Vous êtes sur la page 1sur 12

Southern Cross University

ePublications@SCU
School of Environmental Science and
Management
School of Environmental Science and
Management
1996
Mangroves as indicators of coastal change
F Blasco
University Paul Sabatier
Peter Saenger
Southern Cross University, peter.saenger@scu.edu.au
E Janodet
University Paul Sabatier
ePublications@SCU is an electronic repository administered by Southern Cross University Library. Its goal is to capture and preserve the intellectual
output of Southern Cross University authors and researchers, and to increase visibility and impact through open access to researchers around the
world. For further information please contact epubs@scu.edu.au.
Suggested Citation
Post-print of Blasco, F, Saenger, P & Janodet, E 1996, 'Mangroves as indicators of coastal change', Catena, vol. 27, no. 3-4, pp.
167-178.
Catena home page available at www.elsevier.com/locate/catena
Publisher's version of article available at http://dx.doi.org/10.1016/0341-8162(96)00013-6

Mangroves as Indicators of Coastal Change

F. Blasco*, P. Saenger** and E. Janodet*

* Institute for the International Map of the Vegetation
CNRS/Universit Paul Sabatier
13, avenue du Colonel Roche, 31405 Toulouse Cedex, France

** Centre for Coastal Management
Southern Cross University
P.O. Box 157, Lismore, NSW 2480, Australia.

Keywords

Ecosystems, zonation patterns, coastal stability, sea-level, tidal regime,
Avicennia germinans.

Abstract

In view of the unique biological characteristics of mangroves, it is interesting
to assess the extent to which these ecosystems can be used as indicators of
coastal change or sea-level rise. From recent studies of mangrove mortality
at several locations (including Guiana, Gambia, Cte d'Ivoire, Kenya, India
and Bangladesh), it appears that these coastal ecosystems are so specialized
that any minor variation in their hydrological or tidal regimes causes
noticeable mortality. Each species of mangrove (but particularly those
belonging to the genera Rhizophora, Bruguiera, Sonneratia, Heritiera and
Nypa) occurs in ecological conditions that approach its limit of tolerance
with regard to salinity of the water and soil, as well as the inundation
regime. If the duration of daily immersion were to be modified by tectonic,
sedimentological or hydrological events, the species either readjusts to the
new conditions or succumbs to unsuitable conditions. Consequently, the
use of remote sensing data for mangrove ecosystems offers excellent
potential as a tool for monitoring coastal change.

Introduction

"Mangroves" is an ecological term referring to a taxonomically diverse
assemblage of trees and shrubs that form the dominant plant communities
in tidal, saline wetlands along sheltered tropical and subtropical coasts.
Economically, mangroves are a great source of timber, poles, thatch and
fuel, and the bark is used for tanning materials; some species have food or
medicinal value (Hamilton and Murphy, 1988). In Malaysia, Bangladesh,
India and Thailand, managed forest operations exist which produce
commercial timber, fuelwood, charcoal or woodpulp and subsistence
harvesting of mangroves is widespread in central America and parts of Asia
and Africa. Elsewhere, mangroves are used for honey or fodder production
(Baconguis and Mauricio, 1991).

Although the economic significance of mangroves is being increasingly
recognized, they are still decreasing in area around the world. In some
regions, such losses are due to over-exploitation, clearing and pollution of
mangrove systems (Saenger et al., 1983) but, in other areas, net losses occur
as a result of natural erosion of the shoreline (e.g. Senegal, Benin, Cte
d'Ivoire, Colombia, Brazil and Venezuela). At some localities, such as Guiana
(Blasco, 1991), there are losses due to erosion of established mangroves but
considerable gains on newly deposited sediments with no overall net loss.

There has been considerable recent interest in the probable response of
mangrove shorelines to sea-level and/or climate change (Woodroffe, 1990;
1993; Pernetta, 1993; Snedaker, 1993; Field, 1994), both in terms of the
mangroves themselves and in terms of their use as 'early warning'
ecosystems. In this review, we pose three questions in relation to mangroves
as indicators of regional coastal change. Specifically, (1) What are the unique
biological/ecological characteristics of mangroves? (2) Can mangroves be
used as biological indicators of coastal change or sea-level rise? and (3) Can
mangroves be used to enhance coastal stability and/or reduce coastal
erosion?

What are the unique biological/ecological characteristics of
mangroves?

The development and composition of mangrove communities depend largely
on temperature, soil type, salinity, duration and frequency of inundation,
accretion of silt, tidal and wave energy and such aperiodic factors as cyclone
or flood frequencies (Lugo and Snedaker, 1974; Hutchings and Saenger,
1987). Extensive mangrove communities seem to correlate with those areas
where the water temperature of the warmest month exceeds 24C, and they
are absent from those waters that never exceed 24C throughout the year.
Intertidal, sheltered, low-energy muddy sediments are the most suitable
habitats for mangroves and, under optimal conditions, forests up to 45 m in
height can develop as, for example, in Ecuador, Cameroon, Thailand and
Malaysia. Where less favourable conditions are found, mangrove
communities may reach maturity at heights of around 1-2 m (e.g. on the
arid coasts of the Arabian Gulf, Rajasthan, South Madagascar and Australia
and in cooler waters in New Zealand).

At present, about 150,000 km
2
of mangrove communities occur around the
world (Saenger et al., 1983) and they show a great similarity in structure,
floristics and function throughout their range. Plants of the mangrove
community belong to many different genera and families, most of which are
not closely related to one another phylogenetically. What they do have in
common is a variety of morphological, physiological and reproductive
adaptations that enable them to grow in a particular kind of rather unstable,
harsh and salty environment (Saenger, 1982; Tomlinson, 1986). On the
basis of the common possession of these various adaptations (see below),
approximately eighty species of plants belonging to about thirty genera in
over twenty families are recognized throughout the world as being
mangroves - a very small number of species for a tropical ecosystem. Most of
these species (about sixty) occur in the Indo-Pacific region, while there are
about twenty species in the western hemisphere.

At the generic level, Avicennia and Rhizophora are the dominant plants of
mangrove communities throughout the world, with each genus having
several closely related species in both the east and the west (Macnae, 1968;
Tomlinson, 1986; Ellison, 1991). At the species level, however, only a few
species (such mangrove associates as Thespesia populnea (L.) Soland., the
mangrove fern Acrostichum aureum L. and the swamp hibiscus, Hibiscus
tiliaceus L.) occur in both hemispheres. Such a present-day distribution can
be satisfactorily explained only if the ancestors of the mangroves evolved in
the Early Cretaceous, were dispersed outwards from their centre of origin
both in the Atlantic (e.g. Nypa) and the Indo-Pacific (e.g. Avicennia and
Rhizophora) through the remnants of the Tethys Sea, and subsequently
developed further as two isolated groups with the closure of the
Mediterranean Sea as a dispersal route between the Indo-Pacific and the
Atlantic. Some further spread from the Atlantic to the Pacific through the
still open Panama isthmus explains the similarity between the species on
the Pacific and Atlantic coasts of America (Keay, 1953; Tomlinson, 1986;
Jimenez, 1987).

As indicated earlier, most plants of the mangrove community are
halophytes, well adapted to salt water and fluctuations of tide-level (Clough,
1982). Many species show modified root structures such as stilt or prop
roots which support them on the semi-liquid or shifting sediments while
others have erect root structures (pneumatophores) that facilitate oxygen
penetration to the roots in a hypoxic environment. Salt glands, allowing
excess salt to be excreted through the leaves, occur in several species. Other
species show a range of physiological mechanisms either to exclude salt
from the plants, or to minimize the damage excess salts may cause by
separating the salt from the sensitive enzyme systems of the plant.

The mangrove community is often strikingly zoned parallel to the shoreline,
with a series of different species dominating from open water to the
landward margins (Snedaker, 1982). These zones are the response of
individual mangrove species to the gradients of inundation frequency,
waterlogging, nutrient availability and soil salt concentrations across the
intertidal area (Hutchings and Saenger, 1987; Wolanski et al., 1990). These
zones arise by the successful establishment of seedlings at specific levels in
the intertidal zone (Photo 1). The mechanism of seedling establishment has
been widely discussed. Early studies supported the 'self-planting' theory
(Tomlinson, 1986) where seedlings fall from the parent tree and establish in
the sediment at the base of the parent. Others have suggested that the
'stranding' theory is more generally applicable, where larger propagules are
stranded in deeper water and where only small propagules can successfully
establish in shallow water i.e. high up in the intertidal zone (Rabinowitz,
1975; 1978). Others argue that both strategies occur (Van Spreybroeck,
1992). However, both of these approaches are based on a restricted
distribution of mangrove seedlings in the intertidal zone. In a long-term
study, Saenger (1982) showed that most species broadcast their propagules
widely throughout the mangrove habitat and that seedlings establish very
widely. Several species have well-developed vivipary of their seeds, where the
hypocotyl develops while the fruit is still attached to the tree. These
seedlings are generally buoyant, able to float long distances in the sea and
establish themselves rapidly once stranded in a suitable habitat (Photo 1).
However, with time, seedling mortality is higher outside the optimal zone of
each species - resulting ultimately in distinct zones which reflect physico-
chemical characteristics (Snedaker, 1982; Youssef, 1995) that are still being
identified.

Not surprisingly, the zonation patterns show some similarity between the
eastern and western hemispheres. In the Americas and west Africa,
Rhizophora forms the outermost zone, followed by Avicennia, then
Laguncularia, with a sporadic landward fringe of Conocarpus. In the eastern
hemisphere, Rhizophora forms the outermost zone although at some
localities, Sonneratia and Avicennia may also be present. Avicennia generally
forms monospecific stands behind this outer zone, followed by mixed stands
of Bruguiera, Heritiera and Xylocarpus, with a landward zone of Ceriops
mixed with Lumnitzera and Avicennia.

Given the marked uniformity of zonation, mangrove communities may be
useful in interpreting minor changes in coastal conditions, such as altered
drainage patterns and recent accretion or erosion zones, by the minor
deviations from the regional zonation patterns.

Can mangroves be used as biological indicators of coastal change or
sea-level rise?

Relict mangrove sediments and fossil mangroves, particularly their pollen,
have been used extensively in the interpretation of palaeogeography and
stratigraphy (Churchill, 1973; Muller and Caratini, 1977; Bessedik, 1981;
Mildenhall and Brown, 1987; Grindrod, 1988; Crowley et al., 1990; Ellison,
1991; Woodroffe and Grindrod, 1991). Mangrove pollen are particularly good
markers because their presence in the sediments indicates a warm climate
(with a mean monthly temperature above 16C) and a nearby shoreline at
the time the pollen was deposited.

In addition, the most obvious impacts of relative sea-level rise in coastal
ecosystems are permanent inundation, salinization and coastal erosion
(Snedaker, 1993; Pernetta, 1993; Field, 1994). Each of these parameters
may partially or totally destroy or disrupt the affected mangrove stands.
From studies of mangrove mortality on several continents, it appears that
these ecosystems are so specialised that any minor variations in their
hydrological regimes causes noticeable mortality (Breen and Hill, 1969;
Blasco, 1984; Jimenez et al., 1985). The mortality results from the fact that
each species of mangrove lives in ecological conditions that approach the
limit of tolerance with regard to the salinity of the water and soil, and the
inundation regime. If for some reason (tectonic, sedimentological, or
hydrological) the duration of immersion is modified, the species either
readjusts to the new conditions or succumbs to unsuitable conditions.
Several cases are known of massive mortality of mangroves such as that of
the mangroves of Gambia which, because of a minor (<5 centimetres) but
rapid (<6 weeks) change in the hydrological regime, died over thousands of
hectares between January and March 1982 (Blasco, 1983, Photo 2).

When relative sea-level change is slow (in the order of 10-15 mm.yr
-1
-
Woodroffe, 1990), mangroves react differently. In the delta of the Ganges
(Fig. 1 and 2), there has been a very slow tilting of the coast due to tectonic
uplift in the north-western part (India) and subsidence (caused largely by
compaction of deltaic sediments) in the east (Bangladesh). As these
movements have been very slow, the plants are able to adapt and colonise
those areas which are suitable for them in terms of their ecological
conditions. As a result, species such as Avicennia marina (Forsk.) Vierh. and
Ceriops tagal (Perr.) C.B.. Rob. which tolerate high salt concentrations
(>30) during the dry season, have become predominant in the west of the
delta. In the delta area of Bangladesh, shifts in the position of the land/sea
boundary are further complicated by very high sedimentation rates (around
2,240 x 10
6
tonnes.yr
-1
- Lewis and McConchie, 1994) in some areas and
extensive coastal erosion in others; the high input of fresh water from the
Ganges-Brahmaputra-Meghna river system (around 30.8 x 10
3
m
3
.s
-1
) also
has a major impact on salinity conditions and hence on mangrove species
distribution. Those species (Heritiera fomes Buch.-Ham. and Nypa fruticans
van Wurmb.) which do not tolerate salinities in excess of 15 are confined
to the Bangladesh sections of the delta. Satellite data clearly show this
abrupt vegetation change due to a change in relative sea-levels (Fig. 3).

In contrast with the species-rich systems of Asia, the mangroves of the
Atlantic coast are generally dominated by only two species (Rhizophora
mangle L. and Avicennia germinans L.) which, together, constitute more than
80% of the mangroves. In Guiana, where the upland forests contain around
a thousand species, the shoreline mangroves contain practically only one
(Avicennia germinans L.). Because of the gentle gradient of sediments, this
environmental setting is both selective and fragile. A small change in mean
sea-level would result in a considerable change in the duration of immersion
of the mangroves at any point in the littoral zone, thereby causing plant
mortality. Here again, the rapidity and the scale of changes make the
application of satellite data extremely useful for measuring the extent of
changes and evaluating their consequences.

The study of the littoral zone of Guiana and its mangrove communities has
been augmented by aerial photographs obtained between 1979-87. These
photographs leave no doubt about the scale of displacement of coastal
sediments (Fig. 4) which have been mobilised by tides and currents and
moved along the coast at an average linear rate of 250 to 1250 m.yr
-1
.
Sediment movement rates are seasonally variable and when there are strong
winds (January to April) and the seas are rough, displacement of sediments
is very rapid; it is much slower during the rest of the year.

It appears that erosion commences and intensifies in the littoral zone of
Guiana when the mangroves die. In other words, the new hypothesis is that
the death of mangroves is not a consequence of coastal erosion but one of its
causes, perhaps the principal one. It then remains to explain why the trees
die.

Each Avicennia has several hundred pneumatophores per square metre and
their combined baffling effect favours an acceleration of sedimentation (Bird,
1971; 1986; Wells and Coleman, 1981). Thus, where a large supply of
sediment is available to a mangrove system, an increase in level of the
substrate will rapidly occur. If it is too rapid, it will asphyxiate the trees due
to a lack of oxygen in the sediments. With death and decomposition, the
roots release the sediments which are then susceptible to erosion. Thus,
there are probably two causes of this rapid mangrove change: the rapid
sedimentation from the Amazon and the 'self-destruct' tendency of
mangroves with dense pneumatophores. This explains why the build-up of
sediments at any one point suddenly ceases and is replaced by a phase of
intense erosion. It should be noted that the interaction of biotic and abiotic
factors is rarely expressed as clearly as in this example. If one disappears, so
does the other.

Any point on the coast of the Guianas (from Northern Brazil to Venezuela)
can remain stable while an adjacent section is attacked by waves and, after
a short period of stability, may have sediments returned to it. Finally,
mangroves (practically comprised of only Avicennia germinans L.) regrow on
the alluvium where a new balance has been attained - thus closing the
cycle. Similarly on the Atlantic coast of Africa, where there has been a slight
recession of the littoral zone with losses of tens of metres per year (Awosika
et al., 1993), erosion has often been accompanied by mangrove mortality
(Saenger and Bellan, 1995). The cause of these losses has not been firmly
established.

There is no scientific evidence to date indicating that mangrove mass
mortalities recorded on all continents are related to sea-level rise. An
evaluation of the impacts of this phenomenon, if it takes place at all,
requires careful assessment of possible coastal uplift or subsidence at each
specific site (Hoffman, 1984). Unfortunately, tide-level data are difficult to
use today because there are too few tide-gauging stations, particularly in the
south Atlantic. Where data are available, they are of insufficient duration to
allow significant changes in sea-level to be detected. Estimates of mean sea-
level rise vary around 10-15 cm since the start of this century (Woodroffe,
1990; Awosika et al., 1993). However, the geographic spread of these sea-
level changes is poorly known. Mapping of sea-level variations within several
centimetres is becoming possible with highly accurate altimetric satellites
recently launched (Minster et al., 1991).

The use of remote sensing data for coastal zone monitoring offers two major
possibilities. One, already in widespread use, involves the discrimination
and delineation of floristic units (mainly Rhizophora and Avicennia) and of
mortality within each unit. This is carried out with classical optical data
provided by high resolution systems such as SPOT, Landsat TM or airborne
scanners (Chaudhury, 1989; Hartono, 1991; Blasco et al., 1992).

Although still at an experimental stage, the second possibility for the use of
remote sensing data involves the use of active microwave systems. One of
the first attempts to identify mangrove species from space with radar
imagery was carried out by Gastellu (1987), who was able to map Nypa
fruticans van Wurmb. stands in East Kalimantan (Indonesia) with SIR-A
images (band L) much more easily than with the available panchromatic
aerial photographs. Nypa stands are sensitive to hydrological conditions and
occur in areas of low water salinity (<10ppm) and where the mean daily
inundation period does not exceed 3 hours. Being a sensitive indicator of
hydrological changes, stands of this mangrove palm could be monitored,
with any changes providing specific biological signals of physico-chemical
changes in coastal environmental conditions.

One of the key issues is the simultaneous discrimination between the
vegetation units and the degree of inundation beneath the vegetation
canopy. Such detection appears to be possible with Synthetic Aperture
Radar remote sensing but it has not yet been applied to mangrove
ecosystems (Hess et al., 1990). The experiments that are now being carried
out in coastal Thailand (Khlung mangroves) and in French Guiana with data
provided by ERS 1 and J.ERS 1 are expected to yield new data on the
correlation between changes in tidal inundation and the distribution and
adaptation (adjustment) of mangroves.

Can mangroves be used to enhance coastal stability and/or reduce
coastal erosion?

Mangroves are generally not land-builders (Bird, 1971; 1976) but rather,
they are a group of highly-reactive opportunists which can rapidly colonise
newly deposited and stable intertidal sediments - and, in so doing, they help
to consolidate these recent sediments and may promote further
sedimentation. Along the same lines, mangroves do not prevent coastal
erosion although their elaborate root structures are likely to slow that
process down considerably. Both of these characteristics of mangroves form
the basis of mangrove afforestation programs where mangroves have been
used to stabilise sediments as a means of gaining land from the sea. Thus in
Bangladesh, for example, a World-Bank funded project has assisted in the
stabilisation of around 120,000 hectares of recently deposited Gangetic
sediments over the last ten years (Saenger and Siddiqi, 1993). Similar
afforestation programs are presently underway at numerous locations
including, for example, Australia, Thailand, Vietnam, Philippines and Benin.

In contrast, the mangrove vegetation of Guiana does not appear to
consolidate the sediment, during its 10-30 year life span, to the point where
the coastline is stable (ORSTOM-CNRS, 1979; Wells and Coleman, 1981;
Blasco, 1991). Given the physical constraints on the mangroves of
Bangladesh (large tidal ranges, cyclones, strong currents), the inability of
the Guiana mangroves to maintain themselves on the new sediments where
their growth rates are rapid, is surprising. But, in the Indo-Pacific, when one
species dies, it is immediately replaced by another. The floristic pool in the
Indo-Pacific (with around 60 species) permits this immediate replacement
and the multi-species system appears to have greater physical stability as
well as ecological resilience.

In Guiana, with only the one species of mangrove, additional species (such
as the grass Oryza coarctata (Roxb.) Tak. or other mangroves) could be
introduced to take over the gaps created by the dying Avicennia. Whether
such introductions are feasible is beyond the scope of this review. However,
it does raise the question as to why the flora of the Guiana mangroves is so
depauperate?

Concluding Comments

The potential impacts of sea-level change on mangrove ecosystems have
recently been reviewed (Pernetta, 1993; Snedaker, 1993; Field, 1994) and
further discussion is beyond the scope of the present study. However, the
available scenarios of global change and climatic models suggest an increase
in global mean temperature of 1C by 2025 leading to a rise in global mean
sea-level of about 6 cm per decade (Houghton, 1991; Pernetta, 1993). These
figures refer to mean values at a global level. In the absence of more
accurate scenarios, at sub-regional scales (e.g. Bay of Bengal, Persian Gulf,
Guiana coast) any attempt to anticipate the impacts at a given location
remains impossible.

Most mangroves of the world are organized in almost monospecific belts in
which the Rhizophora and Avicennia groups play essential roles. The
discrimination of these belts from space is now generally feasible. In view of
the well-known ecological properties of mangroves, particularly the tendency
for any change in the daily duration and amplitude of flooding in these
communities to induce sudden and often spectacular mortalities, they
appear to constitute sensitive biological indicators of coastal environmental
changes. Whilst this property has been used for many years by
palynologists, when coupled with satellite technology, it can be used for
monitoring coastal events, almost in real time.

However, the task is not quite so simple. The actual impacts of sea-level
changes on a given mangrove ecosystem would be meaningless if they are
analysed independently of other factors that simultaneously affect the
ecological balance of that particular ecosystem i.e. local tectonics and
sediment dynamics, rainfall changes and changes in the freshwater input,
and frequency and intensity of cyclonic storms.

Despite their obvious economic value, mangroves have often been
considered as wastelands and converted to other forms of land use
(Vannucci, 1988). Gradually, however, opinion has changed and, more often
than not, these communities are now recognized not merely for their
economic value but also as living systems of scientific interest which are
useful indicators of coastal change to which they are particularly vulnerable.
As discussed above, they constitute one means by which to learn more
about the magnitude and mechanisms of changes in coastal phenomena.

Bibliography.

Awosika, L.F., Ibe, A.C. and Shroader, P. 1993. (Eds.), Coastlines of Western Africa.
American Society of Civil Engineers, New York, pp. 398.
Baconguis, S.R. and Mauricio, R.A. 1991. Forage and livestock production in the mangrove
forest. Canopy 16:9-10.
Bessedik, M. 1981. Une mangrove Avicennia L. en Mditerrane occidentale au Miocne
infrieur et moyen. Implications palo-gographiques. C.R. Acad. Sc. (Paris) 293:469-
472.
Bird, E.C.F. 1971. Mangroves as land-builders. Vict. Nat. 88:189-197.
Bird, E.C.F. 1986. Mangroves and intertidal morphology in Westernport Bay, Victoria,
Australia. Mar. Geol. 69:251-271.
Blasco, F. 1983. Mangroves of Gambia and Senegal. Universit de Toulouse and UNSO (New
York). 86 pp.
Blasco, F. 1984. Taxonomic considerations of the mangrove species. In: The mangrove
ecosystem: research methods. Chapter, UNESCO, Paris.
Blasco, F. 1991. Les mangroves. La Recherche 22:444-453.
Blasco, F., Bellan, M.F. and Chaudhury, M.U. 1992. Estimating the extent of floods in
Bangladesh using SPOT data. Remote Sensing and Environment 39:167-178.
Breen, C.M. and Hill, B.J. 1869. A mass mortality of mangroves in the Kosi Estuary. Trans.
Roy. Soc. S. Afr. 38:285-303.
Chaudhury, M.U. 1989. Mangroves in Bangladesh and their analysis using remote sensing
technology. Universit de Toulouse III (France). 206 pp.
Churchill, D.M. 1973. The ecological significance of tropical mangroves in the early Tertiary
floras of southern Australia. Geol. Soc. Aust. Spec. Publ. 4:79-86.
Clough, B.F. 1982. (Ed.), Mangrove Ecosystems in Australia - Structure, Function and
Management. ANU Press.
Crowley, G.M., Anderson, P., Kershaw, A.P. and Grindrod, J. 1990. Palynology of a Holocene
marine transgressive sequence, lower Mulgrave River Valley, north-east Queensland.
Aust. J. Ecol. 15:231-240.
Ellison, J.C. 1991. The Pacific palaeogeography of Rhizophora mangle L. (Rhizophoraceae).
Bot. J. Linn. Soc. 105: 271-284.
Field, C. 1994. (Ed.), Assessment and monitoring of climate change impacts on mangrove
ecosystems. UNEP Regional Seas Reports and Studies No. 154, UNEP, Nairobi, pp. 62.
Froidefond, J.M., Pujos, M. and Andr, X., 1988. TITLE Marine Geology 84:page numbers
Reference incomplete!
Gastellu, J.P. 1987. Radar remote sensing for vegetation with special reference to nipah
delineation on SIR-A image. Indonesian J. Geogr. 17:1-23.
Grindrod, J. 1988. The palynology of Holocene mangrove and saltmarsh sediments,
particularly in northern Australia. Rev. Palaebot. Palynol. 55:229-245.
Hamilton, L.S. and Murphy, D.H. 1988. Use and management of nipa palm (Nypa fruticans,
Arecaeae): a review. Economic Botany 42(2):206-213.
Hartono, M. and Muljosukojo, B. 1991. Monitoring mangrove disappearance by remote
sensing. Indonesian J. Geogr. 21:15-32.
Hess, L.L., Melack, J.M. and Simonett, D.S. 1990. Radar detection of flooding beneath the
forest canopy: a review. Int. J. Remote Sensing 11:1313-1325.
Hoffman, J.S. 1984. Estimates of future sea-level rise. In: Greenhouse effect and sea-level
rise. (Eds. Barth, M.C. and Titus, J.G.), Van Nostrand Reinhold Company, pp. 79-104.
Houghton, J.T., 1991. Scientific assessment of climate change: Summary of the IPCC
Working Group I Report. pp. 23-26. In: Jager, J. and H.L. Ferguson, Climate Change:
Science, impacts and policy. Proceedings of the Second World Climate Conference,
Cambridge University Press.
Hutchings, P. and Saenger, P. 1987. Ecology of Mangroves. Queensland University Press.
Jimenez, J.A., Martinez R. and Encarnacion, L. 1985. Massive tree mortality in a Puerto
Rican mangrove forest. Caribbean Journal of Science 21:75-78.
Keay, R.W. 1953. Rhizophora in West Africa. Kew Bull. 8:121-128.
Lewis, D.W. and McConchie, D. 1994. Practical Sedimentology. Chapman and Hall, New
York, pp. 213.
Lugo, A.E. and Snedaker, S.C. 1974. The ecology of mangroves. Ann. Rev. Ecol. Systematics
5: 39-64.
Macnae, W. 1968. A general account of the fauna and flora of mangrove swamps and forests
in the Indo-West-Pacific region. Adv. Mar. Biol. 6: 73-270.
Mildenhall, D.C. and Brown, L.J. 1987. An early holocene occurrence of the mangrove
Avicennia marina in Poverty Bay, North Island, New Zealand: its climatic and
geological implications. N.Z. J. Bot. 25: 281-294.
Minster, J.F., Lefebvre, M. and Benveniste, J. 1991. Satellite altimetry observations of
ocean dynamic topography. Int. J. Remote Sensing 12:1619-1629.
Muller, J. and Caratini, C. 1977. Pollen of Rhizophora as a guide fossil. Pollen et Spores
19:361-389.
ORSTOM-CNRS, 1979. Atlas des dpartements franais d'outre mer. La Guyana. CEGET,
Bordeaux (France).
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. A Marine Conservation and Development Report. IUCN, Gland, Switzerland,
pp. 46.
Rabinowitz, D. 1975. Planting experiments in mangrove swamps of Panama. In: Proceedings
of the International Symposium on Biololgy and Management of Mangroves. (Eds.
Walsh, G.E., Snedaker, S.C. and Teas, H.J.), Vol.1:385-393, University of Florida,
Gainesville.
Rabinowitz, D. 1978. Early growth of mangrove seedlings in Panama, and an hypothesis
concerning the relationship of dispersal and zonation. J. Biogeogr. 5:113-133.
Saenger, P. 1982. Morphological, anatomical and reproductive adaptations of Australian
mangroves. In: Mangrove ecosystems in Australia (Ed. B.F. Clough), Australian
National University Press, Canberra, pp.153-191.
Saenger, P. and Bellan, M.F. 1995. The mangrove vegetation of the Atlantic coast of Africa.
Universit de Toulouse Press, Toulouse, pp. 96.
Saenger, P. and Siddiqi, N.A. 1993. Land from the sea: the mangrove afforestation program
of Bangladesh. Ocean and Coastal Management 20:23-39.
Saenger, P., Hegerl, E.J. and Davie, J.D.S. (Eds.), 1983. Global Status of Mangrove
Ecosystems. The Environmentalist 3 (Supplement):1-88.
Snedaker, S.C. 1982. Mangrove species zonation: Why? In: Contributions to the Ecology of
Halophytes (Eds. Sen, D.N. and Rajpurohit, K.S.), Tasks for Vegetation Science, Vol.
2, pp. 111-125, W. Junk, The Hague.
Snedaker, S.C. 1993. Impact on mangroves. In: Climatic change in the Intra-Americas Sea.
(Ed. G.A. Maul), Edward Arnold, London, pp. 282-305.
Tomlinson, P.B. 1986. The Botany of Mangroves. Cambridge University Press.
Van Spreybroeck, D. 1992. Regeneration strategy of mangroves along the Kenya coast: a
first approach. Hydrobiologia 247:243-251.
Vannucci Reference quoted but no details in text>
Wells, J.T. and Coleman, J.M. 1981. Periodic mudflat progradation, northeastern coast of
South America: a hypothesis. J. Sediment. Petrol. 51:1069-1075.
Wolanski, E., Mazda, Y., King, B. and Gay, S. 1990. Dynamics, flushing and trapping in
Hinchinbrook Channel, a giant mangrove swamp, Australia. Estuar. Coast. Shelf Sci.
31:555-579.
Woodroffe, C.D. 1990. The impact of sea-level rise on mangrove shorelines. Progress in
Physical Geography 14:483-520.
Woodroffe, C.D. 1993. Sea-level. Progress in Physical Geography 17:359-368.
Woodroffe, C.D. and Grindrod, J. 1991. Mangrove biogeography: the role of Quaternary
environmental and sea-level change. J. Biogeogr. 18:479-492.
Youssef, T. 1995. Ecophysiology of waterlogging tolerance in mangroves. Unpublished PhD
Thesis, Southern Cross University, Australia.

Vous aimerez peut-être aussi