhizophora mangle L., in southwest Puerto Rico. Arboricultural Journal 18, 419-
    429.
Taylor, D.S., Davis, W.P. and Turner, B.J. (1995). Rivulus marmoratus: Ecology of
    distributional patterns in Florida and the central Indian River Lagoon. Bulletin of
    Marine Science 57 (1), 202-207.
Teixeira, C. and Gaeta, S.A. (1991). Contribution of picoplankton to primary production in
    estuarine coastal and equatorial waters of Brazil. Hydrobiologia 209, 117 - 122.
Thangam, T.S. (1990). Studies on marine plants for mosquito control. Ph.D. thesis.,
    Annamalai University, India. 68 pp.
Thangam, T.S. and Kathiresan, K. (1988). Toxic effect of mangrove plant extracts on
    mosquito larvae Anopheles stephensi L.. Current Science 47 (16), 914-915.
Thangam, T.S. and Kathiresan, K. (1989). Larvicidal effect of marine plant extracts on
    mosquito Culex tritaeniorhynchus. Journal of Marine Biological Association of India
    31 (1-2), 306-307.
Thangam, T.S. and Kathiresan, K. (1991). Mosquito larvicidal activity of marine plant
    extracts with synthetic insecticides. Botanica Marina 34 (6), 537-539.
Thangam, T.S. and Kathiresan, K. (1992a). Mosquito larvicidal activity of mangrove plant
    extract against Aedes aegypti. International Pest Control 34 (4), 116-119.
Thangam, T.S. and Kathiresan, K. (1992b). Smoke repellency and killing effect of marine
    plants against Culex quinquefasciatus. Tropical Biomedicine 9, 35-38.
Thangam, T.S. and Kathiresan, K. (1993a). Repellency of marine plant extracts against Aedes
    aegypti. International Journal of Pharmacognosy 31 (4), 321-323.
Thangam, T.S. and Kathiresan, K. (1993b). The mosquito composition and seasonal
    distribution of Culex quinquefasciatus in a coastal town of south India. Tropical
    Biomedicine 10, 175-177.
Thangam, T.S. and Kathiresan, K. (1994). Mosquito larvicidal activity of Rhizophora
    apiculata Blume. International Journal of Pharmacognosy 32, 33-36.
Thangam, T.S. and Kathiresan, K. (1997). Mosquito larvicidal activity of mangrove plant
    extracts and synergistic activity of Rhizophora apiculata with pyrethrum against
    Culex quinquefasciatus. International Journal of Pharmacognosy 35, 1-3.
Thangam, T.S., Srinivasan, K. and Kathiresan, K. (1992). Smoke repellency and killing
    effect of mangrove plants against the mosquito Aedes aegypti Linnaeus. Tropical
    Biomedicine 10, 125-128.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        139


Thollot, P. (1992). Importance of mangroves for the reef fish fauna from New Caledonia.
    Cybium 16 (4), 331-334.
Thomas, E. and Paul, A.C. (1996). Evolution of vivipary in flowering plants: Mini Review.
    Oikos 77, 3-9.
Thorhaug, A. (1990). Restoration of mangroves and seagrasses -- economic benefits for
    fisheries and mariculture. In “Environmental Restoration. Science and strategies for
    restoring the earth” (Berger, J.J., ed), pp. 265-281. Berkeley, USA.
Tietjen, J.H. and Alongi, D.M. (1990). Population growth and effects of nematodes on
    nutrient regeneration and bacteria associated with mangrove detritus from
    northeastern Queensland (Australia). Marine Ecology Progress Series 68, 169-180.
Titus, J.G. and Narayanan, V. (1996). The risk of sea level rise. Climatic Change 33 (2), 151-
    212.
Toledo, G., Bashan, Y. and Soeldner, A. (1995a). Cyanobacteria and black mangroves in
    northwestern Mexico: colonization and diurnal and seasonal nitrogen-fixation on
    aerial roots. Canadian Journal of Microbiology 41 (11), 999-1011.
Toledo, G., Bashan, Y. and Soeldner, A. (1995b). In vitro colonization and increase in
    nitrogen- fixation of seedling roots of black mangrove inoculated by a filamentous
    cyanobacteria. Canadian Journal of Microbiology 1 (11), 1012-1020.
Tomlinson, P.B. (1986). The Botany of mangroves. Cambridge University Press, Cambridge,
    U.K. 413 pp.
Triwilaida, and Intari, S.E. (1990). Factors affecting the death of mangrove trees in the
    Pedada Strait, Indragiri Hilir, Riau, with reference to the site conditions. Bulletin
    Penelitian Hutan 531, 33-48.
Turner, I.M., Tan, H.T.W., Wee, Y.C., Ibrahim, A.B., Chew, P.T. and Corlett, R.T. (1994). A
    study of plant species extinction in Singapore: Lessons of the conservation of tropical
    biodiversity. Conservation Biology 8 (3), 705-712.
Turner, I.M., Gong, W.K., Ong, J.E., Bujang, J.S. and Kohiyama, T. (1995). The architecture
    and allometry of mangrove saplings. Functional Ecology 9 (2), 205-212.
Tussenbroek, B.I (1995). Thalassia testudinum leaf dynamics in a Mexican Caribbean coral
    reef lagoon. Marine Biology 122 (1), 33-40.
Twilley, R.R. and Chen, R. (1998). A water budget and hydrology model of a basin
    mangrove forest in Rookery Bay, Florida. Marine and Freshwater Research 49, 309–
    323.
Twilley, R.R., Solórzano, L. and Zimmerman, R. (1991). The Importance of mangroves in
    Sustaining fisheries and Controlling Water Quality in Coastal Ecosystems. U.S.
    Agency for International Development, Project 8.333.
Twilley, R.R., Chen, R. and Hargis, T. (1992). Carbon sinks in mangroves and their
    implication to carbon budget of tropical ecosystems. Water, Air and Soil Pollution 64,
    265-288.
Twilley, R.R., Bodero, A. and Robadue, D. (1993). Mangrove ecosystem biodiversity and
    conservation in Ecuador. In “Perspectives on Biodiversity: Case Studies of Genetic
    Resource Conservation and Development” (C.S. Potter, J.I. Cohen, and D.
    Janczewski, eds). pp. 105-127. AAAS Press, Washington, D.C.
Twilley, R.R., Pozo, M., Garcia, V.H., Rivera Monroy, V.H., Zambrano, R. and Bodero, A.
    (1997). Litter dynamics in riverine mangrove forests in the Guayas River estuary,
    Ecuador. Oecologia 111 (1), 109-122.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                        140


Twilley, R.R., Gottfried, R.R., Rivera-Monroy, V.H., Zhang, W., Armijos, M.M. and
    Bodero, A. (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.
Tzeng, W.N. and Wang Yu, T. (1992). The temporal and spatial structure, composition and
    seasonal dynamics of the larval and juvenile fish community in the mangrove estuary
    of Tanshui River, Taiwan. Marine Biology 113 (3), 481-490.
Ukpong, I.E. (1994). Soil-vegetation interrelationships of mangrove swamps as revealed by
    multivariate analyses. Geoderma 64 (1-2), 167-181.
Ukpong, I.E. (1995). An ordination study of mangrove swamp communities in West Africa.
    Vegetation 116 (2), 147-159.
Ukpong, I.E. and Areola, O. (1995). Relationships between vegetation gradients and soil
    variables of mangrove swamps in southeastern Nigeria. African Journal of Ecology
    33 (1), 14-24.
Ulken, A., Viquez, R., Valiente, C. and Campos, M. (1990). Marine fungi (Chytridiomycetes
    and Thraustochytriales) from a mangrove area at Punta Morales, Golfo de Nicoya,
    Costa Rica. Review of Biology Tropics 38 (2A), 243-250.
Uncles, R.J., Ong, J.E. and Gong, W.K. (1990). Observations and analysis of a stratification-
    destratification event in a tropical estuary. Estuarine, Coastal and Shelf Science 31
    (5), 651-666.
Vance, D.J. (1992). Activity patterns of juvenile penaeid prawns in response to artificial tidal
    and day-night cycles: A comparison of three species. Marine Ecology Progress Series
    87, 215-226.
Vance, D.J. and Staples, D.J. (1992). Catchability and sampling of three species of juvenile
    penaeid prawns in the Embley river, Gulf of Carpentaria, Australia. Marine Ecology
    Progress Series 87, 201-213.
Vance, D.J., Haywood, M.D.E. and Staples, D.J. (1990). Use of a mangrove estuary as a
    nursery area by postlarval and juvenile banana prawns, Penaeus merguiensis de Man,
    in northern Australia. Estuarine, Coastal and Shelf Science 31 (5), 689-701.
Vance, D.J., Haywood, M.D.E., Heales, D.S., Kenyon, R.A., Loneragan, N.R. and Pendrey,
    R.C. (1996a). How far do prawns and fish move into mangroves? Distribution of
    juvenile banana prawns Penaeus merguiensis and fish in a tropical mangrove forest in
    northern Australia. Marine Ecology Progress Series 131, 115-124.
Vance, D.J., Haywood, M.D.E., Heals, D.S. and Staples, D.J. (1996b). Seasonal and annual
    variation in abundance of postlarval and juvenile grooved tiger prawns Penaeus
    semisulcatus and environmental variation in the Embley River, Australia: a six year
    study. Marine Ecology Progress Series 135, 43-45.
Vance, D.J., Haywood, M.D.E., Heales, D.S., Kenyon, R.A. and Loneragan, N.R. (1997).
    Seasonal and annual variation in abundance of postlarval and juvenile banana prawns,
    Penaeus merguiensis, and environmental variation in two estuaries in tropical
    northeastern Australia: a six-year study. Marine Ecology Progress Series 163, 21-36.
Vannini, M. and Oluoch, A. (1993). Notes on Merguia oligodon (De Man 1888) the Indo
    Pacific semi terrestrial shrimp (Hippolytidae: Natantia). Tropical Zoology 6 (2), 281-
    286.
Vannini, M., and Ruwa, R.K. (1994). Vertical migrations in the tree crab Sesarma leptosoma
    (Decapoda, Grapsidae). Marine Biology 118 (2), 271-278.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      141


Vannini, M., Cannicci, S. and Ruwa, K. (1995). Effect of light intensity on vertical
    migrations of the tree crab, Sesarma leptosoma Hilgendorf (Decapoda: Grapsidae).
    Journal of Experimental Marine Biology and Ecology 185 (2), 181-189.
Vannini, M., Oluoch, A. and Ruwa, K. (1997a). Tree-climbing mangrove decapods of
    Kenyan mangroves. In “Mangrove Ecosystem Studies in Latin America and Africa”
    (B. Kjerfve, L.D. Lacerda and S. Diop, eds), pp. 325-338. UNESCO, Paris.
Vannini, M., Ruwa, R.K., and Cannicci, S. (1997b). Effect of microclimatic factors and tide
    on vertical migrations of the mangrove crab Sesarma leptosoma (Decapoda:
    Grapsidae). Marine Biology 130 (1), 101-108.
Vannucci, M. (1997). Supporting appropriate mangrove management. International News
    Letter of Coastal Management-Intercoast Network, Special edition 1, 1-3.
van Speybroeck, D. (1992). Regeneration strategies of mangroves along the Kenyan coast. In
    “The ecology of mangrove and related ecosystems” (Jaccarini, V. and E. Martens,
    eds), pp. 243-251. Kluwer Academic Publishers, Netherlands.
Van Tussenbroek, B.I. (1995). Thalassia testudinum leaf dynamics in a Mexican Caribbean
    coral reef lagoon. Marine Biology 122 (1), 3-40.
Veenakumari, K., Mohanraj, P. and Bandyopadhyay, A.K. (1997). Insect herbivores and their
    natural enemies in the mangals of the Andaman and Nicobar Islands. Journal of
    Natural History 31 (7), 1105-1126.
Vergara-Filho, W.L., Alves, J.R.P. and Maciel, N.C. (1997). Diversity and distribution f
    crabs (Crustacea, Decapoda, Brachyura) in mangroves of Guanabara Bay, Rio de
    Janeiro, Brazil. In “Mangrove Ecosystem studies in Latin America and Africa” (B.
    Kjerfve, L.D. Lacerda and S. Diop, eds), pp. 155-162. UNESCO, Paris.
Verschelde, D., Muthumbi, A. and Vincx, M. (1995). Papillonema danieli gen. et sp. n., and
    Papillonema clavatum (Gerlach, 1957) comb.n. (Nematoda: Desmodoridae) from the
    Ceriops mangrove sediments of Gazi Bay, Kenya. Hydrobiologia 316 (3), 225-237.
Vervoort, H.C., Richards-Gross, S.E. and Fenical, W. (1997). Didemnimides A-D: novel,
    predator-deterrent alkaloids from the Caribbean mangrove ascidian Didemnum
    conchyliatum. The Journal of Organic Chemistry 62, 1486-1490.
Vethanayagam, R.R. (1991). Purple photosynthetic bacteria from a tropical mangrove
    environment. Marine Biology 110 (1), 161-163.
Vethanayagam, R.R. and Krishnamurthy, K. (1995). Studies on anoxygenic photosynthetic
    bacterium Rhodopseudomonas sp. from the tropical mangrove environment. Indian
    Journal of Marine Sciences 24 (1), 19-23.
Vezzosi, R., Barbaresi, S., Anyona, D. and Vannini, M. (1995). Activity patterns in
    Thalamita crenata (Portunidae: Decapoda): A shaping by the tidal cycles. Marine
    Behaviour Physiology 24 (4), 207-214.
Vikineswary, S., Nadaraj, P., Wong, W.H. and Balabaskaran, S. (1997). Actinomycetes from
    a tropical mangrove ecosystem anti-fungal activity of selected strains. Asia Pacific
    Journal of Molecular Biology and Biotechnology 5 (2), 81-86.
Vrijmoed, L.L.P., Hyde, K.D. and Jones, E.B.G. (1994). Observations on mangrove fungi
    from Macau and Hong Kong, with the description of two new ascomycetes:
    Diaporthe salsuginosa and Aniptodera haispora. Mycological Research 98 (6), 699-
    704.
Vrijmoed, L.L.P., Hyde, K.D. and Jones, E.B.G. (1996). Melaspilea mangrovei sp. nov.,
    from Australian and Hong Kong mangroves. Mycological Research 100 (3), 291-294.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       142


Wafar, S., Untawale, A.G. and Wafar, M. (1997). Litter fall and energy flux in a mangrove
    ecosystem. Estuarine, Coastal and Shelf Science 44, 111-124.
Warkentin, I.G. and Hernandez, D. (1996). The conservation implications of site fidelity: A
    case study involving nearctic-neotropical migrant songbirds wintering in a Costa
    Rican mangrove. Biological Conservation 77 (2-3), 143-150.
Wattayakorn, G., Wolanski, E. and Kjerfve, B. (1990). Mixing, trapping and outwelling in
    the Klong Ngao mangrove swamp, Thailand. Estuarine, Coastal and Shelf Science 31
    (5), 667-688.
Wehrtmann, I.S. and Dittel, A.I. (1990). Utilization of floating mangrove leaves as a
    transport mechanism of estuarine organisms, with emphasis on the decapod
    Crustacea. Marine Ecology Progress Series 60, 67-73.
Wei, X., Lin, M., Teng, J., Chen, S. and Yang, X. (1995). Designing and mapping dynamic
    charts by means of remote sensing for mangrove region--test area in Qinglan Bay,
    Hainan Island. Journal of Oceanography in Taiwan Strait 14 (3), 226-234.
Weinstock, J.A. (1994). Rhizophora mangrove agroforestry. Economic Botany 48 (2), 210-
    213.
Werner, A. and Stelzer, R. (1990). Physiological responses of the mangrove Rhizophora
    mangle grown in the absence and presence of NaCl. Plant, Cell and Environment 13,
    243-255.
West, J.A. (1991a). New algal records from the Singapore mangroves. Gardens Bulletin 43,
    19-21.
West, J.A. (1991b). New records of marine algae from Peru. Botanica Marina 34 (5), 459-
    464.
West, J.A. and Zuccarello, G.C. (1995). New records of Bostrychia pinnata and Caloglossa
    ogasawaraensis (Rhodophyta) from the Atlantic USA Botanica Marina 38 (4), 303-
    306.
West, J.A., Zuccarello, G.C., Pedroche, F.F. and Karsten, U. (1992). Marine red algae of the
    mangroves in Pacific Mexico and their polyol content. Botanica Marina 35 (6), 567-
    572.
Westgate, J.W. (1994). Eocene forest swamp. Research and Exploration 10 (1), 80-91.
Westgate, J.W. and Gee, C.T. (1990). Paleoecology of a middle Eocene mangrove biota
    (vertebrates, plants, and invertebrates) from Southwest Texas. Palaeogeography,
    Palaeoclimatology, Palaeoecology 7-8 (1- 2), 163-177
Wier, A.M., Schnitzler, M.A., Tattar, T.A., Klekowski, E.J., Jr. and Stern, A.I. (1996).
    Wound periderm development in red mangrove, Rhizophora mangle L. International
    Journal of Plant Sciences 157 (1), 63-70.
Whalley, A.J.S., Jones, E.B.G. and Alias, S.A. (1994). The Xylariaceae (ascomycetes) of
    mangroves in Malaysia and South East Asia. Nova Hedwigia 59 (1-2), 207-218.
Wilkinson, C.R. (1996). Global change and coral reefs: Impacts on reefs, economies and
    human cultures. Global Change Biology 2 (6), 547-558.
Williamson, I., King, C. and Mather, P.B. (1994). A comparison of fish communities in
    unmodified and modified inshore habitats of Raby Bay, Queensland. Estuarine,
    Coastal and Shelf Science 39 (4), 401-411.
Woitchik, A.F., Ohowa, B., Kazungu, J.M., Rao, R.G., Goeyens, L. and Dehairs, F. (1997).
    Nitrogen enrichment during decomposition of mangrove leaf litter in an East African
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                      143


   coastal lagoon (Kenya); relative importance of biological nitrogen fixation.
   Biogeochemistry 39 (1), 15-35.
Wolanski, E. and Sarsenski, J. (1997). Larvae dispersion in coral reefs and mangroves.
   American Scientist 85 (3), 236-243.
Wolanski, E., Mazda, Y., King, B. and Gay, S. (1990). Dynamics, flushing and trapping in
   Hinchinbrook Channel, a giant mangrove swamp, Australia. Estuarine, Coastal and
   Shelf Science 31 (5), 555-579.
Wolanski, E., Mazda, Y. and Ridd, P. (1992). Mangrove hydrodynamics. In
   “Tropical Mangrove Ecosystems” (A.I. Robertson and D.M. Alongi, eds). Pp. 43-62.
   American Geophysical Union, Washington D.C.
Wolanski, E., Spagnol, S. and Lim, E.B. (1997). The importance of mangrove flocs in
   sheltering sea grass in turbid coastal waters. Mangroves and Salt Marshes 1, 187-191.
Wong, Y.S., Lam, C.Y., Che, S.H., Li, X.R. and Tam, N.F.Y. (1995). Effect of wastewater
   discharge on nutrient contamination of mangrove soil and plants. Hydrobiologia 295,
   243-254.
Wong, Y.S., Tam, N.F.Y. and Lan, C.Y. (1997a). Mangrove wetlands as wastewater
   treatment facility: A field trial. Hydrobiologia 352 (1-3), 49-59.
Wong, Y.S., Tam, N.F.Y., Chen, G.Z. and Ma, H. (1997b). Response of Aegiceras
   corniculatum to synthetic sewage under simulated tidal conditions. Hydrobiologia
   352 (1-3), 89-96.
Woodroffe, C.D. (1990). The impact of sea-level rise on mangrove shorelines. Progress in
   Physical Geography 14 (4), 483-520.
Woodroffe, C.D. (1992). Mangrove sediments and geomorphology. In “Tropical Mangrove
   Ecosystem” (A.I. Robertson and D.M. Alongi, eds), pp. 7-41. American Geophysical
   Union, Washington DC., USA.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       144




Woodroffe, C.D. (1995). Response of tide-dominated mangrove shorelines in northern
    Australia to anticipated sea-level rise. Earth Surface Processes and Landforms 20 (1),
    65-85.
Woodroffe, C.D. (1999). Response of mangrove shorelines to sea level change. Tropics 8
    (3), 159-177.
Wright, A.E., Forleo, D.A., Gunawardana, G.P., Gunasekera, S.P., Koehn, F.E. and
    McConnell, O.J. (1990). Antitumor tetrahydroisoquinoline alkaloids from the colonial
    ascidian Ecteinascidia turbinata. Journal of Organic Chemistry 55, 4508-4512.
Yang, S., Lin, P. and Tsuneo, N. (1997). Ecology of mangroves in Japan. Journal of Xiamen
    University 36 (3), 471-477.
Yap, Y.N., Sasekumar, A. and Chong, V.C. (1994). Sciaenid fishes of the Matang mangrove
    waters. In “Proceedings of the Third ASEAN-Australian symposium on coastal
    resources” (S. Sundara, C.R. Wilkinson and L.M. Chou, eds), Vol. 2, pp. 491-498.
    Chulalongkorn University, Bangkok, Thailand.
Ye, Y., Lu, C., Wong, Y., Tam N.F.Y., Lin, P., Cui, S., Yang, S. and Li, L. (1997). Methane
    fluxes from sediments of Bruguiera sexangula mangroves during different diurnal
    periods and in different flat zones. Journal of Xiamen University 36 (6), 925-930.
Yonge, C.M. (1957). Enigmonia aenigmatica Sowerby, a motile anomiid (saddle oyster).
    Nature, London 180, 765-766.
Yoshihira, T., Shiroma, K. and Ikehara, N. (1992). Profiles of polypeptides and protein
    phosphorylation in thylakoid membranes from mangroves, Bruguiera gymnorrhiza
    (L.) Lamk. and Kandelia candel Druce. Galaxea 11 (1), 1-8.
Young, B.M. and Harvey, L.E. (1996). A spatial analysis of the relationship between
    mangrove (Avicennia marina var. australasica) physiognomy and sediment accretion
    in the Hauraki Plains, New Zealand. Estuarine, Coastal and Shelf Science 42 (2),
    231-246.
Young, C.M. (1995). Maintenance of diversity in communities dominated by open
    populations: Larval dispersal as a natural mitigator of environmental damage. Bulletin
    of Marine Science 57 (1), 285.
Youssef, T. and Saenger, P. (1996). Anatomical adaptive strategies to flooding and
    rhizosphere oxidation in mangrove seedlings. Australian Journal of Botany 44 (3),
    297-313.
Youssef, T. and Saenger, P. (1998). Photosynthetic gas exchange and accumulation of
    phytotoxins in mangrove seedlings in response to soil physico-chemical
    characteristics associated with waterlogging. Tree Physiology 18 (5), 317-324.
Yu, R.Q., Chen, G.Z., Wong, Y.S., Tam, N.F.Y. and Lan, C.Y. (1997). Benthic macrofauna
    of the mangrove swamp treated with municipal wastewater. Hydrobiologia 347 (1-3),
    127-137.
Zhang, Q., Wen, X., Song, C. and Liu, S. (1996). The measurement and study on
    sedimentation rates in mangrove tidal flats. Tropic Oceanology 15 (4), 57-62.
Zhang, W. and Huang, Z. (1996). Distribution of mangrove in Taiwan and its environment
    significance. Tropical Geography 16 (2), 97-106.
Zhang Y. and Wang, K. (1994). Distribution of mangrove pollen in the sediments from East
    China Sea and South China Sea and its paleoenvironment significance. Oceanologia
    et Limnologia Sinica, 25 (1), 23-28.
BIOLOGY OF MANGROVES AND MANGROVE ECOSYSTEMS                       145


Zhang, Y., Wang, K., Li, Z. and Liu, L. (1997). Studies on pollen morphology of Sonneratia
    genus in China and its paleoecological environment significance. Marine Geology
    and Quaternary Geology 17 (2), 47-52.
Zheng, S., Zheng, D., Liao, B. and Li, Y. (1997). Tideland pollution in Guangdong Province
    of China and mangrove afforestation. Forest Research 10 (6), 639-646.
Zheng, Z. (1991). Pollen flora and paleoclimate of the Chao-Shan plain during the last 50000
    years. Acta Micropalaeontologica Sinica 8 (4), 461-480.
Zhuang, T. and Lin, P. (1993). Soil microbial amount variations of mangroves (Kandelia
    candel) in process of natural decomposition of litter leaves. Journal of Xiamen
    University of Natural Science 32 (3), 365-370.
Zimmermann, U., Zhu, J.J., Meinzer, F.C., Goldstein, G., Schneider, H., Zimmermann, G.,
    Benkert, R., Thuermer, F., Melcher, P., Webb, D. and Haase, A. (1994). High
    molecular weight organic compounds in the xylem sap of mangroves: Implications
    for long-distance water transport. Botanica Acta 107 (4), 218-229.
Zuccarello, G.C. and West, J.A. (1995). Hybridization studies in Bostrychia: 1. B. radicans
    (Rhodomelaceae: Rhodophyta) from Pacific and Atlantic North America.
    Phycological Research 43 (4), 233-240.
by David Bael last modified 07-02-2007 14:43
 

Built with Plone