(David L. Hawksworth, Alan T. Bull) Forest Diversi
(David L. Hawksworth, Alan T. Bull) Forest Diversi
The titles published in this series are listed at the end of this volume.
Forest Diversity and Management
Edited by
David L. Hawksworth
and
Alan T. Bull
123
A C.I.P. Catalogue record for this book is available from the library of
Congress.
Published by Springer,
P.O. Box 17, 3300 AA Dordrecht, The Netherlands.
www.springer.com
Introduction
DAVID L. HAWKSWORTH
The Yellow House, Calle Aguila 12,
Colonia La Maliciosa, Mataelpino,
Madrid 28492, Spain.
E-mail: [email protected]
[1]
Biodiversity and Conservation (2006) 15:1063–1093 Springer 2006
DOI 10.1007/s10531-004-1874-6
-1
Key words: Mixed beech forests, Age-structure, Architecture, Soil, Light regime, Stand history
Abstract. Forest dynamics were analysed in the Upper Vosges mountains of north-eastern France
in two reserve areas, Frankenthal-Missheimle (FM) and Grand Ventron (GV), located in the
Ballons des Vosges Natural Regional Park (Parc Naturel Régional des Ballons des Vosges). Two
plots of 3000 m2 each were established in mixed beech woodlands located just below sub-alpine
beech forests for long-term monitoring. The main aim of the study was to interpret how the
different species populations in mixed-beech woodlands in the Vosges grow and interact over the
long term, and to determine the disturbance history. The study combined vegetation description,
dendrological and structural data, architectural descriptions and drawings and light distribution
and soil analysis. Historical information was also taken into consideration. Soils in the two plots
showed available phosphate P values > 0.14 g kg1, indicating good levels of phosphorus supply
for plants, except for A1/C horizon (1Va soil) which corresponds to a medium-fertility soil.
However, soils were found to be shallow because of the slope, a factor that may limit water
availability for adult trees and seedlings. As the canopy (composed of existing trees) consists of
shade trees, the growth rates for seedlings and saplings (potential trees) depends on the canopy
architecture: when growing in sunlit gaps, saplings reach full daylight (canopy height) in less than
100 years. When developing in shade (suppressed state), saplings may need up to 150 years before
reaching full daylight. Alternating periods of rapid and slow growth explain why some trees present
a wide range of stem diameters and ages in the area leading up to the canopy (some trees are more
than 300-years-old), in contrast with the relatively homogeneous height classes distribution, indi-
cating suppression periods. Trees in the FM and GV plots were found to have different growth
rates. Both study plots developed with similar past disturbance events, the two most important
being at the beginning of the 18th century. In addition, the forests were regularly affected by smaller
disturbances until present. Because of the spatial heterogeneity and large range of ages represented,
the forest stands within the two natural reserve areas are presently considered to be the best-
preserved sites in the upper Vosges, but their situation near the timber line prevents them from
becoming models for forest management at lower altitudes.
Introduction
[3]
1064
Schnitzler 2002). Apart from Russia, where old-growth forests may reach up to
20,000 km2 (Sittler et al. 2000), they are small patches (a mean range of 20–
100 ha) within a larger landscape patchwork of managed forests and various
land uses. For example, France, with 15 million hectares of metropolitan forest
cover, possesses the third largest woody domain in Europe, but only 0.2%
(300 km2) of this total are natural forests (Vallauri and Poncet 2003). Most of
these are concentrated in mountain regions (the Vosges, the Jura, the Pyrenees
and the Alps). Studies have demonstrated that these vestiges cannot remain
completely independent of their managed surroundings, and are unable to
preserve their potential biodiversity (Helle and Järvinen 1986).
Pollen diagrams have, however, demonstrated that even small ‘virgin’ forests
remain stable in composition over hundred of years (Bradshaw and Holmqvist
1999). The main causes of such stability and resilience are the high complexity
in structure and architecture associated with the complexity of biotic interac-
tions, which lead to remarkable resistance to climatological events (White
1978; Franks and McNaughton 1991).
All remaining natural forests urgently need protection in order to preserve
their cultural and scientific value, to protect their wildlife and genetic diversity
and to ensure sites for basic research in ecology. They also provide the nec-
essary reference data for applied research in forest management and environ-
mental monitoring (Leibendguth 1959; Peterken 1996).
The present study looked at several aspects of the forest dynamics and
biodiversity of mixed-beech woodlands of the ‘Parc Naturel Régional des
Ballons des Vosges’ (PNRBV) (Upper Vosges, north-eastern France). The
PNRBV has two natural reserves within its boundaries: the Grand Ventron
(GV), created in 1995 covering over 1647 ha, and the Frankenthal–Missheimle
(FM), created in 1989 with over 746 ha. Their respective elevations (GV: 720–
1204 m; FM : 690–1363 m) correspond to the submontane (400–800) and
montane belts (800–1100 altitude). Both reserves harbour typical, often
endangered, plant communities (Schwoehrer and Despert 1999; Schwoehrer
1999; Untereiner et al. 2002). The impact of human activity has been significant
since the French Revolution and includes logging, local fires and the planting
of non-indigenous spruce since the 1850s (Garnier 1994, 1998). Ungulate
densities are also closely linked to human activity and their populations have
increased considerably during the 20th century (ONF 2000; Heuzé 2002).
One of the main objectives of the creation of the PNRBV, within the
framework of which the present research was carried out, was to contribute to
the conservation and sustainable management of the forests through basic and
applied research and the development of innovative methodologies. For this
reason, our research focused on long-term, forest-monitoring plots situated in
strictly protected woodlands of the FM and GV natural reserves which still
include small stands of nearly natural woodlands (Gilg 1997).
The aim of the study is:
(i) to propose and innovate a sampling protocol for long-term studies,
[4]
1065
Study sites
The Vosges form a long ridge with a continuous crest line, linked to the west
and to the Rhine valley to the east by steep slopes. In the southern Vosges, the
crest oscillates from 1000 to 1425 m. The climate is oceanic (1600 to
3000 + mm rainfall; mean annual temperature of 4 C above 1000 m). The
hercynian bedrock is mainly composed of granite and metamorphic rock,
partially covered with morainic material. They include biotitic granite rock,
locally porphyroid, with acid plagioclase (An10-20), K-Feldspar, quartz and
some apatite (Mansuy 1992). Soils range from acidic browns to podzols
(Souchier 1971; Bonneau et al. 1978).
The mixed-beech woods found between 600 and 1000 m are part of the
Fagion alliance. The main plant community is the Luzulo–Fagetum (Meusel
37) which includes two sub-associations (group Vaccinium myrtillus; group
Festuca altissima, Oberdorfer 1992) typical of Central European mountains
and hills north of the Alps (Ellenberg 1988; Oberdorfer 1991, 1992; Bogen-
rieder 2001).
The three plant communities include the same tree species (Fagus sylv-
atica, Abies alba, Acer pseudoplatanus, Sorbus aucuparia) and shrubs (Rubus
tereticaulis, Rubus idaeus, Lonicera nigra). Beech (Fagus sylvatica) occupies
a central position in the ecology of the Vosgian forests, outcompeting the
other tree species. Silver fir (Abies alba) is regular but suffers from
browsing (Heuzé 2002). Acer pseudoplatanus is competitive in shallow soils
found in rocky habitats. Spruce (Picea excelsa)is rare, despite the impor-
tance of neighbouring plantations. Ground-level flora is dominated by
Vaccinium myrtillus, Luzula luzuloides, Deschampsia flexuosa and Prenanthes
purpurea.
The GV and FM reserves are 8 km apart (Figure 1). They are close (8–
15 km) to a third natural reserve, the Guebwiller (700–950 m), which also
includes some stands of nearly natural woodlands (Renaud et al. 2000).
Permanent plots
[5]
1066
system (GPS) (GV: latitude 4827¢N; longitude 665¢E; the plot is situated near
the ‘Grand Ventron’ farm; FM: latitude 4830¢N ; longitude 710¢E: the plot is
situated below the ‘Trois-Fours’ farm). Metallic boundary markers were buried
(forced in the soil) in the ground in order to pursue the study over the very long
term.
The FM plot is located between 900 and 950 m while the GV plot is 50 m
higher (950–1025 m), very close to the multi-stemmed beech forests that form
the timber-line (Carbiener 1966). Both plots face East. Plot GV is characterized
by a succession of steep slopes (50–60%), and flatter (5–30%), moister zones
while the FM plot exhibits more regular slopes, averaging 65%. Plot FM is
adjacent to a large, permanent gap resulting from the accumulation of boulders
of glacial origin.
The fieldwork started in 2000 and ended in 2003.
Soil data
[6]
1067
In 2002, five soil units were defined in the GV plot. In 2003, three soil-units
were identified and sampled in the FM plot.
All samples were air-dried, passed through a 2 mm sieve and then analysed
for chemical characteristics (analysis performed by INRA Laboratory for Soil
Analysis, Arras, France): residual moisture content at 105 C, organic carbon
and total nitrogen (dry combustion method), ‘available’ phosphorus (extracted
by H2SO4 and NaOH; Duchaufour and Bonneau 1959), and exchangeable
cations (based on cobalthexamine method, Orsiny and Remy 1976). Calcula-
tions were performed for C/N ratio, cation exchange capacity (CEC: sum of
exchangeable cations), base saturation (ration of exchangeable ‘basic’ cations –
Ca2+, Mg2+, K+ and Na+ – to CEC), and exchangeable Mg/Al and Ca/Al
ratios.
Stand characteristics
Ground flora
The ground flora coverage was identified and the coverage of each species esti-
mated, (Braun–Blanquet cover coefficient were converted in percentage cover).
Tree seedling densities (height < 130 cm with d.b.h less than 10 cm) were
identified and quantified per 10 · 10 m quadrant (30 per stand).
[7]
1068
ture in the same tree, by stimulation of resting meristems, Oldeman 1979) may
explain differences in size variables. Fagus sylvatica grows according to Troll’s
model with plagiotropic differentiation in all axes. The flattened and highly
organized leaf layers (monolayers), as well as the plagiotropy are necessary for
intercepting light over large surfaces. Beech is also very flexible, forming shoots
of different lengths in response to environmental conditions (Hallé and Old-
eman 1970; Nicolini 1997; Roloff 1999). Acer pseudoplatanus follows Rauh’s
model, characterized by orthotropic axes and faster growth than the beech.
However, Acer pseudoplatanus is a monolayer, which increases its ability to
intercept sunlight (Hallé et al. 1978). Abies alba follows Massart’s model,
characterized by a specialized plagiotropic organization of the branches that
confer them high individual survival in the lower forest storeys (Edelin 1977).
Architectural criteria also allow us to define the different phases of tree
development. Oldeman (1990) has defined three main phases (‘potential tree’,
‘tree of the present’ and ‘tree of the past’), considered as the principal social
states in the forest ecosystem. Potential trees and trees of the present can be
seen as two distinct phases in the growth of a tree. In potential trees, the growth
in height is relatively more important than growth in stem diameter and crown
extension. Potential trees may grow faster under the high light levels found in
canopy gaps (‘released-growing trees’) or be suppressed by shading from
canopy trees. After successive intervals of suppression and released growth,
many potential trees reach a minimum height above which they cannot be
suppressed. In this case, they develop their architecture by reiteration and
reach the tree of the present phase. The transition between these two growth
phases is gradual, depending on the site and the forest architecture. To dis-
tinguish the two steps in our study, an empirical threshold was set at the height
of 20 m, which corresponds to the minimum needed to no longer be suppressed
on the steep slopes of the Vosges (suppression is visible when trees present
shallow crowns and reiterations along the trunk, and the h/d.b.h. ratio is
greater than 100). Tall tree species that have gone beyond this threshold are
considered full-grown.
For our purposes, living and dead trees of more than 10 cm d.b.h. and height
>1 m 30 were identified in the plot and measured (diameter at breast height,
total height, height of the main fork. Each (living or dead) tree was carefully
drawn according to its exact proportions and morphology. Crown area was
defined using four perpendicular directions, including one facing the slope.
Tree architecture was represented visually by vertical (six, 10 · 50 m parallel
drawings per stand, oriented towards the slope) and horizontal drawings (one
for living and one for dead trees in each plot).
[8]
1069
cycle (‘potential’ versus ‘present’). Thus, tree ring analysis is useful for deter-
mining the influence of ecological conditions within the particular context of
the Upper Vosges.
The ring widths of 156 trees (93 in the GV plot and 63 in FM, including
beech, silver fir and Acer pseudoplatanus were noted. Two cores were extracted
at breast height from each tree, one up-slope and one down-slope. Tree-ring
widths were measured using a computer-assisted device (Becker et al. 1995) and
cross-dated against previously existing site chronologies (data from INRA,
Champenoux), in order to correct for missing or duplicate rings.
Stem diameter growth was calculated by dividing core length by the total
number of ring widths. Growth rate comparisons were done at each site for
each state and each species using non-parametric Mann–Whitney U-tests
(Statistica software).
Since tree growth, architecture and size variables are largely explained by light
dynamics (e.g. Chazdon and Pearcy 1986; Denslow et al. 1990; Vester 1997),
variations in light distribution at points within the canopy were also studied.
Relationships between the tree, the forest and the light level in old-growth
forests have been studied by Koop (1989).
The methodology deals with the architectural parameters of the canopy
(canopy geometry) that are commonly used in the literature (i.e. gap fraction
and foliage area index LAI), and the light variables (i.e. direct and diffuse solar
radiation transmitted throughout the canopy). The simultaneous treatment of
canopy geometry and distribution of incident light (PAR, understood as QPAR,
quantum irradiance, expressed in mol m2 d1, Varlet-Granchet et al., 1989)
was studied using hemispherical canopy photography. For each plot, 25 canopy
photographs were taken in each 100 m2 section, 10 m apart from each other, at
50 cm above the ground, in overcast conditions. Because some photographs
were of poor quality, the final raw data included only 21 photographs for GV
and 24 for FM. The camera was equipped with a fish-eye lens with a view angle
of 180, carefully levelled and oriented for true North. The raw data from each
photograph consisted of a matrix with 18 intervals of 5 zenith angles and 24
sectors of 15 azimuth angles which contained the gap fraction. All values were
corrected for latitude, slope, orientation and topographic mask in the GLA
model (Frazer et al. 1999). Calculations were performed at hourly intervals, then
integrated daily, during the vegetative season (June–September). The radiation
intercepted depended upon values calculated for each canopy element involved,
over the whole hemisphere and along the solar tracks. Hemispherical photo-
graphs therefore present spatial auto-correlations with each other.
Values found in FM and GV will be compared with data obtained using
similar methods in the Guebwiller mixed-beech forest. Data concern one plot
[9]
1070
of 1 ha chosen in a nearly natural forest stand growing in deep soil (slope from
20 to 30)(Renaud et al., 2000; Pierrel 2001).
Results
Soil characteristics
In both plots, soils are dark coloured, stony or gravely, with a fluffy to massive
structure and are finely textured (coarser with depth). Soils are shallow. FM
soils show less variability in type than those in GV. Soils belong mostly to the
Rankosol (Ranker) type, except for 2V soil which belong to Alocrisol and 5V
soil which presents hydromorphic features (located in bench slope). Humus
type ranges from oligo-mull to hemi-moder (Brethes et al. 1992).
Chemical analysis (Table 1) shows a humose trend, with relatively high or-
ganic carbon content, particularly for the 5V soil due to waterlogged condi-
tions. Less organic matter accumulation is observed in FM soils which may be
related to better biodegradability of organic materials. In both stands, rather
low C/N ratios, from 14.9 to 17.7 in uppermost horizons (A11), accounted for a
rapid evolution of plant material added to soil and good nitrogen nutrition for
trees. C/N ratios increasing with depth (1Va, 4V and 1F soils) probably indi-
cate a cryptopodzolisation process, morphologically hidden by the humose
character of soils.
Available phosphorus obtained using the Duchaufour and Bonneau (1959)
method is a good indicator of soil fertility. All samples showed available P2O5
values > 0.14 g kg1, indicating a good phosphorus supply for plants, except
for A1/C horizon (1Va soil) which corresponded to a medium-fertility soil
(Bonneau 1995). No notable differences were observed in regeneration spots, or
between GV and FM plots.
Soils show a medium cation exchange capacity (CEC), mostly related to
organic matter content (organic carbon), decreasing with depth and lower in
FM soils. Low base saturation was always observed in depth, while it exceeded
35% in uppermost horizons in 4V and 3F soils and reached 60% in 2F soil, due
to active nutrient cycling. Among basic exchangeable cations, Ca generally
predominates over Mg and K (low Na). Similarly, among acid exchangeable
cations, Al largely predominates over H (low Mn), except in 1Va, 1Vd, 4V and
2F soils’ uppermost horizons showing a higher H proportion, probably related
to cryptopodzolisation. A plentiful supply of Al on exchange sites is known to
disturb Mg and Ca supply. The lowest Mg/Al and Ca/Al ratios were found in
all deeper horizons (Bw, A1/C and C) except 2F soil, suggesting a possible Mg
and Ca deficiency. On the other hand, all uppermost horizons showed higher
ratios, apart from 3V soil (Mg) and 5Va soil (Ca).
On the whole, no pronounced differences in soil property features were
observed between the two plots.
[10]
Table 1. Chemical characteristics of soils in GV and FM plots.
Stand Soil Horizon Depth Organic C/N Available Exchangeable cations CEC Base Mg/ Ca/ pH
cm carbon g kg1 P2O5 g kg1 saturation% Al Al H 2O
2+ 2+ + + 3+ 2+ +
Ca Mg K Na Al Mn H
cmol+kg1
GV 1Va A11 0–2 14.84 17.7 0.386 1.95 0.59 0.49 0.08 3.99 0.068 3.56 10.73 29.0 0.148 0.490 3.6
1Va A12 2–7 7.91 17.0 0.195 0.35 0.27 0.24 0.05 4.17 0.013 2.09 7.18 12.6 0.064 0.084 3.6
1Va A1/C 7–30 5.79 21.8 0.125 0.04 0.06 0.07 0.03 5.16 0.006 0.33 5.70 3.6 0.012 0.008 4.1
1Vb A11 0–2 18.48 15.4 0.35 0.63 0.38 0.42 0.07 7.29 0.117 0.55 9.47 15.9 0.053 0.086 4.3
1Vc A11 0–2 15.41 16.4 0.316 1.81 0.57 0.45 0.05 6.54 0.159 1.50 11.09 26.0 0.088 0.276 3.7
1Vd A11 0–2 16.23 17.3 0.289 1.33 0.55 0.44 0.08 5.61 0.036 3.58 11.63 20.7 0.098 0.237 3.6
2V A11 0–2 18.03 17.0 0.404 2.84 0.65 0.49 0.05 7.23 0.298 1.60 13.17 30.7 0.090 0.393 3.9
2V A12 2–10 12.65 15.1 0.3 0.26 0.19 0.20 0.05 7.28 0.042 0.33 8.36 8.4 0.026 0.036 4.0
2V Bw 10–25 6.75 15.4 0.184 0.11 0.10 0.11 0.04 4.83 0.027 0.17 5.39 6.9 0.021 0.024 4.3
2V C 25–50 6.61 15.5 0.16 0.07 0.08 0.07 0.03 4.10 0.014 – 4.37 5.9 0.020 0.018 4.5
[11]
3V A11 0–2 29.24 17.2 0.454 0.78 0.28 0.31 0.03 10.65 0.275 0.35 12.67 11.1 0.027 0.073 4.2
3V A12 2–12 17.92 16.0 0.365 0.27 0.27 0.31 0.05 8.23 0.092 0.30 9.52 9.4 0.032 0.032 4.3
3V C 12–40 6.15 14.6 0.219 0.07 0.09 0.08 0.03 4.06 0.028 0.08 4.45 6.3 0.023 0.018 4.5
4V A11 0–2 30.77 17.7 0.503 6.25 1.31 0.75 0.08 6.06 0.168 3.60 18.21 46.0 0.216 1.032 3.7
4V A12 2–10 19.51 17.2 0.404 1.81 0.59 0.46 0.06 8.16 0.028 2.66 13.77 21.3 0.072 0.222 3.6
4V C 10–33 8.06 21.7 0.197 0.14 0.13 0.11 0.04 6.86 0.006 0.50 7.80 5.6 0.020 0.021 4.0
5Va A11 0–2 34.37 17.6 0.54 0.68 0.47 0.49 0.09 11.41 0.140 0.31 13.58 12.7 0.041 0.059 4.5
5Va A12 2–20 5.58 15.9 0.254 0.12 0.10 0.08 0.02 3.24 0.011 0.16 3.74 8.7 0.032 0.038 4.6
5Vb A11 0–2 18.30 14.9 0.438 0.97 0.40 0.39 0.05 9.12 0.289 0.55 11.77 15.4 0.044 0.106 4.3
FM 1F A11 0–2 6.66 15.9 0.412 0.82 0.23 0.28 0.02 4.05 0.214 0.31 5.91 22.8 0.056 0.203 4.4
1F A12 2–15 4.29 15.4 0.208 0.09 0.08 0.09 0.02 4.18 0.033 – 4.50 6.3 0.020 0.022 4.5
1F A1/C 15–25 3.25 17.4 0.286 0.05 0.03 0.04 0.02 2.57 0.010 – 2.72 5.2 0.012 0.020 4.7
2F A11 0–2 12.45 16.6 0.388 4.13 0.99 0.49 0.06 1.99 0.362 1.39 9.42 60.2 0.495 2.073 4.2
2F A1/C 2–10 8.85 14.9 0.317 0.99 0.33 0.34 0.07 4.40 0.036 1.30 7.46 23.2 0.075 0.225 4.0
3F A11 0–2 7.87 16.5 0.46 1.93 0.45 0.32 0.03 3.37 0.263 0.84 7.20 37.9 0.134 0.574 4.2
1071
3F A12 2–20 5.29 14.8 0.333 0.33 0.15 0.13 0.05 4.41 0.110 0.31 5.49 12.1 0.035 0.074 4.3
Data in g kg1 of dry matter, except for available P2O5 in g kg1 of air-dried soil: below detection limits.
1072
Stand characteristics
[12]
Table 2. Densities and volumes of tree species in FM and GV.
Number saplings 4 Number, Fagus Number, Number, Acer Number, Number, Sorbus
sylvatica Abies alba pseudoplatanus Picea excelsa aucuparia
1073
1074
Figure 2. (a) Crown projection map (50 · 60 m) in FM (shaded areas represent regeneration). Fa
for Fagus sylvatica; Ac for Acer pseudoplatanus; Ab for Abies alba. (b) Crown projection map
(50 · 60 m) in VN (shaded areas represent regeneration) Fa for Fagus sylvatica; Ac for Acer
pseudoplatanus; Ab for Abies alba.
[14]
1075
Figure 3a. Examples of vertical profile (10 m · 50 m) in FM (tree species and age are represented
for potential trees and trees of the present. (Fa: Fagus sylvatica; Ab: Abies alba; Ac: Acer
pseudoplatanus).
general, one or two were big and healthy, one to three smaller and the
remaining trunks (big or small) dead. The proportion of dead trunks increased
as the total number of trunks increased.
[15]
1076
was rather complex, with the imbrication of three main layers: 20–25 m, 25–
30 m and some emergents between 35 and 40 m. Plot GV had a nearly equal
number of trees between 20–25 and 25–30 m (Figure 3a, a1), while FM
included more trees in the upper canopy (Figure 3b, b1).
Size-distribution is not related to age-distribution, indicating that shade-
tolerant species cohorts can survive for long periods in a suppressed state.
Correlations are better for older trees in the upper parts of the canopy; i.e.
[16]
1077
Figure 3b. Examples of vertical profile (10 m · 50 m) in GV (tree species and age are represented
for potential trees and trees of the present. (Fa: Fagus sylvatica; Ab: Abies alba; Ac: Acer
pseudoplatanus).
maximum ages are closer to sizes: 40 m high and 111 cm d.b.h. for a 345-year-
old silver fir, 29 m high and 41 cm d.b.h. for a 214-year-old beech; 23 m high
and 54 cm d.b.h. for a 231-year-old Acer pseudoplatanus (Table 3).
Most trees of the present had a rotten heart, particularly the Acer pseudo-
platanus. However, their foliage and axes were well-developed, without any
sign of bark or leaf loss.
Trees can be considered as reaching the stage ‘of the present’ at various ages.
In general, potential trees in FM reach the canopy earlier than those in GV: a
range of 48–162 years in FM compared to a range of 102 (very rare)–223 years
in GV. Thus, under good light conditions, averages of 2.2 mm growth per year
[17]
1078
were recorded for three young 45- to 52-year-old beeches in FM. In this plot,
potential trees and trees of the present also had similar growth rates for stem
diameter while in GV, trees of the present grew significantly (p < 0.001) faster
than potential trees (Table 4).
These data indicate that suppression was more marked in GV, which limits
correlations between stem diameter growth and age (Figure 4). For beech trees,
the Spearman correlation coefficient (Rs) is 0.48 while the correlation is above
0.65 in FM. The silver fir presents similar tendencies (Rs = 0.71 in FM; 0.29 in
GV). Average growth rates of suppressed beech trees in GV are only 0.3–
0.6 mm per year and 0.7 mm for silver fir. This explains why some GV saplings
may be rather old: 119 years for a 4.5 m high silver fir; 135 years for a 7.5 m-
high beech. One potential beech was still in a suppressed state at 215 year of age.
Multi-stemmed beech trees present a broad range of ages (from 20 to
78 years) within one individual.
[18]
1079
Table 3. Ranges of stem diameter and age for beech, silver fir and sycamore in FM and GV.
FM plot GV plot
Potential trees
< 10 m
Fagus sylvatica 6 8–11 54–102 6 5–19 41–135
Abies alba 3 11–79 30–59 2 6 37–119
Acer pseudoplatanus 2 5–7 21–68 1 10 55
10–20 m
Fagus sylvatica 4 13–32 35–52 5 16–32 119–160
Abies alba 3 21–30 54–80 14 14–54 149–226
Acer pseudoplatanus 1 67 77 1 11 66
Dead trees
Dead trees (trees of the past) represented 11.2 and 3.1% of the total volume of
trees in GV and FM, respectively (Table 2). Most of them were silver firs. In
FM no death was recorded among beech trees. Dead trees generated only very
small gaps because they were rarely very large: average stem diameters ranged
from 10–80 cm. The three dead silver fir trees analysed in GV (d.b.h of 13, 13.6
and 25 cm) died at 79, 89 and 191 year of age respectively. Two beech trees
(d.b.h of 25 and 22 cm) died at 145 and 149 years respectively. These trees were
either snapped off at different heights (from 2 to 24 m) or uprooted. Dead trees
presented different degrees of rot. Standing dead trees had woodpecker holes
and were often infected by Fomes fomentarius.
Tree establishment
Tree distribution by species (Fagus sylvatica, Abies alba and Acer pseudoplat-
anus) and age-class (Figure 5) in FM and GV indicates a pattern of estab-
lishment and mortality during the last 350 years. There was a peak in
establishment between 1800 and 1840 in FM (a total of 46 trees in 40 years). In
[19]
1080
Table 4. Stem diameter growth for beech, silver fir and sycamore in FM and GV.
Potential Present
n Growth n Growth
rate rate
FM
Fagus sylvatica 10 1.37 31 1.47 NS
Abies alba 6 1.6 7 1.7 NS
Acer pseudoplatanus 3 1.3 6 1.26 NS
GV
Fagus sylvatica 11 0.64 44 0.95 ***
Abies alba 16 0.83 6 2.02 ***
Acer pseudoplatanus 2 0.59 12 0.96 NS
Potential versus potential per plot
FM GV
n Growth n Growth
rate rate
Fagus sylvatica 10 1.37 11 0.64 **
Abies alba 6 1.6 16 0.83 **
Acer pseudoplatanus 3 1.3 2 0.59 NS
Present versus present per plot
Fagus sylvatica 31 1.47 44 0.95 ***
Abies alba 7 1.7 6 2.02 NS
Acer pseudoplatanus 6 1.26 12 0.96 NS
***p < .001
**p < .01
*p < .05
NS: not significant
GV, the peak occurred 20 years earlier, between 1780 and 1820 (a total of 62
trees). Beech was the colonizing tree in more than 50% of the cases. After 1840,
recruitment was continuous but weak (approximately 2–4 trees per 20 year-
period except in 1920–1940 in FM). Before 1780–1800, trees were very sparse
(probably most of them have died since that time): two silver firs born
respectively in 1640 and 1680, two beech trees born in 1680 and 1720 and one
Acer pseudoplatanus in 1760.
Figure 6a, b illustrate age distribution in the two plots. Trees from 100 to
200 years of age were regularly distributed as small groups of similar ages. Trees
in the 200- 300-year-old category, which had survived stress, pathogens or
windstorms, were scattered as relics among these smaller groups. Trees under
100 years of age were clumped around gaps or at the margins of canopy trees.
The two plots presented differences in canopy geometry and light patterns
(Figure 7a, b; Table 5). Values from FM indicate a more open habitat than
[20]
1081
Figure 4. Age-stem diameter distribution for Fagus sylvatica, Abies alba, Acer pseudoplatanus in
FM and GV. Rs: Spearmann correlation coeffiecient, ***: p < 0.001, **: p < 0.01, *: p < 0.05,
NS: Not significant
[21]
1082
[22]
1083
Figure 6. (a) Tree age and spatial distribution in FM. (b) Tree age and spatial distribution in GV.
[23]
1084
fraction and the total incident light (trans total): (i) for gap fraction, the highest
value in FM reaches 22% compared to 16.5% in GV; (ii) for the total incident
light, the highest value is 14.5% compared to 6.7% in GV.
[24]
1085
[25]
1086
Table 5. Mean and range of values of canopy geometry and of the distribution of light in GV and FM.
n CO% LAI Trans direct Trans Trans diffuse Trans Trans total Trans
m2 m2 mole m2 d1 direct% mole m2 d1 diffuse% mole m2 d1 total%
[26]
GV 21 Mean values 11.5 2.6 0.65 6.1 0.77 7.3 1.42 6.7
Range of extreme values 7–16.5 1.9–3.8 0.07–1.54 0.67–14.3 0.27–1.65 2.6–15.8 0.36–2.77 1.8–11.8
FM 24 Mean values 14.6 2.5 0.29 14.7 0.89 13.5 1.14 14.5
Range of extreme values 9.8–22 1.8–3 0.07–0.91 5.3–43.3 0.5–1.84 5.9–28 0.71–1.8 7.7–24.3
1087
Discussion
The examples given in plots FM and GV indicate that mixed-beech forests are
spatially and temporally heterogeneous. Both plots show wide variations in
stem density, size-class distribution and age distribution that are ecological
traits of woodlands in a nearly natural state.
Despite a higher stem density, the GV site exhibited a lower volume of stem
wood than the FM site.
The importance of deadwood and discontinuities in the distribution of
saplings and seedlings are also ecological features regularly observed in natu-
ral, shady woodlands in Europe and North America (Jones 1945; Lemée 1978;
Mayer and Neumann 1981; Peterken 1996; Schnitzler 2002). These stand
characteristics, combined with the remarkable resistance of mixed-beech for-
ests in the FM and GV reserves (as compared to trees in the surrounding
[27]
1088
[28]
1089
permanent gap further increases the lateral penetration of incident light. This
explains why seedling densities are higher in FM and Acer pseudoplatanus can
regenerate more easily there than in GV. Presence of that specie in FM have a
retroactively impact and directly influences light arrival in the underlayer. A lot
of Acer pseudoplatanus seeds is probably coming from outside, and thus Acer
pseudoplatanus is probably invading the woodlands plots.
Better light conditions also explain why potential trees and trees of the
present have similar growth rates in FM: potential trees reach the canopy in
100 years, and their growth continues thereafter in the canopy at the same
rhythm. In GV many potential trees have grown in shade, and growth rates are
lower: the duration of the potential state lasts lasts 200 years or more. When
potential trees arrive at full light, stem growth increases with the development
of axes and foliage. Beech trees of the present however, grow less rapidly in GV
than in FM which suggests less favorable growth conditions, probably due to
the higher altitude.
The present-day, large range of ages and sizes recorded between trees of the
present in both plots can be interpreted as differences in growth patterns during
tree development: (i.e. alternating suppressed and released-growing periods).
Such growth processes have been recorded in all forests composed of shade
trees (Lemée 1978; Koop and Hilgen 1987; Peters 1992; Korpel 1995; Peterken
1996).
Acer pseudoplatanus presents a different strategy based on its rather low
tolerance for shade. Young Acer pseudoplatanus trees are numerous in open,
rocky areas or at margins, where there is no suppression phase.
Harsher conditions near the summits (for example, only 3 months of age at
900–1000 m) explain why canopy trees are smaller than in forests at lower
altitudes: only 40 m high for 300-year-old silver firs near the crests as compared
to 52–55 m for 180-year-old silver firs in the Guebwiller natural reserve (Re-
naud et al. 2000). In managed stands in the Vosges, some beech trees have been
known to reach 42 m in 120 years (a 2–3 mm annual growth rate between ages
20–70 according to Seynave 1999) compared to only 29 m in height for a 214-
year-old beech in plots of similar density. In the virgin forest of Dobroc (720–
1000 m altitude, granite), Slovakia, there are 45 m beeches that are 230-years-
old (Korpel 1995).
Multi-stemmed beech shape is a particularity of forests growing near the
timber line. Such architecture only occurs under conditions of stress (Carbiener
1966; Peters 1992; Closset 2000). These trees are not lower in stature than
single-stemmed trunks as suggested by Givnish (1984), but they are more
slender than single tree trunk of similar age. Actually, competition between
genetically identical trees has an impact of the lower volume of stem wood.
They form large, very stable individuals in the canopy because a multi-
stemmed growth form ensures better mechanical stability for the tree (Closset-
Kopp and Schnitzler (2000b), thus improving resistance to windthrow.
Multi-stemmed individuals form clusters of genetically similar stems, with the
[29]
1090
potential for separate existence. This explains why stems may have different
sizes and growth rates as related to age and social state.
At the present stage in the evolution of the two plots, old trees are very rare.
The three trees which are more than 300-years-old discovered in the plots are the
only ones recorded in the upper Vosges to date, but clearly more studies could be
done on this subject. The absence of very old trees differentiates forests of the
Upper Vosges from other virgin mixed-beech forests in Europe where there are
silver fir trees more than 400-years-old, and many more beech trees above the age
of 350 (Mayer and Neumann 1981; Korpel 1995; Cenusa 2001; Schnitzler 2002).
Given the present day composition of potential tree, we can predict that
beech will dominate the canopy.
Forest stands in the two reserves represent lesser-managed stands in the Upper
Vosges, but human impacts have nonetheless been multiple, and often irre-
versible. Remnants of more natural forest stands are located near the summits
and on steeper slopes, an inaccessibility which limits the data needed for a
comprehensive analysis of forest dynamics. The present-day surface of strictly
protected forests is also too small and too intermixed with managed forests and
open landscapes to serve as reference points for management principles, be-
cause they are not representative enough of a completely pristine landscape.
Given their rarity in Western Europe, these small areas must however be re-
garded as of utmost importance as a class of woodlands for nature conserva-
tion, research and education. A worthy objective of long-term conservation
efforts would be to re-create more substantial examples of missing types of
mixed-beech forests in the upper Vosges, and in the meantime, to leave un-
managed the remaining forests located in natural reserves.
Acknowledgment
We gratefully acknowledge the Parc Naturel Regional des Ballons des Vosges
for financial support through the study project from C. Schwoehrer. We are
also much indebted to J.L. Dupouey for his invaluable assistance in dendrol-
ogy and the logistical support of his laboratory (INRA Champenoux). We also
wish to express their gratitude to Y. Despert, L. Domergue, C. Kieffer and
P. Behr who have contributed core and data sampling.
References
Becker M. 1985. Le dépérissement du sapin dans les Vosges. Quelques facteurs liés à la détério-
ration des cı̂mes. Revue Forestière Française 37: 281–287.
[30]
1091
Becker M. 1989. The role of climate on present and past vitality of silver fir forests in the Vosges
mountains of northeastern France. Can. J. Forest. Res. 19: 1110–1117.
Bogenrieder A. 2001. Schwarzwald und Vogesen - ein vegetationskundlicher Vergleich. Mitteil-
ungen des Badischen Landesvereins für Naturkunde und Naturschutz 17: 745–792.
Bonneau M. 1995. Fertilisation des forêts dans les pays tempérés. ENGREF, Nancy.
Bonneau M., Faivre P., Gury M., Hétier J.M. and Le Tacon F. 1978. Carte pédologique de la
France à 1/100000. Notice explicative, Saint-Dié. INRA, Service d’Etude des Sols et de la Carte
Pédologique de la France, pp. 11.
Bradshaw R. and Holmqvist B. 1999. Danish forest development during the last 3000 years
reconstructed from regional pollen data. Ecography 22: 53–62.
Brethes A., Brun J.J., Jabiol B., Ponge J.F. and Toutain F. 1992. Typologie des formes d’humus.
In: Baize D. and Girard M.C.R (eds), éférentiel Pédologique, Association Française pour l’Etude
du sol. INRA, Paris, pp. 177–192.
Carbiener R. 1966. La végétation des Hautes Vosges dans ses rapports avec les climats locaux, les
sols et la géomorphologie. Thèse de Doctorat d’Etat, University of Orsay, France.
Cenusa R. 2001. Les montagnes de Calimani. Les forêts vierges de Roumanie, Giurgieu, Donita,
Bandku, Radu, Cenusa, Dissescu, Stoiculescu, Biris. ASBL Forêt wallone, pp.179–184.
Chazdon R.L. and Pearcy R.W. 1986. Photosynthetic light environments in a lowland and tropical
rain forest in Costa Rica. J. Ecol. 72: 533–564.
Closset D. and Schnitzler A. 2000a. Naturalité de la hêtraie d’altitude dans la Réserve naturelle du
Frankenthat Missheimle. University of Metz-Parc Naturel Régional des Ballons des Vosges.
Internal Report 37p.
Closset D. 2000. Sylvigénèse et naturalité de la hêtraie d’altitude dans la Réserve Naturelle du
Frankenthal-Missheimle. DEA Sciences Agronomiques, University of Metz.
COST Action E4 EU 1999. (European corporation in science and technology) Forest Research
Network Final report. Mission, goals, outputs, recommendations, linkage and patterns. Office
for Official Publications of the European Communities, Luxembourg.
Denslow J.S., Schultz J.C., Vitousek P.M. and Strain B.R. 1990. Growth responses of tropical
shrubs to treefall gap environments. Ecology 71: 165–179.
Duchaufour P. and Bonneau M. 1959. Une nouvelle méthode de dosage du phosphore assimilable
dans les sols forestiers. Bulletin de l’Association Française pour l’étude du sol 4: 193–198.
Edelin C. 1977. Images de làrchitecture des conifères. PhD thesis, University of Montpellier.
Ellenberg H. 1988. Vegetation of Middle Europe. Cambridge University Press, Cambridge.
Franks D.A. and McNaughton F. 1991. Stability increases with diversity in plant communities:
empirical evidence from the 1988 Yellowstone drought. Oikos 62: 360–362.
Frazer G.W., Canham C.D. and Lertzman K.P. 1999. Gap Light Analyser (GLA), Version 2.0:
Imaging software to extract canopy structure and gap light transmission indices from true-colour
fisheye photographs, User’s Manual and Program Documentation. Copyright 1999: Simo,
Fraser University, Burnaby, British Columbia, and Institute of Ecosystem Studies, Millbrook,
New York.
Garnier E. 1994. L’homme et son milieu: le massif du Grand Ventron à travers les âges. Parc
Naturel Régional des Ballons des Vosges, Internal Report.
Garnier E. 1998. Jalons pour une histoire de l’environnement: la réserve naturelle du Frankenthal-
Missheimle. Relations des sociétés et du milieu. Parc Naturel Régional des Ballons des Vosges,
Internal report.
Gilg O. 1997. Eléments d’évaluation de la naturalité des écosystèmes forestiers vosgiens. Eléments
conceptuels et méthodologiques, application aux hêtraies-sapinières de la réserve naturelle du
massif du Grand Ventron, DEA, University of Aix Marseille.
Givnish T.J. 1984. Leaf and canopy adaptations in tropical forests. In: Medina E., Mooney H.A.
and Vasques-Yanes C. (eds), Physiological Ecology of Plants of the Wet Tropics. Junk, The
Hague, pp. 51–83.
Hallé F. and Oldeman R.A.A. 1970. Essai sur l’architecture et la dynamique de croissance des
arbres tropicaux. Masson, Paris.
[31]
1092
Hallé F., Oldeman R.A.A and Tomlinson PB. 1978. Tropical Trees and Forests: An Architectural
Analysis. Springer Verlag, Berlin.
Helle T. and Järvinen O. 1986. Population trends in North Finnish land birds in relation to their
habitat selection and changes in forest structures. Oikos 46: 107–115.
Heuzé P. 2002. Impact à moyen terme des grands herbivores sur la hêtraie-sapinière vosgienne.
PhD Thesis, University of Metz.
Jones E. W. 1945. The structure and reproduction of the virgin forest of the north temperate zone.
New Phytologist 44: 130–148.
Koop H. and Hilgen P. 1987. Forest dynamics and regeneration mosaic shifts in unexploited beech
(Fagus sylvatica) stands at Fontainebleau (France). Forest Ecol. Manag. 20: 135–150.
Koop H. 1989. Forest Dynamics. Silvistar: A Comprehensive Monitoring System. Springer Verlag,
Berlin.
Korpel S. 1995. Die Urwälder der Westkarpaten. Gustav Fischer, Stuttgart.
Leibendgut H. 1959. Über Zweck und Methodik der Struktur und Zuwachsanalyse von Urwäldern.
Schweiz Zeitschrift für Forstwesen. 110: 110–124.
Lemée G. 1978. Structure et Fonctionnement des Ecosystèmes Terrestre. Masson, Paris.
Mansuy D. 1992. Les granites et la couverture pédologique dans le bassin du Rouge Rupt. Leurs
participations au contrôle de l’acidification des eaux (Cornimont – Vosges méridionales). PhD
thesis, University of Nancy.
Mayer H and Neumann M. 1981. Struktureller und entwicklungsdynamischer Vergleich der
Fichten-Tannen-Buchen Urwälder Rothwald Niederösterreich und Corkova Urwald. Kroatien
Forstwissenschaftliches Centralblatt 100: 111–132.
Motta R., Nola P. and Piussi P. 2002. Long-term investigations in strict forest reserve in the eastern
Italian Alps: spatio-temporal origin and development in two multi-layered subalpine stands.
J. Ecol. 90: 495–507.
Nicolini E. 1997. Approche morphologique du développement du hêtre (Fagus sylvatica.). PhD
thesis, Université de Montpellier.
Oberdorfer E. 1991. Die hochmontanen Wälder und subalpinen Gebüsche. Der Feldberg im
Schwarzwald. Subalpine Insel im Mittelgebirge. (Landesantstalt für Umweltschutz Baden-
Würtemberg), Karlsruhe.
Oberdorfer E. 1992. Süddeutsche Pflanzen-Gesellschaften. Teil IV: Wälder und Gebüsche. Gustav
Fischer Verlag, Stuttgart.
Office National des Forêts 2000. Evaluation de l’impact des ongulés sur la végétation forestière du
Frankenthal-Missheimle. Parc Naturel Régional des Ballons des Vosges, Internal Report.
Ooisterhius L., Oldeman R.A.A. and Sharik T.L. 1982. Architectural approach to analysis of
North American temperate deciduous forests. Can. J. Forest Res. 12: 835–847.
Oldeman R.A.A. 1974. L’architecture de la forêt guyanaise. Mémoires ORSTOM 73: 73.
Oldeman R.A.A. 1979. Quelques aspects quantifiables de l’arborigenèse et de la sylvigénèse.
Ecologica Plantarum 143: 289–312.
Oldeman R.A.A. 1990. Forests: Elements of Silvology. Springer Verlag, Berlin.
Orsini L. and Remy J.C. 1976. Utilisation du chlorure de cobaltihexamine pour la détermination
simultanée de la capacité d’échange et des bases èchangeables des sols. Science du Sol 4: 268–275.
Peterken G. P. 1996. Natural Woodlands: Ecology and Conservation in Northern Temperate
Regions. Cambridge University Press, Cambridge.
Peters R. 1992. Ecology of beech forests in the northern hemisphere. Ph.D thesis, University of
Wageningen.
Pierrel S. 2001. Détermination du climat lumineux, à l’aide de photographies hémisphériques et de
ses relations avec la répartition de la végétation dans une hêtraie sapinière de la réserve bio-
logique intégrale en forêt domaniale de Guebwiller. DEA. University of Strasbourg, France.
Renaud J.P., Kustner C. and Hauschild R. 2000. La Réserve Biologique Domaniale de Guebwiller
(Haut-Rhin). Présentation générale et résultats d’un premier inventaire réalisé selon le protocole
européen COST. Office National des Forêts, Internal Report.
Roloff A. 1999. Tree vigor and branching pattern. Journal of Forest Science 45: 206–216.
[32]
1093
Schnitzler A. 2002. Ecologie des forêts naturelles d’Europe. Biodiversité, sylvigénèse, valeur pat-
rimoniale des forêts primaires. Lavoisier, Paris.
Schwoehrer C. and Despert Y. 1999. Réserve naturelle du FM, plan de gestion. Parc Naturel
Régional des Ballons des Vosges, Internal Report.
Schwoehrer C. 1999. Réserve naturelle du massif du Grand Ventron, plan de gestion. Internal
Report PNRBV.
Seynave I. 1999. Analyse de la structure de deux peuplements forestiers mélangés équiennes: la
sapinière-hêtraie et la chênaie-hêtraie dans le nord-est de la France. Ph.D thesis, ENGREF,
Nancy.
Sittler B., Tennhardt T. and Shicarts E. 2000. Die Schützgebiete Russlands vor neuen Herausf-
orderungen. Natur und Landschaft 75: 1–9.
Souchier B. 1971. Evolution des sols sur roches cristallines à l’étage montagnard (Vosges). Mémoire
Service Carte géologique. Alsace Lorraine, pp.33.
Untereiner A., Advocat A. Stoehr B. and Despert Y. 2002. Carte de la végétation de la Réserve
naturelle du Frankenthal Missheimle. Parc Naturel Régional des Ballons des Vosges, Internal
Report.
Ulrich E. and Williot B. 1994. Les dépôts atmosphériques en France de 1850 à 1990. ONF, Internal
Report.
Vallauri D. and Poncet L. 2003. La protection des forêts en France métropolitaine. Livre blanc sur
la protection des forêts naturelles en France. Lavoisier, Paris.
Varlet-Granchet C., Gosse G., Chartier M., Sinoquet H., Bonhomme R. and Allirand J.M. 1989.
Mise au point: rayonnement solaire absorbé ou intercepté par un couvert végétal. Agronomie 9:
419–439.
Vester H. 1997. The Trees and the Forest. The role of tree architecture in canopy development; a
case study in secondary forests (Araracuara, Columbia) PhD thesis, University of Wageningen.
White P.S. 1978. Pattern process and natural disturbance in vegetation. Bot. Rev. 45: 229–299.
[33]
Biodiversity and Conservation (2006) 15:1095–1107 Springer 2006
DOI 10.1007/s10531-004-1868-4
-1
Key words: BIOCLIM, Bioclimatic modeling, Climate change, Cloud forest, Fagus, Sierra Madre
Oriental
Abstract. We examined the effects of climate change on the future conservation and distribution
patterns of the cloud forests in eastern Mexico, by using as a species model to Fagus grandifolia
Ehr. var. mexicana (Martı́nez) Little which is mainly located in this vegetation type, at the
Sierra Madre Oriental. This species was selected because it is restricted to the cloud forest,
where it is a dominant element and has not been considered for protection in any national or
international law. It is probably threatened due to the fact that it plays an important social role
as a source of food and furnishing. We used a floristic database and a bioclimatic modeling
approach including 19 climatic parameters, in order to obtain the current potential distribution
pattern of the species. Currently, its potential distribution pattern shows that it is distributed in
six different Mexican Priority Regions for Conservation. In addition, we also selected a future
climate scenario, on the basis of some climate changes predictions already proposed. The
scenario proposed is characterized by +2 C and 20% rainfall in the region. Under this
predicted climatic condition, we found a drastic distribution contraction of the species, in which
most of the remaining populations will inhabit restricted areas located outside the boundaries of
the surrounding reserves. Consequently, our results highlight the importance of considering the
effects of possible future climate changes on the selection of conservation areas and the urgency
to conserve some remaining patches of existing cloud forests. Accordingly, we believe that our
bioclimatic modeling approach represents a useful tool to undertake decisions concerning the
definition of protected areas, once the current potential distribution pattern of some selected
species is known.
Introduction
The cloud forests represent one of the most interesting biological systems in the
Neotropical region (Luna et al. 1999). They are usually rare, vulnerable and
threatened in the world. Its northern distribution limit is the Sierra Madre
Oriental, in the state of Tamaulipas, Mexico (Briones 1991) and its southern
one reaches Argentina (Webster 1995).
[35]
1096
[36]
1097
Figure 1. Model of the potential distribution of Fagus grandifolia var. mexicana, on relationship
to the known records. On the right corner the potential distribution of the species in the state of
Oaxaca is shown.
[37]
1098
Some recent data documenting the wild populations status of the species
have been generated, especially in the states of Tamaulipas, Hidalgo and Ve-
racruz (Williams et al. 2003). In some sites the species is considered extinct,
whereas, in other places there are still some small patches of what used to be a
cloud forest of Fagus grandifolia. So far, the species has not been recorded in
the cloud forests of Querétaro, which is a neighbor state of Hidalgo and San
Luis Potosı́ and bears similar environmental conditions for hosting the species.
Probably the absence of Fagus in Querétaro is due to physiographic differences
as suggested by Cartujano et al. (2002). However, it might be also possible that
the species has been misidentified due to its morphological similarity to
Carpinus sp., Ostrya sp. or Ulmus sp., as has been suggested by López and
Cházaro (1995).
Thus, the purpose of this work is to undertake a comprehensive review of the
current situation of the cloud forests in eastern Mexico by using Fagus gran-
difolia var. mexicana as our species model. Consequently, we attempted to
undertake the following actions: (1) to document the current recorded distri-
bution of the species in Mexico; (2) to obtain the potential distribution patterns
of the species; (3) to assess the effects that the potential distribution pattern of
the species might have, under a climatic change scenario; (4) to evaluate the role
that the Protected Natural Areas and the Priority Regions of Mexico will be
playing for the long-term conservation of cloud forests; (5) to propose a general
strategy for attempting the conservation of the oriental Mexican cloud forests.
Accordingly, the approach of this work includes the utilization of biocli-
matic models that enable to explain the current situation of the eastern cloud
forests of Mexico, on the basis of the potential distribution pattern of a rep-
resentative species (Fagus grandifolia var. mexicana) that is used as a model. In
addition, we present an attempt to assess the future distribution of the cloud
forests, using the species data, once a predicted scenario due to climatic change
is included (Téllez and Dávila 2003).
Methods
The plant geographic distribution information that we used in this analysis was
obtained from the database of the World Information Network of Biodiversity
(REMIB) (http://www.conabio.gob.mx/remib/doctos/remibnodosdb.html).
The herbarium data were obtained from the National Herbarium of Mexico
(MEXU), from 29 specimens that beard geo-referenced information
(i.e. complete latitude, longitude, and elevation). The taxonomical identifica-
tion of the specimens was undertaken by Drs. Shen Shung-Fu and Kevin
Nixon who are important specialists of the Fagaceae. On the other hand, the
information concerning the vegetation structure and ecological attributes of
the species that is included in the discussion of this work was obtained from
relevant literature (Malda 1990; López and Cházaro 1995; Luna et al. 2000;
Williams et al. 2003).
[38]
1099
The bioclimatic modeling approach used in this work was that of the pro-
gram ANUCLIM (Houlder et al. 2000). The program uses mathematically and
statistically interpolated climatic surfaces (digital files in raster format) that
were estimated using the information obtained from a standard network of
meteorological stations. The climatic surfaces or digital files were generated
using thin plate smoothing spline methods in the ANUSPLIN package
(Hutchinson 1991, 1995a, b, 1997; Hutchinson and Gessler 1994). These sur-
faces include long-term monthly mean values of precipitation and temperature
from more than 6200 stations (4000 stations including temperature data and
6000 including precipitation data from the same set of stations). The estimated
mean errors for those surfaces were between 8 and 13% for monthly precipi-
tation values and about 0.4–0.5 C for temperature values. These errors are
similar to those found in the standard meteorological instruments (Nix 1986).
We produced a bioclimatic profile for Fagus grandifolia var. mexicana, using
the program BIOCLIM. The derivation of the bioclimatic profile was based on
selected-simple-matching thresholds. The values for each of the 19 bioclimatic
parameters (Table 1), were assessed by a systematic scanning throughout a grid
of data points. We used the profile to predict potential distribution pattern of
the species. Using the homoclime matching principle, we identified those points
on the climate grid, where the climatic conditions were present within the limits
summarized in the bioclimatic profile of the species (Booth et al. 1987).
We matched the bioclimatic profiles against a grid of data points that con-
tained climatic data from the existing network of stations (bioclimatic
[39]
1100
Results
The results obtained suggest that the present distribution pattern known for
Fagus grandifolia var. mexicana, is indeed correct and complete, due to the fact
that in all cases, the collecting sites fitted within the limits of the potential
distribution area obtained in the analysis (Figure 1). Thus, the species is re-
stricted to the Sierra Madre Oriental from the state of Tamaulipas to southern
Veracruz, as has been stated by Williams et al. (2003). However, on the basis of
the potential distribution assessment of the species, we believe that probably its
southern limit might extend to the state of Oaxaca.
However, field verifications should be done before we can assure it (Figure 1).
The results also point out that the species is restricted to unique climatic condi-
tions in the Sierra Madre Oriental, as it is shown in its bioclimatic profile
(Table 1). Its climatic uniqueness represents the specific spots or areas along the
Oriental Sierra Madre where it can grow. In other words, although we state that
Fagus grandifolia var. mexicana grows along the Sierra Madre, the fact is that it
only grows in some specific areas that have a unique combination of climatic
attributes and do not grow in others that have other climatic features.
[40]
1101
Figure 2. Model of the potential distribution of Fagus grandifolia var. mexicana on relationship to
the Priority Regions for Conservation (CONABIO), once the proposed climate change scenario
was entered.
[41]
1102
Discussion
[42]
1103
[43]
1104
inclusion of only 29 records data for the model generation, might seems not
representative of the species distribution pattern. However the records used
cover, in general terms, all the environmental conditions that theoretically the
species might occupy (the geographic, ecological and altitudinal range of the
taxon).
In addition, natural systems complexity represents a challenge for under-
taking a modeling approach. In particular, the evident limitation of the bio-
climatic models is the lack of inclusion of information concerning biotic
interactions, evolutionary changes, as well as relevant biological processes such
as dispersion (Pearson and Dawson 2003). Consequently, the existence of
certain degree of errors is probably unavoidable.
Also, the bioclimatic data, due to its own nature, shows two kinds of errors:
(1) the omission (= the lack of consideration of the space that is occupied by
the niche; (2) commission (= the consideration of a space that is not occupied
by the niche). Consequently, each algorithm used to model a species ecological
niche, has a combination of commission and omission errors (Peterson and
Vieglais 2001).
Even though, the existence of these errors is recognized, we believe that the
bioclimatic modeling represents a useful tool or starting point for under-
standing the current and potential distribution patterns of animals and plants.
Its usefulness has already been proved for some species at certain scales, in
which this approach has generated relevant information (Pearson and Dawson
2003).
In the case of this study, the model clearly reflects that the spatial climatic
resolution used to correlate it to the species records that were included, enabled
a precise and solid bioclimatic profile of Fagus.
Finally, we believe that with the present biological information, it is fea-
sible and recommendable to carry out a similar exercise for other plant
groups. Endemic species and main elements of plant communities should be
especially important to be submitted to a bioclimatic modeling. By this
means, we can increase the probability of proposing adequate conservation
strategies. In the particular case of this study, the results obtained show that
through the bioclimatic approach, we can be able to focus in long-term
management, planning, and development of new, flexible, and dynamic forms
of wildlife and resource conservation (Nix 1986; Lindenmayer et al. 1991;
Téllez and Dávila 2003).
Acknowledgements
[44]
1105
References
Alcantara O. and Luna V.I. 1997. Florı́stica y análisis biogeográfico del bosque mesófilo de
montaña de Tenango de Doria, Hidalgo, México. Anales del Instituto de Biologı́a, Universidad
Nacional Autónoma de México, Serie Botánica 68: 57–106.
Alcántara O. and Luna V.I. 2001. Análisis florı́stico de dos áreas con bosque mesófilo de montaña
en el estado de Hidalgo, México: Eloxochitlán y Tlahuelompa. Acta Botánica Mexicana 54:
51–87.
Anónimo 2000. Proyecto de Norma Oficial Mexicana PROY-NOM-059-ECOL-2000, protección
ambiental-especies de flora y fauna silvestres de México-Categorı́as de riesgo y especificaciones
para su inclusión, exclusión o cambio. Lista de especies en riesgo. Diario Oficial de la Federa-
ción. de octubre de 2000.
Arriaga L., Espinoza J.M., Aguilar C., Martı́nez E., Gómez L. and Loa E. (coordinadores) 2000.
Regiones Terrestres Prioritarias de México. Comisión Nacional para el Conocimiento y Uso de
la Biodiversidad, México.
Booth T.H., Nix H.A. and Hutchinson M.F. 1987. Grid matching: a new method for homoclime
analysis. Agric. For. Meteorol. 39: 241–255.
Briones O.L. 1991. Sobre la flora, vegetación y fitogeografı́a de la Sierra de San Carlos, Tamau-
lipas. Acta Botánica Mexicana 16: 15–44.
Canziani O.F. and Diaz S. 1998. Latin America. In: Watson R.T., Zinyowera M.C., Moss R.H.
and Dokken D.J. (eds), The Regional Impacts of Climate Change: An Assessment of Vulnera-
bility. Special Report of IPCC Working Group II. Cambridge University Press, Cambridge, UK,
pp. 187–230.
Cartujano S., Zamudio S., Alcantara O. and Luna I. 2002. El bosque mesófilo de montaña en el
municipio de Landa de Matamoros, Querétaro, México. Boletı́n de la Sociedad Botánica de
México 70: 13–44.
Churchill S.P., Griffin III D. and Lewis M. 1995. Moss diversity of the Tropical Andes. In:
Churchill S.P., Balslev H., Forero E. and Luteyn J.L. (eds), Biodiversity and Conservation of
Neotropical Montane Forestes. Proceedings of the Neotropical Montane Forest Biodiversity and
Conservation Symposium. The New York Botanical Garden, 21–26 June 1993, New York,
pp. 335–348.
ESRI (Environmental Scientific Research Institute) 2000. ArcView 3.2. ESRI. Redlands, Califor-
nia, USA.
Giorgi F., Meehl G.A., Kattenberg A., Grassl H., Mitchell J.F.B., Stouffer R.J., Tokiioka T.,
Weaver A.J. and Wigley T.M.L. 1998. Simulated changes in vegetation distribution under global
warning. In: Watson R.T., Zinyowera M.C., Moss R.H. and Dokken D.J. (eds), The Regional
Impacts of Climate Change: An Assessment of Vulnerability. Special report of IPCC working
group II. Cambridge University Press, Cambridge, UK, pp. 427–437.
Houlder D.J., Hutchinson M.F., Nix H.A. and McMahon J.P. 2000. ANUCLIM 5.1 User Guide,
Centre for Resource and Environmental Studies. Australian National University, Australian
Capital Territory, Canberra.
Houghton J.T., Callander B.A. and Varney S.K. 1992. Climate change 1992. The Supplementary
Report to the IPCC Scientific Assessment. Cambridge University Press, Cambridge, UK.
Hutchinson M.F. 1991. The application of thin-plate smoothing splines to continent- wide data
assimilation. In: Jasper J.D. (ed.), BMRC Research Report Series. Bureau of Meteorology,
Melbourne, Australia, pp. 104–113.
Hutchinson M.F. 1995a. Interpolating mean rainfall using thin plate smoothing splines. Int. J.
Geogr. Inform. Syst. 9: 385–403.
Hutchinson M.F. 1995b. Stochastic space-time weather models from ground-based data. Agric.
For. Meteorol. 73: 237–264.
Hutchinson M.F. 1997. ANUSPLIN. Version 4.1. User guide, Centre for Resource and Envi-
ronmental Studies, Australian National University, Australian Capital Territory, Canberra.
[45]
1106
Hutchinson M.F. and Gessler P.E. 1994. Splines – more than just a smooth interpolator. Geoderma
62: 45–67.
Johnston M.C., Nixon K., Nesom G.L., and Martı́nez M. 1989. Listado de plantas vascul-
ares conocidas de la Sierra de Guatemala, Gómez Farı́as, Tamaulipas, México. Biotam 1:
21–33.
Kappelle M., Van Vuuren M.M.I. and Baas P. 1999. Effects of climate change on biodiversity. A
review and identification of key research issues. Biodiv. Conserv. 8: 1383–1397.
Karl T.A. 1998. Regional trends and variation of temperature and precipitation. In: Watson R.T.,
Zinyowera M.C., Moss R.H. and Dokken D.J. (eds), The Regional Impacts of Climate Change:
An Assessment of Vulnerability. Special Report of IPCC Working Group II. Cambridge Uni-
versity Press, Cambridge, UK, pp. 411–425.
Lindenmayer D.B., Nix H.A., McMahon J.P., Hutchinson M.F. and Tanton M.T. 1991. The
conservation of Leadbeater’s possum, Gymnobelideus leadbeateri (McCoy): a case study of the
use of bioclimatic modelling. J. Biogeogr. 18: 371–383.
Little E.L. Jr. 1965. Mexican beech, a variety of Fagus grandifolia. Castanea 30: 167–170.
López M.L. and Cházaro B.M. 1995. Plantas leñosas raras del bosque mesófilo de montaña.
I. Fagus mexicana Martı́nez (Fagaceae). Boletı́n de la Sociedad Botánica de México 57:
113–115.
Luna V.I., Almeida L., Villers L. and Lorenzo L. 1988. Reconocimiento florı́stico y consideraciones
fitogeográficas del bosque mesófilo de montaña de Teocelo, Veracruz. Boletı́n de la Sociedad
Botánica de México 48: 35–63.
Luna V.I., Alcántara A.O., Espinosa O.D.E. and Morrone J.J. 1999. Historical relationships of the
Mexican cloud forests: a preliminary vicariance model applying Parsimony Analysis of Ende-
micity to vascular plant taxa. J. Biogeogr. 26: 1299–1306.
Luna V.I., Alcántara A.O., Morrone J.J. and Espinosa O.D.E. 2000. Track analysis and conser-
vation priorities in the cloud forests of Hidalgo, Mexico. Div. Distribut. 6: 137–143.
Luna V.I., Morrone J.J., Ayala A.O. and Organista D.E. 2001. Biogeographical affinities among
Neotropical cloud forests. Plant Systemat. Evol. 228: 229–239.
Malda G.B. 1990. Plantas vasculares raras, amenazadas y en peligro de extinción en Tamaulipas.
Biotam 2: 55–61.
McNeely J.A., Gadgil M., Leveque C., Padoch C. and Reedford K. 1995. Human influences on
Biodiversity. In: Heywood V.H. and Warton R.T. (eds), Global Diversity Assessment. Cam-
bridge University Press, Cambridge, UK pp. 711–821.
Moguel P. and Toledo M.V.M. 1999. Biodiversity conservation in traditional coffee systems of
Mexico. Conserv. Biol. 13(1): 11–21.
Morrone J.J. and Crisci J.V. 1995. Historical biogeography: introduction to methods. Annu. Rev.
Ecol. Systemat. 26: 373–401.
Morrone J.J. and Espinosa M.D. 1998. La relevancia de los atlas biogeográficos para la conser-
vación de la biodiversidad mexicana. Ciencia (México) 49: 12–16.
Nix H.A. 1986. A Biogeographic analysis of Australian elapid snakes. In: Longmore R. (ed.), Atlas
of the Elapid snakes of Australia. Flora and Fauna. 7: 4–15.
Neilson R.P. 1998. Simulation of regional climate change with global coupled climate models and
regional modelling techniques. In: Watson R.T., Zinyowera M.C., Moss R.H. and Dokken D.J.
(eds), The Regional Impacts of Climate Change: An Assessment of Vulnerability. Special Report
of IPCC Working Group II. Cambridge University Press, Cambridge, UK, pp. 439–456.
Oldfield S.F., Lusty C. and MacKinven A. 1998. The World List of Threatened Trees. World
Conservation Press.
Pearson R.G. and Dawson T.P. 2003. Predicting the impacts of climate change on the distribution
of species: are bioclimate envelope models useful? Global Ecol. Biogeogr. 12: 361–371.
Pérez P.M. 1994. Revisión sobre el conocimiento dendrológico, silvı́cola y un censo de las pob-
laciones actuales del género Fagus en México. Tesis de maestrı́a (Biologı́a). Facultad de Ciencias.
Universidad Nacional Autónoma de México, México, DF, 146 pp.
[46]
1107
Pérez P.M. 1999. Las hayas de México.Monografı́a de Fagus grandifolia spp. mexicana. Univers-
idad Autónoma Chapingo, Chapingo, México, 51 pp.
Peterson A.T. and Vieglais D.A. 2001. Predicting species invasions using ecological niche modeling:
new approaches from bioinformatics attack a pressing problem. BioScience 51: 363–371.
Rzedowski J. 1996. Análisis preliminar de la flora vascular de los bosques mesófilos de montaña de
México. Acta Botánica mexicana 35: 25–44.
Pérez P.M. 1999. Las hayas de México.Monografı́a de Fagus grandifolia spp. mexicana. Univers-
idad Autónoma Chapingo, Chapingo, México, 51 pp.
Shen C.F. 1992. A monography of the genus Fagus Tourn. ex. L. (Fagaceae). Dissertation, City
University of New York, New York.
Téllez V.O. and Dávila A.P. 2003. Protected areas and climate change: a case study of the cacti in
the Tehuacán-Cuicatlán biosphere reserve, México. Conserv. Biol. 17(3): 846–853.
Vovides A.P., Luna V. and Medina G. 1997. Relación de algunas plantas y hongos mexicanos
raros, amenazados o en peligro de extinción y sugerencias para su conservación. Acta Botánica
Mexicana 39: 1–42.
Webster G.L. 1995. The Panorama of Neotropical Cloud Forests. In: Churchill S.P., Balslev H.,
Forero E. and Luteyn J.L. (eds), Biodiversity and Conservation of Neotropical Montane
Forestes. Proceedings of the Neotropical Montane Forest Biodiversity and Conservation Sym-
posium. The New York Botanical Garden, 21–26 June 1993, New York, pp. 53–77.
Williams L.G., Rowden A. and Newton A.C. 2003. Distribution and stand characteristics of relict
populations of Mexican beech (Fagus grandifolia var mexicana). Biol. Conserv. 109: 27–36.
[47]
Biodiversity and Conservation (2006) 15:1109–1128 Springer 2006
DOI 10.1007/s10531-004-2178-6
OLIVARIMBOLA ANDRIANOELINA1,
HERY RAKOTONDRAOELINA2, LOLONA RAMAMONJISOA1,
JEAN MALEY3, PASCAL DANTHU4 and JEAN-MARC BOUVET5,*
1
Silo national des Graines Forestières, Ambatobe BP 5091, Antananarivo, Madagascar; 2PCP Forêts
et Biodiversité/Fofifa DRFP BP 904, Antananarivo, Madagascar; 3Institut des Sciences de l’Evolution
Université de Montpellier II, cc 065, Universite´ Montpellier 2, Place Eugène Bataillon, 34095
Montpellier Cedex 05, France; 4PCP Foreˆts et Biodiversité/Cirad, BP 853, Antananarivo,
Madagascar; 5Cirad-Foreˆt, Campus international de Baillarguet TA10/C, BP 5035, 34398 Mont-
pellier Cedex, France; *Author for correspondence (e-mail: [email protected]; phone:
+33-467593728; fax: +33-467593733)
Key words: Chloroplast microsatellites, Conservation, Gene flow, Genetic structure, Post-glacial
recolonisation, RAPD
Abstract. There is an urgent need to maintain and restore a broad genetic base for the management
of Dalbergia monticola, a very economically important but endangered tree species in Madagascar.
Random amplified polymorphism DNAs (RAPDs) and chloroplast microsatellite markers were
used to quantify the genetic variation and to analyse the geographic distribution of diversity. Ten
locations covering most of the natural range were sampled. Sixty-three RAPD polymorphic and 15
monomorphic loci were obtained from 122 individuals. Genetic diversity was low and very close
among populations and regions. The unrooted neighbour-joining tree exhibited 4 groups, repre-
senting 6% (p = 0.000) of the total variation. The greater part of the variance, 81%, was observed
within populations. A Mantel test suggested that genetic distances between populations were
weakly correlated with geographic distances (R = 0.46, p = 0.12). The three chloroplast micro-
satellite primers assayed on 100 individuals gave 13 chlorotypes. Most of the populations showed 2
or 3 haplotypes. Haplotype diversity for the total population was equal to HeCp = 0.83 and ranged
from 0.00 to 0.80 among the populations. The unrooted neighbour-joining tree exhibited 4 groups
corresponding to the four regions representing 80% (p = 0.0000) of the total variation. Genetic
diversity varies with regions, the north and south being less variable. Chlorotype distribution, the
phylogenetic tree and historical information suggest that putative refugias in the centre-north
region originating from the early Holocene could explain the pattern of variation observed today.
By combining the results obtained at nuclear and organellar loci, a strategy of conservation based
on evolutionarily significant units is proposed.
Introduction
The separation from Gondwana, 158–160 million years ago, has led to high
endemism in Madagascar recognised as one of the most original in the world
(Myers et al. 2000; Briggs 2003). About 80% of the plant species are endemic
and the richness of fauna and flora is great. Present patterns of the Malagasy
[49]
1110
[50]
1111
DNA extraction
DNA was extracted from dried leaves, following the modified protocol described
by (Bousquet et al. 1990). Leaves (100 mg) were ground to a fine powder with a
[51]
1112
Table 1. Characteristics of the Dalbergia monticola populations sampled in the natural range.
mortar and pestle in a 1.5 ml Eppendorf tube under liquid nitrogen. DNA
extraction buffer (5 ml) was added (100 mM Tris–HCl (pH 8.0), 20 mM EDTA,
1.4 M NaCl, 1% PEG 6000, 2% MATAB, 0.5% sodium sulphite). The tube was
then incubated at 74 C for 20 min. Samples were washed with wet chloroform
(CIAA, 24:1) to remove cellular debris and protein. After 15 min of centrifu-
gation at 5000 · g, the liquid phase was transferred to 15 ml tubes. Isopropanol
(5 ml) was added and mixed gently to precipitate the DNA. The resulting DNA
pellets were resuspended in 400 ll of sterile water overnight at 37 C and stored
at 20 C until required.
RAPD methods
[52]
1113
not used so as to ensure good repeatability of the RAPD process and avoid
misscoring.
Data analysis
In the case of RAPD data, amplified DNA marker bands were scored in a
binary manner as either present (1) or absent (0) and entered into a binary
data matrix. Each PCR product was assumed to represent a single locus
because homology is generally high at the intraspecific level. The frequency of
each band and the percentage of polymorphic loci (%P) were calculated in
each population.
Shannon’s diversity index was P used to assess molecular variation. This
parameter, defined as IRAPD = i¼2 i¼1 pi log2 pi, where pi is the frequency of
the RAPD phenotype (presence (1) or absence (0) of the band), is frequently
used in the absence of assumptions concerning the Hardy–Weinberg equilib-
rium (Gillies et al. 1997; Martin and Hernandez Bermejo 2000). It was calcu-
lated for each locus and averaged over loci to provide the degree of variation
within each population, IRAPDpop. Shannon’s index was also estimated for the
whole sample considered as a single population, IRAPDtot. The expected genetic
heterozygosity PHer was estimated with the fixation index F equal to zero.
HeRAPD ¼ 1 ni¼1 p2i (where pi is the frequency of the allele i in a population),
[53]
1114
where mp and ms are pollen and seed migration rates, respectively. In the case
of RAPDs the value is biased, and probably overestimated, due to the
amplification of the cytoplasmic genome.
The association between geographic and genetic distances was estimated as
a Spearman’s rank correlation coefficient (q). The null hypothesis of the
association was tested with the Mantel test using Fstat software (Goudet
2001).
Minimum spanning networks between haplotypes (each network embedding
all minimum spanning trees for a given distance matrix) were computed with the
MINSPNET (Excoffier and Smouse 1994), provided with Arlequin software
version 2000 (Schneider et al. 2000). The distance matrix between haplotypes was
calculated using a distance matrix based on the square of the difference in
microsatellite size with the formula
X
L
Dij ¼ ðail ajl Þ2
l¼1
where aij and ajl give the allele size in base pairs at the lth locus of individuals i
and j, respectively.
[54]
1115
Results
Table 2. Population size (N), number of haplotypes (na), effective number of haplotypes (ne),
Shannon’s index (IRAPD), percent of polymorphic RAPD loci (%P), RAPD diversity (HeRAPD) for
each population. The standard error of each parameter is given between brackets.
[55]
1116
Table 3. Population size (N), number of haplotypes (na), effective number of haplotypes (ne),
Shannon’s index (IRAPD), percent of polymorphic RAPD loci (%P), RAPD diversity (HeRAPD) for
each region. The standard error of each parameter is given in brackets.
Figure 1. Unrooted neighbour-joining tree based on RAPD markers. The four regions are: the
South (S), the Centre (C), the Centre-north (C-N), and the North (N). The tree was drawn with the
Fst matrix given by analysis of molecular variance (Excoffier et al. 1992) with Arlequin software
version 2000 (Schneider et al. 2000).
[56]
1117
centre-north and north. The link of Antsevabe with the cluster of the south is
difficult to explain. This clustering suggested a good relationship between the
geographical distance and the genetic distance. This was partly confirmed by
the Mantel test showing a coefficient of correlation between genetic and geo-
graphic distances moderately high (R = 0.46), but not significantly different
from zero (p = 0.12), this relationship is illustrated in Figure 2a.
Figure 2. Relation between genetic and geographical distances for the RAPD and chsloroplast
microsatellite markers. Matrices of genetic distances were calculated using the AMOVA-derived Fst
(Arlequin software version 2000 (Schneider et al. 2000)).
[57]
1118
Table 4. Allelic characteristics in base pairs for the three loci, allelic combination of each chlor-
otype and frequencies of the chlorotypes present in each population and in the total population of
Dalbergia monticola.
[58]
1119
Table 5. Diversity parameters assessed with chloroplast microsatellite markers for each popula-
tion and the total population of Dalbergia monticola. Population size (N), number of haplotypes
(na), effective number of haplotypes (ne), Shannon’s index (ICp), haplotypic diversity (HCp).
The distribution of the 13 haplotypes across the natural range showed a geo-
graphical structure (Figure 4). Some haplotypes were present in contiguous
populations and regions but none was scattered among distant populations
and regions. The network of haplotypes in Figure 4b shows that haplotype C5
has the highest number of connections and C5 is among the most frequent
(13%). Generally, haplotypes that are closely related to each other, differing by
a single mutation at a microsatellite repeat, are geographically close. For
example, the set of haplotypes C7, C11 and C13 is located in the south,
haplotypes C1, C2, C4, C1 and C5 are in the centre-north, haplotypes C8, C9,
C12, and C10 are in the centre and haplotypes C3, C5 in the north. This
relation of isolation by distance is confirmed by the significant correlation
between genetic distance (Fst) and geographic distance (q = 0.57; p = 0.02).
Figure 2b shows that this linear relationship increased strongly up to 300 km
and then reached a plateau, emphasising that Fst is more or less constant and
close to 1 after this distance.
[59]
1120
Figure 3. Unrooted neighbour-joining tree based on chloroplast microsatellites markers. The four
regions are: the South (S), the Centre (C), the Centre-north (C-N), and the North (N). The tree was
drawn with the Fst matrix given by analysis of molecular variance (Excoffier et al. 1992) with the
Arlequin software version 2000 (Schneider et al. 2000).
Discussion
[60]
1121
2002; Newton et al. 2002; Bouvet et al. 2004). Shannon’s diversity index for
these species ranged from 0.42 to 0.65, whereas it is 0.30 (0.21) for D. monti-
cola. Values of expected heterozygosity are similar to those estimated for
perennial species growing on smaller islands, but the percentage of polymor-
phic loci in Dalbergia monticola is higher (Kwon and Morden 2002). Although
RAPD markers should be considered with caution as they are not a good
predictor of total genetic diversity (Nybom and Bartish 2000), the low values of
estimated diversity parameters in this species could be explained by the limited
range of the species. An island population contains less genetic diversity than a
mainland population (Barett 1998). Although Madagascar is a single large
island the isolation from other sources gene could favour the effect of genetic
drift. Another factor that can explain the low diversity is a recent and rapid
expansion after a bottleneck (Savolainen and Kuittinen 2000).
There are few published studies concerning the diversity of cpSSrs in
angiosperm tree species. The present study shows, however, that the results
varied according to the species and the number of primers used. With 3
primers, 100 individuals distributed in 10 populations, the 13 haplotypes
identified in Dalbergia monticola are consistent with previous studies on an-
giosperms (Palmé and Verdramin 2002; Grivet and Petit 2003). In addition, the
haplotypic heterozygosity HeCp = 0.88 is high compared with estimates for
other tree species (Palmé and Verdramin 2002) and suggest high diversity
compared to a number of temperate tree species studied.
[61]
1122
The distribution of genetic diversity with RAPD markers indicates that most of
the variation is present within populations (81%). This can be explained by
some biological patterns such as long-lived woody perennials and outcrossed
insect pollination species according to Nybom and Bartish (2000). Nested
analysis of variance with RAPDs shows that only 6% of the total variation is
attributed to among-group variation, 16% to among populations within
groups, and 76% among individuals within populations. This is low but sig-
nificant differentiation among regions and populations is closely related to the
correlation (moderately high but not significant) between the geographic and
genetic distance. This pattern of isolation by distance is often observed in tree
species wind or insect-pollinated when long distances within the natural range
are considered (Bekessy et al. 2002), while no significant correlation is seen
when short distances are taken into account (Schierenbeck et al. 1997).
The distribution of the diversity is very different using chloroplast micro-
satellites. We note a strong differentiation between regions (80%) and a low
percentage of variance within populations (16%). This is a classical result for
angiosperm forest trees (Raspé et al. 2000) because of the maternal inheritance
of the chloroplast DNA, and seeds are dispersed over shorter distances com-
pared with pollen (Echt et al. 1998). In Dalbergia monticola, seed dispersal is
mainly barochorous and contributes to this strong structure. The relationship
between genetic distance and geographic distance for cpSSr (Figure 2b)
confirms this strong structure and shows that above 300 km Fst is close to 1.
The different patterns of genetic structure of RAPD and cpSSr can also be
viewed through the relative rate of pollen flow to seed flow. With FstCp = 0.84
and FstRAPD = 0.23, the calculation of r showed that gene flow by pollen is 15
times greater than by seeds. The difference is marked but is smaller than for
temperate species such as Fagus silvatica (r = 84), Quercus robur (r = 286),
Quercus petraea (r = 500) (King and Ferris 1998). This can be explained by
the fact that they are wind-pollinated whereas D. monticola is insect-pollinated.
One of the most consistent factors that strongly influenced the partition of ge-
netic diversity of tree species over the natural range is the last glaciation period
and the subsequent migrations from the refuges (Aide and Riviera 1998; Willis
and Whittaker 2000). In Europe (Verdramin et al. 1998), South America (Dutech
et al. 2000; Bekessy et al. 2002) and the Sudano-Sahelian region of Africa
(Bouvet et al. 2004), the genetic variability of tree species has been analysed in
connection with the last glaciation period, 15–20,000 years ago, but no studies,
to our knowledge, have attempted to establish this relation in Madagascar.
As many tropical regions, the island underwent climate fluctuation during
the Pleistocene and Holocene periods, but also the impact of human activities.
[62]
1123
Three critical periods in the Quaternary should be mentioned. The first was the
Last Glaciation Maximum around 20,000–21,000 years ago which led to an
extremely cold and arid climate. According to (Adams and Faure 1997), during
this episode the rainforest in Madagascar was limited to the north-eastern part
of the country (between latitudes 12 and 16), the high plateaus were covered
with ericoid, graminoid and composite-dominated vegetation and the oriental
escarpment and coastal plain presented tropical woodland characterised by
low, open canopy and usually deciduous trees. Other authors have concluded
that tropical montane vegetation belts must have been vertically displaced 900–
1500 m (Burney 1996). Such a cold-driven displacement of vegetation zones
would have confined the island’s humid forest zones to the relatively small
land area along the east coast, with isolated patches elsewhere (low elevation
humid refugia). From the early Holocene (8000 14C years ago) the rainforest
occupied a broader zone than today. Early Holocene warming led to a gradual
replacement of ericoid vegetation in the mid-elevation with forest in wetter
locations of the eastern escarpment and a rise in the level of Lake Alaotra
(Burney 1996). Some analyses, based on pollen records in high plateau loca-
tions (Lake Itasy, Ankatra, Andasibe) suggested that the rain forest was
present up to elevations of 1800–2000 m (Straka 1996). The third period cor-
responds to human settlements 2000 years ago and the effect of burning which
increased the frequency of fire and resulted in disturbance of the too wet or too
dry ecosystems which generally support a natural fire regime (Burney 1996).
Assuming a limitation of the rainforest in the northern part of the range
during LGM, we can hypothesise that this region may be the zone of putative
refuges for sub-montane rainforest tree species. The assumption cannot be
verified by the results of this study because no sample was collected in the
northern part of the range in the latitude 15. However, our study proposes
another putative refuge. Generally, the highest diversity regions are consid-
ered as putative refuges. The higher diversity in the centre-north region and
the decrease in diversity in the southern and northern limits of the natural
range (Table 6) suggest that the zone of Didy, Ambohijanahary and Ant-
sevabe could be a putative refuge for this species. Phylogenetic trees based on
haplotypes (Figures 4b) and their distribution across the natural range
(Figure 4a) also point to a pattern of migration from the centre-north region.
Table 6. Diversity parameters assessed with chloroplast microsatellite markers for the four main
regions of the natural range of Dalbergia monticola. Population size (N), number of haplotypes
(na), effective number of haplotypes (ne), Shannon’s index (ICp), haplotypic diversity (HCp).
[63]
1124
[64]
1125
Acknowledgements
We are very grateful to colleagues from ‘‘Silo National des Graines Forestières’’
for assistance in collecting samples, to Alexandre Vaillant and Céline Cardi from
Cirad-forêt for their technical assistance in the laboratory, and to the anonymous
reviewers. These results are part of master of science thesis of O. Andrianoelina,
which was financially supported by the Millennium Seed Bank (MSB) project,
managed by the Seed Conservation Department, RBG Kew. The study was
carried out at SNGF, which receives funds from the Malagasy Government, as
well as at Cirad-Forêt, where laboratory work was done. L. Ramamonjisoa, head
of the technique division in SNGF and J.M. Bouvet, head of the Research unit on
forest genetic resources, were responsible for supervision.
References
Adams J.M. and Faure H. 1997. Preliminary vegetation maps of the world since the last glacial
maximum: an aid to archeological understanding. J. Archeol. Sci. 24: 623–647.
[65]
1126
Aide T.M. and Rivera E. 1998. Geographic patterns of genetic diversity in Poulsenia amata
(Moraceae): implications for the theory of Pleistocene refugia and the importance of ripian
forest. J. Biogeogr. 25: 695–705.
Barrett S.C.H. 1998. The reproductive biology and genetics of island plants. In: P.R. Grant (ed),
Evolution on islands. Oxford University Press, Oxford, UK, pp. 19–34.
Bekessy S.A., Allnutt T.R., Premoli A.C., Lara A., Ennos R.A., Burgman M.A., Cortes M. and
Newton A.C. 2002. Genetic variation in the vulnerable and endemic Monkey Puzzle tree,
detected using RAPDs. Heredity 88: 243–249.
Bousquet J., Simon L. and Lalonde M. 1990. DNA amplification from vegetative and sexual tissues
of trees using polymerase chain reaction. Can. J. Forest. Res. 20: 254–257.
Bouvet J.-M., Fontaine C., Sanou H. and Cardi C. 2004. An analysis of the pattern of genetic
variation in Vitellaria paradoxa using RAPD markers. Agroforest. Syst. 60: 61–69.
Briggs J.C. 2003. The biogeographic and tectonic history of India. J. Biogeogr. 30: 381–388.
Bryan G.J., McNicoll J., Ramsay G., Meyer R.C. and De Jong W.S. 1999. Polymorphic simple
sequence repeat markers in chloroplast genomes of solanaceous plants. Theoret. Appl. Genet. 99:
859–867.
Burney D.A. 1996. Climate change and fire ecology as factors in the quaternary biogeography of
Madagascar. In: Lourenço W.R. (ed.), Edition de l’ORSTOM Biogéographie de Madagascar.
Orstrom, Paris, France, pp. 49–58.
Cavers S., Navarro C. and Lowe A.J. 2003. A combination of molecular markers identifies evo-
lutionarily significant units in Cedrela odorata L. (Meliaceae) in Costa Rica. Conserv. Genet. 4:
571–580.
Collevatti R.G., Grattapaglia D. and Hay J.D. 2003. Evidence for multiple lineages of
Caryocar brasiliense populations in the Brazilian Cerrado based on the analysis of chloroplast
DNA sequence and microsatellite haplotype variation. Mol. Ecol. 12: 105–115.
Dutech C., Maggia L. and Joly H.I. 2000. Chloroplast diversity in Vouacapoua americana (Cae-
salpiniaceae), a neotropical forest tree. Mol. Ecol. 9: 1427–1432.
Echt C.S., DeVerno L.L., Anzidei M. and Vendramin G.G. 1998. Chloroplast microsatellites reveal
population genetic diversity in red pine, Pinus resinosa Ait. Mol. Ecol. 7: 307–316.
Ennos R.A. 1994. Estimating the relative rates of pollen and seed migration among plant popu-
lations. Heredity 72: 250–259.
Esselman E.J., Crawford D.J., Brauner S., Stuessy T.F., Anderson G.J. and Silva O.M. 2000.
RAPD marker diversity within and divergence among species of Dendroseris (Asteraceae:
Lactuceae). Am. J. Bot. 87(4): 591–596.
Excoffier L., Smouse P.E. and Quattro J.M. 1992. Analysis of molecular variance inferred from
metric distance among DNA haplotypes: application to human mitochondrial DNA restriction
data. Genetics 131: 479–491.
Gasse F. and Van Campo E. 2001. Late quaternary environmental changes from a pollen and
diatom record in the southern tropics (lake Tritivakely, Madagascar). Paleogeogr. Paleoclimatol.
Paleoecol. 167: 287–308.
Gillies A.C.M., Cornelius J.P., Newton A.C., Navarro C., Hernandez M. and Wilson J. 1997.
Genetic variation in Costa Rican populations of the tropical timber species Cedrela odorata L.,
using RAPDs. Mol. Ecol. 6: 1133–1145.
Grivet D. and Petit R. 2003. Chloroplast DNA phylogeography of the hornbeam in Europe: evidence
for the bottleneck at the outset of the postglacial colonisation. Conserv. Genet. 4: 47–56.
Goudet J. 2001. FSTAT, a program to estimate and test gene diversities and fixation indices
(version 2.9.3). Available from http://www.nil.ch/izea/softwares/fstat.html.
Holsinger K.E. 1996. The scope and the limits of conservation genetics. Evolution 50: 2558–2561.
King R.A. and Ferris C. 1998. Chloroplast DNA phylogeography of Alnus glutinosa (L.) Gaertn.
Mol. Ecol. 7: 1151–1161.
Kwon J.A. and Morden C.W. 2002. Population genetic structure of two rare tree species
(Colubrina oppositifolia and Alphitonia ponderosa, Rhamnaceae) from Hawaiian dry and mesic
forest using random amplified polymorphic DNA markers. Mol. Ecol. 11: 991–1001.
[66]
1127
Lynch M., Pfrender M., Spitze K., Lehman N., Hicks J., Allen D., Latta L., Ottene M., Bogue F.
and Colbourne J. 1999. The quantitative and molecular genetic architecture of a subdivided
species. Evolution 53: 100–110.
Marshall H.D., Newton C. and Ritland K. 2002. Chloroplast phylogeography and evolution of
highly polymorphic microsatellites in lodgepole pine (Pinus contorta). Theoret. Appl. Genet. 104:
367–378.
Martin J.P. and Hernandez Bermejo J.E. 2000. Genetic variation in the endemic and endangered
Rosmarinus tomentosus Huber-Morath & Maire (Labiatae) using RAPD markers. Heredity 85:
434–443.
Moritz C. 1994. Defining significant units for conservation. Trends Ecol. Evol. 9: 373–375.
Myers N., Mittermeier R.A., Mittermeier C.G., da Fonseca G.A.B. and Kent J. 2000. Biodiversity
hotspots for conservation priorities. Nature 403: 853–858.
Nybom H. and Bartish I.V. 2000. Effects of life history traits and sampling strategies on genetic
diversity estimates obtained with RAPD markers in plants. Perspect. Plant Ecol. Evol. Syst. 3:
93–114.
Nesbitt K.A., Potts B.M., Vaillancourt R.E., West A.K. and Reid J.B. 1995. Partitioning and
distribution of RAPD variation in a forest tree species, Eucalyptus globulus (Myrtaceae).
Heredity 74: 628–637.
Newton A.C., Allnutt T.R., Dvorak W.S., Del Castillo R.F. and Ennos R.A. 2002. Patterns of
genetic variation in Pinus chiapensis, a threatened Mexican pine, detected by RAPD and mito-
chondrial DNA RFLP markers. Heredity 89: 191–198.
Newton A.C., Allnutt T.R., Gillies A.C.M., Lowe A.J. and Ennos R.A. 1999. Molecular
phylogeography, intraspecific variation and the conservation of tree species. TREE 14(4):
140–145.
Palme A.E. and Verdramin G.G. 2002. Chloroplast DNA variation, postglacial recolonization and
hybridation in hazel, Corylus avellana. Mol. Ecol. 11: 1769–1779.
Perrier X., Flori A. and Bonnot F. 2003. Data analysis methods. In: Hamon P., Seguin M., Perrier
X. and Glaszmann J.C. (eds), Genetic diversity of cultivated tropical plants Enfield. Science
Publishers, Montpellier,pp. 43–76.
Rajagopal J., Bashyam L., Bhatia S., Khurana D.K., Srivastava P.S. and Lakshmikumaran M.
2000. Evaluation of genetic diversity in the Himalayan poplar using RAPD markers. Silvae
Genet. 49(2): 49–66.
Raspé O., Saumitou-Laprade P., Cuguen J. and Jacquemart A.-L. 2000. Chloroplast DNA hap-
lotype variation and population differentiation in Sorbus aucuparia L. (Rosaceae: Maloideae).
Mol. Ecol. 9: 1113–1122.
Savolainen O. and Kuittinen H. 2000. Small population processes. In: Youg A., Boshier D. and
Boyle T. (eds), Forest Conservation Genetics. Principles and Practises, CSIRO Publishing
Collingwood Victoria, Australia. pp. 94–100.
Schierenbeck K.A., Skupski M., Lieberman D. and Lieberman M. 1997. Population structure and
genetic diversity in four tropical tree species in Costa Rica. Mol. Ecol. 6: 137–144.
Schneider S., Roessli D. and Excoffier L. 2000. Arlequin version 2000: a software for population
genetics data analysis. Genetics and Biometrics Laboratory, Department of Anthropology,
University of Geneva, Geneva, Switzerland.
Straka H. 1996. Histoire de la végétation de Madagascar oriental dans les derniers 100 millénaires.
In: Lourenço W.R. (ed.), Editions de l’ORSTOM, Biogéographie de Madagascar. Orstrom,
Paris, France, pp. 37–47.
Vendramin G.G., Anzidei M., Madaghiele A., Sperisen C. and Bucci G. 1998. Distribution of
genetic diversity in Pinus pinaster Ait. as revealed by chloroplast microsatellites. Theoret. Appl.
Genet. 97: 456–463.
Viard F., El-Kassaby Y.A. and Ritland K. 2001. Diversity and genetic structure in populations of
Pseudotsuga menziesii (Pinaceae) at chloroplast microsatellite loci. Genome 44: 336–344.
[67]
1128
Weising K. and Gardner R.C. 1999. A set of conserved PVR primers for the analysis of simple
sequence repeat polymorphism in chloroplast genomes of dicotyledonous angiosperms. Genome
42: 9–19.
Willis K.J. and Whittaker R.J. 2000. The refugial debate. Science 287: 1406–1407.
Wu J., Krutovskii K.V. and Strauss S.H. 1999. Nuclear DNA diversity, population differentiation
and phylogenetic relatioships in the California closed-cone pines based on RAPD and allozyme
markers. Genome 42: 893–908.
Yeh F.C. and Boyle T.J.B. 1997. Population genetic analysis of co-dominant and dominant
markers and quantitative traits. Belg. J. Bot. 129: 157.
[68]
Biodiversity and Conservation (2006) 15:1129–1142 Springer 2006
DOI 10.1007/s10531-004-3103-8
-1
Key words: Castanea sativa, Coppice stand, Diversity index, Functional trait, Grove
Abstract. Over many centuries, chestnut fruits had an important role as food, while chestnut wood
was used for local purposes. Today sweet chestnut stands are very common around the western
Mediterranean Basin, and it is necessary to analyze the dynamic of plant species diversity in
different chestnut stand types (groves and coppices) to guide management strategies that will allow
the conservation of biodiversity. Our objective was to analyze consequences on plant species
diversity of various management strategies in chestnut stands of three Mediterranean areas, Sal-
amanca (Spain), the Cévennes (France), and Etna volcano (Italy). We found that plant species
diversity is different according to management types; it is higher in groves than in coppice stands.
We also demonstrated that Castanea sativa cultivated groves were characterized by small helio-
phillous therophytes. C. sativa abandoned groves, mixed C. sativa–Quercus pyrenaica coppice
stands, Q. pyrenaica coppice stands, and young C. sativa coppice stands were characterized by
hemicryptophytes with anemochorous dispersal mode and chamaephytes. Medium and old
C. sativa coppice stands (that differ by the shoot age) were characterized by phanerophytes with
zoochorous dispersal mode. Human perturbations maintain a quite high level of species diversity.
In contrast, the abandonment of chestnut stands leads to homogeneous vegetation with decreasing
diversity. One solution could be to maintain a landscape mosaic constituted of diverse chestnut
stands modified by human activities (groves, cultivated or abandoned, and coppice stands). This
could enhance regional plant diversity.
Introduction
Sweet chestnut (Castanea sativa Mill) stands are very common around the
western Mediterranean Basin. Over many centuries, chestnut fruits had an
important role as food for humans and as feed for domestic animals, while
chestnut wood was used for local purposes such as wine barrels, vineyard pegs,
tool handles and carpentry (Arnaud and Bouchet 1995). Today, chestnut
[69]
1130
stands cover large areas particularly in Portugal, Spain, France, Italy and
Greece. Thus, it is necessary to analyze the dynamic of plant species diversity in
different chestnut stand types (groves and coppices) to guide management
strategies that will allow the conservation of biodiversity and at the same time
to optimize productivity and profitability.
The characterization of community response to different management types
in terms of functional traits appears as a promising tool to achieve this goal
(McIntyre et al. 1995; Hadar et al. 1999; Lavorel et al. 1999; Gondard et al.
2003). Indeed, from an ecosystem perspective, species richness (number of
species), which is the conventional metric of biodiversity, is not as important as
functional trait richness. This approach analyzes the functioning of the eco-
system, and its response to abandonment, by focusing on vegetation descrip-
tion defined by functional traits not necessarily linked with taxonomic
attribution (Pillar 1999). Functional traits fall into three biological categories:
morphological traits describing aspect, life history traits indicating plant
behavior in the environment, and regeneration traits (Lavorel et al. 1997). The
use of functional traits for the comprehension and analysis of plant species
dynamics in relation with perturbation is clearly demonstrated by many au-
thors (Dı́az and Cabido 1997; Lavorel and Cramer 1999; McIntyre et al. 1999;
McIntyre and Lavorel 2001; Dı́az et al. 2002; Gondard and Deconchat 2003).
Consequently, our objective was to analyze consequences on plant species
diversity of various management strategies in chestnut stands of three Medi-
terranean areas, Salamanca (Spain), the Cévennes (France), and Etna volcano
(Italy). We hypothesized that, whatever area, species diversity between groves
and coppice stands is different essentially according to dendrometric charac-
teristics and management types. Indeed, groves have, in general, large trees
with regular pruning, understorey cleaning, etc., and coppices have many
shoots without clearing but logging. We assumed that species diversity is
highest in groves. We focused on understorey stratum which is sensitive to
changes of ecosystem conditions (Pregitzer and Barnes 1982; Strong et al. 1991;
Mitchell et al. 1997, 1998) and recognized like a very important component in
ecosystem functioning (Host and Pregitzer 1991; Arsenault and Bradfield 1995;
Brakenhielm and Lui 1998).
The experiment was carried out in three Mediterranean areas, in the Honfrı́a
forest, located in the southern of Salamanca province in Spain, in the Cévennes
in southern France, and on Etna volcano in Italy (Table 1). The Honfrı́a forest
is representative of traditional chestnut (Castanea sativa) management over
many centuries in Spain, but also a model of possible sustainable management
in the future. In this forest, chestnut is considered as a paraclimax species and
the deciduous oak (Quercus pyrenaica) as a climax species. Thus, we selected
five stands that are representative of this forest: a chestnut cultivated grove, a
[70]
1131
Data analyses
The criteria to compare stands were species richness (number of taxa per 100 m2)
and species diversity (Pielou 1975; Magurran 1988). Among the many diversity
indices available, we chose
P the Shannon index (H’), which was recommended by
Pielou (1975): H0 ¼ i¼1;n ðpi log2 ðpi ÞÞ where pi is the abundance ratio of
species (i) in the square, and n is the species number in the square.
Forest stands in the three geographical areas are submitted to different sil-
vicultural management and also to contrasting environment and climate con-
[71]
1132
Table 2. Main characteristics of chestnut stands selected in the Cévennes in France, on Etna
volcano in Italy and in Honfrı́a Forest in Spain. Confidence intervals p = 0.05. For each site, mean
values in the same column followed by different letters are significantly different. p < 0.05, Mann–
Whitney test.
Site Stand Tree age Tree Diameter Shoot density Basal area
(years) height (m) at breast (shoot ha1) (m 2ha1)
height (cm)
ditions, so differences are expected between them. However, due to the low
number of stands analyzed, we used non-parametric test that allows to work
with low size samples. We chose the Mann–Whitney non-parametric test that
allows to compare means pairwise (Falissard 1998).
In each Mediterranean area, we used Correspondence Analysis (CA) and
Canonical Correspondence Analysis (CCA, ter Braak 1987) to quantify the
effects of management types with species functional traits. We performed a
Correspondence Analysis (CA,Greenacre 1984) of plant species observed on
the entire point quadrat set (67 in Honfrı́a forest in Spain, 41 in the Cévennes in
France, and 40 on Etna volcano in Italy) and management types (coppice
stands and groves in Honfrı́a forest and in the Cévennes, and different coppice
stand types on the Etna volcano). We used CCA to determine the fraction of
variance of the species among management types explained by the species and
functional traits. For each Mediterranean area, we carried out the CCA by
confronting the CA table with another table composed by the same species
[72]
1133
60 a
55
50
Species richness
45
40
35
30
25
b
bc b
20 c
15
10
5
0
C. sativa C. sativa C. sativa Mixed C. sativa- Q. pyrenaica
cultivated grove abandoned coppice stand Q. pyrenaica stand
grove stand
Figure 1. Mean species richness in the understorey of the C. sativa and Q. pyranaica stands of the
Honfrı́a forest in the southern of Salamanca province in Spain. Error bars at ±95% confidence
limits. Two different letters between the coppice stands indicated significant statistical difference
(Mann–Whitney non-parametric test, p < 0.05).
Results
In the Honfrı́a forest and the Cévennes, species richness was highest in culti-
vated groves, (Figures 1, 2). On Etna volcano, species richness was highest in
45 a
40
Species richness
35
30 b
25 c cd
20 d
15
10
5
0
Cultivated Abandoned Young Medium Old
Grove Grove Coppice Coppice Coppice
Figure 2. Mean species richness along a successional gradient from cultivated chestnut grove to
old C. sativa coppice stand (Le Cros site in the Cévennes). Error bars at ±95% confidence limits.
Two different letters between the coppice stands indicated significant statistical difference (Mann–
Whitney non-parametric test, p < 0.05).
[73]
1134
45
40
Species richness
35 a
30
25
20 b
bc
15 cd
d
10
5
0
Trisciala Monte Fornazzo Balilla Piano Lepre
Crisimo
Figure 3. Mean species richness in the understorey in the five coppice C. sativa stands on the Etna
volcano in Italy. Error bars at ±95% confidence limits. Two different letters between the coppice
stands indicated significant statistical difference (Mann–Whitney non-parametric test, p < 0.05).
the Trisciala coppice stand (Figure 3), and not significantly different from the
abandoned grove in the Cévennes (p > 0.05). Species diversity was also
highest in cultivated groves and Trisciala coppice stand.
1
Axis 2
(inertia 15%)
Honfría forest - Group 2 0,8
C.sativa coppice stands
Geophyte Honfría forest - Group 3
Arpa, Casa, Celo, Prav,
Qupy, Tesc C. sativa abandoned groves & Mixed
0,6
C. sativa-Q. pyrenaica coppice stands
& Q. pyrenaica coppice stands
Anod, Caof, Capa, Daca, Gehi, Gepi,
0,4 Haha, Jamo, Loco, Meme, Feru, Himu
Figure 4. Ordination in the plane of the two axes of functional traits after a canonical corre-
spondence analysis from a matrix composed by the 67 plant species observed on the line point
quadrat of the 25 plots in the Honfrı́a forest in Spain and a matrix composed by the same plant
species and their functional traits. Groups were identified by an hierarchical ascending classifica-
tion. Plots, and some plant species associated to each group were indicated on the figure. Codes of
plant species are indicated in Appendix 1. The total variation explained by CCA is 39%.
[74]
1135
0,8
The Cévennes - Group 2 Axis 2
The Cévennes - Group 3
C. sativa abandoned (inertia 28%)
groves & young coppice
C. sativa medium and & old
Autochorous coppice stands
stands 0,6
Armo, Clvu, Bepe, Csa, Hede, Qupu, Rops
Dagl, Erci,
Hima, Himu, Barochorous
Pone, Prvu, 0,4 Phanerophyte
Trpr, Veof, > 50 cm
Visa
Chamaephyte 0,2
Zoochorous
Shade tolerant
0
-0,8 -0,6 -0,4 -0,2 0 0,2 0,4 0,6 0,8 Axis 2 1
30-50 cm Hydrochorous Geophyte
Heliophillous (inertia 37%)
Therophyte -0,2
1-30 cm
-0,6
Figure 5. Ordination in the plane of the two axes of functional traits after a canonical corre-
spondence analysis from a matrix composed by the 41 plant species observed on the line point
quadrat of the 25 plots in the Cévennes in France and a matrix composed by the same plant species
and their life traits. Groups were identified by an hierarchical ascending classification. Plots, and
some plant species associated to each group were indicated on the figure. Codes of plant species are
indicated in Appendix 1. The total variation explained by CCA is 46%.
studied. In the Honfrı́a forest, the C. sativa cultivated groves (group 1) were
characterized by small heliophillous therophytes (Figure 4). The C. sativa
coppice stands (group 2) were characterized by shade tolerant phanerophytes
with zoochorous dispersal mode and geophytes. The C. sativa abandoned
groves, mixed C. sativa–Q. pyrenaica coppice stands and Q. pyrenaica coppice
stands (group 3) were composed essentially by hemicryptophytes and
chamaephytes with anemochorous or barochorous dispersal mode.
In the Cévennes, the C. sativa cultivated groves (group 1) were characterized
by therophytes with anemochorous dispersal mode and geophytes (Figure 5).
The C. sativa abandoned groves and the young coppice stands (group 2) were
characterized by heliophillous hemicryptophytes and chamaephytes. The
C. sativa medium and old coppice stands (group 3) were composed more
particularly by phanerophytes with zoochorous dispersal mode.
In the case of C. sativa coppice stands on Etna volcano in Italy, Monte
Crisimo (group 1) were more particularly characterized by therophytes and
chamaephytes (Figure 6), Triciala (group 2) by hemicryptophytes with anem-
ochorous dispersal mode, Piano Lepre (group 3) by geophytes with baroch-
orous dispersal mode, and Balilla and Fornazzo (group 4) by shade tolerant
phanerophytes with zoochorous dispersal mode.
A main trend emerging from our species richness data was higher species
richness in the chestnut cultivated groves than in coppice stands; both in the
[75]
1136
1,5
Etna volcano - Group 2 Axis 2
C. sativa coppice stands (inertia 29%)
Etna volcano - Group 1
Trisciala C. sativa coppice stands
Acli, Brsy, Crle, Himu, Monte Crisimo 1
Heliophillous
Rane, Sivu, Trpu Arth, Avba, Homu, Lasp, 30-50 cm Barochorous
Sete, Stme, Vidi, Vite Geophyte
Anemochorous 0,5
Chamaephyte Etna volcano - Group 3
Hemicryptophyte C. sativa coppice stands
Therophyte Piano Lepre
1-30 cm Hydrochorous Door, Epmi, Lagr, Lave, Leco, Muco, Ptaq
0
-0,8 -0,6 -0,4 -0,2 Autochorous 0 0,2 0,4 0,6 0,8
Axis 1
(inertia 43%)
-0,5
-1,5
Figure 6. Ordination in the plane of the two axes of life traits after a canonical correspondence
analysis from a matrix composed by the 40 plant species observed on the line point quadrat of the
25 plots on Etna volcano in Italy and a matrix composed by the same plant species and their life
traits. Groups were identified by an hierarchical ascending classification. Plots, and some plant
species associated to each group were indicated on the figure. Codes of plant species are indicated in
Appendix 1. The total variation explained by CCA is 49%.
[76]
1137
Acknowledgements
Appendix 1. Functional traits (life form, dispersal mode, plant height and light tolerance) of plant
species observed along the point quadrat line and used in the Canonical Correspondence Analysis
according to available data: Molinier and Müller 1938; Pignatti 1982; van der Pijl 1982; Bonnier
1990; De Bolos et al. 1993. Th – therophyte, G – geophyte, H – hemicryptophyte, Ch – chamae-
phyte, Ph – phanerophyte.
[77]
1138
Appendix 1. Continued.
[78]
1139
Appendix 1. Continued.
[79]
1140
Appendix 1. Continued.
References
Arnaud M.T. and Bouchet M.A. 1995. L’aire écologique du châtaignier (Castanea sativa Mill.) en
Cévennes. Ecologie 26(1): 33–40.
Arsenault A. and Bradfield G.E. 1995. Structural-compositional variation in three age-classes of
temperate rainforests in southern British Columbia. Can. J. Bot. 73: 54–64.
Bonnier G. 1990. La Grande Flore. France, Suisse, Belgique et pays voisins, Belin, Paris.
Brakenhielm S. and Lui Q. 1998. Long-term effects of clear-felling on vegetation dynamics and
species diversity in boreal pine forest. Biol. Conserv. 7: 207–220.
DeBolòs O., Vigo J., Masalles R.M. and Ninot J.M. 1993. Flora Manual dels Paı̈sos Catalans.
Portic S.A. Edition, Barcelone.
Debussche M., Escarré J., Lepart J., Houssard C. and Lavorel S. 1996. Changes in Mediterranean
plant succession: old-fields revisited. J. Veg. Sci. 7: 519–526.
Dı́az S. and Cabido M. 1997. Plant functional types and ecosystem function in relation to global
change. J. Veg. Sci. 8: 463–474.
Dı́az S., McIntyre S., Lavorel S. and Pausas J. 2002. Does hairiness matter in Harare? – Global
comparisons of plant trait responses to disturbance. New Phytol. 154: 7–9.
Escarré J., Houssard C., Debussche M. and Lepart J. 1983. Evolution de la végétation et du sol
après abandon cultural en région méditerranéenne: étude de successions dans les garrigues du
Montpellierais (France). Acta Oecol., Oecol. Plant. 4: 221–239.
Falissard B. 1998. Comprendre et Utiliser les Statistiques dans les Sciences de la vie. Collection
Evaluation et Statistique, Masson, Paris.
[80]
1141
Gilliam F.S., Turrill N.L. and Bethadams M. 1995. Herbaceous-layer and overstory species in
clear-cut and mature central Appalachian hardwood forests. Ecol. Appl. 5: 947–955.
Gondard H. and Deconchat M. 2003. Effects of soil surface disturbances after logging on plant
species diversity. Ann. Forest Sci. 60: 725–732.
Gondard H., Romane F., Grandjanny M., Junqing L. and Aronson J. 2001. Plant species diversity
changes in abandoned chestnut (Castanea sativa) groves in southern France. Biodivers. Conserv.
10: 189–207.
Gondard H., Jauffret S., Aronson J. and Lavorel S. 2003. Plant functional types: a promising tool
for management and restoration of degraded lands. Appl. Veg. Sci. 6: 223–224.
Gounot M. 1969. Méthodes D’étude Quantitative de la Vègétation. Masson, Paris.
Greenacre M.J. 1984. Theory and Applications of Correspondence Analysis. Academic Press,
London.
Grime J.P. and Jarvis B.C. 1975. Shade avoidance and shade tolerance in flowering plants. II.
Effects of light on the germination of species of contrasted ecology. In: Evans G.C., Bainbridge
R. and Rackham O. (eds), Light as an Ecological Factor II. Blackwell Scientific Publications.
Oxford, pp. 525–532.
Hadar L., Noy-Meir I. and Perevolotsky A. 1999. The effect of shrub clearing and grazing on the
composition of a Mediterranean plant community: functional groups versus species. J. Veg. Sci.
10: 673–682.
Houssard C., Escarré J. and Romane F. 1980. Development of species diversity in some Medi-
terranean plant communities. Vegetatio 43: 59–72.
Host G.E. and Pregitzer K.S. 1991. Ecological species groups for upland forest ecosystems of
northwestern Lower Michigan. Forest Ecol. Manag. 43: 87–102.
Kitazawa T. and Ohsawa M. 2002. Patterns of species diversity in rural herbaceous communities
under different management regimes, Chiba, central Japan. Biodivers. Conserv. 104: 239–249.
Lavorel S. 1999. Ecological diversity and resilience of Mediterranean vegetation to disturbance.
Diversity Distribut. 5: 3–13.
Lavorel S. and Cramer W. 1999. Functional response of vegetation to land use and disturbance.
J. Veg. Sci. 10: 604–732.
Lavorel S., McIntyre S. and Grigulis K. 1999. Plant response to disturbance in a Mediterranean
grassland: How many functional groups? J. Veg. Sci. 10: 661–672.
Lavorel S., McIntyre S., Landsberg J. and Forbes T.D.A. 1997. Plant functional classification:
from general groups to specific groups based on response to disturbance. Tree 12(12): 474–478.
Magurran A.E. 1988. Ecological Diversity and its Measurements. Croom Helm, London.
McIntyre S. and Lavorel S. 2001. Livestock grazing in sub-tropical pastures: steps in the analysis of
attribute response and plant functional types. J. Ecol. 89: 209–226.
McIntyre S., Lavorel S., Landsberg J. and Forbes T.D.A. 1999. Disturbance response in vegeta-
tion-towards a global perspective on functional traits. J. Veg. Sci. 10: 621–630.
McIntyre S., Lavorel S. and Tremont R.M. 1995. Plant life history attributes: their relationships to
disturbance response in herbaceous vegetation. J. Ecol. 83: 31–44.
Mitchell R.J., Marrs R.H. and Auld M.H.D. 1998. A comparison study of the seedbanks of
heathland and succession habitats in Dorset, Southern England. J. Ecol. 86: 588–596.
Mitchell R.J., Marrs R.H., Le Duc M.G. and Auld M.H.D. 1997. A study of succession on lowland
heaths in Dorset, Southern England: changes in vegetation and soil properties. J. Appl. Ecol. 6:
1426–1444.
Molinier R. and Müller P. 1938. La Dissémination des Espèces Végétales. Lesot A. (ed.), Paris.
Pielou R.H. 1975. Ecological Diversity. A Wiley-Interscience Publication, New York.
Pignatti S. 1982. Flora d’Italia. Edagricole (ed.), Bologne, 3 volumes, p. 2302.
Pillar V.D. 1999. On the identification of optimal plant functional types. J. Veg. Sci. 10(5): 631–640.
Pregitzer K.S. and Barnes B.V. 1982. The use of ground flora to indicate edaphic factors in upland
ecosystems of the McCormick experimental forest, Upper Michigan. Can. J. Forest Res. 12: 661–
672.
[81]
1142
Romane F., Bacilieri R., Bran D. and Bouchet M.A. 1992. Natural degenerate Mediterranean
forests: Which future? The examples of the holm oak (Quercus ilex.) and chestnut (Castanea
sativa Mill.) coppice stands. In: Teller A., Mathy P. and Jeffers J.N.R. (eds), Responses of Forest
Ecosystems to Environmental Changes. Elsevier Applied Science, London and New York,
pp. 374–380.
Roux M. 1985. Algorithmes de Classification. Masson, Paris.
Rubio A. and Escudero A. 2003. Clear-cut effects on chestnut forest soils under stressful condi-
tions: lengthening of time-rotation. Forest Ecol. Manag., 183 195–204.
Rubio A., Gavilán R. and Escudero A. 1999. Are soil characteristics and understorey composition
controlled by forest management? Forest Ecol. Manag. 113: 191–200.
Strong W.L., Bluth D.J., LaRoi G.H. and Corns I.G.W. 1991. Forest understorey plants as pre-
dictors of lodgepole pine and white spruce site quality in west-central Alberta. Can. J. Forest
Res. 21: 1675–1683.
Tatoni T. and Roche P. 1994. Comparison of old-field and forest revegetation dynamics in
Provence. J. Veg. Sci. 5: 295–302.
terBraak C.J.F. 1987. The analysis of vegetation-environment relationships by canonical corre-
spondence analysis. Vegetatio 69: 69–77.
Tutin T.G., Heywood V.H., Burges N.A., Moore D.M., Valentine D.H., Walters S.M. and Webb
D.A. 1964–1980. Flora Europea. Cambrige University Press.
van der Pijl L. 1982. Principles of Dispersal in Higher Plants. Springer, Berlin, Heidelberg and New
York.
Yorks T.E. and Dabydeen S. 1999. Seasonal and successional understory vascular plant diversity in
second growth hardwood clearcuts of western Maryland, USA. Forest Ecol. Manag. 119: 217–
230.
[82]
Biodiversity and Conservation (2006) 15:1143–1157 Springer 2006
DOI 10.1007/s10531-004-3105-6
-1
Key words: Ecology, Mingling index, Namibia, Spatial diversity, Spatial structure, Uniform angle
index, Woodland savanna
Abstract. The dry woodland savannas of Namibia are of significant socio-economic importance.
The paper tests the suitability of a number of diversity indicators developed for species poor
systems in Europe in the woodland context. The indicators that were tested included the species
specific mingling index, MSp, the measure of surround and the uniform angle index. The simple
application of the methods permit relatively unschooled crews to conduct an enumeration in the
field.The results show that the indicators do not only display current diversity status, but also
reflect the ecological context of the individual species.
[83]
1144
[84]
1145
It is also uncertain whether or not whole stands of any of the above species
will die off and be replaced by others at a different location, or whether the
existing regeneration is sufficient to replace those trees that have died.
[85]
1146
Figure 1. The mingling of black, grey and white ‘trees’ within two square stands (after Gadow,
1999).
[86]
1147
The original measure of mingling and its derivatives are based on the pro-
portion of trees with dissimilar characteristics to those of a selected sample
tree. The species mingling index Mi for a given sample tree, i, using n neigh-
bours is, for example, obtained through:
1X n
ð1Þ Mi ¼ mij ;
n j¼1
[87]
1148
where
1; if the tree is of another species;
mij ¼
0; if the tree is of the same species:
When four neighbours are used to determine Mi the index may obtain one of
five possible values:
0/4 none of the neighbours are of a different species,
1/4 one of the neighbours is of a different species,
2/4 two of the neighbours are of a different species,
3/4 three of the neighbours are of a different species, and
4/4 all of the neighbours are of a different species.
The arithmetic mean (MSp) of the Mi values that were obtained for a par-
ticular species sp provides a measure of the degree of interspersion of the
species in the area. MSp provides a value between 0 and 1.
Values close to 0 indicate that trees of the reference species sp occur in
groups therefore implying a low degree of mingling and high degree of
aggregation. High values of MSp, closer to 1, on the other hand, imply a high
degree of mingling, i.e. trees of the reference species do not occur together.
As is the case when examining the distribution of data around a mean value,
additional information may be extracted from the distribution of Mi values of
individual species.
When the proportion that a species contributes to a stand is known, as
assumed in the studies reported on by Lewandowski and Pommerening (1997)
and Hui et al. (1998) a theoretical distribution of Mi values may be calculated
based on the hypergeometric probability distribution. The distribution reflects
the number of expected Mi values that would be obtained if all trees were
interspersed randomly.
The hypergeometric distribution is used to determine the probability, P, that
a number of trees of a particular species may occur in a given sample of n trees
taken from a population of N trees containing k trees of the species of interest.
The probability that x trees in the sample will be of the species of interest is
then determined after Newmark (1997) as:
k Nk
x nx
P¼ for x ¼ 0; 1; 2; . . . ; n;
N
n
which expands to:
ðNkÞ
k
xðkxÞ
ðnxÞðNkðnxÞÞ
P¼ N for x ¼ 0; 1; 2; . . . ; n:
nðNnÞ
[88]
1149
where
1; if the tree; j; is thicker than the sample tree i;
tij ¼
0; otherwise:
The species specific mean interspersion of tree diameter, TSp, is then the
arithmetic mean of the values of Ti for that species.
Similarly, the interspersion of tree height, Hi, is obtained through:
1X n
ð3Þ Hi ¼ hij ;
n j¼1
where
1; if the tree; j; is higher than the sample tree i;
hij ¼
0; otherwise:
The species specific interspersion of tree height, HSp, is then again deter-
mined as the mean of the values of Hi for the species.
An equivalent measure was used to quantify the interspersion of dead trees
(DSp) by counting the number of dead neighbours for each sample tree.
[89]
1150
The UAI was initially described by Gadow et al. (1998) and later by Stau-
pendahl (2001) to provide a measure of the overall contagion of trees within a
forest stand.
The index is obtained by identifying the n nearest neighbours of a sample
tree. Starting with the closest neighbour and moving in a clockwise direction
around the sample tree the angle, aj, between two adjacent neighbours is
determined in relation to the sample tree. The number of angles smaller than,
or equal to, a given critical angle, a0, are then counted, i.e.
1X n
ð4Þ Wi ¼ wij ;
n j¼1
where
1; if aj a0 ;
wij ¼
0; otherwise:
The critical angle (in degrees) is determined as:
360
ð5Þ a0 ¼ :
Number of neighbours
Four neighbours would therefore be evaluated in terms of a 90 critical
angle1. Since all of the indexes used to measure the interspersion of tree
characteristics were based on four trees, the same neighbours could be used for
the UAI.
A practical advantage of choosing a0 = 90 is that two adjoining sides of a
record book or clipboard may be used to determine whether or not an angle is
greater than or less than the critical angle.
Effectively, the index describes the spatial distribution around a particular
reference tree. If the species of the reference tree is noted we may obtain the
mean value for either for the whole population or for a particular species of
interest.
The mean value of the index is strongly correlated with the nearest neigh-
bour index of dispersion of Clark and Evans (1954) that has long been used in
ecological studies. Together with the number of trees in a stand, the UAI may
be used to estimate the distribution of distances between a tree and its neigh-
bours (Gadow et al. 2003).
This information is generally not available and comparison of observed
index values are compared to the simulation results of Gadow et al. (1998) are
used.
1
More recent studies have shown that this statement needs to be modified; a more suitable critical
angle is 72 (see Gadow et al. 2003).
[90]
1151
Sampling
The extent of the stand was recorded in the field using a Garmin Venture GPS.
The track-log was stored for subsequent mapping. A regular sample grid of one
geographic second was then superimposed on the stand amounting to a sample
point approximately every 30 m at that latitude.
Sampling points were located using a standard GPS receiver. The accuracy
of autonomous GPS readings was considered adequate for the purpose of the
study. While a dense canopy reduces the reliability of a GPS reading within a
stand (Dominy and Duncan 2001), many of the trees in the area had already
shed their leaves and canopy interference was considered negligible after initial
comparison of signal strengths in wooded and in open areas.
Since the enumeration coincided with the war in Iraq it is uncertain whether
GPS readings were affected by selective availability on some days. It was felt,
however, that this was acceptable.
At each sample point the closest tree with a dbh of 5 cm or more was
identified to serve as reference tree. Although trees had, in a few cases, snapped
off below breast height, such trees were nevertheless sampled, since they play a
role in the interspersion of plants.
For each sample tree the four nearest neighbouring trees with a diameter of
greater than 5 cm were determined and compared with the reference tree in
terms of species, mortality, height and diameter, and the UAI was established.
Time was kept short by assigning two persons to each sampling team. While
the enumerator collected the measures, a navigator moved to find the next
sample point.
A total of 1121 sample points were assessed. The data was entered into a
spreadsheet and the indexes were calculated for each species using cross tables.
The species specific indexes are summarized in Table 1. The table also shows a
surrogate species of ‘Dead’ created to record trees that were still standing but
had been burnt beyond a stage where they might be identified. Also, species of
the genus Combretum and Comiphora were lumped, as individual species could
not readily be identified. The row marked ‘overall’ provides the each index as
calculated over the entire data set.
The overall shows a contagion (Wi) greater than 0.6, here indicating a ten-
dency towards non-random (clumped) dispersion of trees (after Gadow et al.
1998). The dispersion around trees of the individual species does not seem to
diverge very much from the mean value of 0.665, if Philenoptera nelsii and
Securidaka longipedunculata are discounted because of their very low overall
occurrence. This is in line with general observations in the field.
The table shows that most of the species have a tendency to aggregate.
[91]
1152
To compare the proportion of a given species within the stand and the value
1 MSp consider Table 2. The table omits those species with very few
observations (less than 5% of the total). The final column in the table reflects
the parameter M proposed by Graz (2004) to determine the degree of inter-
spersion. The value of M is larger than 0 and less than or equal to 1. Values
close to 0 indicates a very low degree of mingling, and 1 indicates a more
random distribution of the species in the stand.
Table 2 shows that B. plurijuga has the highest degree of aggregation fol-
lowed by Terminalia sericea.
More T. sericea seedlings survive in open areas, i.e. away from conspecific
trees (Smith and Grant 1986) where it has the ability to form thickets
(Shakelton 2001). This was evident in the field. T. sericea would colonize gaps
in the canopy, thus causing the aggregation.
The dispersion of B. plurijuga is also shown in Figure 2. The figure shows
that the species occurs in a very limited area. The accompanying graph shows
the relative distribution of Mi values (bar) and the theoretical hypergeometric
distribution. The graph shows a clear difference between the two, due to the
clumping of the species, reflected by the low value of M (Table 2).
The cause of the aggregation of B. plurijuga is uncertain, since the trees had
few larger neighbours as evidenced by the low value of TSp in Table 1. It is
possible that the patch of B. plurijuga is a remnant of a larger stand that has
been subject to high degrees of mortality. This possibility stems from reports by
Von Breitenbach (1968) who suggested that the almost pure stands in the
Caprivi region developed towards mixed stands as a result of fire.
The possibility is corroborated by the high degree of mortality (DSp) asso-
ciated with the species (see Figure 3). The dead trees within the B. plurijuga
Table 1. Mean of the various indicators for each of the identified species.
[92]
1153
Table 2. Comparing the proportion P(Sp) that a species contributes to the population with
(1 MSp).
P ðpÞ
Species of sample tree N P(Sp) MSp 1 MSp M ¼ 1M Sp
patch, shown in figure 3, are generally large trees. This is not evident from the
indexes but supports the suggestion by Von Breitenbach cited above.
Actual tree mortality may be caused directly by repeated burning of the
stem, as well as changes in the osmotic potential of the top-soil that is caused
by the accumulation of ash in the upper soil layers (Mitlöhner pers. comm.)
In contrast to B. plurijuga, P. angolensis is interspersed almost randomly
according to Table 2 and in Figure 2. As is evident in the figure, the observed
distribution of Mi values (bars) follow the theoretical distribution much more
closely than those of B. plurijuga. It must be noted, that P. angolensis occurs
comparatively seldom within the B. plurijuga patch. This exclusion from the
patch is more pronounced for B. africana. The reason or cause for this is not
readily apparent. Outside this patch B. africana is more aggregated resulting in
the lower value of M.
The random distribution of P. angolensis is probably a reflection of the
regeneration requirements of the species. Vermeulen (1990) reports that
P. angolensis is especially sensitive to competition in the seedling and estab-
lishment phases. The species therefore often regenerates in areas that have been
cleared by human or other action. Other species would then establish them-
selves later.
The interspersion of trees of different size is reflected in the columns TSp
(diameter specific) and HSp (height specific) in Table 1. Preliminary simulation
results have shown that a random interspersion of tree sizes would result in an
overall average of TSp = 0.5 and HSp = 0.5. The table shows, therefore, that
size classes are not interspersed randomly.
P. angolensis, S. rautanenii and B. plurijuga need to be highlighted. The low
values of TSp and HSp for these species imply that few neighbouring trees are
larger than the reference tree. This is supported by general observations in the
field. The species therefore dominate in the area in which they occur. It also
reflects the regeneration requirements of P. angolensis noted previously, but
highlights the importance of further research into the demography of the other
two species.
[93]
1154
Figure 2. The dispersion of Baikiaea plurijuga, Pterocarpus angolensis and Burkea africana, within
the study area. High values of Mi are shown in large circles and vice versa. The graphs depict the
observed relative distribution of Mi values (bars), and the theoretical hypergeometric distribution
(lines) of the values that would indicate a completely random interspersion of the species.
Table 1 also shows a similarity between the values of TSp and HSp of the
individual species. Unpublished data shows a high degree of correlation be-
tween the dbh and height of B. africana (r2 = 0.8352), as well as for P. an-
golensis (r2 = 0.7317) for nearby stands.
[94]
1155
Figure 3. The aggregation of dead trees within the stand. The degree of interspersion is reflected
by the size of the points, with a high degree of aggregation shown by larger points.
Differences between the two indexes are due to the number of species
found in the stand, and the differences in their respective diameter height
relationships. A larger difference occurs for the Comiphora species, however,
reflecting the squat form of the trees; a relatively thick-trunked but short
tree.
Conclusions
[95]
1156
Acknowledgements
I would like to thank my sister Ms. H. Riehmer and Ms. R. Haipinge as well as
the late Mr. H. Roth for their assistance with data collection in the field. My
sincere thanks also to the Directorate of Forestry, Namibia, for allowing me to
use the Kanovlei Forest Station as a base, and the Polytechnic of Namibia who
funded the field work. I would particularly like to thank Prof. K. von Gadow,
Institute of Forest Management, Univeristy of Goettingen for comments.
References
Albert M. and Gadow K.v. 1998. Assessing biodiversity with new neighborhood-based parameters.
In: Proceedings of the International Conference on Data Management and Modelling Using
Remote Sensing and GIS for Tropical Forest Land Inventory. Jakarta, Indonesia, Oct. 26–29,
1998.
Amakali M. 1992. Updated isohyetal rainfall map for Namibia. Unpublished Report. Department
of Water Affairs, Namibia.
Büschel D. 1999. A study of resource utilization: a case from Namibia, Mpungu constituency,
Kavango district, northern Namibia. CRIAA SA-DC Report, Windhoek, Namibia.
Chakanga M. 1995. Forest Cover Reconnaissance Mapping Project. Directorate of Forestry,
Republic of Namibia, Windhoek.
Childes S.L. 1984. The population dynamics of some woody species in the Kalahari sand vegetation
of the Hwange National Park. M.Sc. thesis, University of the Witwatersrand, South Africa.
Clark P.J. and Evans F.C. 1954. Distance to Nearest Neighbor as a Measure of Spatial Rela-
tionships in Populations. Ecology 35(4): 445–453.
Coetzee M.E. 2001. NAMSOTER, A SOTER database for Namibia. Agroecological Zoning
Programme, Ministry of Agriculture, Water and Rural Development, Namibia.
Crerar S.E. and Church J.T. 1988. Evaporation map for south West Africa/Namibia. Hydrological
Report No. 11/1/8/1/H1, Department of Water Affairs, Namibia.
Dominy N.J. and Duncan B. 2001. GPS and GIS methods in an African rain forest: applications to
tropical ecology and conservation. Conserv. Ecol. 5(2): 6. URL: http://www.consecol.org/vol5/
iss2/art6.
Department of Water Affairs. 1971. Consolidated report on reconnaissance surveys of soils of
northern and central south West Africa in terms of their potential for irrigation. Report TS/30/
71, Department of Water Affairs, Windhoek, Namibia.
Department of Water Affairs. 1991. Groundwater investigation in Kavango and Bushmanland,
Namibia. Unpublished Report. Department of Water Affairs, Windhoek, Namibia.
De Pauw E. and Coetzee M.E. 1999. Production of an agro-ecological zones map of Namibia (first
approximation), Part 2: Results. Agricola 10: 8–68.
Gadow K.v. 1999. Waldstruktur und Diversität. Alg. Forst-u. J.-Ztg. 170: 117–122.
Gadow K.v., Hui G.Y. and Albert M. 1998. Das Winkelmass – ein Strukturparameter zur Bes-
chreibung der Individualverteilung in Waldbeständen. Centralblatt das gesamte Forstwesen.
115(1): 1–10.
Gadow K.v., Hui G.Y., Chen B.W. and Albert M. 2003. Beziehungen zwischen Winkelmass und
Baumabständen. Forstw. Cbl. 122: 127–137.
Giess W. 1998. A preliminary vegetation map of Namibia (3rd revised edition). Dinteria 4: 1–112.
Graz F.P. 1996. Management of a Pterocarpus angolensis population under the influence of fire and
land use. M.Sc. thesis, University of Stellenbosch, South Africa.
[96]
1157
Graz F.P. 1999. A preliminary terrain feature classification of the Okavango region, Namibia. S.
Afr. J. Survey. Geo-Inf. 1: 123–129.
Graz F.P. 2003. Fire damage to Schinziophyton rautanenii (Schinz) trees in North-Eastern Namibia.
Dinteria 28: 39–43.
Graz F.P. 2004. The behavior of the species mingling index MSp in relation to species dominance
and dispersal. Eur. J. Forest Res. 1: 87–92.
Hui G.Y., Albert M. and Gadow K.v. 1998. Das Umgebungsmaß als Parameter zur Nachbildung
von Bestandesstrukturen. Forstw. Cbl. 117: 258–266.
Lee R.B. 1973. Mogongo: the ethnography of a major wild food resource. Ecol. Food Nutr. 2: 307–
321.
Leger S. 1997. The Hidden Gift of Nature. DED, German Development Service, Berlin, Germany.
Lewandowski A. and Pommerening A. 1997. Zur Beschreibung der Waldstruktur – Erwartete und
beobaachtete Arten-Durchmischung. Forstw. Cbl. 116: 129–139.
Mitlöhner R. 1997. Pflanzeninterne Potentiale als Indikatoren für den tropischen Standort. Shaker
Verlag.
Newmark J. 1997. Statistics and Probability in Modern Life, 6th ed. Sauders College Publishing,
Fort Worth.
NFSP. 1996. Namibia Forestry Strategic Plan. Directorate of Forestry, Windhoek, Namibia.
Ollikainen T. 1992. Study on wood consumption in Namibia. Internal Report, Directorate of
Forestry, Windhoek.
Pielou E.C. 1977. Mathematical Ecology. John Wiley & Sons, New York, 384 pp.
Rutherford M.C. 1981. Survival, regeneration and leaf biomass changes in woody plants following
spring burns in Burkea africana – Ochna pulchra savanna. Bothalia 13: 531–552.
Shackleton C.M. 2001. Managing regrowth of an indigenous savanna tree species (Terminalia
sericea) for fuelwood: the influence of stump dimensions and post-harvest coppice pruning.
Biomass Bioenergy 20: 261–270.
Smith T.M. and Grant K. 1986. The role of competition in a Burkea africana – Terminalia sericea
savanna. Biotropica 18: 219–223.
Staupendahl K. 2001. Das flächenbezogene Winkelmass Wf – Ein Index zur quantitativen Bes-
chreibung der horizontalen Baumverteilung. In: Akca A. et al. (HRSG). Waldinventur,
Waldwachstum und Forstplanung – Moderne Technologien, Methoden und Verfahrensweisen.
Festschrift K. von Gadow. Zohab-Verlag, Göttingen. pp. 101–115.
Vermeulen W.J. 1990. A monograph on Pterocarpus angolensis. SARCCUS Standing Committee
for Forestry, Pretoria, South Africa.
Von Breitenbach F. 1968. Long-term plan of forestry development in eastern Caprivi Zipfel.
Unpublished Report. George, South Africa.
Wilson B.G. and Witkowski E.T.F. 2003. Seed banks bark thickness and damage in age and size
structure (1978–1999) of the African savanna tree Burkea africana. Plant Ecol. 167: 151–162.
Yeaton R.I. 1988. Porcupines, fires and the dynamics of the tree layer of the Burkea africana
savanna. J. Ecol. 76: 1017–1029.
[97]
Biodiversity and Conservation (2006) 15:1159–1177 Springer 2006
DOI 10.1007/s10531-004-3509-3
-1
Abstract. To counteract an increasing biodiversity decline, parks and protected areas have been
established worldwide. However, many parks lack adequate management to address environmental
degradation. To improve management strategies simple tools are needed for an assessment of
human impact and management effectiveness of protected areas. This study quantifies the current
threats in the heavily fragmented and degraded tropical rainforest of Kakamega, western Kenya.
We recorded seven disturbance parameters at 22 sites in differently managed and protected areas of
Kakamega Forest. Our data indicate a high level of human impact throughout the forest with
illegal logging being most widespread. Furthermore, logging levels appear to reflect management
history and effectiveness. From 1933 to 1986, Kakamega Forest was under management by the
Forest Department and the number of trees logged more than 20 years ago was equally high at all
sites. Since 1986, management of Kakamega Forest has been under two different organizations, i.e.
Forest Department and Kenya Wildlife Service. The number of trees logged illegally in the last
20 years was significantly lower at sites managed by the Kenya Wildlife Service. Finally, logging
was lower within highly protected National and Nature Reserves as compared to high logging
within the less protected Forest Reserves. Reflecting management effectiveness as well as protection
status in Kakamega Forest, logging might therefore provide a valuable quantitative indicator for
human disturbance and thus an important tool for conservation managers. Logging might be a
valuable indicator for other protected areas, too, however, other human impact such as e.g. hunting
might also prove to be a potential indicator.
Introduction
Recent decades have seen a serious biodiversity decline due to habitat loss and
alteration especially of tropical forests leading to a profound species-extinction
crisis (Heywood 1995; Pimm et al. 1995; Whitmore 1997). Thus, much of
tropical biodiversity is unlikely to survive without effective protection (Pimm
et al. 1995; Myers et al. 2000). To counteract the anthropogenic impact and
conserve biodiversity and ecosystem processes parks and protected areas have
been established worldwide. Some studies demonstrate that parks can indeed
provide basic safeguard against land-clearing in the context of high land-use
pressure (Brunner et al. 2001). However, more often parks appear to lack
[99]
1160
[100]
1161
to protected area managers. The primary tool for RAPPAM is the rapid
assessment questionnaire which covers management planning, input and pro-
cesses, and the identification of future threats and past pressures (Ervin 2003c).
However, quantitative and objective approaches are still urgently required for
the assessment of threat status and management effectiveness of protected
areas to provide reliable, scientifically sound data.
In this paper we present results from a survey quantifying human impact and
evaluating management effectiveness in Kakamega Forest, western Kenya.
Kakamega Forest is one of the last remaining indigenous forests of Kenya
situated in an agricultural area with a high human density of more than
175 individuals per km2 (Tsingalia 1988). Like many other countries Kenya
harbours an on-going conflict between forest conservation and land use needs
of its increasing population (Tsingalia 1988; Wass 1995). This has put a long-
term pressure on Kakamega Forest leading to its severe reduction and
fragmentation in the last century. Additionally, it has suffered increasing
degradation through both, extensive commercial and local exploitation of
timber (Tsingalia 1988; Fashing et al. 2004; Mitchell, 2004). Large-scale
commercial logging was reduced in the last decades, mostly through official
presidential decree banning all indigenous tree species exploitation in the forest
in the early 1980s (Tsingalia 1988; Mitchell, 2004), the transfer of the northern
part of the forest under the rigorous authority of the Kenya Wildlife Service
(KWS), the establishment of forest stations and ranger patrols, and through
tourism and long-term research (e.g. Zimmerman 1972; Cords 1987; Mutangah
1996; Fashing et al. 2004). However, illegal activities including logging, fuel-
wood collection and extraction of bark for medicinal purposes occur to this
day and appear to be heterogeneously throughout the forest with some sites
providing more protection than others (Kiama and Kiyiapi 2001; Fashing et al.
2004; Fashing, in press). Our study presents a quantitative assessment of the
current threats in the main forest block of Kakamega Forest and its fragments
comprising areas of different management regimes and different protection
priorities. In order to evaluate effectiveness of conservation measures we asked
how differently managed and protected areas differ in their level of human
impact. With this assessment we aim to provide a quantitative, simple site-level
monitoring tool and a first guidance to management planners and decision
makers on problems related to human impact and management in Kakamega
Forest.
Study site
[101]
1162
Figure 1. Satellite image (channel 5 of Landsat 7 ETM+, 05 Feb 2001) of Kakamega main forest
and its five fragments in western Kenya with official forest boundaries as gazetted in 1933 (dashed line)
and official boundaries of National and Nature Reserves (white line). Coordinates in UTM 36 N.
[102]
1163
Management history
Kakamega Forest is a highly fragmented and disturbed forest and has been
continually exploited for many years due to the high surrounding population
pressure (Kokwaro 1988; Wass 1995). The main forest block gazetted in 1933
by the FD to control human activities covered 23,777 ha (Kokwaro 1988, for
original forest boundaries see Figure 1). The FD aimed mostly at provision of
timber for local communities and commercial demand. Clear-felling of indig-
enous forest to make way for fast-growing exotic tree and softwood plantations
was extensive under colonial forest service. Especially the southern parts of
Kakamega were exploited until the late 1980s (Bennun and Njoroge 1999;
Mitchell, 2004). Clearance for settlement and tea plantations slowed over the
1980s as forest protection was better enforced, but more areas were cleared
south of the Yala river (Brooks et al. 1999). Furthermore, selective logging was
intense in the past and the general trend of timber extraction showed a con-
tinued rise from 1933 to 1981 (Tsingalia 1988; Mutangah 1996). Consequently,
the main agents of forest degradation have been mostly logging and extraction
of commercially valuable timber, followed by charcoal burning, cattle grazing,
shamba system farming, hunting for bush-meat, tree debarking and removal of
dead trees for firewood (Oyugi 1996; Mitchell, 2004). In the early 1980s a
presidential decree banned all indigenous tree species exploitation, leading to a
halt of commercial logging, however, tree poaching and other illegal activities
still exist.
[103]
1164
Site No. Site name Main forest/ Area Transect Area Management Protection
fragmenta (ha)b length (m) surveyed (ha) regimec statusd
Methods
Disturbance survey
In February and April 2002 and in June and July 2003, disturbance surveys
were carried out at 22 forested sites in Kakamega main forest and its peripheral
fragments (for a complete list of all sites see Table 1). Twelve of the 22 sites
[104]
1165
Figure 2. Location of 22 disturbance survey sites in Kakamega Forest (left) and the number of
trees logged per hectare in the last 20 years for both, trees 610 cm in diameter and >10 cm in
diameter for each site, respectively (right).
chosen for surveys were close to the 12 sites where Mutangah (1996) carried out
his disturbance surveys in 1992/1994 (see Table 1); an additional 10 new sites
where chosen where Mutangah (1996) had not carried out any surveys in the
past (e.g. in the fragments Kisere, Malava and Kaimosi). This was done in
order to obtain a large sample size of representative sites distributed over the
whole Kakamega Forest. The sites chosen were not necessarily near points of
easy access (Figure 2); in fact, many of the sites are located in the centre of the
forest (e.g. No. 5, 9, 10). With many trails running through the whole of
Kakamega Forest, it is easily accessible for local exploitation. At each site
except for the smallest fragment (Kaimosi), transects were run at least 1000 m
in length (Table 1). Transects sometimes followed existing trails, e.g. at Col-
obus site we chose some of the former overgrown monkey research transects
established by Gathua in 1996 (Fashing and Gathua, in press). In all other
cases where trails did not exist, we made our way through undergrowth along a
line. Surveys included recording any of seven disturbance parameters in a belt
[105]
1166
Data analysis
[106]
1167
t-tests (t) for normally distributed data with an adjustment in case variances
were unequal and Mann–Whitney U-tests (U) for non-normally distributed
data. We correlated the different disturbance parameters calculating non-
parametric pairwise Spearman correlations.
We compared our data from 12 sites (i.e. site No. 5, 7, 8, 11–13, 15–17, 19–
21) with data from Mutangah (1996) who quantified the same disturbance
parameters in 1992/1994 at the same 12 sites. Although the sites were the same,
transects were not; thus, data are not dependent data. For comparisons of the
two data sets we calculated Pearson and Spearman rank correlations for
normally and non-normally distributed data, respectively.
Data analysis was carried out using JMP (1995).
Results
Evidence for human impact was found at all our 22 sites with a median
number of 21.1 disturbance events per hectare (q1 = 9.8, q3 = 44.6, range
1.8–81.5, n = 22). The sites Salazar II (No. 7) situated in the northern Ka-
kamega National Reserve and managed by the KWS as well as Yala (No. 19)
situated in the Yala Nature Reserve managed by the FD showed lowest
disturbance levels with 2.8 and 4.9 disturbances per hectare, respectively
(Table 2). The site Ishiru (No. 21) had highest disturbance levels with 68.5
disturbances per hectare; is situated at the southern forest edge and is
managed by the FD.
Of all seven disturbance parameters, logging of trees was the most wide-
spread at all 22 sites (Table 2), thus providing the most useful indicator of
forest disturbance in this study. Over a total survey area of 56.6 ha, we
found 1023 logged trees from 68 species. The most frequent tree species
logged were Funtumia africana (2.1 trees logged per hectare), Prunus africana
(1.3), Celtis new species name: Celtis gomphophylla (1.0), Trilepisium mad-
agascariensis (0.9), Diospyros abyssinica (0.8) and Aningeria altissima (0.7).
The average diameter of tree stumps was 27.0 cm ± 9.6 (if not otherwise
noted mean ± 1 SD).
[107]
1168
Table 2. Summary data of disturbance survey for all 22 sites.
Site No. trees No. trees Total No. trees No. trees No. charcoal No. sawing No. honey No. No. cattle Total no. trees
No. logged/ha logged/ha logged/ha with signs of kilns/ha pits/ha gathering paths/ha tracks/ha logged/ha in
>20 years <20 years debarking/ha sites/ha 1992/1994
Figure 3. Number of trees logged more than 20 years ago (a, b) and in the last 20 years (c, d) for
sites under management by the Kenya Wildlife Service (KWS, n = 6) and the Forest Department
(FD, n = 16) (a, c) and for sites having high protection priority (i.e. situated within National/
Nature Reserves, n = 10) and low protection priority (i.e. situated within Forest Reserves, n = 12)
(b, d). Given are medians, quartils, minimum and maximum values and significance levels. n.s., not
significant; * p < 0.05; **p < 0.005.
20 years) with fewer trees logged at sites managed by the KWS and with high
protection priority (Figures 3c, d; management regime, t: t = 19.21, df = 1,
p = 0.0004; protection status, U: Z = 2.54, p = 0.0111).
[109]
1170
levels (Figure 2). Tree cutting for polewood/timber (tree diameter >10 cm)
had low levels at KWS sites, but high levels at FD sites (Figure 2). Both,
firewood collection and logging for polewood/timber was significantly higher
at FD sites as compared to KWS sites (firewood/ha: KWS median = 0,
q1 = 0, q3 = 0.3, range 0–1.2, FD median = 5.5, q1 = 1.6, q3=17.3, range
0–26.0; U: Z = 3.04, p = 0.0024; polewood/ha: KWS 2.9 ± 2.6; FD
16.9 ± 12.7; t: F = 16.54, df = 1, p = 0.0007).
Similarly, significant differences in the number of logged trees for both
firewood and polewood/timber, were found between sites of high and low
protection priority (firewood/ha: high protection median = 0, q1 = 0,
q3 = 2.0, range 0–19.0, low protection median = 10.0, q1 = 1.8, q3 = 17.3,
range 0–26.0; U: Z = 2.55, p = 0.011; polewood/ha: high protection
median = 3.2, q1 = 1.5, q3 = 8.1, range 0–26.5, low protection med-
ian=15.4, q1 = 7.4, q3 = 26.3, range 2.5–47.5; U: Z = 2.61, p = 0.0092).
[110]
1171
Discussion
Our data do not only show the current status quo of the human impact on
Kakamega Forest, but also reflect its management history in the last
20 years. Before 1986, when all of Kakamega Forest was managed by the
FD, Colobus, Buyangu and Salazar sites (No. 3, 4, 6, 7) in the northern part
were well known for intensive commercial logging through timber companies
(Tsingalia 1988; Mitchell, 2004). Correspondingly, the number of trees logged
more than 20 years ago appears to be equally high at those sites as compared
to others. In 1986, the KWS took over the northern part of Kakamega
Forest as a National Reserve and the changes in management appear to have
resulted in changes in logging numbers in the last 20 years. Illegal tree
poaching was reduced at sites under KWS management probably due to
tightened security, whereas FD sites still experience higher tree poaching rates
today. Furthermore, FD sites show various other local threats such as e.g.
charcoal burning and cattle grazing. Under FD management, sites with high
protection priority such as Yala and Isecheno Nature Reserves (No. 14, 15,
19) still show lower overall threat levels as compared to sites with low pro-
tection priority. For example, Fashing et al.’s (2004) results of a long-term
study of tree populations in Kakamega Forest indicate that their study plots
in Isecheno remained relatively undisturbed over the last 20 years. A decrease
of pioneer species density by 21% in these sites are taken as evidence that the
[111]
1172
forest is maturing towards a climax forest and that at least the conservation
measures applied to Isecheno appear to have succeeded Fashing et al. 2004).
Nevertheless, prospects for other severly disturbed sites are assumed to be
bad, as is the general prognosis for Kakamega Forest if protection efforts are
not increased and illegal exploitation by local people remains high, particu-
larly on its periphery (Cords and Tsingalia 1982; Kokwaro 1988; Tsingalia
1988; Fashing et al. 2004).
How do the two conservation boards KWS and FD differ in their man-
agement aims and strategies? The overall aim of the KWS is ‘to conserve,
protect and sustainably manage the wildlife resources’ and its areas are set
aside for conservation and tourism only (Wass 1995). People are not allowed
to collect any forest products and these policies are strictly enforced through
regular patrols by up to five game rangers (E.W. Kiarie, personal commu-
nication). The overall aim of the FD is to ‘enhance conservation and pro-
tection of indigenous forest, to improve the production of timber and
fuelwood and to establish a framework for the long-term development
forestry’ (Wass 1995). Some sites are also set aside for conservation, how-
ever, some used to be plantations of exotic tree species or mixtures of
indigenous species, while others experienced enrichment planting (A. Oman,
personal communication). Logging, tree debarking and charcoal burning is
prohibited, whereas fuelwood collection was licenced until recently (A.
Oman, personal communication). It appears that the FD has been largely
restricted in its capacity to implement conservation policies effectively due to
the lack of adequate resources in contrast to the better funded KWS,
leading to insufficient levels of staffing, patrols, weaponry etc. These differ-
ences in resources might have led to different disturbance levels as found in
our survey.
Besides overall funding and the number of staff, other potential factors
associated with management regime and effectiveness might be e.g. accessibility
to the forest or proximity of the forest to neighbouring settlements, population
density, community relations and compensation programs to locals. In a recent
assessment of the impact of anthropogenic threats on 93 protected areas of 22
tropical countries park effectiveness was shown to correlate most strongly with
density of guards i.e. the more guards the higher effectiveness (Brunner et al.
2001). Furthermore, effectiveness correlated with the level of deterrence of
illegal activities in the parks and with the degree of border demarcation and
existence of direct compensation programs for local communities (Brunner et
al. 2001). However, it did not correlate with enforcement capacity (i.e. a
composite variable of training, equipment and salary), accessibility, budget,
number of staff working on economic development or education, or the local
involvement of communities in park management (Brunner et al. 2001). To
obtain more information on the factors influencing management effectiveness
in Kakamega Forest more studies are highly recommended following the
RAPPAM guidelines.
[112]
1173
The high human impact on Kakamega Forest especially along the western
and eastern edge of the main forest block indicates an imminent danger of
further fragmentation. The main forest block might fall into two separate
forest blocks, i.e. Kakamega National Reserve in the North and Isecheno
Nature Reserve in the South. To prevent this from happening in the near
future, we strongly recommend following the management plan of forest
zoning as outlined by the Kenya Indigenous Forest Conservation Pro-
gramme (KIFCON 1994; Wass 1995): establishing a protection zone to
provide a core for biodiversity conservation extending from the North to
South; setting up a rehabilitation zone with enrichment planting where
degradation has reached high levels; and establishing a subsistence use zone
flanking the protection zones where local people are allowed to extract
forest products. This forest zoning aims both, to maintain as much indig-
enous forest cover as possible and to permit optimal use of forest resources
on a sustainable basis (Wass 1995). We recommend placing the protection
zone under strict KWS management as our survey indicates that areas of
Kakamega Forest managed by the KWS appear to hold surprisingly low
disturbance levels despite high land-use pressure.
The degradation and logging levels in the suggested subsistence zones are
already alarming, so that we suggest enrichment planting there. Finally,
encouragement of on-farm-forestry projects might provide resources in the
long-term and thus might relieve the subsistence use zone. This is supported
by the fact that a tree nursery run by the local grassroot conservation
organization KEEP (Kakamega Environmental Education Program) at Is-
echeno Forest Station has been successfully nursing seedlings of both,
indigenous and exotic tree species, for sale to local farmers. Beyond con-
servation measures for the main forest block, high protection priority must
also be given to the low-disturbance sites central Ikuywa (No. 17), Yala
(No. 19) and the 400 ha fragment of Kisere (No. 2). Kisere Nature Reserve
is of particurlar conservation significance because it has been relatively
undisturbed in the past and still harbours species-rich forest communities
that include the rare DeBrazza monkeys (Cercopithecus neglectus) (Muriuki
and Tsingalia 1990; Chism and Cords 1997). Although managed by KWS, it
appears to have experienced increasing disturbance levels in the last few
years (N. Saijita and C. Analo, personal communication). This might be due
to the lack of ranger outposts in Kisere (KWS headquarters is at Buyangu),
and the fact that the number of rangers (10–20) might not be sufficient to
cover both, Buyangu and Kisere. Therefore, an immediate increase of reg-
ular ranger patrols to control logging more effectively is highly recom-
mended, as suggested by previous authors (Kokwaro 1988; Mutangah 1996;
Chism and Cords 1997).
[113]
1174
[114]
1175
Acknowledgements
The study was funded by the German Federal Ministry of Education and
Research within the framework of BIOTA East Africa (01LC0025 / subpro-
jects E03, E10 & E11). We thank the Kenyan Ministry for Education and
Research for the permission to carry out research in Kakamega Forest, and the
KWS and FD for granting us access to their reserves. We appreciate infor-
mation on management by E.W. Kiarie (Senior Warden, Kakamega National
Reserve) and A. Oman (Assistant District Forest Officer, Kakamega). We
highly acknowledge field assistance by C. Analo and N. Saijita. G. Schaab
kindly provided maps and data on forest patch sizes. We wish to thank the
paperclub at Mainz, N. Mitchell, L. Todt, H.Todt, and two anonymous ref-
erees for comments on earlier drafts of the manuscript and K. Boehning-Gaese,
H. Dalitz and M. Kraemer for overall support.
References
Barlow J., Haugaasen T. and Peres C.A. 2002. Effects of ground fires on understorey bird
assemblages in Amazonian forests. Biol. Conserv. 105: 157–169.
Bennun L. and Njoroge P. 1999. Important Bird Areas in Kenya. The East Africa Natural History
Society, Nairobi, Kenya.
Blackett H.L. 1994. Forest inventory report no. 3: Kakamega. Kenya Indigenous Forest Conser-
vation Programme, Nairobi, Kenya.
Bowles I.A., Rice R.E., Mittermeier R.A. and da Fonseca G.A.B. 1998. Logging and tropical forest
conservation. Science 280: 1899–1900.
Brooks T.M., Pimm S.L. and Oyugi J.O. 1999. Time lag between deforestation and bird extinction
in tropical forest fragments. Conserv. Biol. 13: 1140–1150.
Brunner A.G., Gullison R.E., Rice R.E. and da Fonseca G.A.B. 2001. Effectiveness of parks in
protecting tropical biodiversity. Science 291: 125–128.
Chapman C.A. and Chapman L.J. 1997. Forest regeneration in logged and unlogged forests of
Kibale National Park, Uganda. Biotropica 29: 396–412.
[115]
1176
Chism J. and Cords M. 1997. De Brazza’s monkeys Cercopithecus neglectus in the Kisere National
Reserve, Kenya. Afr. Primates 3: 18–22.
Cochrane M.A. and Laurance W.F. 2002. Fire as a large-scale edge effect in Amazonian forests.
J. Trop. Ecol. 18: 311–325.
Cords M. 1987. Mixed species association of Cercopithecus monkeys in the Kakamega Forest,
Kenya. Univ. Cali. Publ. Zool. 117: 1–109.
Emerton L. 1991. Summary of the Current Value of use of Kakamega Forest. Kenya Indigenous
Forest Conservation Programme, Nairobi, Kenya.
Ervin J. 2003a. Protected area assessment in perspective. BioScience 53: 819–822.
Ervin J. 2003b. Rapid assessment of protected area management effectiveness in four countries.
BioScience 53: 833–841.
Ervin J. 2003c. Rapid assessment and prioritization of protected area management (RAPPAM)
methodology. World Wide Fund for Nature, Gland, Switzerland. Available at http://
www.panda.org/downloads/forests/rappam.pdf.
Fashing P.J. Mortality trends in the African cherry (Prunus africana) and the implications for
colobus monkeys (Colobus guereza) Kakamega Forest, Kenya. Biodiv. Conserv., in press.
Fashing P.J., Forrestel A., Scully C. and Cords M. 2004. Long-term tree population dynamics and
their implications for the conservation of the Kakamega Forest, Kenya. Biodiv. Conserv. 13:
753–771.
Fashing P.J. and Gathua M. Spatial variability in the vegetation structure and composition of an
East African rain forest. Afr. J. Ecol. in press.
Heywood V.H. (ed.) 1995. Global Biodiversity Assessment. Cambridge University Press, Cam-
bridge, UK.
Hockings M. 2003. Systems of assessing the effectiveness of management in protected areas.
BioScience 53: 823–832.
Hockings M., Stolton S. and Dudley N. 2000. Evaluating effectiveness: a framework for assessing
the management of protected areas. IUCN, Gland, Switzerland. Available at http://wcpa.
iucn.org/pubs/pdfs/evaluating_effect.pdf.
JMP 1995. Statistics and Graphics Guide. Version 3.1. SAS Institute Inc., Cary, NC.
Kiama D. and Kiyiapi J. 2001. Shade tolerance and regeneration of some tree species of a tropical
rain forest in western Kenya. Plant Ecol. 156: 183–191.
KIFCON 1994. Kakamega Forest – The offical guide. Kenya Indigenous Forest Conservation
Programme, Nairobi, Kenya.
Kowkaro J.O. 1988. Conservation status of the Kakamega Forest in Kenya: the easternmost relict
of the equatorial rain forests of Africa. Monogr. Syst. Bot. Missouri Bot. Garden 25: 471–489.
Laurance W.F. 1998. A crisis in the making: responses of Amazonian forests to land use and
climate change. Trends Ecol. Evol. 13: 411–415.
Laurance W.F. 2001. Tropical logging and human invasions. Conserv. Biol. 15: 4–5.
Lindenmayer D.B., Margules C.R. and Botkin D.B. 2000. Indicators of biodiversity for ecologi-
cally sustainable forest management. Conserv. Biol. 14: 941–950.
Margules C.R. and Pressey R.L. 2000. Systematic conservation planning. Nature 405: 243–253.
Margoluis R. and Salafsky N. 1998. Measures of Success: Designing, Managing, and Monitoring
Conservation and Development projects. Island Press, Washington, DC.
Mitchell N. 2004. The exploitation and disturbance history of Kakamega Forest, western Kenya.
In: Bleher B. and Dalitz H. (eds), BIOTA Report No. 1. Bielefelder Ökologische Beiträge 20.
Muriuki J.W. and Tsingalia M.H. 1990. A new population of de Brazza’s monkey in Kenya. Oryx
24: 157–162.
Mutangah J.G. 1996. An investigation of vegetation status and process in relation to human
disturbance in Kakamega Forest, western Kenya. Ph.D thesis, University of Wales, UK.
Mutangah J.G., Mwangangi O. and Mwaura P.K. 1992. Kakamega: A Vegetation Survey Report.
Kenya Indigenous Forest Conservation Programme, Nairobi, Kenya.
Myers N., Mittermeier R.A., Mittermeier C.G., da Fonseca G.A.B. and Kent J. 2000. Biodiversity
hotspots for conservation priorities. Nature 403: 853–858.
[116]
1177
Noss R.F. 1990. Indicators for monitoring biodiversity: a hierarchical approach. Conserv. Biol. 4:
355–364.
Oyugi J.O. 1996. Kakamega Forest is dying. East Afr. Nat. Hist. Soc. Bull. 26: 47–49.
Parrish J.D., Braun D.P. and Unnasch R.S. 2003. Are we conserving what we say we are? Mea-
suring ecological integrity within protected areas. BioScience 53: 851–860.
Pimm S.L., Russell G.J., Gittleman J.L. and Brooks T.M. 1995. The future of biodiversity. Science
269: 347–350.
Plumptre A.J. 1996. Changes following 60 years of selective timber harvesting in the Budongo
Forest Reserve, Uganda. Forest Ecol. Manage. 89: 101–113.
Putz F.E., Blate G.M, Redford K.H, Fimbel R. and Robinson J. 2001. Tropical forest management
and conservation of biodiversity: an overview. Conserv. Biol. 15: 7–20.
Rice R.E., Gullison R.E. and Reid J.W. 1997. Can sustainable management save tropical forests?
Sci. Am. 276: 44–49.
Robinson J.G., Redford K.H. and Bennett E.L. 1999. Wildlife harvest in logged tropical forest.
Science 284: 595–596.
Salafsky N. and Margoluis R. 1999. Thread reduction assessment: a practical and cost-effective
approach to evaluating conservation and development projects. Conserv. Biol. 13: 830–841.
Salafsky N., Margoluis R., Redford K.H. and Robinson J.G. 2002. Improving the practice of
conservation: a conceptual framework and research agenda for conservation science. Conserv.
Biol. 16: 1469–1479.
Sparrow H.R., Sisk T.D., Ehrlich P.R. and Murphy D.D. 1994. Techniques and guidelines for
monitoring neotropical butterflies. Conserv. Biol. 8: 800–809.
Struhsaker T.T. 1997. Ecology of an African Rain Forest: Logging in Kibale and the Conflict
Between Conservation and Exploitation. University of Florida Press, Gainesville, Florida.
Tsingalia M.H. 1988. Animals and the regeneration of an African rainforest tree. Ph.D. Thesis,
University of California, Berkely.
van Schaik C.P., Terborgh J. and Dugelby D. 1997. The silent crisis: the state of the rainforest
nature reserves. In: Kramer R., van Schaik C.P. and Johnson J. (eds), The Last Stand: Protected
Areas and the Defense of Tropical Biodiversity. Oxford University Press, New York, pp. 64–89.
Wass P. 1995. Kenya’s Indigenous Forests – Status, Management and Conservation. IUCN,
Gland, Switzerland.
Whitmore T.C. 1997. Tropical forest disturbance, disappearance and species loss. In: Laurence
W.L. and Bieregaard R.O.Jr. (eds), Tropical Forest Remnants: Ecology, Management, and
Conservation of Fragmented Communities. University of Chicago Press, Chicago, pp. 3–12.
Wilkie D.S., Sidle J.G. and Boundzanga G.C. 1992. Mechanized logging, market hunting, and a
bank loan in Congo. Conserv. Biol. 6: 570–580.
Zimmerman D.A. 1972. The avifauna of the Kakamega Forest, western Kenya, including a bird
population study. Bull. Am. Museum of Nat. Hist. 149: 257–339.
[117]
Biodiversity and Conservation (2006) 15:1179–1191 Springer 2006
DOI 10.1007/s10531-004-4693-x
-1
Key words: Batophila acutangula, Biodiversity, Insect diversity, Larix kaempheri, Leaf beetles,
Species composition, Species richness, Sphaeroderma tarsatum
Abstract. Species richness and composition of the Chrysomelidae (Coleoptera) were studied in
larch (Larix kaempheri [Lamb.] Carrière) plantations, secondary forests, and primary forests. In
addition, the effects of forest management practices, such as thinning and long rotation, were
examined in the larch plantation. The species richness of Chrysomelidae was higher in the larch
plantation than in the secondary forest or in the primary forest. Among the larch plantations, the
species richness in old-aged plantations was higher than that in middle-aged plantations. The
composition of the beetle assemblages in the larch plantation differed from that in the secondary
forest or in the primary forest. Exosoma akkoae (Chujo), Batophila acutangula Heikertinger, and
Calomicrus nobyi Chujo were caught with a bias toward the larch plantation. Longitarsus succineus
(Foudras) and Sphaeroderma tarsatum Baly were caught more in the secondary forest and the
primary forest, respectively. More B. acutangula and S. tarsatum were caught in stands where their
host plants occurred at higher rates. Species richness of understory plants was an important factor
for chrysomlid species richness, and frequency of host occurrence affected the number of indi-
viduals of leaf beetles examined. It seems that forest types and forest management practices affect
host plants as well as Chrysomelidae, and that these effects on the host plants also influence
chrysomelid assemblages.
Introduction
[119]
1180
Quercus crispula Blume, (41 and 33% of the total area, respectively). Larch
plantations were established for timber production; thinning has been conducted
twice within a 45-year period in this area, in order to enhance growth of domi-
nant trees. Long rotations have been adopted for some larch plantations to
obtain high quality wood. Secondary forests were formerly used to produce
firewood, charcoal, etc. Now, however, they are mostly left unattended.
Because larch plantations and secondary forests occupy large areas, their
role in conserving biodiversity is important. Primary forests, though frag-
mented today, should also be considered important because of the potential
presence of rare indigenous species inhabiting in them. Three forest types (larch
plantation, secondary forest, and primary forest) and two management prac-
tices frequently used in larch plantations (thinning and long rotations) were
chosen for this study.
Leaf beetles (Chrysomelidae, Coleoptera) are small in size (less than 1.5 cm
in general) and their antennas are usually less than half the length of their
bodies. They distribute worldwide and about 500 species have so far been
reported in Japan (Kimoto and Takizawa 1994). Though they are related with
cerambycid beetles which are saproxylic in their larval stage, leaf beetles feed
on leaves, stems or roots in both larval and adult stages and are regarded as
phytophagous herbivores. The diversity of Cerambycidae, saproxylic beetles,
was investigated in this area, and it was reported that higher species richness of
the beetles was observed in secondary forests than in larch plantations or
primary forests, and that thinning increased cerambycid diversity in larch
plantations (Ohsawa 2004). This time, leaf beetles were chosen for the study of
diversity to clarify the effects of different forest types and forestry activities on
the phytophagous herbivores. There has been one report published on the
subject of the diversity and conservation of Chrysomelidae (Greatorex-Davies
and Sparks 1994). In this study, the diversity of leaf beetles and some other
insects was investigated in the rides of woodlands, and it has been reported that
both species richness and abundance of leaf beetles declined with increasing
levels of shade, and that rides must be actively managed to keep light levels
high if species richness and abundance are to be maintained.
The purpose of this study was to compare the species richness and compo-
sition of Chrysomelidae, phytophagous herbivores, in larch plantations, sec-
ondary forests, and primary forests, and assess the effects of forest practices,
such as thinning and long rotations, on the diversity of this family in order to
obtain information for conservation of leaf beetles in forest area.
Methods
Study site
[120]
1181
A Malaise trap was set in each of the 46 stands to capture insects, and leaf
beetles were separated from among those trapped. To minimize forest edge
effects and micro-topographical differences within each stand, efforts were
made to set each trap in a typical spot for the stand on a simple slope (or in a
flat area if there was no slope) in the interior of the stand.
Leaf beetles were captured in three 14-day periods: in middle to late June,
in middle to late July, and in early to mid-August. Because of high eleva-
tion, June to August is an appropriate period for capturing many beetles
including Chrysomelidae, according to the results of an investigation
conducted on insects through spring to autumn in this area (Ohsawa
unpublished data).
Dry specimens were prepared for all species of leaf beetles trapped, and were
kept in the Yamanashi Forest Research Institute.
Environmental factors
[121]
1182
Statistical analyses
One way ANOVA with Bonferroni’s post-hoc test (Bonferroni test) was con-
ducted to compare the number of species, the number of individuals, and
Shannon-Wiener diversity index (H¢, Magurran 1988) among the forest types.
Logarithmic transformation was performed for the number of individuals
before Bonferroni test. Simple regression analysis was conducted to identify the
environmental factors that affected the species richness of leaf beetles. The
regression analysis was also applied to the relationship between the numbers of
two chrysomelid species and occurrence of their host plants.
Detrended correspondence analysis (DCA) was applied to compare the
composition of leaf beetles in all investigated stands. The analysis was conducted
with the software package, PC-ORD (McCune and Mefford 1999). Bonferroni
test was applied to compare DCA scores (the first and second axes) among forest
types. To define the meaning of the first two axes of DCA, simple correlation
coefficients were calculated between the two axes and stand-structure parame-
ters, such as the number of species, the number of individuals, H¢, larch plan-
tation vs. natural broad-leaved forest, altitude, gradient, openness of the canopy,
the number of vascular plants (<2 m and 2 m in height, for 11 stands) and
forest age (for larch plantations),
Indicator species analysis (Dufrêne and Legendre 1997) was carried out to
compare species of Chrysomelidae among the forest types and to identify
species associated with each forest type.
[122]
1183
Results
In this study, 1303 individuals of 41 species were captured. The species with the
highest number of individuals captured was Aphthona perminuta Baly (635),
followed by S. tarsatum (201), Exosoma akkoae (Chujo) (92), Zipangia lewisi
(Jacoby) (54), Syneta adamsi Baly (51), B. acutangula (47), Cryptocephalus
amiculus Baly (30), Stenoluperus nipponemsis (Laboissiere) (29), Pseudoliprus
nigritus (Baly) (26), and Zeugophora annulata (Baly) (25).
The average number of species, the average number of individuals, and the
average H¢ in each forest type are shown in Table 1. The number of species and
H¢ in the larch plantation was significantly higher than those in the secondary
forest or in the primary forest (Bonferroni test, p < 0.05). The number of
individuals was not significantly different among three forest types.
In the larch plantation, the number of species and that of individuals were
significantly higher in the old plantation than in the middle-aged plantation
(Bonferroni test, p < 0.05, Table 2). H¢ was not significantly different among
three types of larch plantations.
Environmental factors
Table 1. Numbers of chrysomelid species, numbers of individuals, and diversity index (H¢) in
three forest types.
[123]
1184
Table 2. Numbers of chrysomelid species, numbers of individuals and diversity index (H¢) in three
types of larch plantations.
Species composition
Table 3. Relationship between number of species and environmental factors (simple regression
analysis).
[124]
1185
Figure 1. Detrended correspondence analysis ordination (first and second axes) performed on leaf
beetles in 46 stands in central Japan.
Discussion
Species richness
Species richness of leaf beetles in the larch plantation was significantly higher
than that in the secondary forest or in the primary forest. Nagaike (2002)
Table 4. Simple correlation coefficients for two chrysomelid speices between the number of
individuals and occurrence of host plants.
S. tarsatum B. acutangula
[125]
1186
Table 5. Simple correlaton coefficients between scores of detrended corespondence analysis (first
two axes) and stand-structure parameters.
studied diversity of vascular plants in this area and reported that species
richness was significantly higher in larch plantations than in secondary forests.
This result in higher species richness of vascular plants coincides with higher
species richness of Chrysomelidae in the larch plantation than in the secondary
forest. In this study, species richness of leaf beetles was higher in stands with
higher species richness of understory plants (<2 m in height), though the
relationship between species richness of leaf beetles and species richness of the
plants (2 m) was not clear. Because leaf beetles feed on leaves, stems, or roots
of various plants, and because host plants of the beetles vary with the beetle
species, it seems that forests with more plant species enable chrysomelid species
to inhabit in them in higher numbers.
Among larch plantations, the old larch plantation had a higher species
richness of leaf beetles than did the middle-aged plantation. This result cor-
responds with the result of regression analysis that the higher number of
chrysomelid species was caught in older larch plantations. Species richness of
vascular plants was higher in the old plantation than in the middle-aged
plantation, though the difference was not significant (Nagaike et al. 2003).
They also found that plant species composition was different between the
middle-aged plantation and the old plantation. This difference may have
resulted in a higher species richness of Chrysomelidae in the old plantation
than in the middle-aged plantation. Further investigations will be necessary to
clarify this point. Thinning did not cause significant difference in species
richness of leaf beetles, though the recently thinned plantation had wider
openness of the canopy than the middle-aged plantation did (Bonferroni test,
p < 0.01).
S. tarsatum and B. acutangula had a significant Ind Val for the primary
forest and the larch plantation, respectively. However, more individuals of
both beetles were found in stands with higher occurrences of host plants. For
[126]
1187
Table 6. Indicator value (Ind Val) of chrysomelid species for three forest types.
[127]
1188
Table 7. Indicator value (Ind Val) of chrysomelid species for larch plantation types.
Species composition
As the results of DCA showed that scores in the second axis was higher in the
larch plantation than in the secondary forest or in the primary forest, the
composition of Chrysomelidae in the larch plantation differed from that in the
secondary forest or in the primary forest.
Six chrysomelid species had significant Ind Vals in relation to the three forest
types in this study. E. akkoae, B. acutangula, and C. nobyi each showed a
significant Ind Val for the larch plantation. B. acutangula and C. nobyi feed on
Rubus spp. and Clematis apiifolia DC., respectively (Kimoto and Takizawa
[128]
1189
1994). The host plants for E. akkoae are unknown. L. succineus having
a significant Ind Val for the secondary forest feeds on Artemisia spp. S. tarsatum
had a significant Ind Val for the primary forest and feeds on bamboo grass. It is
interesting that host plants of these chrysomelid species having significant Ind
Vals for certain forest types in this study were not dominant tall trees, but
understory herbs.
Conclusion
The larch plantation had a higher species richness of Chrysomelidae than did
other two natural forests. However, the composition of chrysomelid assem-
blage in the larch plantation was different from that in the primary forest which
is considered to have indigenous species in the area. Among the three types of
[129]
1190
larch plantation, the species richness of Chrysomelidae was higher in the old
plantation than in the middle-aged plantation. Long rotations of larch plan-
tations seem favorable for chrysomelid diversity.
No difference in chrysomelid species composition between the secondary
forest and the primary forest could be detected by DCA. Since the primary
forest which is thought to contain indigenous species has diminished and
fragmented, the secondary forest is expected to maintain these leaf beetles.
However, considering that the secondary forest contained only 43% (6/14
species) of species caught in the primary forest in this study, the forest type may
not be sufficient to maintain the beetle assemblage in the primary forest.
This study showed that species richness of understory plants is an important
factor for chrysomelid species richness, and that the frequency of host occur-
rence significantly affects the number of individuals. Because forest types affect
plant diversity (Qian et al. 1997; Nagaike 2002; Nagaike et al. 2003), it is likely
that both forest types and forest management practices affect host plants as
well as Chrysomelidae, and that these effects on the host plants also influence
chrysomelid assemblages.
Acknowledgements
We wish to thank Mr. Akihiro Ohashi, Gifu Forest Science Institute, and
Dr. Masahiro Isono, Forestry and Forest Products Research Institute, Japan,
for identifying several leaf beetles and helpful advice. We are also grateful to
Mr. Masao Osawa for correction of English for the manuscript and helpful
advice. We deeply appreciate technical assistance given me by Mrs Miki Saso,
and the helpful advice and encouragement by members of Yamanashi Forest
Research Institute.
References
Dufrêne M. and Legendre P. 1997. Species assemblages and indicator species: the need for a flexible
asymmetrical approach. Ecol. Monogr. 67: 345–366.
Fridman J. and Walheim M. 2000. Amount, structure, and dynamics of dead wood on managed
forestland in Sweden. Forest Ecol. Manage. 131: 23–36.
Greatorex-Davies J.N. and Sparks T.H. 1994. The response of Heteroptera and Coleoptera species
to shade and aspect in rides of coniferised lowland woods in Southern England. Biol. Conserv.
67: 255–273.
Green P. and Peterken G.F. 1997. Variation in the amount of dead wood in the woodlands of the
Lower Wye Valley, UK in relation to the intensity of management. Forest Ecol. Manage. 98:
229–238.
Kimoto S. and Takizawa H. 1994. Leaf Beetles (Chrysomelidae) of Japan. Tokai University Press,
Tokyo.
Kirby K.J., Reid C.M., Thomas R.C. and Goldsmith F.B. 1998. Preliminary estimates of fallen
dead wood and standing dead trees in managed and unmanaged forests in Britain. J. Appl. Ecol.
35: 148–155.
[130]
1191
Magurran A.E. 1988. Ecological Diversity and its Measurement. Princeton University Press,
Princeton, NJ.
Martikainen P., Siitonen J., Punttila P., Kaila L. and Rauh J. 2000. Species richness of Coleoptera
in mature managed and old-growth boreal forests in southern Finland. Biol. Conserv. 94:
199–209.
McCune B. and Mefford M.J. 1999. PC-ORD. Multivariate Analysis of Ecological data, version 4.
MjM Software Design, Gleneden Beach, Oregon.
Nagaike T. 2002. Differences in plant species diversity between conifer (Larix kaempferi) planta-
tions and broad-leaved (Quercus crispula) secondary forests in central Japan. Forest Ecol.
Manage. 168: 111–123.
Nagaike T., Hayashi A., Abe M. and Arai N. 2003. Differences in plant species diversity in
Larix kaempferi plantations of different ages in central Japan. Forest Ecol. Manage. 183:
177–193.
Ohsawa M. 2004. Species richness of Cerambycidae in larch plantations and natural broad-leaved
forests of the central mountainous region of Japan. Forest Ecol. Manage. 189: 375–385.
Qian H., Klinka K. and Sivak B. 1997. Diversity of the understory vascular vegetation in
40 year-old and old-growth forest stands on Vancouver Island, British Columbia. J. Veg. Sci. 8:
773–780.
ter Steege H. 1993. HEMIPHOTO, A Programme to Analyze Vegetation Indices, Light Quality
from Hemispherical Photographs. The Tropebos Foundation, Wageningen.
[131]
Biodiversity and Conservation (2006) 15:1193–1217 Springer 2006
DOI 10.1007/s10531-004-8230-8
-1
ANDREAS HEMP
Department of Plant Physiology, Universität Bayreuth, Universitätsstr. 30, 95440 Bayreuth (e-mail:
[email protected]; phone: 0921-552630; fax: 0921-552642)
Abstract. Natural flora, vegetation, diversity and structure of 62 traditional coffee–banana plan-
tations on Kilimanjaro were investigated and compared with the other vegetation formations on
this volcano on basis of over 1400 plots following the method of Braun-Blanquet. The vegetation of
the so-called Chagga homegardens belongs floristically to the formation of ruderal vegetation
forming two main communities that are determined by altitude. These coffee–banana plantations
maintain a high biodiversity with about 520 vascular plant species including over 400 non-culti-
vated plants. Most species (194) occurring in the Chagga homegardens are forest species, followed
by 128 ruderal species, including 41 neophytes. Typical of the agroforestry system of the Chagga
homegardens is their multilayered vegetation structure similar to a tropical montane forest with
trees, shrubs, lianas, epiphytes and herbs. Beside relicts of the former forest cover, which lost most
of their former habitats, there are on the other hand (apophytic) forest species, which were directly
or indirectly favoured by the land use of the Chagga people. High demand of wood, the intro-
duction of coffee varieties that are sun-tolerant and low coffee prizes on the world marked endanger
this effective and sustainable system.
Introduction
Humans have continuously inhabitated the slopes of Mt. Kilimanjaro for the
last 2000 years (Schmidt 1989). However, during the last decades the human
population increased dramatically. According to ethnographic studies of
Widenmann (1899) 50,000–60,000 Chagga people were estimated to live on Mt.
Kilimanjaro in 1895. In 2002 the census counted 1,053,204 people (National
Bureau of Statistics 2003). As such, the population has multiplied 20 times
since 1895. Most of the population is concentrated at an altitude between 1000
and 1800 m, with densities varying from 500 to 1000 people per km2 in some
areas (FAO 1986; Timberlake 1986). Here, within the submontane zone, a very
remarkable kind of land use prevails: dense ‘banana forests’ with a scattered
upper tree layer, the so-called Chagga homegardens, in Chagga language
‘vihamba’ (the term ‘homegarden’ refers to the small size and subsistence-level
of the farms, cp. Nair 1993). Due to this sustainable and well developed
agroforestry system (cp. e.g. Fernandes et al. 1984) degradation in this vege-
tation belt is rare, despite the enormous population. In their homegardens the
[133]
1194
Chagga use four vegetation layers. Under a tree layer, which provides fire-
wood, fodder and shadow, banana trees (in about 25 varieties, cp. Simmonds
1966) are grown and under the bananas coffee trees, and under these vegeta-
bles. This multilayer system maximizes the use of limited land. The area is
irrigated by a network of canals fed by main furrows originating from the
montane forest. Rough estimates give over one thousand furrows of varying
lengths and capacities (Ramsay 1965) some dating back to the 17th century
(Hemp and Winter 1999). This farming system evolved over several centuries
and did not change much over the last decades compared with the land uses in
the lower zones.
This densely populated coffee–banana belt stretches on the climatically most
favourable zone of the southern and south-eastern slopes (Figure 1) over an
area of 1000 km2. The Chagga live within their homegardens in single dwell-
ings; villages as such do not exist. But, along the main roads centres with
church, village council, schools and some shops are situated. The average size
of a homegarden varies between 0.5 and 1.7 ha (O’Kting’ati and Kessy 1991;
Mdoe and Wiggins 1997). Livestock – cattle, goats, sheep and pigs – and
sometimes poultry are kept in stalls. Women and children spend a great part of
the day collecting grass along paths, fields and forest edges and on steep
meadow slopes. Pasture farming is rare in submontane zone due to intensive
agriculture. Therefore (based on the nature of their components) the Chagga
homegardens can be classified as an agrisilvicultural system (cp. Nair 1993).
Bee-keeping plays an important role.
The chagga homegardens were the subject of different studies dealing mainly
with socio-economic (Clemm 1963; Brewin 1965; Fernandes et al. 1984;
O’Kting’ati and Kessy 1991; Mdoe and Wiggins 1997) or ethnobotanical and
ethnozoological aspects (O’Kting’ati et al. 1984; Hemp 1999; Hemp 2001). The
array of cultivated species was already described in detail by the first scientists
on Kilimanjaro e.g. Volkens (1897) or Widenmann (1899). The aim of the
present study is to describe natural flora, vegetation and structure of the
Chagga homegardens and to highlight their function for biodiversity and as a
refuge area for natural plant species. For this purpose the species composition
of this man-made habitat is compared with all vegetation formations of Mt.
Kilimanjaro as presented by Hemp (2001a). In a parallel study the function of
the Chagga homegardens as a habitat of endangered and endemic grasshopper
species was investigated on the same plots (Hemp in press).
Study area
Mt. Kilimanjaro, a relic of an ancient volcano, rising from the savanna plains
at 700 m elevation to a snow-clad summit of 5895 m altitude is located 300 km
south of the equator in Tanzania on the border with Kenya. Its climate is
characterized by a bimodal rainfall pattern with the long rains from March to
May forming the main rainy season, and the short rains centred around the
[134]
1195
Figure 1. Land use and vegetation cover of the study area with location of plots (dots).
month of November of the small rainy season. The foothills of the southern
slopes receive an annual rainfall of 800–900 mm and the lower slopes at
1500 m receive 1500–2000 mm. The forest belt between 2000 and 2300 m re-
ceives partly over 3000 mm (Hemp 2001a), which is more than on other high
mountains of East Africa. In the alpine zone the precipitation decreases to
200 mm.
According to the different climatic conditions several vegetation zones are
apparent on the southern slopes of Mt. Kilimanjaro (Figure 1). Between 700
and 1000 m a.s.l. the dry and hot colline savanna zone stretches around the
mountain base, where most areas are farmed with maize, beans and sunflowers,
in West Kilimanjaro with wheat. Around Lake Chala at the eastern foot of the
mountain, and around Ngare Nairobi of West Kilimanjaro, savanna grass-
lands are still intact. The main cultivation zone with its coffee–banana plan-
tations, the actual study area, is located between 1000 and 1800 m. Natural
forests cover an area of about 1000 km2 on Mt. Kilimanjaro. In the lower parts
of the southern slope the montane forests are characterized by the tree Ocotea
usambarensis and higher up in the cloud forest zone by Podocarpus latifolius,
Hagenia abyssinica and Erica excelsa. On the drier northern slope the vege-
tation zonation starts with Croton-Calodendrum forests, Cassipourea forests at
midaltitudes and Juniperus forests at higher altitudes. At around 3100 m the
forests are replaced by Erica bush. At an altitude of about 3900 m the Erica
heathlands grade into Helichrysum cushion vegetation that reaches up to
4500 m. Higher altitudes are very poor in vegetation while the highest
[135]
1196
elevations of Kibo peak are covered with glaciers. For a more detailed
description of these vegetation types see Hemp (2001a).
Methods
Data have been collected since 1996. Over 1400 relevés (plots) of all vegetation
types were produced on the whole mountain using the method of
Braun-Blanquet (1964), 62 of them representing banana gardens (location see
in Figure 1). Special attention was given to homogeneity and representation of
the stands. The relevé size was chosen with respect to the minimum area, which
was 1000 m2 in the banana gardens.
In the relevés a herb layer (<1 m tall), a shrub layer (1–10 m tall) and a tree
layer (>10 m tall) were differentiated. Climbers (stranglers, scramblers,
tendrillar plants, hook and root climbers and twiners, cp. Lind and Morrison
1974) reaching into the shrub and tree layer were defined as lianas. Epiphytes
(accidental, casual, hemi- and holo-epiphytes, cp. Kress 1986) were treated as
an additional layer. To portray the structure of the banana gardens, a stand-
profile diagram according to Hammen et al. (1989) was produced.
The relevés were clustered according to floristic similarity and the resulting
plant communities were united into nine formations: rocks; ruderal vegetation;
vegetation of trampled grounds; grasslands; salt marshes; freshwater swamps;
forest clearings; forests, and; heathlands. Constancy tables of these formations
(except the vegetation of trampled ground) are presented by Hemp (2001a).
Based on their constancy in the different formations (or vegetation classes)
character species of these formations were determined.
pH was measured in the main root horizon in 5–10 cm depth of selected
plots using a WTW pH-metre (pH 330). Two parallel samples were taken and
measured in distilled water and 0.01 M CaCl2 solution, respectively.
Nomenclature follows FTEA (1952–2003) and (Beentje 1994).
Results
Based on Landsat ETM images from the year 2000 (source: USGS/UNEP-
GRID-Sioux Falls) extend and distribution of the Chagga homegardens on
Kilimanjaro was determined. Excluding other vegetation types of the sub-
montane cultivated zone (e.g. riverine forests, meadows, maize fields etc.)
banana fields cover an area of about 675 km2, which is about two thirds of the
extend of the montane forest belt.
[136]
[137] Table 1. Vegetation of the Chagga homegardens.
1197
Table 1. Continued.
1198
[138]
1199
[139]
1200
Hemp 2001a). Ruderal species form in respect of species number and vegeta-
tion cover of the herb layer the most important ecological group. Widespread
ruderal species (character species of the vegetation class) that can be found as
well on fallow arable land, waste places in towns and roadsides from the colline
savanna area up to the montane zone are e.g. Bidens pilosa, Oxalis corniculata,
Commelina benghalensis and Galinsoga parviflora. Phyllanthus odontadenius,
Didymodoxa caffra and Celosia schweinfurthiana occur in different ruderal
vegetation types within the submontane zone. Finally, a characteristic weed
species confined to the banana gardens on Kilimanjaro is Oxalis latifolia.
Based on the floristic composition, two distinct plant communities were
distinguished. Diagnostic species of community 1 are the ruderal species
Euphorbia heterophylla, Amaranthus hybridus, Sida acuta, Blainvillea gayana
and Malvastrum coromandelianum. Impatiens walleriana, Drymaria cordata and
Plectranthus parvus are diagnostic for community 2; the latter species is re-
stricted in the study area to this community. Each of these two main com-
munities can be divided into a vicarious vegetation unit of the southern and the
(north-) eastern slope, respectively, governed by the rainfall regime. This
partition into a wet southern and drier northern slope is also reflected by other
vegetation formations such as forests (cp. Hemp in press a).
Determining factors for the differentiation of the two main communities 1
and 2 are altitude and factors related to altitude: community 1b of the southern
slope occurs between 800 m and about 1300 m, receiving 900–1580 mm rain-
fall, and community 2b between 1300 and 1800 m, receiving 1580–2200 mm
rainfall. Mean annual temperature varies between 23.4 and 18.8 C in com-
munity 1b and between 18.8 and 16.1 C in community 2b (climate data from
Hemp in press a). As a result of the increasing precipitation mean pH-values
decrease from 6.4 (CaCl2) and 6.7 (H2O), respectively in community 1b to 5.5
(CaCl2) and 6.0 (H2O), respectively in community 2b. As in the forests of
Kilimanjaro (Hemp in press a, b) the influence of precipitation is reflected by
the distribution of epiphytes and pteridophytes. Vascular epiphytes are nearly
missing in community 1b and start above 1300 m in community 2b. The same
holds for pteridophytes. Mean cover values of epiphytes (including mosses and
lichens) in both communities are 0 and 8%, respectively.
On the drier eastern slope the lowland community 1a occurs between 1100
and about 1500 m, and community 2b is less distinctly developed than on the
wetter southern slope, occurring only in a narrow strip above 1500 m. Such a
rising and condensation of vegetation zones due to the ‘Mass-
enerhebungseffekt’ was also observed in the forest vegetation on the dryer
leeward slope of Kilimanjaro (Hemp in press a).
Diversity
Five hundred and twenty-three vascular plant species were recorded in the
relevés. These are about three quarter of the species occurring in the ruderal
[140]
1201
Figure 2. Vascular plant species richness in the main vegetation formations (as presented with
constancy tables by Hemp 2001a) of Mt. Kilimanjaro, based on the evaluation of about 1400
vegetation plots.
vegetation formation on Kilimanjaro (Figure 2). With over 700 species this
formation holds rank three in respect of species richness after the forests and
grasslands. Mean vascular plant species number per plot is 54.
Figure 3 shows the floristic composition of the banana fields in relation to
the different vegetation formations on Kilimanjaro. Most species (194)
occurring in the Chagga homegardens are forest species, followed by 128
ruderal species. Cultivated plants contribute 19% to this floristic spectrum with
99 species.
A characteristic feature of the ruderal vegetation on Kilimanjaro is the high
contribution of neophytes. In the Chagga homegardens nearly a quarter of the
species is introduced, with 99 cultivated species and 41 neophytes (Figure 4,
Table 2).
Vegetation structure
[141]
1202
Figure 3. Floristic composition of the banana fields in respect of the different vegetation formations
on Kilimanjaro.
Figure 4. Share of cultivated, neophytic and indigenous plants in the Chagga homegardens.
layer this number adds to 82) are remnants of the former forest cover (Table 3).
Most widespread are Albizia schimperiana, Rauvolfia caffra, Cordia africana,
Commiphora eminii and Margaritaria discoidea. Nearly all banana fields are
[142]
1203
Name Frequency
Ageratrum conyzoides 57
Oxalis corniculata 45
Galinsoga parviflora 39
Oxalis latifolia 38
Euphorbia heterophylla 33
Conyza sumatrensis 32
Amaranthus hybridus ssp. hybridus 31
Tagetes minuta 22
Galinsoga quadriradiata 20
Solanum nigrum 20
Richardia scabra 14
Nicandra physalodes 12
Lantana camara 8
Tridax procumbens 8
Adiantum raddianum 7
Euphorbia hirta 7
Bryophyllum pinnatum 5
Lepidium bonariense 5
Senna occidentalis 5
Stachytarpheta jamaicensis 5
Galium aparine 4
Ipomoea hederifolia 4
Lagascea mollis 4
Lantana camara 4
Sonchus oleraceus 4
Acanthospermum hispidum 3
Senna bicapsularis 3
Apium leptophyllum 2
Argemone mexicana 2
Caesalpinia decapetala 2
Chenopodium ambrosioides 2
Tradescantia zebrina 2
Acmella oleracea 1
Brassica juncea 1
Crambe hispanica 1
Datura stramonium 1
Ipomoea nil 1
Sambucus nigra 1
Senna septentrionalis 1
Solanum seaforthianum 1
Spermacoce assurgens 1
covered by at least some trees (Figure 7); mean coverage of the tree layer is
28%, mean species number 4.5.
Twenty-nine species were found growing epiphytically in the plots (Table 4).
Twenty-two were holo-epiphytes, mainly restricted in their occurrence to the
epiphyte layer of the forests on Kilimanjaro (cp. Hemp 2001b). Most wide-
[143]
1204
Figure 5. Profile (27 · 2.5 m) and ground plan (27 · 5 m; bold lines indicate the area used for the
profile) of a typical Chagga homegarden in Kidia (Old Moshi) at 1400 m a.s.l. Exposition: south
west, inclination: 25. An open light upper canopy is formed by Albizia schimperiana var. amani-
ensis, on which epiphytes such as the fern Drynaria volkensii and Telphairia pedata, a liana with oil-
containing seeds, find habitats. Bananas form a dense upper shrub layer of 4–6 m height, coffee
trees a lower shrub layer of 1.5–2 m, intermingled with 1–1.5 m high Coco Yam (Colocasia
esculenta). The lower side of the banana field borders a road; here Dracaena fragans is planted as a
hedge.
[144]
1205
Figure 6. Growth form spectrum of the Chagga homegardens; (a) species number of the respective
stratum in the vegetation plots; (b) species number of all representatives of a growth form, e.g. of
trees including young trees occurring in the shrub and herb layer or e.g. of herbs excluding young
trees etc.
agricultural crop plants such as three Dioscorea and Passiflora species and the
Cucurbitaceae Telphairia pedata belonging to this growth form.
One hundred twenty-six species were encountered in the shrub layer of the
plots. Excluding young trees, 63 species remain (Table 6). Similar to the trees,
epiphytes and lianas most of the shrubs in the Chagga homegardens were forest
species. However, in the shrub layer the most important cultivated plants oc-
curred: Different varieties of Musa · sapientium (dessert bananas) and
M. · paradisiaca (cooking bananas) and Coffea arabica. Bananas form a dense
(mean cover value 50%) upper shrub layer of about 4–6 m height and coffee
trees a lower layer of 1.5–2 m. Already Volkens (1897) and Widenmann (1899)
reported that the bananas on Kilimanjaro are most luxuriant with heights of 6–
8 m in the area of Kibosho and Kilema. This may be due to the fact that the
bedrocks in these areas of the wet central southern slope consist of rhomb
porphyry instead of porous tuff and ashes as in the adjacent regions of the
southern slope (Downie and Wilkinson 1972) and that the eastern slope
receives less precipitation.
Four hundred nine species were found in the herb layer of the homegardens.
Excluding young trees, shrubs and lianas 304 herbs and grasses occurred in the
plots. In contrast to the other strata the main species group of the herb layer
consisted of representatives of ruderal vegetation. This is shown in Figure 8. In
this figure species of grasslands and wet habitats are lumped together into
‘semi-natural’ species, as such habitats are mostly influenced by cutting and
grazing of cattle. It appears that the tree and epiphyte layer are the most
[145]
1206
Name Frequency
Grevillea robusta t 37
Albizia schimperiana 36
Persea americana f 21
Rauvolfia caffra 18
Cordia africana 14
Mangifera indica f 13
Commiphora eminii ssp. zimmermannii 11
Margaritaria discoidea 9
Markhamia lutea 8
Bridelia micrantha 7
(Citrus aurantium) f 7
Croton macrostachyus 7
Albizia petersiana 6
Alangium chinense 5
Ficus sur 5
Casearia battiscombei 4
Ficus thoningii 4
Jacaranga mimosiaefolia t 4
Milicia excelsa 4
Syzygium cumini f 4
Turraea robusta 4
Albizia gummifera 3
(Annona muricata) f 3
Annona reticulata f 3
Artocarpus heterophyllus f 3
(Azaridachta indica) x 3
(Cassia siamea) o 3
Cassia spectabilis o 3
(Citrus limon) f 3
Cupressus lusitanica t 3
Cussonia holstii 3
Eriobotrya japonica f 3
Ficus vallis-choudae 3
Olea capensis ssp. welwitschii 3
Trema orientalis 3
Croton megalocarpus 2
(Deinbollia kilimandscharica) 2
Ficus lutea 2
Prunus africana 2
Trichilia emetica 2
(Allophylus abyssinicus) 1
(Bersama abyssinica) 1
(Bombacopsis glabra) 1
Ceiba pentandra x 1
Celtis africana 1
(Celtis mildbraedii) 1
Cinnamomum verum x 1
(Citrus paradisi) f 1
Citrus sinensis f 1
[146]
1207
Table 3. Continued.
Name Frequency
(Clausena anisata) 1
Cordia goetzei 1
Diospyros abyssinica 1
(Ehretia cymosa) 1
Ekebergia capensis 1
Erythrina abyssinica 1
Eucalyptus spec. t 1
(Euclea divinorum) 1
Ficus exasperata 1
Kaya anthotheca t 1
Lecaniodiscus fraxinifolius 1
((Lepidotrichilia volkensii)) 1
(Macaranga capensis) 1
(Maesopsis eminii) t 1
Manihot glaziovii x 1
(Maytenus undata) 1
Melia azedarach x 1
(Mystroxylon aethiopicum) 1
(Neoboutonia macrocalyx) 1
((Ochna holstii)) 1
(Olea europaea ssp. africana) 1
Phoenix reclinata 1
(Polyscias fulva) 1
Psidium guajava f 1
Rapanea melanophloeos 1
((Sorindeia madagascariensis)) 1
Synadenium cf. volkensii 1
Syzygium cordatum 1
Syzygium guineense 1
(Tabernaemontana stapfiana) 1
(Tabernaemontana ventricosa) 1
Teclea nobilis 1
(Vepris spec.) 1
f = fruit tree; o = ornamental tree; t = timber tree; x = introduced tree of other uses; no label:
indigenous tree; in brackets = only found in the shrub layer; in two brackets = only found as
young tree in the herb layer.
natural vegetation strata in the Chagga homegardens whereas the herb layer is
dominated by (128) ruderal species followed by 117 forest species and 64
cultivated plants. Plants of the semi-natural grasslands and wet habitats
contributed 51 and 15 species, respectively.
Discussion
[147]
1208
Figure 7. Typical Chagga homegarden in Kidia (Old Moshi) with an open tree canopy and a dense
banana undergrowth.
[148]
1209
Name Frequency
Pleopeltis macrocarpa 28
Asplenium aethiopicum 21
Asplenium theciferum 12
Lepisorus excavatus 8
Drynaria volkensiii 7
Peperomia tetraphylla 6
Drynaria volkensii 5
Rhipsalis baccifera 5
Impatiens walleriana 4
Phragmanthera usuiensis 4
Loxogramme abyssinica 3
Polystachia simplex 3
Erianthemum dregei 2
Loranthaceae spec. 2
Plicosepalus curviflorus 2
Rangaeris amaniensis 2
(Achyranthes aspera) 1
Aerangis amaniensis 1
Asplenium megalura 1
Asplenium sandersonii 1
Asplenium strangeanum 1
Chamaeangis sarcophylla 1
(Didymodoxa caffra) 1
(Dorstenia zansibarica) 1
(Drymaria cordata) 1
Ficus thoningii 1
(Pilea tetraphylla) 1
(Poa schimperana) 1
(Setaria homonyma) 1
in brackets: accidental epiphytes
[149]
1210
Dioscorea lecardii c 21
(Glycine wightii) 13
Dioscorea bulbifera c 12
Passiflora edulis c 8
Telphairia pedata c 7
Passiflora laurifolia c 6
Zehneria scabra 6
Toddalia asiatica 5
Ficus thoningii 5
Smilax anceps 4
Dioscorea alata c 4
Passiflora quadrangularis c 3
Solanecio angulatus 3
Thunbergia alata 3
Vigna spec. 3
Cyphostemma masukuense ssp-masukuense 3
Cyphomandra betacea 3
Clerodendron johnstonii 3
Basella alba 2
Caesalpinia decapetala 2
Cissus oliveri 2
Cyphostemma kilimandscharicum 2
Diplocyclos schliebenii 2
(Ipomoea obscura) 2
Momordica foetida 2
Pentarrhinum insipidum 2
(Stephania abyssinica) 2
(Teramnus labialis ssp. labialis) 2
Vigna membranacea 2
(Adenia gummifera) 1
Adenia spec. 1
(Begonia meyeri-johannis) 1
Clematis brachiata 1
Dahlbegia lactea 1
(Diplocyclos palmatus) 1
Embelia schimperi 1
Helinus mystacinus 1
Ipomoea hederifolia 1
Ipomoea indica 1
Ipomoea nil 1
Landolphia buchananii 1
(Merremia palmata) 1
(Merremia spec.) 1
Mondia whytei 1
Monstera spec. c 1
Oreosyce africana 1
Paullinia pinnata 1
Piper nigrum c 1
(Rourea thomsonii) 1
Rubus rosifolius 1
Sechium edule c 1
Solanum seaforthianum c 1
c = cultivated; in brackets = only found in the herb layer.
[150]
1211
Name Frequency
[151]
1212
Table 6. Continued.
Name Frequency
Rubus rosifolius i 1
Sambucus nigra 1
Senna septentrionalis 1
Senna spec. 1
(Sesbania sesban) 1
Solanecio cydonifolius i 1
Solanecio mannii i 1
Sterculia cf. stenocarpa i 1
Tephrosia vogelii i 1
Vernonia lasiopus i 1
Vernonia myriantha i 1
Ximenia americana var. caffra i 1
Ziziphus mucronata i 1
i = indigenous species; in brackets = only found as young shrub in the herb layer.
Figure 8. Forest species, semi-natural species (i.e. species of grasslands and wet habitats), cultivated
and ruderal species (including species of trampled ground) with their share in the different strata of
the Chagga homegardens.
the forest cover. Only in deep inaccessible gorges remnants of floristically very
rich forests can be found, resembling the intermediate forests of the Eastern
Arc. Mts. (a chain of ancient crystalline mountains in East Africa), e.g. of the
Pare and Usambara Mts. Here many species, which were believed to be en-
demic to the Eastern Arc. Mts, were encountered on Kilimanjaro (Hemp 2002).
These findings suggest a rich forest flora inhabiting the southern slopes of Mt.
Kilimanjaro in former times. Thus, the lower diversity of the forest flora and
[152]
1213
Figure 9. Records of Christella dentata on Mt. Kilimanjaro, at the base of the UTM grid. The scale
of the squares is 4 km2. This apophytic fern, which thrives naturally in riverine forests was able to
spread over the whole submontane banana plantation belt due to the network of over one thousand
furrows. Gaps within the distribution area inside the banana plantations (especially in its eastern
part) are mostly due to lack of data.
the lower share of endemic plant species of Kilimanjaro can be explained by the
widely destroyed submontane (intermediate) forest rather than by the higher
age and greater ecoclimatic stability of the Eastern Arc. Mts. as suggested by
e.g. Rodgers and Homewood (1982) or Fjeldså et al. (1997). This is corrobo-
rated by the fact, that forest inhabitants, which become less affected by forest
devastation like Saltatoria, have similar numbers of endemic forest species in
the submontane and montane zone on Mt. Kilimanjaro (including Mt. Meru)
and the East Usambara Mts. (Hemp, C. unpub. data). Endemic grasshopper
species like Ixalidium sjöstedti, Parepistaurus deses and Altiusambilla modicicrus
have coped with the habitat change from forest to plantations (Hemp and
Hemp 2003).
On the other hand there are forest species, which were directly or indirectly
favoured by the Chagga people. An example of an intentionally dispersed
forest plant is Dracaena fragrans. This shrub or small tree has lost almost all
natural habitats on Kilimanjaro, except very few submontane river gorges (cp.
Hemp in press a). However it is one of the most characteristic species in the
banana plantations, where it is used as a hedge plant (cp. Figure 5). Dracaena
[153]
1214
Figure 10. Tschibo estate at Mweka, southern slope of Kilimanjaro in November 2003. Hundreds
of old trees were felled to grow new coffee varieties that do not need shade.
[150]
1215
the Pare Mountains and Mt. Meru, which shows nearly exact the same floristic
and structural composition (Hemp unpub. data).
The Chagga homegardens maintain not only a high biodiversity, they are an
old and very sustainable way of land use that meets several different demands.
Beside crop production, the sparse tree layer provides people with fire wood,
fodder and timber. But the high demand of wood, low coffee prizes on the
world marked and the introduction of coffee varieties that are sun-tolerant
endanger this effective system (Hemp et al. in press). In some areas of the
mountain (e.g. on the eastern slopes) the trees in the banana fields are very
scattered or already missing. A very bad example in this respect was observed
at Mweka on the central southern slope: Here a large foreign coffee company
felled hundreds of old trees in November 2003 to grow coffee (Figure 10). In
order to reduce the pressure on the forest, it is necessary, to support the tree
planting in the Chagga homegardens with their unique agroforestry system.
Similar to environmental programmes for farmers in the European Union (e.g.
for the protection of wetlands or dry meadows), there should be a programme
that rewards farmers to have a certain share of their land covered by trees. It
can be estimated that a homegarden supplies 1/4 to 1/3 of the fuelwood
requirements of a family (Fernandes et al. 1984). As the banana belt is nearly as
extensive as the forest reserve, this will of course have major effects in terms of
forest protection and the water balance. In combination with new marketing
and farming strategies for growing organic coffee through traditional methods
an advertising campaign should be started especially in European countries
where the awareness of environmental problems is high. A concept with this
idea was already proposed to the OECD (Hemp 2003).
Acknowledgements
References
Beentje H.J. 1994. Kenya Trees, Shrubs and Lianas. National Museums of Kenya, Nairobi.
Braun-Blanquet J. 1964. Pflanzensoziologie. Springer, Wien.
[151]
1216
Brewin D.R. 1965. Kilimanjaro agriculture. Tanganyika Notes Records 64: 115–117.
Clemm V.M. 1963. Agricultural productivity and sentiment on Kilimanjaro. Econ. Bot. 18: 99–121.
Downie C. and Wilkinson P. 1972. The Geology of Kilimanjaro. University of Sheffield.
FAO 1986. Rehabilitation of traditional schemes (Phase I), Tanzania AG: TCP/URT/4522, Mis-
sion Report, Rome.
Fernandes E.C.M., Okting’ati A. and Maghembe J. 1984. The Chagga homegardens: a multistoried
agroforestry cropping system on Mt. Kilimanjaro (Northern Tanzania). Agroforest. Syst. 2: 73–
86.
Fjeldså J., Ehrlich D., Lambin E. and Prins E. 1997. Are biodiversity ‘hotspots’ correlated with
current ecoclimatic stability? A pilot study using the NOAA-AVHRR remote sensing data.
Biodivers. Conserv. 6: 401–422.
FTEA 1952–2003. Flora of Tropical East Africa. Royal Botanic Garden, Kew.
Hammen T.V.D., Mueller-Dombois D. and Little M.A 1989. Manual of Methods for Mountain
Transect Studies. IUBS, Paris.
Hemp A. and Winter J.C. 1999. Ergebnisse der ethnobotanischen Forschung am Kilimanjaro.
Bayreuther Forum Ökologie 64: 117–144.
Hemp A. 1999. An ethnobotanical study on Mt. Kilimanjaro. Ecotropica 5: 147–165.
Hemp A. 2001a. Ecology of the pteridophytes on the southern slopes of Mt. Kilimanjaro. Plant
Biol. 3: 493–523.
Hemp A. 2001b. Life form and strategies of forest ferns on Mt. Kilimanjaro. In: Gottsberger G.
and Liede S. (eds), Life forms and Dynamics in Tropical Forests. Dissertationes Botanicae 346,
pp. 95–130.
Hemp A. 2002. Ecology of the pteridophytes on the southern slopes of Mt. Kilimanjaro. Part I:
Altitudinal distribution. Plant Ecol. 159: 211–239.
Hemp A. 2003. Climate impacts and responses on Mount Kilimanjaro. In: Agrawala S. and
Moehner A. (eds), Development and Climate Change in Tanzania: Focus on Mount Kilimanj-
aro. OECD, Paris.
Hemp A. in press a. Continuum or zonation? Altitudinal gradients in the forest vegetation of Mt.
Kilimanjaro Plant Ecol.
Hemp A. in press b. Ecology and altitudinal zonation of pteridophytes on Mt. Kilimanjaro. In
Proceedings of the XVIIth AETFAT Congress 21–26 September 2003.
Hemp A. in press c. The impact of fire on diversity, structure and composition of Mt. Kilimanjaro’s
vegetation. In Proceedings of the GMBA Symposium 19–24 August 2002 in Moshi, Tanzania.
Hemp A., Lambrechts C. and Hemp C. (in press). Global Trends in Africa: The Case of Mt.
Kilimanjaro. UNEP, Nairobi.
Hemp C. and Hemp A. 2003. Saltatoria coenoses of high-altitude grasslands on Mt. Kilimanjaro,
Tanzania (Orthoptera: Saltatoria). Ecotropica 9: 71–97.
Hemp C. in press. The Chagga home gardens – relict areas for endemic Saltatoria species (Insecta:
Orthoptera) on Mt. Kilimanjaro. Biol. Conserv.
Hemp C. 2001. Ethnozoological research on invertebrates on Mt. Kilimanjaro, Tanzania. Eco-
tropica 7: 139–149.
Kress W.J. 1986. The systematic distribution of vascular epiphytes: an update. Selbyana 9: 2–22.
Lind E.M. and Morrison M.E.S. 1974. East African Vegetation. Longman, London.
Mdoe N. and Wiggins S. 1997. Returns to smallholder dairying in the Kilimanjaro region, Tan-
zania. Agr. Econ. 17: 75–87.
Nair P.K.R. 1993. An introduction to agroforestry. Kluwer, Dordrecht.
National Bureau of Statistics and Central Census Office. 2003. 2002 Population and Housing
census. Genaral report. The United Republic of Tanzania, Dar es Salaam.
O’Kting’ati A. and Kessy J.F. 1991. The farming system on Mount Kilimanjaro. In: Newmark
W.D. (ed), The Conservation of Mount Kilimanjaro. The IUCN Tropical Forest Programme,
pp. 71–80.
O’Kting’ati A., Maghembe J.A., Fernandes E.C.M. and Weaver G.H. 1984. Plant species in the
Kilimanjaro agroforestry system. Agroforest. Syst. 2: 177–186.
[154]
1217
Ramsay J.C. 1965. Kilimanjaro – sources of water supplies. Tanganyika Notes Records 64: 92–94.
Rikli M. 1903/1904. Die Anthropochoren und der Formenkreis des Nasturtium palustre DC.
Berichte der Zürcherischen Botanischen Gesellschaft 8: 71–82.
Rodgers W.A. and Homewood K.M. 1982. Species richness and endemism in the Usambara
mountain forests, Tanzania. Biol. J. Linn. Soc. 18: 197–242.
Schmidt P.R. 1989. Early exploitation and settlement in the Usambara Mountains. In: Hamilton
A.C. and Bensted-Smith R. (eds), Forest Conservation in the East Usambara Mountains Tan-
zania. The IUCN Tropical Forest Programme, pp.75–78.
Simmonds N.W. 1966. Bananas, 2nd edn. London.
Sukopp H. and Kowarik I. 1987. Der Hopfen (Humulus lupulus L.) als Apophyt der Flora Mit-
teleuropas. Natur und Landschaft 62: 373–377.
Sukopp H. and Langer A. 1996. Campanula rapunculoides – ein Apophyt in der Vegetation
Mitteleuropas. Verhandlungen der Gesellschaft für Ökologie 25: 261–276.
Timberlake L. 1986. Krisenkontinent Afrika. Hammer, Wuppertal.
Volkens G. 1897. Der Kilimandscharo. Darstellung der allgemeineren Ergebnisse eines fünfzehn-
monatigen Aufenthalts im Dschaggalande. Reimer, Berlin.
Widenmann A. 1899. Die Kilimandscharo-Bevölkerung. Anthropologisches und Ethnographisches
aus dem Dschaggalande. Petermanns geographische Mitteilungen, Ergänzungs-Heft 129, 104 pp.
[155]
Biodiversity and Conservation (2006) 15:1219–1252 Springer 2006
DOI 10.1007/s10531-005-0768-6
-1
Abstract. Until recently, patterns of species richness and endemism were based on an intuitive
interpretation of distribution maps with very limited numerical analyses. Such maps based solely on
taxonomic collections tend to concentrate on collecting efforts more than biodiversity hotspots,
since often the highest diversity is found in well-collected areas. During the last decades, there has
been an overwhelming concern about the loss of tropical forest biological diversity, and an
emphasis on the identification of biodiversity hotspots in an attempt to optimise conservation
strategies. Furthermore, the concept of sites of high diversity, or hotspots, has attracted the
attention of conservationists as a tool for conservation priority settings. With the development of
GIS tools, geostatistics, phytosociological and multivariate analysis software packages, more rig-
orous numerical analyses of distributional and inventory data can be used for assessing conser-
vation priorities. In the Campo-Ma’an rain forest, inventory data from 147 plots of 0.1 ha each and
7137 taxonomic collections were used to examine the distribution and convergence patterns of strict
and narrow endemic species. We analysed the trends in endemic and rare species recorded, using
quantitative conservation indices such as Genetic Heat Index (GHI) and Pioneer Index (PI), to-
gether with geostatistic techniques that help to evaluate and identify potential areas of high con-
servation priority. The results showed that the Campo-Ma’an area is characterised by a rich and
diverse flora with 114 endemic plant species, of which 29 are restricted to the area, 29 also occur in
southwestern Cameroon, and 56 others that are also found in other parts of Cameroon. Although
most of the forest types rich in strict and narrow endemic species occur in the National Park, there
are other biodiversity hotspots in the coastal zone and in areas such as Mont d’Eléphant and Massif
des Mamelles that are located outside the National Park. Unfortunately, these areas, supporting 17
strict endemic species that are not found in the park, are under serious threat and do not have any
conservation status for the moment. Taking into consideration that with the growing human
population density, pressure on these hotspots will increase in the near future, it is suggested that
priority be given to the conservation of these areas and that a separate management strategy be
developed to ensure their protection.
[159]
1220
Introduction
Central African rain forests are among the top conservation priority areas
in the world (Davis et al. 1994; Heywood and Watson 1995; Myers et al.
2000). The Campo-Ma’an rain forest, in the southern part of Cameroon,
falls under the Guineo–Congolian Regional Centre of Endemism that is
reported to be species-rich with high levels of endemism (White 1979, 1983;
Davis et al. 1994). It is situated in the middle of the Atlantic Biafran forest
zone that extends from Southeast Nigeria to Gabon and the Mayombe area
in Congo (Letouzey 1968, 1985). The vegetation in the Campo-Ma’an area
is determined by climate, especially rainfall, altitude, soils, proximity to the
sea and human disturbance (Tchouto 2004). The structure and composition
of the forest, as well as the vegetation types change progressively from the
mangrove or coastal forest on sandy shorelines through the endemic low-
land evergreen forest rich in Caesalpinioideae with Calpocalyx heitzii and
Sacoglottis gabonensis, to the submontane forest on hilltops and the mixed
evergreen and semi-deciduous forest in the drier Ma’an area. Other vege-
tation types/sub-types include swamps, seasonally flooded forests, riverine
and secondary forests. The forest in the Ma’an area is described as tran-
sitional between the coastal evergreen forest and the semi-deciduous forest
of the interior.
In view of the rich and diverse flora of the Campo-Ma’an rain forest, as
well as its high level of endemism, it has been identified as one of the key
conservation sites in Cameroon (Gartlan 1989; Foahom and Jonkers 1992).
The Campo-Ma’an area is a Technical Operational Unit (TOU) that com-
prises a National Park, five forest management units, two agro-industrial
plantations, and a multi-uses zone (Tchouto 2004). Despite the low popu-
lation density, there are many stakeholders and different types of land use.
Activities such as logging, industrial and shifting agriculture exert varying
ecological impact on the forest ecosystem. This has led to deforestation,
habitat fragmentation and alteration of the coastal forests. With the
increasing destruction of natural ecosystems, it is important to identify
biodiversity hotspots and conservation priorities in order to enable an
effective management. To achieve this, we need to study the species com-
position and species distribution, so that we can target conservation re-
sources and efforts to rich and diverse areas with a high number of endemic
species. Endemism is commonly regarded as an important criterion for
assessing the conservation value of a given area. In this study, forest
inventory data and taxonomic collections will be used to examine the dis-
tribution and convergence patterns of strict and narrow endemic species. We
will use new quantitative conservation indices such as GHI (Genetic Heat
Index) and Pioneer Index (PI) to analyse trends in endemic and rare species
in the various forest types. Finally, geostatistic analysis and techniques
will help to evaluate and identify potential areas of high conservation
priority.
[160]
1221
Study area
The study was conducted in the Campo-Ma’an rain forest in south Cameroon.
The site covers about 7700 km2 and it is located between latitudes 210¢–
252¢ N and longitudes 950¢–1054¢ E. Following the FAO classification sys-
tem, soils in the Campo-Ma’an area are generally classified as Ferrasols and
Acrisols (Franqueville 1973; Muller 1979; van Gemerden and Hazeu 1999).
They are strongly weathered, deep to very deep and clayey in texture (except at
the seashores and in river valleys where they are mainly sandy), acid and low in
nutrients with pH (H2O) values generally around 4. The topography ranges
from undulating to rolling in the lowland area, to steeply dissect in the more
mountainous areas. In the Campo area, altitudes are mostly low, ranging from
sea level to about 500 m. In the eastern part, which is quite mountainous, the
altitude varies between 400 and 1100 m and the rolling and steep terrain brings
about a more variable landscape.
The area has a typical equatorial climate with two distinct dry seasons
(November–March and July to mid-August) and two wet seasons (April–June
and mid-August to October). The average annual rainfall generally decreases
with increasing distance from the coast, ranging from 2950 mm/year in Kribi
and 2800 mm in Campo to 1670 mm in Nyabissan in the Ma’an area. The
average annual temperature is about 25 C and there is little variation between
years. The hydrography of the area shows a dense pattern with many rivers,
small river basins, fast-flowing creeks and rivers in rocky beds containing many
rapids and small waterfalls. Generally, the area has a low population density of
about 10 inhabitants per square kilometre and is sparsely populated (ca. 61,000
inhabitants) with most people living around Kribi, along the coast, and in
agro-industrial and logging camps (ERE Développement 2002; de Kam et al.
2002). Despite the low population density, there are few employment oppor-
tunities. The local people are very poor and so far rely solely on the forest
resources to meet their basic needs. As a result, local pressure on the Campo-
Ma’an rain forest is increasing and there are several activities that are carried
out in the area with varying ecological impacts on the forest ecosystem. These
activities include agriculture, logging, poaching and hunting.
Botanical assessment
Sampling was carried out between 2000 and 2003 in 147 plots of 0.1 ha
(50 m · 20 m) in representative and homogeneous vegetation types (see Table 6
for an overview of the plots per vegetation type). In each 0.1-ha plot, all trees,
shrubs, herbs and lianas with DBH ‡1 cm (diameter at breast height, about 1.3 m
above ground) were measured, recorded and identified. These plots were estab-
lished in undisturbed forests or matured secondary forests within 12 vegetation
[161]
1222
Table 1. Star categories and GHI weight classes as defined for Cameroon.
Black (BK) 27 Species which are only found in Campo-Ma’an area (strictly
endemic) or near endemics (species which also occur in some
localities around Campo-Ma’an such as Bipindi, Edea-Kribi,
Lolodorf or southern part of Cameroon). Urgent attention to
conservation of population is needed.
Gold (GD) 9 Cameroon endemics, rare and threatened Lower Guinea
endemics. Cameroon has definitely responsibility for preserving
these species.
Blue (BU) 3 Lower Guinea and Guineo–Congolian endemics which are
widespread internationally but rare in Cameroon, or vice versa.
Scarlet (SC) 1 Common but under serious pressure from heavy exploitation.
Exploitation needs to be curtailed if usage is to be sustainable.
Protection of all scales vital.
Red (RD) 1 Common but under pressure from exploitation.
Pink (PK) 1 Common and moderately exploited.
Green (GN) – Widespread Guineo–Congolian, pantropical and tropical
African species that are not under pressure. No particular
conservation concern.
Adapted from Hawthorne and Abu-Juam (1995), Hawthorne (1996) and Tchouto et al. (1998).
types ranging from coastal forest, mangrove, swamp, lowland evergreen forest,
mixed evergreen and semi-deciduous forest to submontane forest at higher ele-
vations (800–1100 m above sea level). Most of the plots were located in the
National Park and the forest management units which are less affected by human
activities. Furthermore, in each representative vegetation type, a provisional
plant species checklist was made in the field with information on their growth
form, guild and frequency. A guild refers to a group associated with a common
way of life (Table 2). For unknown species, a voucher specimen was collected.
The study also involved the collection of fertile specimens encountered in plots,
vegetation types and specific habitats such as exposed rocks and riverbanks. The
geographic co-ordinates of each plot, sample or specimen were recorded using the
Global Positioning System (GPS, Garmin 12XL model with estimated precision
of ±10 m). These co-ordinates were used for mapping main vegetation types,
species distribution, and biodiversity hotspots. A duplicate of each specimen was
mounted and preserved in the Kribi Herbarium. Others duplicates were sent to
the National Herbarium in Yaounde, Cameroon (YA) and the Nationaal Her-
barium Nederland, Wageningen University Branch (WAG) for further identifi-
cation and preservation.
[162]
1223
priority such as endemic, rare, new and threatened species was carried out
using existing floras and monographs (Keay and Hepper 1954–1972; Aubré-
ville and Leroy 1961–1992; Aubréville and Leroy 1963–2000; Lebrun and Stork
2003; Satabié and Leroy 1980–1985; Satabié and Morat 1986–2001), the IUCN
(2002) red list categories, and the WCMC (1998) world list of threatened trees.
On the basis of this information, a list of 141 species of high conservation
values was produced with information on their habit, guild, star category
(Table 1) and chorology. In this list, priority was given to taxa that are strictly
endemic to the Campo-Ma’an area. Followed by species that are endemic to
southwestern Cameroon (also occurring in Bipindi and Lolodorf areas) or
Cameroon and Lower Guinea endemics (especially if they reach their northern
or southern limit of distribution in Campo-Ma’an). Furthermore, species that
reach their northern or southern limit of distribution in the Campo-Ma’an area
were also included in the list.
A star rating system, based on the work of Hawthorne and Abu-Juam (1995)
and Hawthorne (1996) in Ghana, Cable and Cheek (1998) and Tchouto et al.
(1998) in Cameroon, was used to define the conservation status of each species
recorded (Table 1). The factors considered when categorising species into star
categories are their distribution, ecology, local abundance, taxonomy, life
history, interaction with ecosystem parameters and economic importance
(Hawthorne 1996). Therefore, species that are endemic, rare, threatened, or
likely to represent a scarce genetic resource, are more valuable than others are.
Hence, forests richer in such species receive a higher score than others.
The GHI concept was developed by Hawthorne (1996) to express the con-
servation value of a given forest, and the PI concept to express the level of
[163]
1224
Table 3. Forest condition classes showing the degree of human disturbance on the natural forest
cover.
[164]
1225
Geostatistical analysis
Conservation indices such GHI and PI are likely to vary throughout a region.
Geostatistics (Isaaks and Srivastava 1989; Webster and Oliver 2001) were applied
to quantify the spatial distribution of GHI within the Campo-Ma’an forest.
Geographic analyses were done using ILWIS software (ILWIS 2001) and
GSTAT package (Pebesma and Wesseling 1998) of R software (R Development
Core Team 2002). The semivariance was calculated for GHI data on a minimum
lag distance of 1250 m and each lag distance class contained at least 105 pairs of
points. The semivariogram parameters (nugget, sill and range) were computed
using the GSTAT fit variogram function. During the study of GHI spatial var-
iability, the main objective was to obtain a map from point observations. Since
this also required the estimation of a value at un-visited locations, the technique
commonly used is known as kriging (Isaaks and Srivastava 1989). The semi-
variogram function was then used to extrapolate the GHI values in the Campo-
Ma’an forest at 100 m · 100 m grid, using ordinary kriging. The output map was
reclassified into five classes of conservation value (Hawthorne 1996).
Results
A plant species checklist made of 2297 species of vascular plants, ferns and fern
allies was generated from inventory data and from 2348 herbarium specimens and
4789 ecological specimens collected in the various plots. They belonged to 851
genera and 155 families. More than 67% of the specimens were identified at species
level, 28% at generic level, 4% at family level and 1% remained unidentified. The
20 most important families and genera are shown in Tables 4 and 5. In terms of
growth form, tree species contributed for 26% to the total number of 2297 species
recorded, followed by herbs (24%), shrubs (23%) and climbers (17%), respec-
tively. About 72% of the total number of species recorded was also found in the
Campo-Ma’an National Park and the remaining 28% were only found in the
coastal forest and the semi-deciduous forests located outside the park.
In addition to a list of 92 threatened species (Appendix 2) recorded in IUCN
(2002) and WCMC (1998), a list with 141 plant species of high conservation
priorities was produced, with information on their growth forms, guild, cho-
rology and star categories (Appendix 1). Only species that are endemic to
Cameroon and species that reach their northern or southern limit of distri-
bution are included in this list. The Campo-Ma’an area has about 114 endemic
species, 29 of which are only known from the area, 29 only occur in the
southwestern part of Cameroon, and 56 near endemics that also occur in other
parts of Cameroon (Appendix 1 and Figure 1).
Shrubs contributed for 38% of the 114 endemic species (Appendix 1), herbs
29%, trees 20% and climbers 11%. Moreover, 540 species (23% of the total
[165]
1226
[166]
1227
Figure 1. Distribution of 114 strict and narrow endemic plant species recorded in the Campo-
Ma’an area (gray circle). Black circle represents the distribution of 17 threatened strict endemics
that are not found in the National Park. The size of the circle represents the relative density of
endemics at a given point.
More than 57% of the plots have a high GHI score with the highest score
recorded in the submontane forest (GHI = 294.4) and the lowest score in
mangrove (GHI = 3.1). As shown in Figure 2, the submontane forest had the
highest average GHI score of 214.7, followed by the lowland evergreen forest
rich in Caesalpinioideae with Calpocalyx heitzii and Sacoglottis gabonensis
[167]
1228
Figure 2. The average Genetic Heat Index (GHI = bars) and average Pioneer Index (PI = line)
for the various vegetation types as defined in Table 6.
(GHI = 194.1). The mangrove and the coastal forest on sandy shorelines had
the lowest average GHI score (GHI = 3 and 120.2, respectively). The average
PI was very high in the mangrove forest (PI = 125), coastal forest on sandy
shorelines (PI = 66.9) and in the forest rich in Aucoumea klaineana (PI = 60).
Generally, there was a significant decrease in average GHI with increasing
average PI (Figure 2). As shown in Figure 3, there was a very strong significant
negative correlation between the average GHI scores and the PI scores
recorded in the various vegetation types (F1–10 = 111.71, R2 = 0.918,
p < 0.0001). However, the correlation was rather weak with a low explanatory
factor when the analysis was carried out using all plots as individual data
points (F1–45 = 94.00, R2 = 0.393, p<0.0001). Most of the forest types within
the National Park were virtually undisturbed or less than 25% disturbed
(Figure 4). The coastal forest between Campo and Kribi, as well as the forests
around Massif des Mamelles, Mont d’Eléphant, agro-industrial plantations,
logging concessions and settlements were much more affected by human
activities (Figure 4). These forests were often more than 25% disturbed by
human activities and were characterised by a high PI scores (Figures 2 and 4).
Geostatistical results
The analysis of the spatial structure of the dataset did not show any preferential
spatial trend. Therefore, an omni-directional analysis of the semivariance (best
described by a spherical model) was applied. Figure 5 shows the semivariogram
and its characteristics. The GHI variable showed a strong spatial dependence
[168]
1229
Figure 3. Correlation between the average GHI scores and the average PI scores for the various
vegetation types.
within a range of 10,500 m. The nugget (645) was low compared to the total
variance or sill (3700). This suggests that more than 82% (100*(Sill-Nugget)/Sill)
of the semivariance of GHI could be modelled by the variogram over a range of
10 km. The output map of the ordinary kriging (Figure 6) was reclassified into
[169]
1230
Figure 5. Spherical variogram model for GHI in the Campo-Ma’an rain forest (estimated from
147 points of 0.1 ha each).
Figure 6. Ordinary kriging map showing the distribution of GHI scores and conservation hot-
spots within the Campo-Ma’an rain forest The following GHI values are defined for the various
conservation classes (Hawthorne 1996): Very high conservation value for GHI >200; High con-
servation value (150 ‡ GHI <200); Moderate conservation value (100 ‡ GHI <150); Low con-
servation value (50 < GHI <100) and very low conservation value (GHI <50).
[170]
1231
five GHI classes, partitioning the conservation value of the Campo-Ma’an forest.
This partition showed that 1% of the area was characterised by a very high
conservation value, 45% by a high conservation value, 30% by an average
conservation value, 15% by a low conservation value and 9% by a very
low conservation value. A considerable portion of the National Park and the
forests around Massif des Mamelles and Mont d’Eléphant was characterised by a
high conservation value, with highest values found in Dipikar Island, Massif des
Mamelles, Mont d’Eléphant and in the submontane forest on hilltops. The for-
ests in the Ma’an area, around Campo and agro-industrial plantations, near
villages and along the roads had a low conservation value. Similar patterns were
observed for the distribution of strict and narrow endemic species (Figure 1).
Discussion
The Campo-Ma’an area is characterised by a rich and diverse flora with more
than 2297 species of vascular plants, ferns and fern allies. The site has about
114 endemic plant species out of which 29 are strictly endemic to the site. The
number of endemic plant species is relatively high considering the size of the
area, and more than 75% of the current vegetation cover was characterised
[171]
1232
Table 6. Number of plots, number of species and number of stem/ha recorded within the various
vegetation types for all plants with DBH ‡1 cm.
either by very high, high or average GHI values (Figure 6). Furthermore, the
distributions of strict and narrow endemic species showed a high concentration
of these species in the submontane forest between Ebianemeyong and Akom II,
in Dipikar Island, and in the forests in and around Massif des Mamelles, Lobe,
Mont d’Eléphant and Zingui. Surprisingly, the mixed evergreen and semi-
deciduous forest in the Ma’an area showed a relatively low concentration of
these species (Figure 1). The explanation for the high occurrence of endemics
might stem partly from the fact that the area falls within a series of postulated
rain forest refugia in Central Africa (Hamilton 1982; White 1983; Maley 1987,
1989, 1990, 1993, 1996; Sosef 1994, 1996). In such refugia, the unique combi-
nation of climatic and geological histories, contemporary ecological factors,
and inherent biological properties of taxa and their combinations, may have
contributed to survival and/or speciation (Barbault and Sastrapradja 1995;
Hawksworth and Kalin-Arroyo 1995). Furthermore, the Campo-Ma’an area
forms part of the Guineo–Congolian Regional Centre of Endemism (White
1983). All families endemic to this biogeographic region are also found in the
Campo-Ma’an area (White 1983). They include Hoplestigmataceae, Huaceae,
Lepidobotryaceae, Medusandraceae, Pandaceae, Pentadiplandraceae and
Scytopetalaceae. Moreover, 82% of endemic genera cited by White (1983) also
occur in the area.
[172]
1233
Considering the fact that the occurrence of endemic species contributes sig-
nificantly to the conservation value of a forest, it is important to study their
distribution and abundance prior to any conservation initiatives. This is mainly
due to the fact that strict and narrow endemic species are restricted to small
areas, and are therefore highly vulnerable to human disturbance and other
forms of environmental changes (Myers 1988; Williams 1993; Heywood and
Watson 1995). A study carried out in the Campo-Ma’an rain forest has re-
vealed that the submontane forest, the lowland evergreen forest rich in Cae-
salpinioideae with Calpocalyx heitzii and Sacoglottis gabonensis, and the
lowland evergreen forest rich in Caesalpinioideae are richer in strict and nar-
row endemics compared to the other forest types found in secondary forest and
along the coast. This is confirmed by the high average GHI scores recorded in
these forest types (Figure 2). Most of these forest types were located in the
National Park and the lowland evergreen forests around Massif des Mamelles
and Mont d’Eléphant. They were virtually undisturbed or less than 25% dis-
turbed by human activities (Figure 4). This implies that the Massif des
Mamelles and the Mont d’Eléphant areas represent other biodiversity hot-
spots, located outside of the Park (Figure 6).
There was a strong significant negative correlation between the average GHI
scores and the average PI scores recorded in the various vegetation types. Most
plots located near settlements and in secondary forests were characterised by a
low conservation value with low GHI scores and high PI scores. This con-
firmed that disturbed forests are rich in pioneer species but poor in plant
species with high conservation priority. It is worth reiterating that a consid-
erable portion of the Campo-Ma’an area has been selectively logged at least
twice during the past 30 years. Although logging damages were moderate and
had low impact on the total forest biodiversity, it has created forest gaps that
allowed the development of many pioneer species. This might have contributed
to the high average PI scores registered in the coastal forest types. In these
areas with conflicts between human and conservation activities, there is an
urgent need to develop participatory approaches to sustainable natural re-
source management that integrates the objectives of conservation with local
development.
Threatened species
During the selection of species of high conservation priority, taxa were chosen
on a global rather than a Cameroonian or a Campo-Ma’an perspective of
conservation importance. Of the 29 strict endemic species that are only known
from the Campo-Ma’an area, 17 were not recorded in the National Park
illustrating the need for conservation activities outside the park. Although
these 17 strict endemics are not immediately threatened with extinction, the
[173]
1234
most threatened are probably those occurring in the coastal zone and in areas
located at the vicinity of large agro-industrial plantations, since these areas are
heavily exploited. As shown in Figure 4, their habitats are fragmented and
degraded because these areas are surrounded by farms and heavily disturbed
forests.
Considering the fact that extinct species are taxa that are no longer known to
exist in the world after repeated search in their type localities (WCMC 1998;
IUCN 2002), we cannot yet talk about extinction because no attempt has been
made to search for these species. Furthermore, only 67% of the total amount
of specimens collected was identified at species level. However, with the
ongoing speed of forest degradation noticed in the coastal area, eight of these
strict endemics (Beilschmiedia dinklagei, Deinbollia macroura, Ledermanniella
batangensis, Psychotria aemulans, P. batangana, P. dimorphophylla, P. oligo-
carpa, and Strychnos canthioides) that are only known from the coastal zone
can be categorised as endangered species. While the nine others that are located
inland around Efoulan, Fenda, Massif des Mamelles, Mont d’Eléphant and
Zingui can be categorised as vulnerable. They are Afrotrewia kamerunica,
Bulbophyllum alinae, Begonia montis-elephantis, Calvoa stenophylla, Dorstenia
dorstenioides, Guaduella mildbraedii, Hypolytrum sp. nov., Scaphopetalum ac-
uminatum and S. brunneo-purpureum. Some of them so far are only known
from type specimens or from a few collections made in the type locality before
the 60s. Others such as Afrotrewia kamerunica, Begonia montis-elephantis and
Hypolytrum sp. nov. have a restricted range with a small and restricted pop-
ulation. Furthermore, habitat fragmentation may convert a previously more
continuous population structure to a metapopulation structure, with local
populations becoming so small that they may have a substantial risk of
extinction (Hawksworth and Kalin-Arroyo 1995).
[174]
1235
Caesalpinioideae and the mixed evergreen and semi-deciduous forests are also
well represented. So far the National Park is the only area with a legal con-
servation status. It is a permanent state forest that is protected by law and
solely used for forest and wildlife conservation. However, its boundaries have
not been marked, the management plan has not yet been produced and pro-
tection is weak. Therefore, it is of urgent need to demarcate the boundary of
the park, to reinforce its protection, and to complete and implement its
management plan as soon as possible.
[175]
1236
human pressure that has led to the destruction of most of its natural vegetation
(Figure 4). However, it is worth mentioning that some rare endemic species
such as Deinbollia macroura, Psychotria batangana, P. dimorphophylla, P. oli-
gocarpa, and Strychnos canthioides are so far only known from this zone.
Furthermore, there is an impressive network of rivers and streams in the
Campo-Ma’an area that presents a number of very specialised riparian habi-
tats. Our study confirmed that the Lobe, Bongola, Memve’ele waterfalls and
Ntem basin (Boucle du Ntem) support a rich riparian flora with many endemic
and rare rheophile species (Cusset 1987; Thomas and Thomas 1993). Most of
the endemic rheophytes are of the genus Ledermanniella in the Podostemaceae
family. These rheophytes which are found on exposed rocks in streambeds, are
seasonally submerged by fast-flowing water, and normally reproduce in drier
periods when the water level recedes. The Ntem basin is also reported to
constitute an important refuge for wildlife and fish fauna because of the
presence of many rare species of freshwater fishes (Vivien 1991; Matthews and
Matthews 2000; Djama 2001). Therefore, it is suggested to develop a separate
management strategy in order to protect these riparian habitats.
Conclusion
[176]
1237
permanent state forest which is protected by law and should be solely used for
forest and wildlife conservation, its boundaries have not been marked, the
management plan has not yet been produced and protection is weak. It is,
therefore, of urgent need to demarcate its boundary, reinforce its protection,
and complete and implement its management plan as soon as possible. Fur-
thermore, in view of the fact that pressure on these fragmented hotspots is
likely to increase in the future with the growing human population density, it is
suggested that a separate management strategy be developed to ensure the
protection of these biodiversity hotspots and their endemic species.
Acknowledgements
This study was carried out in the framework of the Campo-Ma’an Biodiversity
Conservation and Management Project, Cameroon, and was financially sup-
ported by Tropenbos International, The Netherlands. We will like to thank G.
Achoundong, J.M. Onana, B. Sonke, L. Zapfack and P. Mezili at the National
Herbarium, Cameroon, and F.J. Breteler and C.C.H. Jongkind at the Na-
tionaal Herbarium Nederland, Wageningen University Branch, who assisted in
plant identification. The staff of Campo-Ma’an Project is also acknowledged
with gratitude for their assistance and support during the fieldwork. Particular
thanks are for my field assistants Elad Maurice and Ossele Mathilde for their
enthusiastic support and cooperation. We will also like to extend our sincere
thanks to all chiefs and village representatives, for their active participation in
the organisation and collection of field data.
[177]
Appendix 1 List of 141 plant species that are either strictly endemic to the Campo-Ma’an area (only found in Carapo-Ma’an) or near endemic (also occur in
1238
the western parts of south Cameroon or other parts of Cameroon).
1 Acanthaceae Stenandrium thomense sb GD Hb Cam Akom II, Dipikar Island, Western and South Cameroon
(Milne-Redh.) Vollesen
2 Annonaceae Monanthotaxis elegans sb GD Sh Sw-Cam Akom II, Dipikar Island, Massif des Mamelles, Bipindi and
(Engl. and Diels) Verdc. Lolodorf
3 Annonaceae Monodora zenkeri Engl. sb GD Sh Cam Massif des Mamelles, Bipindi and Lolodorf
and Diels*
4 Apocynaceae Callichilia monopodialis sb GD Sh Cam Ma’an, South, Centre and East Cameroon
(K.Schum.) Stapf*
5 Apocynaceae Landolphia flavidiflora Np GD Lwcl Cam Efoulan, Bipindi, Makak and Mt. Cameroon
(K. Schum.) Persoon*
6 Apocynaceae Petchia africana sb BK Sh Sw-Cam Campo, Bipindi and Lolodorf
Leeuwenb.*
[178]
7 Apocynaceae Tabernaemontana hallei sb GD Sh Lg Northern limit of distribution, from Gabon to Akom II, Ono-
(Boiteau) Leeuwenb. yong and Ma’an
8 Araceae Culcasia bosii Ntepe-Nyame sb BK He Sw-Cam Massif des Mamelles, Dipikar Island, Ma’an and Bipindi
9 Araceae Culcasia panduriformis sb GD Hb Cam Bifa, Zingui, Akom II, Dipikar Island, Bipindi, Mt Cameroon
Engl. and Krause and Eseka
10 Aristolochiaceae Pararistolochia preussii Np GD Swcl Cam Dipikar Island, Ebolowa and Mt. Cameroon
(Engl.) Hutch. & Dalziel
11 Balsaminaceae Impatiens hians Hook.f. var. sb bu Hb Lg Northern limit of distribution, from Gabon to Bipindi, Zingui
bipindensis (Gilg)
Grey- Wilson
12 Balsaminaceae Impatiens gongolana sb bu Hb Lg Northern limit of distribution, from Gabon to Ebianemeyong
N.Hallé
13 Begoniaceae Begonia anisosepala Hook.f. sb bu Hb Lg Northern limit of distribution, from Gabon to Bipindi, Zingui
and Grand Batanga
14 Begoniaceae Begonia clypeifolia Hook.f. sb bu Hb Lg Northern limit of distribution, from Congo, Gabon to Mvini
and Efoulan
15 Begoniaceae Begonia elaeagnifolia Hook. ep bu Ep Lg Northern limit of distribution, from Gabon to Mvini, Efoulan
f. and around Kom River
16 Begoniaceae Begonia heterochroma Sosef sb bu Hb Lg Northern limit of distribution, from Gabon to Lolabe and
around Kribi
17 Begoniaceae Begonia mbangaensis Sosef sb BK Hb Sw-Cam Akom II, Efoulan, Bipindi and Lolodorf
18 Begoniaceae Begonia microsperma Warb. sb GD Hb Cam Ebianemeyong, Ma’an, South-west and South Cameroon
19 Begoniaceae Begonia montis’-elephantis sb BK Hb Campo-Ma’an Rare species, only known from a small population on Mt
J.J.de Wilde* d’Eléphant
20 Begoniaceae Begonia zenkeriana Smith sb BK Hb Sw-Cam Campo, Massif des Mamelles, Dipikar Island, Bipindi and
and Wassh. Lolodorf
21 Burseraceae Aucoumea klaineana Pierre Pi bu Tr Lg Northern limit of distribution, from Gabon to Ma’an and
Ebianemeyong
22 Burseraceae Dacryodes buettneri (Engl.) Np bu Tr Lg Northern limit of distribution, from Gabon to Ma’an and
Lam. Ebianemeyong
23 Capparaceae Ritchiea simplicifolia Oliv. sb BK Sh Cam Lobe, Campo, Kienke, Dipikar Island, Bipindi, Lolodorf and
var. Caloneura (Gilg) Kers Ebolowa
[179]
24 Celastraceae Pristimera luteoviridis (Ex- Np BK Swcl Campo-Ma’an Rare species, only known from few collections on Mt d’Elé-
ell) N.Hallé var. kribiana phant and Dipikar Island
N.Hallé
25 Chrysobalanaceae Dactyladenia cinera (Engl. sb BK Tr Sw-Cam Rare species, only known from type specimens (Bipindi) and a
ex de Wild) Prance and record from Grand Batanga
F.White**
26 Chrysobalanaceae Dactyladenia icondere (Ba- sb bu Sh Lg Northern limit of distribution, from Congo, Gabon to Grand
ill.) Prance and F.White Batanga, Campo and Dipikar Island
27 Combretaceae Combretum cinnabarinum np bu Lwcl Lg Northern limit of distribution, from Gabon to Bipindi and
Engl. and Diels Dipikar Island
28 Cyperaceae Hypolytrum sp. nov. ined.* sb BK Hb Campo-Ma’an New species only known from Mont d’Eléphant
29 Dichapetalaceae Dichapetalum altescandens np bu Lwcl Lg Northern limit of distribution, from Gabon to Efoulan and
Engl. * Zingui
30 Dichapetalaceae Dichapetalum cymulosum np GD Lwcl Cam Grand Batanga, Campo, Bipindi, Lolodorf and Douala
(Oliv.) Engl.*
31 Dichapetalaceae Dichapetalum librevillense np bu Lwcl Lg Northern limit of distribution, from Gabon to Mt d’Eléphant
1239
Pellegr.* and Campo
Appendix 1 (Continued)
1240
No. Family Species Guild Star Habit Chorology Notes
32 Dichapetalaceae Dichapetalum oliganthum np BK Lwcl Sw-Cam Grand Batanga, Campo, Mt d’Eléphant, Kribi, Longi and
Breteler* Lolodorf.
33 Dichapetalaceae Tapura tchoutoi Breteler sb BK Sh Campo-Ma’an Rare species, only known from few collections around Bifa and
Dipikar Island
34 Dryopteridaceae Lastreopsis davalliaeformis sb bu He Lg Northern limit of distribution, from Gabon to Bipindi and
(Tardieu) Tardieu* Zingui
35 Ebenaceae Diospyros alboflavescens sb BK Tr Sw-Cam Rare species, only known from few collections from Bifa,
(Gürke) F.White Zingui and Bipindi
36 Ebenaceae Diospyros soyauxii Gürke sb Bu Tr Lg Northern limit of distribution, from Gabon to Campo and
and K. Schum. Zingui
37 Euphorbiaceae Afrotrewia kamerunica Pax sb BK Sh Campo-Ma’an Rare species, only known from Massif des Mamelles
and Hoffm.*
38 Gnetaceae Gnetum buchholzianum Engl. np GD Hcl Cam Dipikar Island, Onoyong, Ma’an, Littoral, South-west and
[180]
1241
-Caesalpinioideae (Harms) J.Léonard
Appendix 1 (Continued)
1242
No. Family Species Guild Star Habit Chorology Notes
62 Leguminosae Plagiosiphon multijugus sb GD Tr Cam Akom II, Dipikar Island, Ma’an, Bipindi and Kribi-Edea areas
-Caesalpinioideae (Harms) J.Léonard
63 Leguminosae Tetraberlinia moreliana sb bu Tr Lg Northern limit of distribution, from Gabon, Bidou and Mt.
-Caesalpinioideae Aubrév.* D’Eléphant
64 Liliaceae Chlorophytum petrophyllum sb GD Hb Cam Bifa, Dipikar Island, Mvini, Littoral and South Cameroon
K.Krause
65 Loganiaceae Mostuea neurocarpa Gilg sb bu Sh Lg Northern limit of distribution, from Gabon to Bifa, Campo and
Dipikar Island
66 Loganiaceae Strychnos canthioides Leeu- np BK Lwcl Campo-Ma’an Rare species, only known from few collections around Grand
wenb. * Batanga and Lolabe
67 Loganiaceae Strychnos elaeocarpa Gilg ex ri GD Tr Cam Akom II, Dipikar Island, Ebianemeyong, Onoyong, Bipindi,
Leeuwenb. Lolodorf, Kribi-Edea and South-west Cameroon
68 Loganiaceae Strychnos mimfiensis Gilg ex np GD Lwcl Cam Dipikar Island, Mvini, Ma’an, Bipindi, Masok, Douala-Edea-
[182]
1243
Appendix 1 (Continued)
1244
No. Family Species Guild Star Habit Chorology Notes
90 Orchidaceae Vanilla africana Lindley np BK Hcl Sw-Cam Campo, Massif des Mamelles, Mt d’Eléphant and Bipindi
subsp. cucullata
(Kraenzlin and K. Shum.)
Szlachetko and Olszewski *
91 Podostemaceae Ledermanniella annithomae rh BK Hb Campo-Ma’an Rare species, only known from Memve’ele water falls
C. Cusset*
92 Podostemaceae Ledermanniella batangensis rh BK Hb Campo-Ma’an Rare species, only known from Lobe water falls
(Engl.) C. Cusset*
93 Podostemaceae Ledermanniella bosii C.Cus- rh BK Hb Campo-Ma’an Rare species, only known from the Ntem Basin, Bongola, Lobe
set and Memve’ele waterfalls
94 Podostemaceae Ledermanniella boumiensis rh bu Hb Lg Northern limit of distribution, from Gabon to the Bongola and
C. Cusset Memve’ele water falls
95 Podostemaceae Ledermanniella kamerunensis rh BK Hb Campo-Ma’an Rare species, only known from the Bongola water falls in
[184]
111 Rubiaceae Psychotria aemulans sb BK Sh Campo-Ma’an Rare species, only known from few collections around Grand
K. Schum.** Batanga
112 Rubiaceae Psychotria batangana sb BK Sh Campo-Ma’an Rare species, only known from few collections around Grand
K. Schum.* Batanga
113 Rubiaceae Psychotria camerunensis sb GD Sh Cam Akom II, Bifa, Ma’an, Bipindi, Lolodorf, Centre and South
Petit Cameroon
114 Rubiaceae Psychotria dimorphophylla ri BK Sh Campo-Ma’an Rare species, only known from few collections from Grand
K. Schurm.* Batanga and Lobe
115 Rubiaceae Psychotria lanceifolia sb BK Sh Sw-Cam Rare species, only known from Akom II, Onoyong, Bipindi and
K.Schum. Lolodorf
116 Rubiaceae Psychotria oligocarpa sb BK Sh Campo-Ma’an Rare species, only known from few collections around Grand
K.Schum.* Batanga
117 Rubiaceae Psychotria sadebeckiana sb GD Sh Cam Akok, Bifa, Campo, Dipikar Island, Kom, Mvini and South
K.Schum.var. elongata Petit Cameroon
118 Rubiaceae Psychotria sadebeckiana sb GD Sh Cam Akom II, Dipikar Island, Massif des Mamelles, Mvini, and
K.Scham. var. sadebeckiana South Cameroon
1245
Appendix 1 (Continued)
1246
No. Family Species Guild Star Habit Chorology Notes
119 Rubiaceae Tricalysia amplexicaulis sb GD Sh Cam Dipikar Island, Massif des Mamelles, Centre and South Cam-
Robbr. eroon
120 Rubiaceae Tricalysia talbotii sb GD Sh Cam Ebianemeyong, Mvini, Centre and South Cameroon
(Wemham) Keay
121 Rubiaceae Vangueriella laxiflora sb GD Swcl Cam Mvini, Nkoelon, Centre and South Cameroon
(K.Schum.) Verdc.
122 Sapindaceae Deinbollia macroura Gilg ex sb BK Sh Campo-Ma’an Rare species, only known from few collections around Campo
Radlkofer*
123 Sapindaceae Deinbollia mezilii D.W.Tho- sb BK Sh Campo-Ma’an Rare species, only known from Bifa, Massif des Mamelles and
mas and D.J.Harris Dipikar Island
124 Sapindaceae Deinbollia pycnophylla Gilg sb bu Sh Lg Northern limit of distribution, from Gabon to Dipikar Island
ex Radlk.
125 Scytopetalaceae Pierrina zenkeri Engl. sb GD Sh Cam Bifa, Campo, Ebianemeyong, Ma’an, Nyabissan, Littoral and
[186]
South Cameroon
126 Scytopetalaceae Rhaptopetalum sessilifoIium sb BK Sh Sw-Cam Rare species, only known from few collections around Efoulan
Engl.* and Bipindi
127 Sterculiaceae Cola fibrillosa Engl. and sb BK Tr Sw-Cam Rare species, only known from few collections around Dipikar
Krause Island and Bipindi
128 Sterculiaceae Cola letouzeyana Nkongm. sb GD Sh Cam Akora II, Dipikar Island, Ebianemeyong, Onoyong, Centre and
South Cameroon
129 Sterculiaceae Cola praeacuta Brenan and sb GD Sh Cam Bifa, Dipikar Island, Massif des Mamelles, South and South-
Keay west Cameroon
130 Sterculiaceae Scaphopetalum acuminatum sb BK Sh Carapo-Ma’an Rare species, only known from few collections from Efoulan
Engl. and K. Krause* and Fenda
131 Sterculiaceae Scaphopetalum brunneo-pur- sb BK Sh Campo-Ma’an Rare species, only known from few collections from Fenda and
pureum Engl. and K. Kra- Zingui
use**
132 Sterculiaceae Scaphopetalum zenkeri sb BK Sh Sw-Cam Akom II, Dipikar Island, Ebianemeyong, Bipindi and Lolodorf
K-Schum.
133 Thymelaeaceae Dicranolepis glandulosa sb GD Sh Cam Akom II, Dipikar Island, Grand Batanga, Campo, Littoral
H.H.W.Pearson South, and South-west Cameroon
134 Urticaceae Urera gravenreuthi Engl. pi GD Hcl Cam Dipikar Island, Ma’an, Littoral, South and South-west Cam-
eroon
135 Violaceae Allexis zygomorpha sb BK Sh Cam Coastal forest between Edea and Campo, Bidou, Akok, Longi,
Achoundong and Onana* Bipindi and Lolodorf
136 Violaceae Rinorea campoensis sb BK Sh Campo-Ma’an Rare species, only known from Campo, Dipikar Island, Lobe
M. Brandt ex Engl. and Massif des Mamelles
137 Violaceae Rinorea microglossa Engl.* sb BK Sh Sw-Cam Efoulan, Bipindi, Lolodorf, Centre and South Cameroon
138 Violaceae Rinorea sp. nov. 1 ined.* sb GD Sh Cam Coastal forest between Kribi and Campo, Dipikar Island, and
Douala-Edea-Kribi regions
139 Violaceae Rinorea sp. nov. 2 ined.* sb GD Sh Cam Kienke, Massif des Mamelles, Dipikar Island, Kribi, Kribi-
Edea, Douala- Yaounde, and Eseka regions
140 Zingiberaceae Aulotandra kamerunensis Sb BK Hb Sw-Cam Rare species, only known from few collections from Ebia-
Loes. nemeyong, Nyabissan and Bipindi
141 Zingiberaceae Renealmia densispica Sb BK Hb Sw-Cam Rare species, only known from few collections from Dipikar
[187]
Those species that reach their northern or southern limit of distribution in the Campo-Ma’an area are also included in the list.
*Species strictly endemic to the Campo-Ma’an area that were not recorded in the National Park.
**Species for which the status or range needs more investigation.
Guild: ep, epiphyte; np, non-pioneer light demanding; pi, pioneer; rh, rheophyte; ri, riverine; sb, shade-bearer; and sw, swamp.
Star: as defined in Table 1.
Habit: Ep, epiphyte; Hb, herb; Hcl, herbaceous climber; He, hemi-epiphyte; Lwcl, large woody climber; Swcl, small woody climber; Pa, parasite; Sh, shrub;
and Tr, tree.
Chorology: Campo-Ma’an, strict endemic to Campo-Ma’an; Sw-Cam, endemic to southwestern part of Cameroon; Cam, endemic to Cameroon; Lg, Lower
Guinea endemic (especially those species that reach either their northern or southern limit of distribution in the Campo-Ma’an area).
1247
Appendix 2 IUCN (1994) threat categories for 92 plant species recorded in the Campo-Ma’an area that are listed in The IUCN (2002) Red List of Threatened
1248
Species and The World List of Threatened Trees (WCMC 1998).
1249
63 Meliaceae Entandrophragma utile (Dawe and Sprague) Sprague np Tr Gc VU A1cd
64 Meliaceae Guarea cedrata (A.Chev.) Pellegr. np Tr Gc VU A1c
Appendix 2 (Continued)
1250
No. Family Species Guild Habit Chorology IUCN/WCMC
65 Meliaceae Guarea thompsonii Sprague and Hutch. np Tr Gc VU A1c
66 Meliaceae Khaya anthotheca (Welw.) C. DC. np Tr Gc VU A1cd
67 Meliaceae Khaya ivorensis A.Chev. np Tr Gc VU A1cd
68 Meliaceae Lovoa trichilioides Harms np Tr Gc VU A1cd
69 Meliaceae Turraeanthus africanus (Welw. ex C DC.) Pellegr. sb Tr Gc VU A1cd
70 Moraceae Milicia excelsa (Welw.) C.C.Berg pi Tr Tra LR/nt
71 Myrtaceae Eugenia kameruniana Engl. sb Sh Cam CR A1c
72 Ochnaceae Lophira alata Banks ex Gaertn.f. pi Tr Gc VU A1cd
73 Ochnaceae Testulea gabonensis Pellegr. np Tr Lg EN A1cd
74 Rhizophoraceae Anopyxis klaineana (Pierre) Engl. np Tr Gc VU A1cd
75 Rubiaceae Hallea stipulosa (DC.) Leroy sw Tr Gc VU A1cd
76 Rubiaceae Nauclea diderrichii (De Wild. And T.Durand) Merrill pi Tr Gc VU A1cd
77 Rutaceae Vepris heterophylla Letouzey sb Sh Gc EN A1c, B1 + 2c
[190]
References
Aubréville A. and Leroy J.-F. (eds) 1961–1992. Flore du Gabon. Muséum National d’Histoire
Naturelle, Paris.
Aubréville A. and Leroy J.-F. (eds) 1963–2000. Flore du Cameroun, vol. 1–20. Muséum National
d’Histoire Naturelle, Paris.
Barbault R. and Sastrapradja S. 1995. Generation, maintenance and loss of biodiversity. In
Heywood V.H. and Watson R.T. (eds), Global Biodiversity Assessment. Cambridge University
Press, UNEP, pp. 192–294.
Cable S. and Cheek M. 1998. The plants of Mount Cameroon. A conservation checklist. Royal
Botanic Gardens, Kew.
Cusset C. 1987. Flore du Cameroun. In: Satabié B. and Morat Ph. (eds), Podostemaceae, vol. 30.
Muséum National d’Histoire Naturelle, Paris.
Davis S.D., Heywood V.H. and Hamilton A.C. 1994. Centres of Plant Diversity. A Guide and
Strategy for their Conservation. WWF, IUCN.
de Kam M., Fines J.-P. and Akogo G. (eds) 2002. Schéma Directeur pour le developpement de
l’Unité Technique Opérationnelle de Campo-Ma’an, Cameroun. Campo-Ma’an Series 1, .
Djama T. 2001. Inventaire quantitatif des poissons dans l’UTO Campo-Ma’an. Projet Campo-
Ma’an, Kribi, Cameroon.
ERE Développement 2002. Etude socio-économique dans l’UTO de Campo-Ma’an. Rapport final
phase 2: résultats d’énquetes auprés des ménages. SNV/Projet Campo-Ma’an.
Foahom B. and Jonkers W.B.J. 1992. A Programme for Tropenbos Research in Cameroon. Final
report. Tropenbos-Cameroon Programme (Phase I). The Tropenbos Foundation, Wageningen.
Franqueville A. 1973. Atlas regional Sud-Ouest 1. République du Cameroun. ORSTOM, Yaoundé,
Cameroun.
Gartlan J.S. 1989. La conservation des ecosystemes forestiers du Cameroun. IUCN.
Hamilton A.C. 1982. Environmental History of Africa: A Study of the Quaternary. Academic
Press, London.
Hawks worth D.L. and Kalin-Arroyo M.T. 1995. Magnitude and distribution of biodiversity. In:
Heywood V.H. and Watson R.T. (eds), Global Biodiversity Assessment. UNEP, Cambridge
University Press, Cambridge, pp. 108–191.
Hawthorne W.D. 1996. Holes and the sums of parts in Ghanian forest: regeneration, scale and
sustainable use. Proc. R. Soc. Edin. 104: 75–176.
Hawthorne W.D. and Abu-Juam M. 1995. Forest protection in Ghana with particular reference
to vegetation and plant species. The IUCN Forest Conservation Programme.
Heywood V.H. and Waston R.T. 1995. Global Biodiversity Assessment. UNEP, Cambridge
University Press, Cambridge.
ILWIS 2001. ILWIS 3.0 Academic: User’s Guide. ITC, Enschede.
Isaaks E.H. and Srivastava R.M. 1989. An Introduction to Applied Geostatistics. Oxford Uni-
versity Press, New York.
IUCN 1994. IUCN Red List Categories. IUCN, Gland.
IUCN 2002. 2002 IUCN Red list of Threatened Species. IUCN Conservation Monitoring Centre,
Cambridge.
Keay R.W.J. and Hepper F.N. (eds) 1954. Flora of West Tropical Africa. 2nd edn., 3 vols. Crown
Agents, London.
Lebrun J.P. and Stork A.L. 2003. Tropical African flowering plants-ecology and distribution,
vol. 1. Annonaceae-Balanitaceae. Conservatoire et Jardin Botanique de la Ville de Genéve.
Letouzey R. 1968. Etude phytogeographique du Cameroun. Le Chevalier, Paris.
Letouzey R. 1985. Notice de la carte phytogéographique du Cameroun, vol. 1–5. Institut de la
Carte Internationale de la Végétation, Toulouse.
Maley J. 1987. Fragmentation de la forêt dense humide Africaine et extention des biotopes mon-
tagnards du Quaternaire recent. Paleoecol. Afr. 18: 307–309.
[191]
1252
Maley J. 1989. Late Quaternary climatic changes in the African rainforest: the question of forest
refuges and the major role of sea surface temperature variations. In Lewen M. and Sarthein M.
(eds), Paleoclimatology and Paleometeolorogy: Modern and Past Patterns of Global Atmo-
spheric Transport. Kluwer, London, pp. 585–616.
Maley J. 1990. Histoire récente de la forêt dense humide africaine: essai sur le dynamisme de
quelques formations forestières. In: Lanfranchi R. and Schwartz D. (eds), Paysage quaternaire de
l’afrique centrale Atlantique pp.367–382.
Maley J. 1993. The climatic and vegetational history of the equatorial regions of Africa during the
upper Quaternary. In: Shaw T., Sinclair P., Andah B. and Okpoko A. (eds), The Archeology of
Africa: Food, Metals and Towns, pp. 43–52.
Maley J. 1996. Le cadre paléoenvironnemental des refuges forestiers africains : quelques données et
hypothéses. In: der Maesen L.J.G., der Burgt X.M. and de Rooy J.M. (eds), The Biodiversity of
African Plants. Kluwer Academic Publishers, Dordrecht, pp. 519–535.
Matthews A. and Matthews A. 2000. Primate Population and Inventory of Large and Medium
Sized Mammals in the Campo-Ma’an Project Area in Southwest Cameroon, including Man-
agement Recommendations. Projet Campo-Ma’an, Kribi, Cameroun.
Muller J.P. 1979. Carte des sols du Cameroun. Atlas de la République Unie du Cameroun.
Myers N. 1988. Threatened biotas ‘hotspots’ in tropical forests. Environmentalist 8: 187–208.
Myers N., Mittermeir R.A., Mittermeir C.G., da Fonseca G.A.B and Kent J. 2000. Biodiversity
hotspots for conservation priorities. Nature 403: 853–858.
Pebesma E.J. and Wesseling C.G. 1998. GSTAT: a program for geostatistical modelling, prediction
and simulation. Comput. Geosci. 24: 17–31.
R Development Core Team 2002. The R Environment for Statistical Computing and Graphics:
Reference Index. The R Foundation for Statistical Computing, Vienna. Version 1.6.2.
Satabié, B. and Leroy J.-F. 1980. Flore du Cameroun, vol. 21–28. Muséum National d’Histoire
Naturelle, Paris.
Satabié B. and Morat Ph. 1986–2001. Flore du Cameroun, vol. 29–37. Muséum National d’Histoire
Naturelle, Paris.
Sosef M.S.M. 1994. Refuge Begonias. Taxonomy, phylogeny and historical biogeography of
Begonia sect. Loasibegonia and sect. Scutobegonia in relation to glacial rain forest refuges in
Africa. PhD Thesis, Wageningen Agricultural University.
Sosef M.S.M. 1996. Begonias and African rain forest refuges : general aspects and recent progress.
In: der Maesen L.J.G., der Burgt X.M. and de Rooy J.M. (eds), The Biodiversity of African
Plants. Kluwer Academic Publishers, Dordrecht, pp. 602–611.
Tchouto P., Pouakouyou D. and Acworth J. 1998. Rapid Botanical Survey methodology. Tech-
nical report to DFID/Mount Cameroon Project. Limbe Botanic Garden and Herbarium.
Tchouto M.G.P. 2004. Plant diversity in a Central African rain forest: implifications for biodi-
versity conservation in Cameroon. Phd Thesis, Wageningen University.
Thomas D.W. and Thomas J.C. 1993. Botanical and ecological survey of the Campo-Ma’an area.
A report to the World Bank, Washington.
Van Gemerden B.S. and Hazeu G.W. 1999. Landscape ecolofical survey (1:100,000) of the Bipindi-
Akom II-Lolodorf region, Southwest Cameroon. Tropenbos-Cameroon Document 1.
Vivien J. 1991. Faune du Cameroun. Guide des mammiféres et poisons. GICAM, Cameroun.
WCMC 1998. The World List of Threatened Trees. World Conservation Press, .
Webster R. and Oliver M.A. 2001. Geostatistics for Environmental Scientists. Wiley & sons,
Chichester.
White F. 1979. The Guineo-Congolian region and its relationships to other phytochoria. Bull. Jard.
Nat. Belg./Bull. Nat. Plant. Belg. 49: 11–55.
White F. 1983. The Vegetation of Africa. UNESCO, Paris.
Williams P.H. 1993. Measuring more of biodiversity for chosing conservation areas, using taxo-
nomic relatedness. In: Moon T.-Y. (ed.), International Symposium on Biodiversity and Con-
servation. Korean Entomological Institute, Seoul, pp. 199–227.
[192]
Biodiversity and Conservation (2006) 15:1253–1270 Springer 2006
DOI 10.1007/s10531-005-0772-x
-1
Abstract. Species diversity is a function of the number of species and the evenness in the abundance
of the component species. We calculated diversity and evenness profiles, which allowed comparing
the diversity and evenness of communities. We applied the methodology to investigate differences
in diversity among the main functions of trees on western Kenyan farms. Many use-groups (all
trees and species that provide a specific use) could not be ranked in diversity or evenness. No use-
group had perfectly even distributions. Evenness could especially be enhanced for construction
materials, fruit, ornamental, firewood, timber and medicine, which included some of the most
species-rich groups of the investigated landscape. When considering only the evenness in the dis-
tribution of the dominant species, timber, medicine, fruit and beverage ranked lowest (>60% of
trees belonged to the dominant species of these groups). These are also use-groups that are mainly
grown by farmers to provide cash through sales. Since not all communities can be ranked in
diversity, studies that attempt to order communities in diversity should not base the ordering on a
single index, or even a combination of several indices, but use techniques developed for diversity
ordering such as the Rényi diversity profile. The rarefaction of diversity profiles described in this
article could be used in studies that compare results from surveys with different sample sizes.
Introduction
[193]
1254
Study area
All trees (woody perennials) were censused using Beentje (1994) as the key
reference. For each tree species encountered on a farm, its abundance (the total
number of trees) and uses (see below) were recorded by participatory interviews
with household informants involving farm walks, tree counting by the inter-
viewer and data recording on a species-by-species basis.
[194]
1255
Households listed all the products or services (uses) that are provided by
the different species encountered on their farm. Free responses were obtained
on tree uses (primary and all secondary uses) that were postcoded during
data entry and checking. In total, 60 use categories were recorded, but
analyses were only conducted for the 12 use categories that occurred on more
than 20% of farms. Because it is possible that some informants could have
forgotten some uses of particular species, information was adjusted
(increasing species–farm-use combinations from 6859 to 7526) by always
including all trees of a species in a use-group if more than 50% of farmers
and minimum five farmers with the species mentioned the use.
Use-groups were defined as all the trees encountered in the survey that
provided one particular type of use. These use-groups were analysed by
matrices with as rows the information collected on a particular farm and as
columns the various species encountered. Each cell of a use-group matrix
provides the number of trees of each species that was used for a particular
function on a particular farm.
The Rényi diversity profile is one of the techniques for diversity ordering that
were specifically designed to rank communities from low to high diversity.
Rényi diversity profile values (Ha) are calculated from the frequencies of each
component species (proportional abundances pi = abundance of species i/
total abundance) and a scale parameter (a) ranging from zero to infinity
(Tóthmérész 1995; Legendre and Legendre 1998) as:
P a
ln pi
Ha ¼
1a
It can be demonstrated that values of the Rényi profile at the respective
scales of 0, 1, 2 and ¥ are related to species richness S, the Shannon diversity
index H, the Simpson diversity index D1 and the Berger–Parker diversity
index d1 (Magurran 1988; Legendre and Legendre 1998; Shaw 2003):
H0 ¼ lnðSÞ
X
H1 ¼ H ¼ pi log pi
X
1
H2 ¼ lnðD1 Þ ¼ ln ðp2i Þ
H1 ¼ lnðd1 Þ ¼ lnðp1
max Þ
[195]
1256
[196]
1257
[197]
1258
Results
Figure 1 shows the Rényi diversity ordering of all trees and the 12 most fre-
quent use-groups. By examining the values at scales 0, 1, 2, and ¥, species
richness and values of Shannon, Simpson, and Berger–Parker diversity indices
can be inferred. The many intersections in the figure show the difficulties to
order most groups in diversity. Figure 2 shows many groups with intersecting
evenness profiles, which is an indication that many groups cannot be ranked in
evenness. Table 1 provides some more precise statistics for certain profile
values than can be inferred from the figures, and also gives some parameters
that describe the pattern of the various diversity profiles (see below).
We differentiated between five pools of use-groups based on their total
species richness (H0) (which is an ordering of use-group only based on rich-
ness): (i) pool A – very high richness, including all trees and firewood; (ii) pool
B – high richness, including shade; (iii) pool C – medium richness, including
Figure 1. Diversity profiles based on the Rényi series Ha for all trees and trees belonging to
particular use-groups.
[198]
1259
Figure 2. Evenness profiles for all trees and trees belonging to particular use-groups.
Table 1. Values for specific profile values of the diversity and evenness profiles depicted in
Figures 1 and 2 with for all trees and for the most frequent use-groups in western Kenya.
All trees 5.16 2.79 2.38 1.78 2.37 0.09 1.61 1.98
Firewood 5.04 2.69 2.29 1.71 2.35 0.10 1.54 1.92
Shade 4.43 3.00 2.60 2.04 1.43 0.24 1.43 3.84
Medicine 4.06 1.59 0.79 0.41 2.47 0.08 0.15 0.77
Ornamental 3.97 1.34 0.94 0.68 2.63 0.07 0.37 1.12
Timber 3.89 1.11 0.77 0.48 2.78 0.06 0.36 0.61
Boundary 3.53 1.95 1.69 1.17 1.58 0.21 0.97 1.44
Soil fertility 3.30 1.44 1.08 0.64 1.86 0.16 0.17 1.57
Charcoal 3.26 1.29 0.95 0.70 1.97 0.14 0.29 1.41
Fruit 3.22 1.08 0.61 0.32 2.13 0.12 0.24 0.41
Construction 3.00 0.66 0.43 0.24 2.33 0.10 0.17 0.32
Fodder 1.95 1.61 1.40 0.87 0.33 0.72 0.44 1.63
Beverage 1.39 0.50 0.38 0.22 0.89 0.41 0.04 0.44
H¥;l and H¥;u are 95% CI limits for H¥. See methods for the formulas.
[199]
1260
Figure 3 shows the average profile values and associated 95% CI for subs-
amples of 47 farms (the number of farms with fodder, the least frequent use-
group). Figure 3 indicates that, on average, similar profiles are obtained as for
the full sample. Diversity profiles, especially medicine and shade, had large
95% CI, however. The 95% CI still allow classifying use-groups in the richness
[200]
1261
Figure 3. Averages of rarefied diversity profiles based on the Rényi series Ha for all trees and trees
belonging to particular use-groups calculated from 1000 random subsamples of 47 farms, with 95%
CI. All profile values were calculated for the same scales, but groups were presented at different
scales for better discrimination.
[201]
1262
Figure 4. Accumulation curve for the Shannon diversity index for all trees and trees belonging to
particular use-groups. The vertical reference indicates the sample size of 47 farms depicted in
Figure 3.
groups where stable values were obtained at the respective scale in Figure 6
(for example timber, fruit, construction, and beverage). However, these CI
were among the smallest. Large CI corresponded to non-linear patterns for the
Berger–Parker index (Figure 6, see discussion above). Medicine which had a
non-linear negative pattern for the Berger–Parker index had a relatively small
CI, however.
Discussion
[202]
1263
Figure 5. Accumulation curve for H2 for all trees and trees belonging to particular use-groups.
The vertical reference indicates the sample size of 47 farms depicted in Figure 3.
[203]
1264
Figure 6. Accumulation curve for H¥ for all trees and trees belonging to particular use-groups.
The vertical reference indicates the sample size of 47 farms depicted in Figure 3.
It has been observed for a long time in ecological research that sample size has
an influence on species richness (e.g., Arrhenius 1921). Since diversity is
influenced by richness, sample size also has an effect on diversity. We were able
to investigate the influence of sample size on diversity by studying the range of
values observed in subsets of the data by using a randomisation approach
similar to the randomisation approach to calculate species accumulation curves
(Kindt et al. 2001). This approach allowed rarefaction to the same number of
sample units (farms in our example), thus removing the effects from differences
in sample size. This approach could be useful in other studies that compare
results obtained from surveys with different sample sizes, especially since the
approach includes a diversity ordering technique which is necessary for an
accurate description of the diversity of a community.
Magurran (1988) indicated that a Rényi series value with larger scale
parameter value has reduced sensitivity to sample size. Gimaret-Carpentier
et al. (1998) observed that the Simpson index reached stable values at lower
sample sizes compared to the Shannon index.
We observed such asymptotic patterns both for H2 and H¥ for those groups
where stable values were obtained for the Shannon index (timber, fruit,
[204]
1265
construction, and beverage). Shade and fodder, however, showed the other
extreme with increasing profile values with increasing sample size, and not
asymptotic values.
Hayek and Buzas (1997) proposed to investigate accumulation patterns of
H, ln(E) and ln(E)/ln(S) to choose the best abundance distribution model
(‘SHE analysis’). Reaching asymptotic values for these statistics with
increasing sample size would indicate that the species abundance distribution
corresponds to respectively a log-series, a broken-stick or log-normal
distribution. Based on our results, we could therefore conclude that most use-
groups (and especially timber, fruit, construction and beverage) have abun-
dance distributions more typical of the log series distribution (asymptotic H),
while the distributions of shade and fodder corresponded more to a log-
normal distribution (asymptotic ln(E)/ln(S), figure not included). (A pre-
liminary analysis found that shade, medicine and ornamental conformed to
the log-normal distribution, whereas all trees, firewood, timber, and bound-
ary did not – for other use-groups, unreliable results were obtained for chi-
square goodness-of-fit tests.) We suggest expanding the SHE analysis to the
complete diversity profile, as in Figure 3. For example, for a pure log-series
(and geometric series), the Simpson and Berger–Parker indices are also
expected to reach constant values (May 1975), which was a phenomenon that
we could observe most clearly for the groups with the clearest asymptotic H1
pattern. Since asymptotic H1 and H2 are only expected for the log-series
distribution, the findings of Gimaret-Carpentier et al. (1998) described above
most likely correspond to studies of systems more conform to this distribu-
tion. For the same reason, the observation of reduced sensitivity to sample
size with increasing scale provided by Magurran (1988) could only apply to
the log-series.
The basic aim of our study was to plan for diversification. The diversity and
evenness profiles were calculated on several use-groups. By differentiating
between several use-groups, use-groups of lower diversity or evenness were
identified and a benchmark dataset was created to measure the impact of future
diversification. By identifying use-groups of lower relative diversity (such as
beverage and construction), priority can be given to these use-groups for
diversification. Such approach can be described as a coldspot approach that
focuses on subregions of lowest diversity in a study area. For example, it would
be less efficient for diversification to add new species to firewood than adding
new species to construction. With the rarefaction procedure, comparisons are
made only for farmers that are already growing trees for a use-group. This
approach should be preferred when there is no scope for increasing the number
of farmers for each use-group: this is a better benchmark to plan diversification
efforts (see below).
[205]
1266
The value of using diversity profiles lays not only in determining which use-
groups are more diverse. Diversity profiles also allow discrimination between
richness and evenness contributions to diversity. If the intervention would
attempt to increase diversity for two groups with intersecting diversity profiles,
then for one group improvement of richness could be attempted, while
improvement of evenness would be the target for the other group.
The lack of evenness in the distribution of the dominant species and the
many steep decreases in profiles with increasing scale parameter value showed
that diversity could be increased substantially in many use-groups by targeting
evenness, rather than targeting richness. No group had perfectly evenly dis-
tributed species. Evenness increment could be achieved by encouraging farmers
to establish trees in more even numbers (influencing the demand for tree
germplasm) or by more species-even germplasm distribution (influencing the
supply of tree germplasm). The analysis shows use-groups with steep diversity
and evenness profiles where such diversity improvements would be most useful:
i.e. the construction, fruit, ornamental (although this group will probably not
constitute a priority to farmers), firewood, timber and medicine groups. When
considering the frequency of the dominant species only (not how evenly this
species is distributed), timber, medicine, fruit and beverage are the groups with
frequencies larger than 60%. Interestingly, these are also the use-groups that
could be categorised as providing more cash income to farmers (‘high value
trees’ that could be selected to be of higher priority for domestication for this
reason).
The analysis of diversity did not include all aspects that could influence
decisions on alterations in tree species composition on farms. Such factors
include potential differences in importance attributed by farmers to different
use-groups so that diversification of a more important but more diverse use-
group could be given priority over that of a less diverse but also less important
use-group. Another aspect is potential differentiation among farms in alpha
diversity, so that farms with low diversity for a specific use-group could be
targeted, rather than only targeting use-groups with lower diversity at the
survey level (Kindt et al. 2004).
Effective diversity planning will require that the relative abundances of the
composing species are analysed within use-groups that were prioritised for
diversification. When planning for increments in evenness, the potential for
increasing the abundances of rare species should be investigated. Where it is
not possible to add new trees to a particular farm or landscape, the potential of
substituting some trees of the dominant species with trees of the rare species
should be explored. Participatory research should investigate why some species
occur in higher numbers than other species: substitution of common species by
rare species may be easier when differences in abundance are not related to
differences in farmer preference, but caused by factors such as differences
between species in natural regeneration, historic promotion by development
agencies, or erosion in local knowledge. Since all species within a particular
use-group are used for that particular purpose, there is a definite potential for
[206]
1267
increasing evenness, but such increments may require balancing diversity with
differences in preference – some species may need to be promoted for the
insurance that they bring at a cost for short-term productivity.
McNeely and Scherr (2002) describe that a new type of agriculture is needed
that leads to increased food security and conservation gains since human
population density and biodiversity are positively correlated in many areas.
Their book provides examples of innovative landscape management strategies
that successfully combined both objectives by applying ecoagriculture strate-
gies. Our study documented that many tree species have been integrated in
farming systems already (conform Ecoagriculture Strategy 4 of mimicking
natural habitats by integrating productive perennial plants). Since careful
scrutiny of species identities of the various use-groups (such detailed study was
beyond the scope for this article, but is a logical next step as described above)
showed that the dominant species in most use-groups were exotic species
(exceptions were boundary demarcation, fodder and medicine), whereas the
majority of rare species were indigenous species, increments of evenness could
result in increments of indigenous tree species in the farming landscape (since
not only exotic species are planted, increasing abundances for indigenous
species entails small relative changes in planting practices by local communi-
ties). Diversification could therefore result in improved conservation, although
the links between development and conservation goals need to be explored
carefully (Adams et al. 2004).
Landscape diversification could also consider population structure and
geneflow of particular species, especially for indigenous species. Where less
frequent species are promoted, interventions should attempt to ensure that
population sizes are large enough to avoid substantial genetic erosion. Diver-
sification planning could for example consider corridors in farmland between
natural populations to reduce genetic erosion (O’Neill et al. 2001). Such con-
siderations of genetic diversity may indicate limits to diversification (not all
species could potentially be maintained at large enough population sizes for a
more even distribution of species) that could be incorporated in planning of
diversification (finding a more even species abundance distribution that avoids
too small population sizes for each species, possibly with a smaller number of
species). Alternatively, species can be maintained at very small population sizes
given that new genetic diversity is regularly introduced from genetically diverse
seed sources.
Ecological reasons for diversification within a use-group could include
minimizing the chances of pest and disease outbreaks. Promotion of single
species for a particular use-group should especially be avoided since several
pest outbreaks on agroforestry species have been experienced after large-
scale promotions of monoculture agroforestry technologies (Atta-Krah et al.
2004). Ecological research has indicated that biodiversity can affect ecosys-
tem function, but that differences in species function are conditions for
positive effects of biodiversity on ecosystem stability and productivity (e.g.,
Hector et al. 1999; Loreau et al. 2001, 2002; Tilman et al. 2001). Natural
[207]
1268
Acknowledgements
References
Adams W.M., Aveling R., Brockington D., Dickson B., Elliott J., Hutton J., Roe D., Vira B. and
Wolmer W. 2004. Biodiversity eradication and the eradication of poverty. Science 306: 1146–
1149.
Arrhenius O. 1921. Species and area. J. Ecol. 9: 95–99.
Atta-Krah K., Kindt R., Skilton J.N. and Amaral W. 2004. Managing biological and genetic
diversity in tropical agroforestry. Agroforest. Syst. 61: 183–194.
Beentje H.J. 1994. Kenya Trees, Shrubs and Lianas. National Museums of Kenya, Nairobi 722 pp.
Belaoussoff S., Kevan P.G., Murphy S. and Swanton C. 2003. Assessing tillage disturbance on
assemblages of ground beetles (Coleoptera: Carabidae) by using a range of ecological indices.
Biodivers. Conserv. 12: 851–882.
Bradley P.N. 1991. Woodfuel, Women and Woodlots, vol. 1. Macmillan Education Ltd, London
and Basingstoke 338 pp.
Bradley P.N., Chavangi N. and Van Gelder A. 1985. Development research and energy planning in
Kenya. Ambio 14: 228–236.
Colwell R.K. 1997. Estimates: Statistical Estimation of Species Richness and Shared Species from
Samples. University of Connecticut, Storrs.
Dougall T.A.G. and Dodd J.C. 1997. A study of species richness and diversity in seed banks and its
use for the environmental mitigation of a proposed holiday village development in a coniferized
woodland in south east England. Biodivers. Conserv. 6: 1413–1428.
[208]
1269
Gimaret-Carpentier C., Pelissier R., Pascal J.P. and Houllier F. 1998. Sampling strategies for the
assessment of tree species diversity. J. Veg. Sci. 9: 161–172.
Gotelli N.J. and Colwell R.K. 2001. Quantifying biodiversity: procedures and pitfalls in the
measurement and comparison of species richness. Ecol. Lett. 4: 379–391.
Hayek L.-A.C. and Buzas M.A. 1997. Surveying Natural Populations. Columbia University Press,
New York xvi, 563 pp.
Hector A., Schmid B., Beierkuhnlein C., Caldeira C., Diemer M., Dimitrakopoulos P.G.,
Finn J.A., Freitas H., Giller P.S., Good J., Harris R., Högberg P., Huss-Danell K., Joshi J.,
Jumpponen A., Körner C., Leadley P.W., Loreau M., Minns A., Mulder C.P., O’Donovan
G., Otway S.J., Pereira J.S., Prinz A., Read D.J., Scherer-Lorenzen M., Schulze E.D.,
Siamantziouras A.D., Spehn E.M., Terry A.C., Troumbis A.Y., Woodward F.I., Yachi S. and
Lawton J. 1999. Plant diversity and productivity in European grasslands. Science 286: 1123–
1127.
Hill M.O. 1973. Diversity and evenness: a unifying notation and its consequences. Ecology 54: 427–
431.
Hubbell S.P. 2001. The Unified Neutral Theory of Biodiversity and Biogeography. Princeton
University Press, Princeton xiv + 375 pp.
ICRAF 1997. ICRAF Medium-Term Plan 1998–2000. International Centre for Research in
Agroforestry, Nairobi 77 pp.
Jaetzhold R. 1982. Farm Management Handbook of Kenya: Natural Resources and Farm Man-
agement Information vol. iia – West Kenya (Nyanza and Western Provinces). Kenya Ministry of
Agriculture, 397 pp.
Kindt R. 2001. RenyiAccum. Program to Calculate Average Values and Ranges for the Rényi
Diversity Series Associated with Random Site Sequences. International Centre for Research in
Agroforestry (ICRAF), Nairobi.
Kindt R. 2004. Biodiversity Analysis Functions for R (Biodiversity.R). ICRAF, Nairobi.
Kindt R., Degrande A., Turyomurugyendo L., Mbosso C., Van Damme P. and Simons A.J. 2001.
Comparing species richness and evenness contributions to on-farm tree diversity for data sets
with varying sample sizes from Kenya, Uganda, Cameroon, and Nigeria with randomized
diversity profiles. In: IUFRO Conference on Forest Biometry, Modelling and Information
Science, 26–29 June 2001. University of Greenwich, UK. URL: http://cms1.gre.ac.uk/confer-
ences/iufro/proceedings/ (last accessed 4–2004).
Kindt R. and Lengkeek A.G. 1999. Tree diversity on farm – use it or lose it. In: National Workshop
on Agricultural Biodiversity Conservation, 27–29 January 1999. Intermediate Technology
Development Group (ITDG, Nairobi, pp 75–85.
Kindt R., Simons A.J. and Van Damme P. 2004. Do farm characteristics explain differences in tree
species diversity among western Kenyan farms? Agroforest. Syst. 63: 63–74.
Legendre P. and Legendre L. 1998. Numerical Ecology. Elsevier Science BV, Amsterdam 853 pp.
Loreau M., Naeem S., Inchausti P., Bengtsson J., Grime J.P., Hector A., Hooper D.U., Huston
M.A., Raffaelli D., Schmid B., Tilman D. and Wardle D.A. 2001. Biodiversity and ecosystem
functioning: current knowledge and future challenges. Science 294: 804–807.
Loreau M., Naeem S. and Inchausti P. 2002. Biodiversity and Ecosystem Functioning: Synthesis
and Perspectives. Oxford University Press, Oxford 304 pp.
Magurran A.E. 1988. Ecological Diversity and Its Measurement. Princeton University Press,
Princeton, N.Jx, 179 pp.
Magurran A.E. and Henderson P.A. 2003. Explaining the excess of rare species in natural species
abundance distributions. Nature 422: 714–716.
May R.M. 1975. Patterns of species abundance and diversity. In: Cody M.L. and Diamond J.M.
(eds), Ecology and the Evolution of Communities. The Belknap Press of Harvard University
Press, Cambridge and London, pp. 81–120.
McGill B.J. 2003. A test of the unified neutral theory of biodiversity. Nature 422: 881–885.
McNeely J.A. and Scherr S.J. 2002. Ecoagriculture: Strategies to Feed the World and Save Wild
Biodiversity. Island Press, Washington 323 pp.
[209]
1270
Mishra B.P., Tripathi O.P., Tripathi R.S. and Pandey H.N. 2004. Effects of anthropogenic dis-
turbance on plant diversity and community structure of a sacred grove in Meghalaya, northeast
India. Biodivers. Conserv. 13: 421–436.
O’Neill G.A., Dawson I.K., Sotelo-Montes C., Guarino L., Current D., Guariguata M. and Weber
J.C. 2001. Strategies for genetic conservation of trees in the Peruvian Amazon basin. Biodivers.
Conserv. 10: 837–850.
Purvis A. and Hector A. 2000. Getting the measure of biodiversity. Nature 405: 212–218.
Ricotta C. 2003. On parametric evenness measures. J. Theor. Biol. 222: 189–197.
Ricotta C. and Avena G.C. 2002. On the information-theoretical meaning of Hill’s parametric
evenness. Acta Biotheor. 50: 63–71.
Rousseau D., Van Hecke P., Nijssen D. and Bogaert J. 1999. The relationship between diversity
profiles, evenness and species richness based on partial ordering. Environ. Ecol. Stat. 6: 211–223.
Shaw P.J.A. 2003. Multivariate Statistics for the Environmental Sciences. Hodder Arnold, London
ix + 233 pp.
Slik J.W.F., Verburg R.W. and Keßler P.J.A. 2002. Effects of fire and selective logging on tree
species composition of lowland dipterocarp forest in East Kalimantan, Indonesia. Biodivers.
Conserv. 11: 85–98.
Taillie C. 1979. Species equitability: a comparative approach. In: Grassle J.F., Patil G.P., Smith
G.K. and Taillie C. (eds), Ecological Diversity in Theory and Practice. International Cooperative
Publishing House, Fairland, pp. 51–62.
Tilman D., Reich P.B., Knops J., Wedin D., Mielke T. and Lehman C. 2001. Diversity and
productivity in a long-term grassland experiment. Science 294: 843–845.
Tóthmérész B. 1995. Comparison of different methods for diversity ordering. J. Veg. Sci. 6: 283–
290.
Van Noordwijk M. and Ong C.K. 1999. Can the ecosystem mimic hypotheses be applied to farms
in African savannahs? Agroforest. Syst. 45: 131–158.
Zavaleta E.S. and Hulvey K.B. 2004. Realistic species losses disproportionately reduce grassland
resistance to biological invaders. Science 306: 1175–1177.
Zilihona I.J.E., Niemela J. and Nummelin M. 2004. Effects of a hydropower plant on Coleopteran
diversity and abundance in the Udzungwa Mountains, Tanzania. Biodivers. Conserv. 13: 1453–
1464.
[210]
Biodiversity and Conservation (2006) 15:1271–1301 Springer 2006
DOI 10.1007/s10531-005-2576-4
-1
Key words: Conservation value, Dipterocarp forests, Gene bank, Molave forest, Native species,
Species richness, Tropical rain forest, Vascular plant species
Abstract. The Philippines are one of the most important biodiveristy hotspots on earth. Due to the
extraordinary rate of environmental destruction, leaving only 3% of the land with primary forest,
this biodiversity is at high risk. Despite that situation information on Philippine forest vegetation is
fragmentary and focused on trees. This study aimed at analysing forest remnants in the Leyte
Cordillera on the Island of Leyte, and at evaluating their role as refuge to the largely destroyed
lowland forest vegetation. A total of 49 plots (100 m2 each) between 55 and 520 m a.s.l. were
studied. All vascular plant species except epiphytes were included. Records include 685 taxa from
289 genera and 111 families, representing nearly 8% of the known Philippine vascular plant species.
More than half (52%) of the species are Philippine endemics. A number of 41 tree species, or 6% of
all taxa recorded, are included in the IUCN red list, either as vulnerable, endangered, or critically
endangered. Life form composition was dominated by phanerophytes (65.3%), followed by lianas
and chamaephytes (17.1 and 16.9%, respectively). The most common families were the Rubiaceae
with 35 and the Euphorbiaceae with 32 species. All five Philippine dipterocarp forest types as well
as the molave forest type were represented by typical tree species. The area provides an important
gene bank of the highly threatened Philippine lowland forest vegetation and is of high value for
biodiversity conservation. Additionally, it can play an important role as seed source of valuable tree
species for the increasing initiatives to rehabilitate and reforest degraded land with native species.
Introduction
The destruction of tropical rain forests is still continuing at high rates (FAO
2003). This process, especially threatens the earth’s biodiversity hotspots such as
the Philippines (Myers et al. 2000; Brooks et al. 2002). Despite this, there are only
very few studies worldwide which aimed at the documentation of the total plant
species richness of such sites. Most inventories were restriced to selected life forms
such as ground herbs (e.g. Kiew 1987; Poulsen and Balslev 1991; Poulsen 1996) or
trees of a defined minimum diameter (e.g. Valencia et al. 1994; Lieberman et al.
1996; Newbery et al. 1996; Rennolls and Laumonier 2000; Slik et al. 2003).
Vascular plant species composition of tropical lowland forests was studied in
Ghana on 0.5- and 1-ha plots by Hall and Swaine (1981), in Amazonia on 0.02-ha plots
by Takeuchi (1960) and on 10 non-contiguous plots of 0.1 ha by Duivenvoorden
[211]
1272
(1994), in Ecuador on 0.1-ha plots of three lowland forest types by Gentry and
Dodson (1987), and in stratified plots with a total area of about 2 ha in Puerto
Rico by Smith (1970). In Southeast Asia, Kochummen et al. (1992) studied the
trees and shrubs (>1 cm diameter at breast height (dbh)) in a 50-ha plot in the
Pasoh Forest Reserve in Malaysia. The most comprehensive study including all
vascular plants as well as mosses was conducted by Whitmore et al. (1985) on a
single 100-m2 plot in the lowland rain forest of Costa Rica. However, no study
representing a complete inventory of vascular plant species richness of any site of
lowland rain forest in Southeast Asia was found.
The Philippines are among the most seriously depleted tropical countries
with only 3% of the land area still covered by primary forest (Myers et al.
2000). From 1990 to 2000, the Philippines lost 1.4%, or 89,000 ha, of the forest
area annually (FAO 2003). At the same time, the Philippine archipelago is one
of the most important biodiversity hotspots on earth (Myers et al. 2000) with
high proportions of endemic plant and animal species (Heaney and Regalado
1998). The endemism rate of plants was estimated to be 39% (Davis et al.
1995), but for certain taxa, it can be much higher. For example, 11 of 12 species
of pitcher plants (Nepenthes spp.) known from the Philippines are endemic
(Cheek and Jebb 2001). Similarly, there are high rates of endemism among the
fauna. Referring to terrestrial vertebrates, 64% of the archipelago’s land
mammals are endemic, as well as 44% of the breeding land birds, 68% of the
reptiles, and ca. 78% of the amphibians (Heaney and Regalado 1998). Most of
them depend on forest ecosystems.
Despite the ecological uniqueness on the one hand and the extensive
destruction on the other, the study of Philippine forest vegetation has been
neglected (Tan and Rojo 1989; Kartawinata 1990; Soerianegara and Lemmens
1994). Much of the current knowledge is still based on studies conducted in the
early 20th century (Whitford 1906; Whitford 1911; Brown and Mathews 1914;
Brown 1919), which were mainly dealing with timber trees under economical
aspects. Recent studies focused on the vegetation of montane and submontane
forest types on different islands. However, in most cases (Aragones 1991;
Pipoly and Madulid 1998; Proctor et al. 1998; Hamann et al. 1999) these were
largely restricted to trees of a defined size, which usually is ‡10 cm dbh. Buot
and Okitsu (1997) only considered woody plants higher than 1.3 m, and Ingle
(2003) those of at least 5 cm dbh. The only data without size limitations are
provided by Gonzales-Salcedo (2001) from Mt. Amuyao, Luzon, at elevations
of 1600–1800 a.s.l. and by Gruezo (1998) from the highly degraded vegetation
of Pagbilao Grande Island. No study dealing with lowland forest vegetation
was found in the literature.
In order to provide more substantial information on species richness and
composition of Philippine lowland forests, we analysed forest remnants in the
rugged foothills of the Leyte Cordillera. The island of Leyte is located in the
central part of the Philippine archipelago and represents a typical example of
the environmental situation in the Philippines. In 1987, the remaining forest
cover of Leyte was 12%, and in 1994 only 2% of the island’s area have been
[212]
1273
estimated to be primary forest (Dargantes and Koch 1994). More recent data
(DENR 1998) show that about 40% of the land area of Leyte is covered by
grassland and barren land, resulting from abandoned cultivation and grazing
land that marginalised in productivity through erosion and leaching. Another
40% of the island’s area is under coconut plantations. The remaining area is
composed of settlements, agricultural land and forest. In the view of this sit-
uation, the objectives of this study were (a) to analyse the vascular plant species
composition and diversity of selected plots of mature primary forest and (b) to
evaluate the role of the study area as refuge to lowland forest vegetation and its
significance for conservation and as a gene bank.
Study area
[213]
1274
Figure 1. The Philippine archipelago, the Island of Leyte, and the location of the study area.
The soil type in the primary forest between 370 and 520 m a.s.l. is a haplic
Andosol with rudic phase (FAO/UNESCO 1988) overlying basaltic and
andesitic breccia (Zikeli 1998). The soil at lower elevations (100 m a.s.l.) has
been classified as haplic Alisol (FAO/UNESCO 1988) over basalt (Asio 1996).
Climate
[214]
1275
the year. The standardization of the rainfall pattern in northern Leyte com-
pared to that of southern Leyte might be explained by the protective effect of
the Island of Samar off the northeast-coast of Leyte, although Samar’s
mountains are lower than those of Leyte (ca. 850 m a.s.l.).
Local climatic conditions have been analysed from a 23-years period (1976–
1998) of record by using data from the PAGASA (Philippine Atmospheric,
Geophysical and Astronomical Services Administration) weather station on
the campus of the Leyte State University (7 m a.s.l.), ca. 1–3 km west of the
study plots. The annual average temperature is 27.4 C and the average annual
precipitation is 2586 mm. Highest precipitation occurs during November to
January. Lowest rainfall is observed between March and May. On average, all
months receive at least 100 mm precipitation, i.e. there is no dry month
according to the definition of Walsh (1996). However, drought periods (i.e. less
than 50 mm of monthly precipitation according to Walsh 1996) of up to
4 months have been recorded during El Niño Southern Oscillation events. The
general rainfall patterns and the climatic conditions measured at the PAGASA
station are more similar to climatic type II with its clear impact of the northeast
monsoon than to climatic type IV, implying that neither the mountains
of Samar nor the Leyte Cordillera itself causes a distinct rain shadow west
of Mt. Pangasugan.
Orographic rainfalls are an important factor in the Leyte Cordillera, espe-
cially in the vicinity of the mountain summits. The summit of Mt. Pangasugan
is often observed being cloud covered, and during field work heavy rainfalls
have been experienced, while the coastal plain did not receive any precipitation.
An important climatic feature of the area are typhoons. Leyte lies at the
southern margin of the typhoon tracks entering the Philippines, and is hit at a
rate of five typhoons in three years, mainly during the summer months (Parong
1984; cited in Kintanar 1984).
Vegetation analysis
Field studies were conducted in 1997 and 1998. The attempt to identify a
minimum area in mature primary forest failed due to the heterogeneity of the
vegetation. A plot size allowing a reasonable number of replications proved to
be 100 m2. Where relief conditions and homogeneity of the vegetation allowed,
plots were arranged along a catena from ridge to river bank. The 100-m2 plots
were generally designed as quadrats, but on narrow ridges and river banks,
other rectangular design was used due to relief constraints. A total of 49 plots
was established, with 15 on the ridge, 21 on slopes, and 11 on riverbanks. Two
plots were established in ca. 6-year-old land slide successions, with one of them
located ca. 2 km south of the main study area.
The vegetation analysis procedure was based on a ‘nested quadrat
design’ (Kent and Coker 1992). All plants >2.5 m were identified from
the total plots (100 m2). On central subplots of 25 m2, all plants £ 2.5 m
[215]
1276
as well as the lianas were considered. Records included epiphytic and climbing
plants on the stem bases of trees up to a height of 2.5 m. Crown epiphytes were
not included in the analysis, but epiphytes found on the ground were identified
and added to the species lists. From species which could not be identified in the
field, voucher specimens were collected. Tall trees were sampled with the help
of a tree climber. However, no samples were taken from erect and climbing
palms (rattans), because this would have been destructive and the chance of
identification was very low due to the lack of fertile specimens. Therefore, most
palms had to be distinguished as morphospecies.
Taxa were assigned to life forms on the base of field observations, or with the
help of literature information in the case of juveniles. Life form classification
followed Ellenberg and Mueller-Dombois (1967). Plant samples collected in
this study were deposited at the Department of Plant Breeding, Herbarium,
Leyte State University, ViSCA, Baybay, Leyte, 6521A, Philippines.
Data analysis
Species richness, diversity and evenness were determined for each of the 49
plots. Only those plants rooting within the plots were considered in the
[216]
1277
analysis. The Shannon-Index (H¢) was used as a robust and simple diversity
measure (Magurran 1988). For the analysis of species dominance patterns,
Evenness (E) based on the Shannon-Index was calculated for each of the plots.
To assess the area’s value as a refuge to Philippine tree species, the char-
acterization of forest types by Whitford (1911) was used. His classification and
characterization is based on the occurrence of typical tree species and tree
species combinations. He often used vernacular names or typical families or
genera as e.g. ‘Apitong’ for Dipterocarpus spp. to characterise his forest types.
For many of these vernacular names a scientific species could not be assigned
with certainty, and therefore, were not used for comparisons. Whitford (1911)
pointed out that the description of his forest types was based on a still frag-
mentary knowledge of Philippine forests. Most of his ‘typical’ tree species –
with few exceptions such as mangroves – occur in the other forest types as well.
For example, many species of the dipterocarp forest types occur at wet local-
ities in the Molave forest (limestone forest). On the other hand, the typical
Molave forest species also exist in the dipterocarp forest types, especially on
dry sites. Of such reasons, Whitford’s (1911) forest types are primarily related
to the major habitat conditions in the Philippines and do not represent real
plant associations. The comparison of the species recorded in this study with
Whitford’s (1911) forest types merely demonstrates the diverse habitat condi-
tions in the present study area. Unfortunately, not much work has been con-
ducted so far to improve Whitford’s system, and information on species
composition of the undergrowth vegetation, which might be especially valuable
to characterise habitat conditions (Schulze and Whitacre 1999), is still missing.
Results
From the 49 plots, a total of 685 taxa was recorded. Of these, 58.3% were
identified to species level, 86.2% to genus level, and 96.7% to family level. The
remaining 3.3% of the taxa could only be assigned to higher taxonomic levels.
All taxa identified to species level are listed in the Appendix. Species inventory
was clearly dominated by angiosperms, accounting for 92.1% of all species.
The pteridophytes represented 7.5% of the species. Only three species of
gymnosperms (Podocarpus rumphii, Gnetum gnemon, G. latifolium) were found
(Table 1).
More than half (52%) of all species identified are endemic to the Philippines,
including one endemic genus (Greeniopsis, Rubiaceae). The most common
families were the Rubiaceae (35 species) and the Euphorbiaceae (32 species),
followed by the herbaceous family of Araceae and the erect and climbing palms
(Arecaceae) with 28 species each. The Meliaceae and Moraceae included 27
species each (Figure 2). The ratio between the number of genera and the
number of species ranged between 1:1.5 (Anacardiaceae) and 1:6.7 (Moraceae).
The frequency of taxa was low. Nearly half (48.5%) of all taxa were recorded
from only one of the 49 plots, and nearly one third (30.5%) of the taxa were
[217]
1278
Gymnosperms Angiosperms
Figure 2. The 20 most common plant families recorded from 49 plots (100 m2 each) in the study
area. Figures in brackets indicate the ratio between the number of genera and the number of
species.
[218]
1279
represented by only one single individual. Very few species showed high fre-
quencies as e.g. the two tree species, Calophyllum blancoi (present in 32 plots)
and Dacryodes rostrata (present in 31 plots), which was due to a high rate of
juveniles.
The average number of species per plot was 47 and ranged between 17 and
80. Shannon diversity (H¢) reached values between 2.2 and 3.9, and evenness
(E) ranged between 0.64 and 0.98. The species–area curve for all plots shows a
steady increase of species numbers with only a weak tendency to level off
(Figure 3). The flattening of the curve at its beginning is the result of the river
bank vegetation which was comparatively species poor and homogenous. The
species–area curve starts to rise again with the addition of the slope plots.
Life form composition is clearly dominated by phanerophytes (65.3% of all
taxa), followed by lianas (17.1%) and chamaephytes (16.9%). Geophytes were
rare (0.7%) and largely represented by few ground orchids. Hemicryptophytes
and therophytes were absent (Figure 4). Epiphytes were not the focus of this
study and are therefore not included in the calculation of life form composi-
tion. A rough estimate of epiphyte contribution to the area’s species inventory
is ca. 10%. The most conspicuous epiphytic plant group observed were orchids.
Many of the vegetation clusters observed in the tree crowns were composed
of the accumulation of orchid bulbs belonging to a single species (e.g.
Grammatophyllum multiflorum).
The following taxa occurring in the study plots have been classified by
Soepadmo (1995) as endangered and economically important lowland forest
genera in SE Asia: Anisoptera, Dipterocarpus, Parashorea, Shorea, Vatica
(Dipterocarpaceae), Artocarpus (Moraceae), Mangifera (Anacardiaceae), and
Calamus (Arecaceae). Additionally, 41 of the tree species recorded are listed as
endangered for the Philippines by IUCN (2000). Of these, 23 are classified as
vulnerable, one as endangered, and 17 as critically endangered (see Appendix).
Figure 3. Species–area curve for 49 plots (100 m2 each) in the study area.
[219]
1280
Figure 4. Life form spectrum (after Ellenberg and Müller-Dombois 1967) of species recorded
from 49 plots (100 m2 each) in the study area.
Discussion
The 685 taxa recorded from the 49 plots account for nearly 8% of the ca. 8900
vascular plant species so far described for the Philippines (Davis et al. 1995).
Although the plots were not contiguous and species numbers can therefore be
expected to be higher than in contiguous plots (Whitmore 1985) this figure is
high, considering the small overall study area (4900 m2 in total). Only very few
datasets cover tropical lowland forest vegetation comprehensively and are
therefore suitable for comparison. The only study using the same plot size was
conducted by Whitmore et al. (1985) in the tropical lowland rain forest of
Costa Rica, who analysed a plot of 100 m2, considering all vascular plants.
They recorded a total of 233 species, including 59 (25%) epiphyte species. In
the present study, the highest number of species recorded from a single 100-m2
plot was 80 and thus much lower than the number found by Whitmore et al.
(1985). However, in the present study the vegetation up to 2.5 m tall as well as
the lianas were collected from subplots of 25 m2, and crown epiphytes were
excluded. Despite this, the maximum number of vascular plant species on 100-
m2 plots in the study area can expected to be clearly lower than the number of
233 species recorded by Whitmore et al. (1985).
An estimate of the overall vascular plant species richness of Mt. Pangasugan
area, including mossy forest as well as the different stages of succession, results
in 1500–2000 species. This estimate is based on the very conservative
assumption that 50% of the lowland forest species was recorded in this study,
and that the mossy forest has a similar species richness as a 1-ha plot studied by
[220]
1281
Representation of taxa
The rugged relief of the study area represents a broad spectrum of Philippine
habitats. The comparison of the tree species recorded from our study with the
typical tree species composition of the forest types described by Whitford
(1911) showed a high degree of correspondence. Many tree species typical of
the five dipterocarp forest types as well as the Molave type (Figure 5) were
present. From the 18 tree species listed by Whitford (1911) as typical for the
Lauán-hagághak, 15 were also present in our study area. Originally, this forest
type is established on lowland plains on wet soils (Whitford 1911), but was
transformed into rice fields in the study area. However, tree species repre-
senting this type of forest still occur on the banks of the small creeks at low
[221]
1282
Figure 5. Comparison of the number of characteristic species of the different lowland forest types
in the Philippines (after Whitford 1911) with the number of respective species recorded in this
study.
elevations. The typical tree species of other forest types were also well repre-
sented (Figure 5). The high number of Molave type species (50% of the typical
species as mentioned by Whitford 1911) in the study area is remarkable, as this
forest represents dry limestone areas (Whitford 1911). This is another indica-
tion that the area’s vegetation might be strongly influenced by drought periods.
[222]
1283
Conservation value
Kochummen et al. (1992) stated that comparatively small areas might represent
high numbers of a regional flora. They found that their 50-ha plot in the Pasoh
Forest Reserve (Malaysia) included 25% of all trees and shrubs (‡ 1 cm dbh)
of the Malay Peninsula. In our study, an overall sample area of approximately
half hectare included ca. 8% of all Philippine vascular plant species. Given the
small area considered as well as the fact that neither the successional vegetation
nor the mossy forest is included, the representation of Philippine flora in the
Mt. Pangasugan area is clearly higher than 8%.
The proportion of 52% endemic taxa recorded in this study is clearly higher
than the proportion of 39% stated as an average for the Philippines (Davis
et al. 1995). This result agrees with Ashton (1993) who stated that the south-
eastern part of the Philippines is especially rich in endemic plants. The area’s
endemism might be even higher than 52%, as a number of taxa could not be
identified. For example, only 3 of the 16 rattan species (Arecaceae) recorded,
which generally show a high degree of endemism (Dransfield 1990), could be
assigned to a scientific name. Two of them were Philippine endemics.
Another aspect referring to the conservation value of the area is the
occurrence of 41 tree species in the red list of IUCN (2000). However, from the
species recorded from this study, other than trees are not represented in the red
list. Despite this, it can be expected that many of the non-tree taxa recorded are
threatened by habitat destruction as well. For example, no rattans are listed by
IUCN although this plant group is still heavily exploited and shows high rates
of endemism. The IUCN red list seems to have a strong focus on well known
and economically important tree species. This is supported by the fact that only
dipterocarps are classified as critically endangered, although many other tree
species are more rare in the study area. This was e.g. true for the valuable tree
species Heritiera sylvatica (Sterculiaceae) and Xanthostemon verdugonianus
(Myrtaceae) which were known by local farmers from only one mature tree
each in the entire western foothills of Mt. Pangasugan.
Taken together, the Mt. Pangasugan region on Leyte represents a unique
refuge for a high number of species, which are characteristic of all Philippine
[223]
1284
dipterocarp forest types and the molave type. In view of the large areas of
degraded land in the Philippines, the conservation value of the Mt. Pangasugan
region is very high and represents an important gene bank of the Philippine
forest vegetation.
Acknowledgements
This study was part of the ViSCA-gtz Applied Tropical Ecology Program (PN
95.2290.5–001.00) and partly funded by the Tropical Ecology Support Pro-
gram (TÖB) of GTZ (PN 90.2136.1–03.107). We are grateful to the President
of the Leyte State University, Dr. P.P. Milan and her staff for their support and
help. We are also grateful to the Cienda San Vicente Farmer Association
(CSVFA) and their community organiser Marlito Bande, who made the
extensive field trips possible. Special thanks to the curator of the Philippine
National Herbarium, Dr. Madulid, and his staff, as well as to the director of
the National Herbarium of the Netherlands, Leiden branch, Prof. Baas and his
staff who helped to put species identification on a firm ground. We are also
very grateful to Leonardo Co, Nina Ingle, David and Luze Bicknell, Franz
Seidenschwarz, and B.C. Tan for various assistance. We also like to thank the
two anonymous referees for their comments and constructive criticism.
Appendix
Species list of the vascular plant species found in 49 plots (100 m2 each) in the
foothills of the Leyte Cordillera at Mt. Pangasugan, Leyte, Philippines.
The list includes only those species which could be identified to species level.
Some species recorded outside the plots are provided in addition.
Numbers in brackets following the species name indicate the first voucher
specimen collected of this species.
Life form classification of species is based on observations of mature indi-
viduals in the study area, or from species descriptions in literature. Life form
definitions follow Ellenberg and Mueller-Dombois (1967) with a minor revision
by Richter (1997). MacP, Macrophanerophyte (>20–50 m); MesP, Meso-
phanerophyte (>5–20 m); MiP, Microphanerophyte (>2–5 m); NP, Nano-
phanerophyte (>1–2 m); NP herb, herbaceous Nanophanerophyte; Ch,
Chamaephyte ( £ 1 m); Ch frut, fruticose Chamaephyte; Ch suff, suffruticose
Chamaephyte; Ch herb, herbaceous Chamaephyte; G rhiz, rhizome Geophyte;
PL, Phanerophytic Liana; r PL, root PL; st PL, winding PL; el PL, tendril PL; d
PL, spread climber; E, Epiphyte.
[224]
1285
Species classified by IUCN (2000) as endangered are listed along with their
status in bold letters. Short definitions of the status are:
CR, critically endangered (‘… facing an extremely high risk of extinction in
the wild in the immediate future …’); EN, endangered (‘… not critically
endangered but facing a very high risk of extinction in the wild in the near
future …’); VU, vulnerable (‘… not critically endangered or endangered but
facing high risk of extinction in the wild in the medium-term future …’).
For comprehensive definitions and criteria of classification see www.
iucnredlist.org/search-basic.html
I.
Aceraceae
Acer laurinum Hassk. (1221) MacP
Actinidiaceae
Saurauia cf. denticulata C.B. Rob. (1078) MiP
Saurauia samarensis Merr. (235) MiP
Alangiaceae
Alangium longiflorum Merr. (1331) VU MesP
Amaranthaceae
Deeringia polysperma (Roxb.) Moq. (2214) Ch herb
Anacardiaceae
Dracontomelon dao (Blco.) Merr. & Rolfe (660) MacP
Dracontomelon edule (Blco.) Skeels Mes
Koordersiodendron pinnatum (Blco) Merr. (162) MacP
Mangifera altissima Blco. (971) VU MacP
Rhus taitensis Guill. (818) MesP
Semecarpus cuneiformis Blco. (538) MiP
Annonaceae
Alphonsea arborea (Blco.) Merr. (1009) MesP
Anaxagorea javanica Bl. (1509) MiP
Artabotrys cf. rolfei Vid. (2159) el PL
Cananga odorata (Lamk.) Hook. f. & Thoms. MesP
Goniothalamus elmeri Merr. (327) MiP
Meiogyne virgata (Bl.) Miq. MesP
Papualthia cf. lanceolata (Vid.) Merr. (206) MesP
Popowia pisocarpa (Bl.) Endl. (1054) MesP
Apocynaceae
Alstonia macrophylla Wall. ex. G. Don (1774) MesP
Alstonia scholaris (L.) R. Br. MesP
Kibatalia blancoi (Rolfe) Merr. (467) MesP
Lepiniopsis ternatensis Val. (2220) MesP
Tabernaemontana pandacaqui Poir. (140) MicP
Voacanga globosa (Blco.) Merr. (218) MiP
Araceae
Alocasia cf. zebrina Schott ex van Houtte (1677) Ch herb
Amorphophallus paeoniifolius (Dennst.) Nicolson (1924) NP herb
Costus speciosus (J. Konig) Sm NP herb
[225]
1286
[226]
1287
[227]
1288
[228]
1289
[229]
1290
[230]
1291
[231]
1292
[232]
1293
[233]
1294
[234]
1295
[235]
1296
II.
Aspidiaceae
Didymochlaena cf. truncatula (Sw.) J. Sm. (2056) Ch herb
Aspleniaceae
Asplenium nidus L. (1902) Ch herb
Asplenium tenerum Forst. (2096) Ch herb
Athyriaceae
Diplazium asperum (Bl.) Milde (1809) Ch herb
Diplazium esculentum (Retz.) Sw. (1846) Ch herb
Cyatheaceae
Cyathea cf. contaminans (Hook.) Copel. MesP
Davalliaceae
Davallia solida (G. Forst.) Sw. (1462) Ch herb
Davallia trichomanoides Bl. var. lorrainii (Hance) Holttum (222) Ch herb
Hymenophyllaceae
Trichomanes javanicum Bl. (1042) Ch herb
Lindsaeaceae
Lindsaea lucida Bl. ssp. lucida (533) Ch herb
Sphenomeris chinensis (L.) Maxon Ch herb
Tapeinidium pinnatum (Cav.) C.Chr. (1267) Ch herb
Lomariopsidaceae
Bolbitis cf. guoyana (Gaudich.) Ching (2016) Ch herb
Bolbitis guoyana (Gaudich.) Ching Ch herb
Bolbitis heteroclita (Presl) Ching (1049) r PL
Lomogramma cf. copelandii Holttum (1851) r PL
Lomogramma copelandii Holttum r PL
Teratophyllum arthropteroides (Christ) Holttum (2084) Ch herb
Teratophyllum cf. articulatum (J. Sm. ex Fèe) Mett. (516) Ch herb
Osmundaceae
Osmunda banksiaefolia (Pr.) Kuhn (1261, 1392) Ch herb
Polypodiaceae
Drynaria quercifolia (L.) J. Sm E
Leptochilus cf. decurrens Bl. Ch herb
Microsorum cf. longissimum J. Sm. ex Fée (964) Ch herb
Microsorum membranifolium (R. Br.) Ching Ch herb
Microsorum punctatum (L.) Copel. (1821) Ch herb
Microsorum scolopendria (Burm. f.) Copel. (1445) Ch herb
Pyrrosia cf. lanceolata (L.) Farwell PL
Microsorum plukenetii (Presl) M.G. Price (1860)
Pteridaceae
Pteris cf. pellucida Presl Ch herb
Pteris ensiformis Burm. f. (1806) Ch herb
Pteris longipinnula Wall. (334) Ch herb
Schizaeaceae
Lygodium auriculatum (Willd.) Alst. et Holtt. (1974) st PL
Lygodium circinnatum (Burm. f.) Sw. (1603) st PL
[236]
1297
References
Anonymous 1993. Flora of Taiwan Vol. 3, Angiosperms, Dicotyledons, , 2nd edn. Editorial
Committee of the Flora of Taiwan, Department of Botany, National Taiwan University, Taipei,
Taiwan, ROC.
Anonymous 1994. Flora of Taiwan Vol. 1, Pteridophyta, Gymnospermae, 2nd edn. Editorial
Committee of the Flora of Taiwan, Department of Botany, National Taiwan University, Taipei,
Taiwan, ROC.
Anonymous 1996. Flora of Taiwan Vol. 2, Angiospermae, 2nd edn. Editorial Committee of the
Flora of Taiwan, Department of Botany, National Taiwan University, Taipei, Taiwan, ROC.
Aragones E.G. Jr. 1991. Vegetation-soil pattern along altitudinal gradient in the western slopes of
Mt. Banahaw, Luzon, Philippines: I. The forest communities and changes in forest composition
with altitude. Sylvatrop 1(1): 15–45.
Ashton P.S. 1993. Philippine phytogeography. Asia Life Sciences 2(1): 1–8.
Asio V. 1996. Characteristics, weathering, formation and degradation of soils from volcanic rocks
in Leyte, Philippines. Hohenheimer Bodenkundliche Hefte 33 Universität Hohenheim Institut
für Bodenkunde und Standortslehre, Stuttgart, 209 pp.
Barrera A., Aristorenas I. and Tingzon J.A. 1954. Soil Survey of Leyte Province, Philippines. Soil
Survey Report No. 18, Bureau of Printing, Manila, 103 pp.
Bremer H. 1995. Boden und Relief in den Tropen: Grundvorstellungen und Datenbank. Gebrüder
Bornträger, Berlin, Stuttgart, 324 pp.
Bremer H. 1999. Die Tropen – Geographische Synthese einer fremden Welt im Umbruch. Gebrüder
Bornträger, Berlin, Stuttgart, 428 pp.
Brooks T.M., Mittermeier R.A., Mittermeier C.G., da Fonseca G.A.B., Rylands A.B., Konstant
W.R., Flick P., Pilgrim J., Oldfield S., Magin G. and Hilton-Taylor C. 2002. Habitat loss and
extinction in the hotspots of biodiversity. Conservation Biology 16(4): 909–923.
[237]
1298
Brown W.H. 1919. Vegetation of Philippine Mountains: The Relation between the Environment
and Physical Types at Different Altitudes. Bureau of Science, Manila, 434 pp.
Brown W.H. and Mathews D.M. 1914. Philippine dipterocarps forests. Philippine Journal of
Science A 9(5): 413–561.
Buot I.E. Jr. and Okitsu S. 1997. Woody species composition in the altitudinal zones of the mossy
forest of Mt. Pulog, Luzon, Philippines. Flora Malesiana Bulletin 12(1): 6–11.
Burger D. 1972. Seedlings of Some Tropical Trees and Shrubs Mainly of South East Asia. Pudoc,
Wageningen, 399 pp.
Cain S.A., de Oliveira Castro G.M., Pires J.M. and da Silva N.T. 1956. Application of some
phytosociological techniques to Brazilian rain forest. American Journal of Botany 43: 911–941.
Cheek M. and Jebb M. 2001. Nepenthaceae. Flora Malesiana Vol. 15, Nationaal Herbarium
Nederland, Universiteit Leiden Branch, the Netherlands, 164 pp.
Corner E.J.H. 1988. Wayside Trees of Malaya. Vol. 1 and 2. Malayan Nature Society, Kuala
Lumpur, 774 pp.
Cromer D.A.N. and Pryor L.D. 1942. A contribution to rain-forest ecology. Proceedings of the
Linnean Society 67: 249–268.
Dargantes B.B. and Koch W. 1994. Case studies on the occupation and cultivation of the forest
lands of Leyte, Philippines. Annals of Tropical Research 16: 13–29.
Davis S.D., Heywood V.H. and Hamilton A.C. (eds) 1995. Centres of Plant Diversity – A Guide
and Strategy for their Conservation, Vol. 2: Asia, Australasia and the Pacific. IUCN Publications
Unit, Cambridge, UK.
de Guzman E.D., Umali R.M. and Sotalbo E.D. 1986. Guide to Philippine Flora and Fauna, Vol.
3: Dipterocarps, Non-Dipterocarps. Natural Resources Management Center, Ministry of Nat-
ural Resources and University of the Philippines, 414 pp.
DENR 1998. Philippine atlas. Department of Environment and Natural Resources (DENR),
Quezon City, Philippines.
DENR and UNEP 1997. Philippine Biodiversity: An Assessment and Action Plan. Bookmark Inc.,
Makati City, The Philippines, 298 pp.
Dransfield J. 1990. Outstanding problems in Malesian palms. In: Baas P., Kalkmann K. and
Geesink R. (eds), The Plant Diversity of Malesia, Proceedings of the Flora Malesiana Sympo-
sium, Leiden, the Netherlands, August 1989. Kluwer Academic Publishers, Dordrecht-Boston-
London, pp. 17–25.
Duivenvoorden J.F. 1994. Vascular plant species counts in the rain forests of the middle Caquetá
area. Colombian Amazonia. Biodiversity and Conservation 3: 685–715.
Ellenberg H. and Mueller-Dombois D. 1967. A key to Raunkiaer plant life forms with revised
subdivisions. Berichte des geobotanischen Instituts Rübel 37: 56–73.
FAO 2003. State of the World’s Forests 2003. FAO, Rome, Italy. www.fao.org/DOCREP/005/
Y7581E/Y7581E00.HTM, accessed 25 February 2004.
FAO/UNESCO 1988. Soil Map of the World, Revised Legend. World Soil Resources Report 60.
FAO, Rome.
Gentry A.H. and Dodson C. 1987. Contribution of nontrees to species richness of a tropical rain
forest. Biotropica 19(2): 149–56.
Gonzales-Salcedo P.V. 2001. Floral diversity and vegetation zones of the northern slope of
Province, Luzon, Philippines. Asia Life Sciences 10(2): 119–157.
Gruezo W.S. 1998. Species alpha diversity of Pagbilao Grande Island vegetations, Quezon Prov-
ince, Philippines. Asia Life Sciences 7(1): 39–92.
Hall J.B. and Swaine M.D. 1981. Distribution and Ecology of Vascular Plants in a Tropical Rain
Forest: Forest Vegetation in Ghana. Dr. W. Junk Publishers, The Hague.
Hamann A., Barbon E.B., Curio E. and Madulid D.A. 1999. A botanical inventory of a submontane
tropical rainforest on Negros Island, Philippines. Biodiversity and Conservation 8: 1017–1031.
Heaney L. and Regalado J.C. Jr. 1998. Vanishing Treasures of the Philippine Rain Forest: A
Comprehensive Introduction to the Biodiversity of the Philippines. The Field Museum Uni-
versity of Chicago Press, Chicago, 88 pp.
[238]
1299
Henderson M.R. 1974a. Malayan Wild Flowers: Monocotyledons, reprint of the 1954 edition. The
Malayan Nature Society, Kuala Lumpur, Malaysia, 357+27 pp.
Henderson M.R. 1974b. Malayan Wild Flowers: Dicotyledons, Reprint. The Malayan Nature
Society, Kuala Lumpur, Malaysia, 478 pp.
Ingle N. 2003. Seed dispersal by wind, birds, and bats between Philippine montane rainforest and
successional vegetation. Oecologia 134: 251–261.
IUCN 2000. Red List – Philippines. www.iucnredlist.org/search-basic.html. Accessed 14.01.2004.
Kartawinata K. 1990. A review of natural vegetation studies in Malesia, with special reference to
Indonesia. In: Baas P., Kalkmann K. and Geesink R. (eds), The Plant Diversity of Malesia,
Proceedings of the Flora Malesiana Symposium, Leiden, the Netherlands, August 1989. Kluwer
Academic Publishers, Dordrecht-Boston-London, pp. 120–132.
Keller R. 1996. Identification of Tropical Woody Plants in the Absence of Flowers and Fruits.
Birkhäuser Verlag, Basel, Boston, Berlin, 229 pp.
Keng H. 1983. Orders and Families of Malayan Seed Plants, 3rd edn. Singapore University Press,
National University of Singapore, 441 pp.
Keng H. 1990. The Concise Flora of Singapore: Gymnosperms and Dicotyledons. Singapore
University Press, National University of Singapore, Singapore, 222 pp.
Kent M. and Coker P. 1992. Vegetation Description and Analysis. Belhaven Press, London, 363
pp.
Kiew R. 1987. The herbaceous flora of Ulu Endau, Johore-Pahang, Malaysia, including taxonomic
notes and descriptions of new species. Malayan Nature Journal 41(2 and 3): 201–234.
Kintanar R.L. 1984. Climate of the Philippines. PAGASA, Quezon City, Philippines, September
1984, 38 pp.
Kochummen K.M., LaFrankie J.V. and Manokaran N. 1992. Floristic composition of Pasoh
Forest Reserve, a lowland rainforest in peninsular Malaysia. In: Kheong Y.S. and Win L.S.
(eds), In Harmony with Nature, Proceedings of the International Conference on Conservation of
Tropical Biodiversity, Kuala Lumpur, Malaysia, 12–16 June 1990. Malayan Nature Society,
Kuala Lumpur, Malaysia, pp. 545–554.
Lieberman D., Lieberman M., Peralta R. and Hartshorn G.S. 1996. Tropical forest structure and
composition on a large-scale altitudinal gradient in Costa Rica. Journal of Ecology 84: 137–152.
Magurran A.E. 1988. Ecological Diversity and Its Measurement. Chapman and Hall, London, 179
pp.
Manokaran N. and Kochummen K.M. 1990. A re-examination of data on structure and floristic
composition of hill and lowland dipterocarp forest in peninsular Malaysia. Malayan Nature
Journal 44: 61–75.
Meijer W. 1959. Plantsociological analysis of montane rainforest near Tjibodas, West Java. Acta
Botanica Neerlandica 8: 277–291.
Merrill E.D. 1912. A Flora of Manila. Department of the Interior, Bureau of Science, Manila, 491
pp.
Myers N., Mittermeier R.A., Mittermeier C.G., da Fonseca G.A.B. and Kent J. 2000. Biodiversity
hotspots for conservation priorities. Nature 403: 853–858.
Newbery D.McC., Campbell E.J.F., Proctor J. and Still M.J. 1996. Primary lowland dipterocarp
forest at Danum Valley, Sabah, Malaysia. Species composition and patterns in the understorey.
Vegetatio 122: 193–220.
Ng F.S.P. (ed.) 1978. Tree Flora of Malaya Vol. 3. Forest Department, Ministry of Agriculture and
Lands, Malaysia, Longman Singapore, 339 pp.
Ng F.S.P. (ed.) 1989. Tree Flora of Malaya Vol. 4. Forest Department, Ministry of Agriculture and
Lands, Malaysia, Longman Malaysia, 548 pp.
Ng F.S.P. 1991. Manual of Forest Fruits, Seeds and Seedlings. Vols. I and II. Malayan Forest
Record No. 34. Forest Research Institute Malaysia, Kepong, 52109 Kuala Lumpur, Malaysia,
997 pp.
NSCB RU-8 2001. Provincial/City statwatch: Regional Unit VIII – Eastern Visayas. National
Statistical Coordination Board, www.nscb.gov.ph. Accessed 20.11.2003.
[239]
1300
Pancho J.V. 1983. Vascular Flora of Mount Makiling and Vicinity, Part 1. Kalikasan, The Phil-
ippine Journal of Biology, Supplement No. 1. New Mercury Printing Press, Quezon City,
Philippines, 476 pp.
Pipoly J.J.III and Madulid D.A. 1998. Composition, structure and species richness of a submon-
tane moist forest on Mt. Kinasalapi, Mindanao, Philippines. In: Dallmeier F. and Comiskey J.A.
(eds), Forest Biodiversity Research, Monitoring and Modeling: Conceptual Background and Old
World Case Studies. UNESCO and The Parthenon Publishing Group Limited, Carnforth, UK,
UNESCO, France, pp. 591–600.
Poulsen A.D. 1996. Species richness and density of ground herbs within a plot of lowland rainforest
in north-west Borneo. Journal of Tropical Ecology 12: 177–190.
Poulsen A.D. and Balslev H. 1991. Abundance and cover of ground herbs in an Amazonian rain
forest. Journal of Vegetation Science 2: 315–322.
Proctor J., Argent G.C. and Madulid D.A. 1998. Forests of the ultramufic Mt. Giting-Giting,
Sibuyan Island, the Philippines. Edinburgh Journal of Botany 55(2): 295–316.
Raunkiaer C. 1934. The Life-forms of Plants and Statistical Plant Geography. Oxford University
Press, Oxford.
Rennolls K. and Laumonier Y. 2000. Species diversity structure analysis at two sites in the tropical
rain forest of Sumatra. Journal of Tropical Ecology 16: 253–270.
Richards P.W. 1996. The Tropical Rain Forest: An Ecological Study. Cambridge University Press,
Great Britain, 575 pp.
Richter M. 1997. Allgemeine Pflanzengeographie. B.G. Teubner, Stuttgart, 256 pp.
Santos J.V., de Guzman E.D. and Fernando E.S. 1986. Guide to Philippine Flora and Fauna. Vol. IV:
Bamboos, Grasses, Palms. Natural Resources Management Center, Ministry of Natural Resources
and University of the Philippines, JMC Press, Incorporated, Quezon City, Philippines, 256 pp.
Schulze M.D. and Whitacre D.F. 1999. A classification and ordination of the tree community of
Tikal National Park, Peten, Guatemala. Bulletin of the Florida Museum of Natural History
41(3): 169–297.
Sist P. and Saridan A. 1998. Description of the primary lowland forest of Berau. In: Bertault J.-G.
and Kadir K. (eds), Silviculture research in a lowland mixed dipterocarp forest of East Kali-
mantan, the contribution of STREK project. CIRAD-Forêt Publication, pp. 51–93.
Slik J.W.F., Poulsen A.D., Ashton P.S., Cannon C.H., Eichhorn K.A.O., Kartawinata K.,
Lanniari I., Nagamasu H., Nakagawa M., van Nieuwstadt M.G.L., Payne J., Purwaningsih,
Saridan A., Sidiyasa K., Verburg R.W., Webb C.O. and Wilkie P. 2003. A floristic analysis of the
lowland dipterocarp forests of Borneo. Journal of Biogeography 30: 1517–1531.
Smith R.F. 1970. The vegetation structure of a Puerto Rican rain forest before and after short-term
irradiation. In: Odum H. (ed.), A Tropical Rain Forest. Oak Ridge, Atomic Energy Commission,
USA, pp. D103–140.
Soepdadmo E. 1995. Plant diversity of the Malesian tropical rainforest and its phytogeographical
and economic significance. In: Primack R.B. and Lovejoy T.E. (eds), Ecology, Conservation, and
Management of Southeast Asian Rainforests. Yale University, pp. 19–40.
Soepadmo E. and Wong K.M. (eds) 1995. Tree Flora of Sabah and Sarawak. Vol. 1. Forest
Research Institute Malaysia, Sabah Forestry Department, and Sarawak Forestry Department,
Ampang Press Sdn. Bhd., Kuala Lumpur, Malaysia, 513 pp.
Soepadmo E., Wong K.M. and Saw L.G. (eds) 1996. Tree Flora of Sabah and Sarawak. Vol. 2.
Forest Research Institute Malaysia, Sabah Forestry Department, and Sarawak Forestry
Department. Ampang Press Sdn. Bhd., Kuala Lumpur, Malaysia, 443 pp.
Soerianegara I. and Lemmens R.H.M.J. (eds) 1994. Plant Resources of South-East Asia No. 5(1):
Timber Trees: Major Commercial Timbers. Prosea Foundation, Bogor, Indonesia and Pudoc-
DLO, Wageningen, the Netherlands, 610 pp.
Takeuchi M. 1960. A estrutura da vegetaçao na Amazonia III: A mata da campina na regiâo do
Rio Negro. Boletim do Museo Paraense Emilio Goeldi 8: 1–13.
Tan B.C. and Rojo J.P. 1989. The Philippines. In: Campbell D.G. and Hammond H.D. (eds),
Floristic Inventory of Tropical Countries – The Status of Plant Systematics, Collections, and
[240]
1301
Vegetation, Plus Recommendations for the Future. New York Botanical Garden, New York, pp.
44–62.
Turner I.M. 1994. The taxonomy and ecology of the vascular plant flora of Singapore: a statistical
analysis. Bot. J. Linn. Soc. 114: 215–227.
Valencia R., Balslev H. and Paz Y. Miño C.G. 1994. High tree alpha-diversity in Amazonian
Ecuador. Biodiversity and Conservation 3: 21–28.
van Balgooy M.M.J. 1997a. Malesian Seed Plants. Vol. 1: Spot-characters. Rijksherbarium /Hortus
Botanicus, Leiden, 154 pp.
van Balgooy M.M.J. 1997b. Malesian Seed Plants. Vol. 2: Portraits of Tree Families. Rijksher-
barium /Hortus Botanicus, Leiden, 307 pp.
van Steenis C.G.G.J. (ed.) 1950-ongoing. Flora Malesiana. Martinus Nijhoff Publishers, Dordr-
echt, the Netherlands.
Walsh R.P.D. 1996. Climate. In: Richards P.W. (ed.), The Tropical Rainforest. Cambridge Uni-
versity Press, pp.159–205.
Whitford H.N. 1906. Vegetation of the Lamao Forest Reserve. Philippine Journal of Science I:
373–431 and 637–682.
Whitford H.N. 1911. The Forests of the Philippines. Part I: Forest Types and Products. Depart-
ment of the Interior Bureau of Forestry, Bulletin No. 10, Manila, 94 pp.
Whitmore T.C. (ed.) 1983a. Tree Flora of Malaya. Vol. 1. Corrected Reprint of the 1972 Edition.
Forest Department, Ministry of Agriculture and lands, Malaysia Longman Malaysia, 473 pp.
Whitmore T.C. (ed.) 1983b. Tree Flora of Malaya. Vol. 2. Corrected Reprint of the 1972 Edition.
Forest Department, Ministry of Primary Industries, Malaysia, Longman Malaysia, 442 pp.
Whitmore T.C. 1985. Tropical Rain Forests of the Far East. Corrected Reprint of the 1984 Edition.
Clarendon Press, Oxford, 352 pp.
Whitmore T.C., Peralta R. and Brown K. 1985. Total species count in a Costa Rican tropical rain
forest. Journal of Tropical Ecology 1: 375–378.
Wilkie P., Argent G., Cambell E. and Saridan A. 2004. The diversity of 15 ha of lowland mixed
dipterocarp forest, Central Kalimantan. Biodiversity and Conservation 13(4): 695–708.
Zamora P.M. and Co L. 1986. Guide to Philippine Flora and Fauna. Vol. II: Economic Ferns,
Endemic Ferns, Gymnosperms. Natural Resources Management, Center Ministry of Natural
Resources, and University of the Philippines, JMC Press, Incorporated, Quezon City, Philip-
pines, 273 pp.
Zikeli S. 1998. Nutrient status and nutrient cycles of the tropical rainforest, Mt. Pangasugan, Leyte,
Philippines. Diploma thesis, Martin - Luther - University Halle - Wittenberg, Landwirtschaf-
tliche Fakultaet, Institut für Bodenkunde und Pflanzenernaehrung, 100 pp.
[241]
Biodiversity and Conservation (2006) 15:1303–1318 Springer 2006
DOI 10.1007/s10531-005-3873-7
-1
Introduction
#
Supported by the important directional item of the Chinese Academy of Sciences (KSCX2-SW-
104)
[243]
1304
during the tertiary period or even before (Zhou 1992). The genus has only three
extant species and Trigonobalanus doichangensis is the only one distributed in
China (Hsu et al. 1981). T. doichangensis is most closely related to the genus
Quercus L. (Zhou 1992) and has a high scientific value for studies on Fagaceae
phylogeny, continental drift theory and global environmental changes.
Due to heavy exploitation for fuel wood, vegetation destruction in China
and its restricted distribution, T. doichangensis is seriously threatened and is
being pushed to the verge of extinction (Fu 1992). Hence, it was listed as a
national rare and endangered plant in 1984 and has also been proposed as a
second-ranked plant for national protection in China (Anon 1999). So far, its
phylogeny (Wang and Zhang 1988; Nixon and Crepet 1989; Wu and Xiao
1989; Liao et al. 1998; Wang et al. 1998) and the community floristic elements
(Li 1994) have been comprehensively studied. However, there are only a few
recent reports on its conservation biology (Zhou et al. 2003; Zhou, 2003; Sun
et al. 2004).
It is well known that a practical conservation strategy for an endangered
plant is based on an understanding of the threats it faces. In this study we
describe the status of the current distribution, population ecology, reproduc-
tive biology and genetic variation of T. doichangensis. Our goal was to provide
comprehensive information relevant to the conservation and restoration of this
species.
Methods
Data analysis from the general surveys and ecological plotting studies, deter-
mined the threat category by assessment of current distribution and population
size following to the IUCN Red List Categories and Criteria (IUCN 2001).
Investigations were performed in the flowering and fruiting season. Plot sizes of
10 · 20 m, 20 · 20 m, 30 · 30 m and 50 · 60 m were applied in accordance with
[244]
1305
the population size. The characteristics recorded were: (1) Site conditions, (2)
Plant height and crown size (m·m), trunk base diameter and diameter at breast
height of all T. doichangensis in the plot, (3) Habitat quality, (4) The main
accompanying higher plants and their abundance, and (5) Regeneration
capacity of T. doichangensis from seeds.
DNA extraction
Total DNA was extracted from dried leaves using a modified CTAB method
(Doyle and Doyle 1987). Modifications included incubation in CTAB at 65 C
for 1.5 h; a second extraction in chloroform–isoamyl alcohol; repeated wash-
ings in 70% ethanol without ammonium acetate; a final washing in 100%
ethanol and the addition of RNase to the TE buffer. DNA was visualized on
1.0% agarose gels.
Data analysis
The individuals were scored for the presence (1) or absence (0) of amplified
bands. The data matrix of RAPD bands was statistically analyzed using
POPGENE (Yeh et al. 1997). The parameters of the genetic diversity and
genetic structure are the percentage of polymorphic loci P(%), Shannon’s in-
dex of diversity (I), Nei’s gene diversity (h), the coefficient of gene differen-
tiation (Gst) and an estimate of gene flow (Nm).
[245]
1306
Reproduction biology
Results
[246]
1307
Plots
Altitude (m) 1020 1450 1550 1590 1730
Slope N, 30 degrees W, 40 degrees SE, 40 degrees S, 30 degrees S, 30 degrees
Coverage (%) >90 >90 >95 >95 >95
Size (m2) 800 400 3000 600 400
TD no. in the plot 195 140 41 101 118
TD no. per 100 m2 24 35 2 17 30
TD Avg. height (m) 5.1 5.6 17.9 6.8 14.4
Hst/Shst of TD (m) 13.5/0.2 9.5/0.3 35/0.4 32/0.28 20/0.2
% of TD ‡20 m / / 51 6 0.8
% of TD ‡4–20 m 59 73 44 55 92
% of TD 2–3 m 15 21 / 23 5
% of TD £ 1.0 m 26 6 5 16 3
% of dead TD / / / 17 15
Avg. height of DP (m) / / / 1.5 10 3.5 18
Accompanier no. 45 30 25 27 25
Main companions (TP) 1 2 3 2 3 4 5 4 5 4 5
[247]
1308
these individuals were mostly replaced by other native plant species, alien plant
invaders, such as Eupatorium adenophorum (Ageratina adenophora) and Eup-
atorium odoratum (Chromolaena odorata), and agricultural crops. Occasionally,
some seedlings or young trees appeared around scattered individuals inside the
evergreen broadleaf forests. However, other plants such as Castanopsis ca-
lathiformis, Castanopsis echinocarpa, Lithocarpus fenestratus, Schima wallichii
and Anneslea fragrans had already dominated in the vegetation. The formation
of Type IsI is due to heavy cutting and vegetation destruction. The isolated
individuals are found in the most endangered habitat and their genetic diversity
is easily reduced by further human activities, grazing animals, farming and
further biotic invasion.
(2) Type SW (Sprouting woods): Type SW is the result of fuel wood cutting by
indigenous people. Investigations show that the indigenous ethnic groups of
Dai, Wa and Laku are familiar with T. doichangensis and have realized that the
tree can sprout easily after top-cutting and thinning and thus they have adopted
the methods of ‘alternate cutting or thinning cutting’ for the primitive sustain-
able use of the tree as fuel wood. As a result of these practices, plants of
T. doichangensis in these woods showed some unique characteristics in tree
shape, tree height structure and associated floristic composition (Table 1). In
this community the tallest T. doichangensis was about 13 m and the average
height was around 5 m. Some 60% of T. doichangensis reached the reproductive
phase and about 25.6% of the individuals were below 1 m in height. T. doi-
changensis was the dominant species in woods, and 50 species of accompanying
higher plants were present. The most important trees were Vaccinum bractea-
tum, Castanopsis hystrix, Lithocarpus fenestratus, Craibiodendron stellatum,
Anneslea fragrans, Ternstroemia gymnanthera and Schima wallichii. Vernonia
parishii, Arthraxon lanceolatus, Carex baccans and Zingiber striolatum were the
typical herbaceous plants. Among epiphytes, Phymatodes lucida, Vanda coerulea
and Eria pannea were also commonly found on the trunk. As ontogenesis of
T. doichangensis was prevented by cutting, most trees could not complete their
natural growth and their reproductive capacity was relatively restrained.
(3) Type MDF (Mono-dominant forest): Plants of T. doichangensis in type
MDF are found as small mono-dominated patches scattered in the secondary
evergreen broadleaf forest. The tallest plant of T. doichangensis was about
10 m and the average height was around 5–6 m. Approximately 70% of the
trees in the plots were mature with a height of ‡4 m. Around 20% of the plants
were 2–3 m in height, while seedlings and young trees represented only 6% of
the total population. Accompanying higher plants were represented by some 30
species and most of the woody species were the same as in Type SW, but there
were far fewer herbaceous plants and epiphytes. Osyris wightiana, Viburnum
cylindricum, Broussonetia papyrifera and Phyllanthus emblica were present in
the community. T. doichangensis in Type MDF showed vigorous growth and a
strong regenerative ability.
(4) Type CDF (Co-dominant forest): Plants of T. doichangensis in Type CDF
often formed a mosaic of mono-dominant patches in the primitive evergreen
[248]
1309
One hundred and fifty seven bands from 100 to 2900 bp were generated by the
16 selected primers, with each primer producing 6–13 bands with an average of
9.81 bands. Eighty three of the total bands were polymorphic and accounted
for 52.87% (Table 2). These findings show that genetic diversity at the species
level was abundant. Table 2 also indicates that the percentages of polymorphic
bands in both populations of Lancang and Chiang-Rrai were far lower than
those in the 3 other populations of Cangyuan, Menglian and Ximeng.
Therefore, the genetic variation within both populations of Lancang and
Chiangrai was lower that that in others.
Table 3 shows that the effective number of alleles (ne), Shannon’ s index of
diversity (I) and Nei’ s gene diversity (h) were 1.2646, 0.2431 and 0.1595,
[249]
1310
Table 2. The sequences of random oligonucleotide primers and the number of polymorphic bands.
OPC05 TCGTCTGCCC 9 4
OPC14 TGCGTGCTTG 10 5
OPC15 GACGGATCAG 11 6
OPJ 09 TGAGCCTCAC 11 6
OPM02 ACAACGCCTC 6 0
OPM06 CTGGGCAACT 8 5
OPM13 GGTGGTCAAG 10 4
OPM15 GACCTACCAC 12 5
OPM16 GTAACCAGCC 13 7
OPM20 AGGTCTTGGG 10 4
OPN14 TCGTGCGGGT 12 9
OPN20 GGTGCTCCGT 7 3
OPS03 CAGAGGTCCC 11 6
OPV06 ACGCCCAGGT 10 9
OPV10 GGACCTGCTG 9 4
OPV15 CAGTGCCGGT 8 6
[250]
1311
In T. doichangensis the ovary has 3 locules with 2 ovules per locule. Obser-
vation on the ovary in transverse section showed that 6 ovules in each ovary
were generated at an early stage, and at the developmental anaphase 1 of 2
ovules per chamber were well developed, while another one was aborted.
Meanwhile, both axile and parietal placentas were observed, and thus
[251]
1312
1 1.9
2 8.1
3 18.3
4 22.9
5 22.1
6 22.4
7 22.2
8 20.2
9 20.9
10 21.8
11 37.8
12 37.2
13 22.7
14 21.4
15 18.3
Notes: 1 agar 7; 2 agar 7 + sucrose 100; 3 agar 7 + sucrose 150; 4 agar 7 + sucrose 200;
5 agar 7 + sucrose 250; 6 agar 7 + sucrose 200 + boric acid 100; 7 agar 7 + sucrose
200 + boric acid 150; 8 agar 7 + sucrose 200 + boric acid 200; 9 agar 7 + sucrose 200 +
boric acid 300; 10 agar 7 + sucrose 200 + boric acid 400; 11 agar + sucrose 200 + calcium
nitrate 100; 12 agar 7 + sucrose 200 + calcium nitrate 200; 13 agar 7 + sucrose 200 +
calcium nitrate 300; 14 agar 7 + sucrose 200 + calcium nitrate 400; 15 agar 7 + sucrose
200 + calcium nitrate 500 (All units are g/L).
[252]
1313
Discussion
[253]
1314
Deficient) in The World List of Threatened Trees (Oldfield et al. 1998) and its
threat category has not yet been evaluated by using any version of IUCN Red
List Categories and Criteria.
Based on our observations and data analyses, the threat category of
T. doichangensis in China is proposed as EN (Endangered) [EN, B1a,
B1b(i,ii,iii)] (IUCN 2001). It is therefore considered to be facing a very high
risk of extinction in the wild, because: (1) Its extent of occurrence is estimated
to be less than 5000 km2, (2) Its populations are severely fragmented and (3)
There is a continuing decline in the extent of occurrence, area of occupancy
and quality of habitat. Judging by the limited distribution of the species in
northern Thailand, and its only postulated occurrence in Myanmar, it is likely
that EN will also be the Global threat category.
Identifying threats
[254]
1315
are already fragmented (Li et al. 2003). However, long-term species survival
depends on the maintenance of sufficient genetic variability within and among
populations to accommodate new selection pressures brought about by envi-
ronmental changes (Elena et al. 2003), and populations with low genetic var-
iation have a high risk of extinction (Li et al. 2003; Ledig et al. 2002).The
decline of seed viability of Metasequoia glyptostroboides, is caused by its poor
genetic base and inbreeding, and its effective seed germination relies on suitable
conditions (Liao and Zhou 1989). Accordingly, lower levels of genetic diversity
and intense genetic differentiation in T. doichangensis will not be able to re-
spond as well to changes in abiotic or biotic environmental conditions as can
other fagaceous species. The low seed germination and plant establishment
may also partly contribute to its present genetic status caused by habitat
fragmentation and population isolation.
We may conclude that endangerment of T. doichangensis is not caused by a
single factor. It seems that habitat destruction, over-exploitation and repro-
duction barriers, are the most likely factors. However, it may be a combination
of historic factors such as climate changes, habitat degeneration, a poor genetic
base and physiological stress. Therefore, a practical conservation strategy for
T. doichangensis is urgently needed.
Conservation considerations
Protection and restoration of natural habitats is the best and cheapest method
of preserving the biological diversity and stability of global ecosystem (Lande
1988). Undoubtedly, the long-term survival of T. doichangensis is dependent on
habitat conservation. However, only 1 of the 4 extant populations of T. doi-
changensis in China has been legally preserved by government ownership and
others are facing a high risk of habitat disappearance. Perhaps more than most
endangered plant species in China, the unique characteristics of T. doichang-
ensis, population isolation and the complex community interaction, exemplify
the importance of habitat preservation and in-situ conservation. Nevertheless,
both in-situ and ex-situ measures are needed for preserving T. doichangensis
and its genetic diversity, and following aspects should be particularly consid-
ered:
(1) Preservation of the habitat and population of T. doichangensis in
Canyuan must be reinforced. The Canyuan population is the only 1 of the 4
populations in China which has been well protected inside a national natural
reserve. Therefore, the population is a vital resource for further research into
the origin and evolution, eco-biological characteristics and genetic realities of
T. doichangensis. Furthermore, the population also has a great value in pop-
ulation restoration research.
(2) As T. doichangensis populations in Menglian, Lancang and Ximeng are
still exposed to high habitat destruction and cutting for fuel wood, it is essential
that new in-situ conservation sites in these area should be urgently planned.
[255]
1316
These new sites will play an important role in habitat recovery and population
restoration.
(3) Although cross-planting is often controversial in plant conservation
planning, it is proposed that seedling cross-planting between different popu-
lations of T. doichangensis should be considered. At least, the potential impacts
of this measure should be studied to determine if low gene flow would be
enhanced and if overall diversity would be increased.
(4) As young T. doichangensis plants propagated from seeds can tolerate
temperatures below 2 C at Kunming (Sun et al. 2004), it may be cultivated
for fuel wood and as a landscaping plant in northern parts of the Tropic of
Cancer. However, the mixed planting of trees propagated from various pop-
ulations is essential.
(5) Ex-situ efforts need to be undertaken to preserve genetic diversity and
multiply specimens. This applies in particular to those populations outside of
the well-protected national nature reserve. Ideally, ex-situ sites will be close to
nature reserves or to botanical gardens. Propagation from seeds is preferred,
since it would be the least detrimental to the extant populations and would
include the widest range of genetic diversity. Optimally, seeds should be taken
from many individuals from each of the populations.
References
Anon 1999. List of National Key Protected Wild Plants (First Group). The Order of National
Forestry Bureau and Agriculture Ministry of China 4: 2–13.
Chen X.Y., Wang X.H. and Song Y.C. 1997. Genetic diversity and differentiation of Cycloba-
lanopsis glauca populations in east China. Acta Botanica Sinica 39: 149–155.
Crepet W.L. and Nixon K.C. 1989. Earliest megafossil evidence of Fagaceae: phylogenetic and
biogeographic implications. American Journal of Botany 76: 842–855.
Doyle J.J. and Doyle J.L. 1988. Isolation of plant DNA from fresh tissue. Focus 12: 13–15.
EBYV (Editorial Board of Yunnan Vegetation). 1987. Vegetation of Yunnan. Science Press,
Beijing, China.
Elena T., Jose M.I. and Cesar P. 2003. Genetic structure of an endangered plant, Antirrhinum
microphyllum (Scrophulariaceae): Allozyme and RAPD analysis. American Journal of Botany
90: 85–92.
Fiedler P.L. and Ahouse J.J. 1992. Hierarchies and cause: toward an understanding of rarity in
vascular plant species. In Fielder P.L. and Jain S.K. (eds), Conservation biology: the theory and
practice of nature conservation, preservation and management. Chapman and Hall, New York,
USA, pp. 23–47.
Forman L. 1964. Trigonobalanus: a new genus of Fagaceae. Kew Bulletin 17: 381–396.
Fu L.G. 1992. China Plant Red Data Book, Vol.1. Science Press, Beijing, China.
He S.A., Hao R.M. and Tang S.J. 1996. A study on the ecological factors of endangering mechanism
of Liriodendron chinensis (Hemsl.) Sarg. Journal of Plant Research and Environment 5: 1–8.
Hsu Y.C., Wang C.J, Wu C.Y. and Li H.W. 1981. Trigonobalanus Forman-A new recorded genus
of Fagaceae in China. Acta Botanica Yunnanica 3: 213–215.
Hu S.Y. 1982. Embryology of Angiosperm. Higher Education Press, Beijing, China.
Huang H.Y., Ma S.B. and Li L. 2001. The genesis of microspore and the formation of male
gametophyte in Dysosma versipellis. Bulletin of Botanical Research 21: 561–566.
IUCN 2001. IUCN Red List Categories and Criteria: Version 3.1. Information Press, Oxford, UK.
[256]
1317
Jiang Z.G., Ma K.P. and Han X.G. 1997. Conservation Biology. Sciences and Technology Press,
Hangzhou, Zhejiang, China.
Johri B.M., Ambegaokar K.B. and Srivastava P.S. 1992. Comparative Embryology of Angio-
sperms, Vol. I. Springer-Verlag, Berlin, Heidelberg.
Kaul R.B. and Abbe E.C. 1989. Inflorescense architecture and evolution in the Fagaceae. Journal
of Arnold Arboretum 65: 375–401.
Kimura M. and Crow J.F. 1964. The number of alleles that can be maintained in a finite popu-
lation. Genetics 49: 725–738.
Lande R. 1988. Genetics and demography in biological conservation. Science 241: 1455–1460.
Ledig F.T., Hodgeskiss P.D. and Jocob-Cervantes V. 2002. Genetic diversity, mating system, and
conservation of a Mexican subalpine relict, Picea Mexicana Martinez. Conservation Genetics 3:
113–122.
Li T.Q. and Cao H.J. 1986. Microsporogenesis and development of male gametophyte of Camellia
chrysantha (Hu) Tuyama. Journal of Beijing Forestry University 8: 30–34.
Li J. 1994. A preliminary study on the floristic elements of the community of Formanodendron
doichangensis. Acta Botanica Yunnanica 16: 17–24.
Li J., Chen K.Y. and Li B.S. 1997. Preliminary study on genetic structure of alpine oaks in Tibet.
Journal of Beijing Forestry University 19: 93–98.
Li Z.Z., Gong J.J., Wang Y. and Huang H.W. 2003. Spatial structure of AFLP genetic diversity of
remnant populations of Metasequoia glyptostroboides (Taxodiaceae). Biodiversity Science 11: 265–275.
Liao H.M., Gou G.Q. and Ye N.G. 1998. A study on seedling morphology and anatomy and
systematic position of Trigonobalanus doichangensis Form. Journal of Wuhan Botanical Re-
search 16: 223–226.
Liao S.Z. and Zhou W.L. 1989. Seed germination experiment of Metasequoia glyptostroboides.
Sichuan Forestry Science and Technology 10: 71–74.
Nixon K.C. and Crepet W.L. 1989. Trigonobalanus (Fagaceae): Taxonomic status and phylloge-
netic relationships. American Journal of Botany 76: 828–841.
Oldfield S., Lusty C. and MacKinven A. 1998. The World List of Threatened Trees. World
Conservation Press, Cambridge, UK.
Pan Y.Z., Gong X. and Liang H.X. 2003. A study on the embryology of endangered plant
Manglietia aromatica. Journal of Wuhan Botanical Research 21: 1–8.
Pang X.A., Liu X., Liu H., Wu C., Wang J.Y., Yang S.X. and Wang Q.F. 2003. The geographic
distribution and habitat of the Isoetes plants in China. Biodiversity Science 11: 288–294.
Rabinowitz D. 1981. Seven forms of rarity. In: Synge H. (ed.), The Biological Aspects of Rare Plant
Conservation. Wiley, New York, USA, pp. 205–217.
Simberloff D. 1988. The contribution of population and community biology to conservation
sciences. Annual Review of Ecology and Systematics 19: 473–511.
Soepadmo E. 1972. Fagaceae. In: Van Steenis C.G.G.J. (ed.), Flora Malesiana, Ser.1, Vol.7, Pt.
2:265–403.
Sun W.B., Zhou Y., Zhao J.C. and Chen G. 2004. Current distribution,population attributes and
biological characters of Trigonobalanus doichangensis in relation to its conservation. Acta Eco-
logia Sinica 24: 352–348.
Wang P.L., Pu F.T. and Zheng Z.H. 1998. Palynological evidence for taxonomy of Trigonobalanus
(Fagaceae). Acta Phytotaxonomica Sinica 36: 238–241.
Wang P.L. and Zhang J.T. 1988. On the pollen morphology and systematic position of Trigono-
balanus doichangensis. Acta Phytotaxonomica Sinica 26: 44–46.
Whelan E.D.P. 1974. Discontinuities in the callose wall, inter-meiocyte connections, and cytomixis
in angiosperm meiocytes. Canadian Journal of Botany 53: 1219–1224.
Wu S.M. and Xiao S.Q. 1989. Comparative anatomical studies on the woods of Castanea Mill. and
Trigonobalanus Forman in China. Guihaia 22: 567–571.
Xia R.X., Ma M.T. and Mao Q.S. 1989. Observation of some features of staminate inflorescence
and pollen of the major chestnut cultivars in Luotian county. Journal of Huazhong Agricultural
University 8: 189–191.
[257]
1318
[258]
Biodiversity and Conservation (2006) 15:1319–1338 Springer 2006
DOI 10.1007/s10531-005-3875-5
-1
Key words: Biodiversity conservation, Geographical information systems, Ghyll woodlands, The
Weald, Woodland characterisation
Abstract. Ghylls are linear valley features cut into the sandy beds of the Weald of south-eastern
England. The ghyll’s indigenous woodlands are highly species rich at the small scale, support
distinctive assemblages of cryptogamic plants, and are unique to south-east England. Field surveys
were carried out for 48 ghyll woodlands in the Weald with a GIS used to examine the ecology,
landform and conservation status of the ghyll woodlands. The data were analysed using spatial and
multi-variate techniques in order to identify sub-groups or ghyll woodland communities based
upon species composition, topography and geology. The ghylls are shown to be reasonably uniform
for canopy vegetation type and structure and for their geological and soil characteristics. However,
analysis shows that geomorphology, understorey and field layer variability may act as stronger
indicators of site conditions and character. Further analysis focused on the level and extent of
nature conservation protection that these unique and ancient systems receive. The study concludes
that despite their ecological importance and potentially international significance, ghyll woodlands
are poorly understood and protected.
Introduction
[259]
1320
1
Coppicing is the productive management of woodland. Trees are cut down and encouraged to
grow again from the stump. Coppicing produces a large number of thin stems, which are harvested
on a regular cycle of about 5–15 years.
[260]
1321
(Gallois 1965). This variability in geology and unusual micro-climate has clear
implications for ecological diversity, and is considered to influence the high
levels of bryophyte diversity observed in ghyll systems (Rose and Patmore
1997).
The study presented within this paper constitutes the first attempt to use
geographical information system (GIS) approaches and contemporary surveys
to analyse and characterise these unique systems. The analysis focuses on the
geomorphological and ecological characteristics of the ghylls, but also con-
siders how the ghyll systems relate to existing conservation provision. GIS-
based cadastral approaches and multi-variate statistical techniques are used to
compare and analyse environmental, management and species information
derived from field surveys undertaken for a sample of ghyll woodlands within
southeast England.
Methodology
GIS techniques coupled with recent field survey data and secondary data
derived from digital maps have been used in an inductive study to analyse a
range of environmental and ecological characteristics for ghyll woodlands
within the central Weald.
The study used boundary data for 1130 ghylls in the Weald. The development
of the GIS required the collation of large amounts of secondary data held in
digital and non-digital form. Data, such as ghyll woodland boundaries, nature
conservation designations, topography and soil associations, were obtained in
digital form. The ghyll woodland boundary data were compiled from field
reports held at the regional Biological Records Centre. The statutory nature
conservation designations examined included the boundaries of Sites of
Special Scientific Interest (SSSI), Special Areas of Conservation (SAC),
National Nature Reserves (NNR), and also ancient woodland (AW). The
resultant GIS was structured around geographic and ecological landscape
features.
Field survey
This study has used data from field surveys of 48 ghyll woodland systems
throughout East and West Sussex (Figure 1). Using a uniform walkover survey
approach, field survey data recorded both biotic and abiotic characteristics of
the ghylls (Table 1). The data were used to construct a GIS database con-
taining site-specific information on the ecology, landform and management of
[261]
1322
48 ghyll woodlands randomly selected from 1130 ghyll woodlands. Where there
were substantial differences in the vegetational or geomorphological structure
in individual ghylls, multiple surveys were undertaken so as to represent this
internal variation. The variables included in the survey reflected those con-
sidered important in characterising ghylls (Table 1), and were integrated in a
GIS and analysed using spatial and multi-variate techniques to identify sub-
groups or types of ghyll systems.
Table 1. Field survey information for the ghyll woodland sample in the Weald.
Characteristics Descriptors
[262]
1323
[263]
1324
Results
Spatial patterns
Analysis indicates that ghyll woodlands cover 9332 ha of East and West Sus-
sex, and are divided into 1130 patches (Table 2). The mean patch size (or
habitat area) of ghyll woodland fragments is 8 ha. The ghyll woodlands have a
high density of patches with 12 patches per 100 ha in the area of the Weald
examined (Table 2). The mean patch shape calculations show that the ghyll
woodland habitats (Shape Index = 17) are not uniform in shape (i.e. patch
shapes move away from being regular shapes such as squares or circles). The
calculation of mean nearest neighbour distance (NND) between ghyll wood-
land habitats gives values of 217 m (±250.5). A number of ghylls are however
joined by contiguous areas of other woodland (managed and unmanaged).
Thus, care must be taken to ensure that conclusions are not drawn regarding
potential fragmentation effects (Wiens 1989; Bailey et al. 2002).
Conservation designations
Table 2. Spatial statistics for the Ghyll woodlands within the Central Weald of East and West
Sussex (raster format).
[264]
1325
Figure 2. The relationship between ghyll woodland systems and English Nature landscape
character areas.
designations such as area of outstanding natural beauty (AONB) and the HW.
Comparison of the distribution of ghyll woodlands and LCA classification
(English Nature 2002) reveals that 92% of ghyll systems are associated with the
HW LCA. A further 6% are found in the Low Weald LCA and, as might be
anticipated, only 2% are found in the Romney Marsh LCA (Figure 2).
GIS data and field surveys revealed that Quercus robur – Pteridium aquilinum
– Rubus fruticosus woodland (NVC – W10) and Fraxinus excelsior – Acer
campestre – Mercurialis perennis woodland (NVC – W8) were the most
common woodland community types found. Of the woodland surveyed, 47%
was found to be predominantly W10, and a further 43% dominated by W8
woodland. The remaining areas were comprised of W7, W12, W15 and some
pockets of W4, W6 and W16 woodland (Rodwell 1998). Analysis using multi-
variate statistical approaches (Cluster analysis and PCA) was performed on
the digital survey records. This provided a descriptive means of assessing
similarity and differentiation both within individual ghylls and between dif-
ferent ghyll woodland sites. Cluster analysis was initially undertaken for the
[265]
1326
three vegetation components surveyed; the canopy, the understorey and the
field layer.
The canopy
The canopy data shows that overall there is a moderately good level of
similarity in the composition of the canopy within the sampled ghyll wood-
land systems (Figure 3a). Values of between 38 and 55% similarity were
obtained when comparing the overall data set and the various sub-groupings.
When clusters were grouped, analysis showed an increase in the similarity
values with groups 1–3 and 8 between 50 and 60% similar whilst groups 4–7
showed similarity values greater than 60% (see appendices). Within these
groups, one major anomaly was seen at Fairlight Glen, Hastings which was
clearly identified as an outlier. This is because the canopy of this southern,
more coastal, section of Fairlight Glen ghyll is solely dominated by Salix
species.
Further analysis identified some clear ghyll woodland associations, all
showing comparatively high similarity values (>64%). The data may suggest
that there are some ‘canopy’ based associations within these ghyll woodlands.
For example, Group 1 is shown to be 42% similar to the other ghyll woodlands
surveyed yet, there is a moderately high level of canopy similarity within the
group itself (56%) (see appendices).
More specifically, two classes are identified and, of these, the two Brick Kiln
ghyll woodland areas are identical in canopy structure. Furthermore, the
analysis picks out two, somewhat individual, ghyll woodlands based upon
canopy composition, Kiln Wood (Lower ghyll) (no. 11) and Sandyden Wood
(no. 49). These woodlands are shown as only 38% similar to all other ghyll
woodlands surveyed and have no comparative ghylls at the group or class level
(Figure 3a).
The understorey
The data shows that, generally, there is similarity in the composition of the
understorey within the ghyll woodland systems (Figure 3b). Values of between
31 and 61% similarity were obtained when comparing the overall data set and
the various sub-groupings provided through cluster analysis. The analysis at
the group level showed a substantial increase in the similarity values with all
groups, apart from group 1, greater than 60% similar. These findings may
indicate some potential categories or communities.
Analysis at the class level shows strong association within the ghyll
woodland systems, with similarity values around 75–100%. The analysis
shows that when the understorey vegetation is examined many of the ghyll
sites display little difference (e.g. Marline Valley, Batemans and Northlands
[266]
1327
Figure 3. Cluster analysis dendrogram of the ghyll woodland canopy (a), under-storey (b), and
field (c) survey data. A complete table of percentage similarity values derived from cluster analysis
is provided in the appendix.
[267]
1328
Wood). It is, however, interesting to note that the analysis shows that
Marline valley does have some zonation within the ghyll – reflected in the
fact that two surveys were conducted at the site. Marline valley (away from
stream) and Marline Valley (near stream) are identical, as are Marline Valley
(by watercourse) and Marline Valley (away from watercourse). When com-
paring these two groups the analysis only shows a similarity value of 47%
between them (Figure 3b).
The analysis isolates Waterfall Wood as relatively distinctive. When the GIS
database and survey sheets are examined, this separation appears to be an
artefact of the understorey composition. Waterfall wood is the only ghyll with
the understorey solely dominated by Alnus species. Highams Ghyll and
Sandyden Wood also have Alnus species present within the understorey;
however, closer inspection shows that Highams also has Crataegus monogyna
dominant and Ilex aquifolium occasional within the understorey, whilst
Sandyden Ghyll has Sambucus nigra frequent within the understorey and
Corylus avellana occasional.
The analysis of the field layer data using cluster analysis shows that some data
sets appear very dissimilar with values ranging from 21 to 39% similarity
(derived from 11 site surveys). In contrast, 48 of the remaining sites show high
levels of similarity (Figure 3c). When considering the cluster analysis for field
data, the dendrogram clearly shows that those sites to the right of the figure are
on the whole largely dissimilar whilst those sites to the left of the figure reveal
moderately good levels of similarity, >50% (Figure 3c).
When grouped, over half of the sites show similarity levels of 62%. Fur-
thermore, within group 1, for example, the associations show 13 sites with
80% similarity and a further 15 sites with 81% similarity (see appendices).
Despite these high levels of field layer similarity within the majority of ghyll
systems, there are additional ghyll systems which are shown to be relatively
distinctive in character and field layer composition. Sandyden Wood and
Waterfall Wood are shown to be 50% similar to each other in presence/
absence, but are only 39% similar to all other ghyll woodlands surveyed.
Wicks Copse and Tilsmore Wood are also outliers with similarity values of
25 and 21% respectively (Figure 3c). Wren’s Warren is also distinct, sepa-
rated from other ghylls by the presence of Vaccinium spp. within the field
layer (see appendices).
Cadastral analysis via the GIS demonstrates that the ghyll woodlands occupy a
range of soil types. At a general level, GIS analysis indicates that 85% of the
[268]
1329
ghyll woodlands are associated, to some extent, with brown soils and partic-
ularly with argillic brown earths of the Curtisden association (Association 5.72i
of Jarvis et al. 1984). This association is widespread throughout the HW,
occurring on siltstones and sandstones. In this association, slope relief is typ-
ically strong with moderately to steeply sloping valleys [ghylls] separated by
gently sloping interfluves. Soils of the Curtisden association are usually mod-
erately deep on gentle or moderate slopes. Characteristically, the soils comprise
slowly permeable compact subsoil, which is subject to seasonal water logging.
Springs and flushes are often common on the sloping ground of permeable and
impermeable strata (Jarvis et al. 1984).
Ghyll woodlands are also commonly found on surface-water gley soils
(Stagnogley Associations 7.11e, 7.11i and 7.12b of Jarvis et al. 1984), which
are linked, again to varying degrees, with 55% of ghyll systems. The
Wickhams 1 and 5 Associations are extensive on the Low Weald. The
Wickham Associations have slowly permeable subsoils and are often water-
logged for prolonged periods in winter. Often heavily wooded, these soils are
naturally acidic and can have good reserves of available water (Jarvis et al.
1984). Other associations are linked with the ghyll systems but their
importance is low. Podzolic soils, in particular gley podzolic soils, are
associated with 4% of the ghyll systems. Only 1% are associated with ground
water gley soils.
Cluster analysis was undertaken for the geological components of the surveys,
including surrounding geology, 1st, 2nd, 3rd Geological Beds (in chronological
sequence), height at top of ghyll, and height difference within the ghyll valley.
The geological and topographical data shows substantial level of similarity
within the ghyll woodland systems with 89% of the sites being >68% similar
(Figure 4). All reside on the clays and sands of the Hastings Beds. Beyond this
level of similarity there are some additional sub-groupings with particularly
strong associations yielding similarity values >80% (Figure 4), which reflect
the presence of ghylls on Ashdown sands, Tunbridge sands, Wadhurst clays
and Fairlight clays.
Figure 4a also shows that some sites appear relatively distinctive in com-
parison to the majority of ghylls surveyed. Site 41 (Guestling Wood) shows
low similarity values of 28% to all other ghyll systems examined. Two of the
Fairlight Glen surveys (Sites 46 and 27) show good similarity when compared
to each other (87%), but are only 41% similar to all other systems. Equally,
Courtlands Wood (Site 5) and Little Iwood (Site 10) are similar to one
another (90%) but are only 54% similar to all other survey sites. Examina-
tion of the GIS database and survey sheets suggests that these differences are
related to surrounding geology, strata and average levels of fall along the
ghyll. Further cluster analysis was undertaken for the geological components
[269]
1330
Figure 4. Cluster analysis dendrogram of the ghyll woodland geological (a) and geological and
vegetation survey data combined (b).
and the vegetational data from the ghyll woodland surveys combined (Figure
4b). The dendrogram repeats patterns observed within the previous cluster
analysis showing a strong association (>62%) between the majority of ghyll
woodland sites surveyed (over 80% of sites). In addition, the dendrogram
isolated the same sites as being distinctive including Fairlight Glen, Guestling
Wood, Courtlands Wood, Little Iwood and Marline Valley (Figure 4b). This
similarity of outcomes may be associated with the dominance of geological
data within the analysis of both data matrices and the relatively weak
[270]
1331
PCA analysis of the geological and vegetational survey data reinforces the
conclusions drawn from the cluster analysis. Analysis shows high levels of
association between the majority of ghyll systems based upon the vari-
ables (shown by the main grouping on the ordination, Figure 5). This
suggests that the majority of ghyll systems have similar geological and
vegetational characteristics and little or no separation is present along
either axis.
It is however, interesting to note that there are two ghylls separated from
the main grouping along the first component showing 24% of the variation
(eigenvalue = 2.419) (Figure 5). Both Fairlight Glen and Marline Valley
(near watercourse) are separated along this axis and examination of the GIS
database suggests that this relates to strong differences in their geological
Figure 5. PCA analysis ordination of the ghyll woodland geological and vegetational survey data.
[271]
1332
beds, valley form and average degree of fall along the ghyll. Equally, a
further three ghylls are separated from the main grouping along the second
component showing 20% of the variation (eigenvalue=2.002). Guestling
Wood, Courtlands Wood and Little Iwood are separated here and the survey
data again points to geological conditions although some differences are
evident within their National Vegetational Classifications (Rodwell 1998)
(Figure 5).
Discussion
[272]
1333
[273]
1334
Conclusion
Acknowledgements
Special thanks go to Neil Carrett for his assistance and input in the project, and
Colin Reader (Habitat Management Services) and John Patmore (University
of Brighton) for conducting the surveys. We would also like to thank David
Saunders and Laetitia Tual (East Sussex County Council) for their support and
provision of additional data for this project.
GIS analysis was performed on ArcView 3.x software, Environmental Sys-
tems Research Institute and FragStats (ver 2), McGarigal and Marks (1992).
[274]
1335
Statistical analysis performed using WinSTAT 3.1, Kalmia Co. Inc., 1995.
Terrain data supplied by EDINA Edinburgh University, 2002; soil data sup-
plied by SILSOE, Cranfield University, 2002; forestry and woodland data
supplied by Forestry Commission, 2002; English Nature GUI, 2002, and
Sussex Biological Records Centre, 2001.
Appendices
[275]
1336
[276]
1337
References
Bailey S.A., Haines-Young R.H. and Watkins C. 2002. Species presence in fragmented landscapes:
modelling of species requirements at the national level. Biological Conservation 108: 307–316.
Brandon P. 1977. The Sussex Landscape. Hodder and Stoughton, London.
Burke D.M. 1998. Edge and fragment size effects on the vegetation of deciduous forests in Ontario,
Canada. Natural Areas Journal 18: 45–53.
Burnside N.G., Carrett N.J., Metcalfe D. and Waite S. 2002. GIS analysis of Ancient and Ghyll
Woodlands within East Sussex. Biogeography & Ecology Research Group, University of
Brighton, Brighton.
Cleere H. and Crossley D. 1985. The Iron-industry of the Weald. Leicester University Press,
Leicester.
Countryside Commission. 1994. The High Weald. Countryside Commission, Northhampton.
English Nature. 2002. Landscape Character Areas, Data prepared by English Nature GUI. URL
http://www.english-nature.org.uk/.
Ernoult A., Bureau F. and Poudevigne I. 2003. Patterns of organisation in changing landscapes:
implications for the management of biodiversity. Landscape Ecology 18(3): 239–251.
ESRI. 1998. ArcView 3.2, Software System. Environmental Systems Research Institute Inc.
Everitt B.S., Landau S. and Leese M. 2001. Cluster Analysis, 4th ed. Arnold, London.
[277]
1338
Forestry Commission. 2001. National Inventory of Woodland and Trees: England. Forestry
Commission, Edinburgh.
Fowler J., Cohen L. and Jarvis P. 1999. Practical Statistics for Field Biology. Wiley, Chichester,
West Sussex.
Gallois R.W. 1965. British Regional Geology, 4th ed. Department of Scientific and Industrial
Research, London, UK.
Gkaraveli A., Williams J.H. and Good J.E.G. 2001. Fragmented native woodlands in Snowdonia
(UK): assessment and amelioration. Forestry 74(2): 89–103.
Haines-Young R. and Chopping M. 1996. Quantifying landscape structure: a review of landscape
indices and their application to forested landscapes. Progress in Physical Geography 20: 418–445.
Harrison S. and Bruna E. 1999. Habitat fragmentation and large-scale conservation: what do we
know for sure? Ecography 22: 225–232.
Hodgetts N.G. 1997. A National and International Context for the Cryptograms in the Weald with
Reference to Current and Future Conservation Initiatives for UK Cryptograms. In: Jackson A.
and Flanagan M. (eds), Conservation of Cryptogams in the Weald. Proceedings of the Work-
shop held at Wakehurst Place 2nd May 1996. Royal Botanic Gardens, Kew pp.1–18.
Jarvis M.G., Allen S.J., Fordham S., Hazelden J., Moffat A.J. and Sturdy R.G. 1984. Soils and
their Use in South East England. Soil Survey, Harpenden.
Kent M. and Coker P. 1992. Vegetation Description and Analysis a Practical Approach. John
Wiley and Sons, Chichester.
Kirby K.J. 2003. Woodland conservation in privately-owned cultural landscapes: the English
experience. Environmental Science and Policy 6(3): 253–259.
McGarigal, K. and Marks B.J. 1994. FRAGSTATS Spatial Pattern Analysis Program for Quan-
tifying Landscape Structure, Version 2. Forest Science Department, Oregon State University,
Corvallis.
Peterken G.F. 1981. Wood anemone in central Lincolnshire: an ancient woodland indicator?
Transactions Lincolnshire Natural Union 20: 78–82.
Peterken G.F. 1993. Woodland Conservation and Management, 2nd ed. Chapman and Hall,
London.
Rackham O. 1980. Ancient Woodland. Arnold, London.
Ratcliffe D.A. 1968. An ecological account of the Atlantic bryophytes in the British Isles. New
Phytologist 67: 365–430.
Reid C.M., Kirby K.J. and Cooke R. 1996. A Preliminary Assessment of Woodland Conservation
in England by Natural Area. English Nature, Peterborough, UK.
Rodwell J.S. 1998. British Plant Communities Volume 1: Woodlands and Scrub. Cambridge
University Press, Cambridge.
Rose F. 1995. The Habitats and Vegetation of Sussex. The Booth Museum of Natural History,
Brighton Borough Council.
Rose F. and Patmore J.M. 1997. Gill Woodlands in the Weald. English Nature, Peterborough.
SSLRC. 2001. Sussex Soil Associations. Data prepared by National Soil Resources Institute
(SSLRC), Digital Data Copyright (c) Cranfield University, 2001 (SSLRC Project Code: JP7017v/
9).
Sussex Biodiversity Partnership. 2000. Woodland Habitat Action Plan. Sussex Biodiversity Part-
nership, Sussex.
Waller M.P. and Marlow A.D. 1994. Flandrian vegetational history of southeastern England –
stratigraphy of the Brede Valley and pollen data from Brede Bridge. New Phytologist 126(2):
369–392.
Waite S. 2000. Statistical Ecology: A Practical Guide. Prentice Hall, Harlow, UK.
Wiens J.A. 1989. The Ecology of Bird Communities, Vol. 2. Cambridge University Press, New
York.
Woodland Trust. 2000. Woodland Biodiversity: Expanding Our Horizons. Woodland Trust,
Grantham, Lincs.
Woolridge S.W. and Goldring F. 1962. The Weald, 3rd ed. Collins, London. UK.
[278]
Biodiversity and Conservation (2006) 15:1339–1351 Springer 2006
DOI 10.1007/s10531-005-4875-1
-1
Key words: Ardisia crenata var. bicolor, Differentiation, Forest fragmentation, Genetic diversity,
Population size, RAPD markers
Abstract. Due to the long generation times and high densities, dominant tree species usually did
not respond consistently with theoretical predictions to the recent fragmentation. Genetic struc-
tures of shrubs and herbs, especially those with low densities, may be more sensitive to forest
fragmentation. We studied the genetic structure of a self-compatible subshrub, Ardisia crenata var.
bicolor (Myrsinaceae) in a recently fragmented landscape. Ten RAPD primers used for analysis
generated a total of 76 bands. We found that A. c. var. bicolor had relatively low species-level
(P95 = 63.2%; H = 0.106; Shannon diversity index (SI) = 0.246) and within-population diversity
(P95 = 5.346.1%; H = 0.0260.175; SI = 0.0320.253), and significant population differenti-
ation (GST = 0.445). Significantly positive relationships were found between measures of diversity
(P95, H and SI) and the log of estimated population size. No significant relationship was observed
between Nei’s genetic distance and spatial distance of pairwise populations, indicating no isolation-
by-distance. Given most species of forests are shrubs and herbs with short generation times, our
observation indicated that distinct genetic consequences of recent fragmentation may be expected
for quite a number of plant species.
Introduction
[279]
1340
tiation have been found in some systems (e.g., Raijmann et al. 1994; Hall
et al. 1996; Morden and Loeffler 1999; Frankham et al. 2002). However,
these responses to increased fragmentation are unlikely to be common, and
other factors may influence the genetic consequences of fragmentation. First,
most studies were conducted on long-lived species in a recently fragmented
landscape and there were not sufficient time for bottleneck and inbreeding to
take action (Young et al. 1993; Cardoso et al. 1998). Long-lasting, dormant
seed banks also can buffer against genetic effects for decades or centuries
(Morris et al. 2002). Second, many plant species studied were dominant
species with high density. Thus, populations in fragmented habitats were
large enough to maintain relatively high genetic diversity. Thirdly, some
species are naturally rare species, and have evolved mechanisms to overcome
the disadvantages of small population size. Thus, genetic consequences of
recent fragmentation may not be detectable for a long time (England et al.
2002).
Shrubs and herbs constitute the main part of species composition of
forests. Therefore, genetic effects observed in dominant species of forests
might not be general for most forest species. Shrubs and herbs – especially
those with low densities – may be more genetically sensitive to forest frag-
mentation because of their much shorter life span. A shorter life span means
they pass many generations even in recently fragmented habitats and, given
their low densities, they therefore experience large declines in population size
even if the fragmentation is not serious. However, much fewer studies have
been conducted on herbs and shrubs than on tree species. For example,
populations of herbaceous Swertia perennis in small, isolated habitats had
reduced genetic variability and the highest within-population inbreeding
coefficients (Lienert et al. 2002). In the herb, Scutellaria montana, popula-
tions that were less than 100 individuals tended to have lower proportions of
polymorphic loci than that of populations more than 100 individuals (Cru-
zan 2001).
Ardisia (Myrsinaceae) is a tropical and subtropical genus and includes
about 200 species. Coral ardisia, A. crenata, native to Japan to north India, is
an insect-pollinated and self-compatible evergreen subshrub (Cheon et al.
2000). In China, a variant, A. crenata var. bicolor, was identified according to
the purple color of the lower side of its leaves, whereas some researchers
thought it as a distinct species, i.e. A. bicolor. A. crenata var. bicolor is a small
upright-growth shrub. Although outcrossing rate of A. c. var. bicolor was
estimated to be about 1 based on allozyme using Ritland’s (1990) MLT
program (Chen et al. 2001), bag-pollination treatments indicated that it is self-
compatible (unpublished data). This species can reproduce vegetatively via
rhizome, but spreading to a short distance, usually less than 1 m (personal
observations).
In the present study, populations of A. c. var. bicolor in a fragmented
landscape were selected to determine whether there is a relationship between
population size and the level of variation and to evaluate the degree of
[280]
1341
Methods
Population sampling
The study sites were located in Tiantong Forest Park (TFP) and adjacent
areas (Figure 1). TFP was distributed by evergreen broad-leaved forests
(EBLFs) dominated by Fagaceae species, such as Castanopsis fargesii, Ca.
carlesii, Ca. sclerophylla, Lithocarpus glaber, L. henryi, Cyclobalanopsis nubi-
um, and species of Theaceae (Schima superba) and Lauraceae (Machilus
thunbergii) (Song and Wang 1995). Around TFP, there were EBLFs fragments
of previous continuous forests or recovered from abandoned or unmanaged
plantations. These fragmented EBLFs were usually dominated by Cycloba-
lanopsis glauca, Cy. gilva, Ca. sclerophylla, L. glaber, Ca. carlesii, M. thun-
bergi. Surrounding these EBLFs, there were Cunninghamia lanceolata
plantations, Phyllostachys pubescens forests, and shrubs dominated by Quer-
cus fabra and bamboos.
In TFP and adjacent areas, A. c. var. bicolor usually appears in forests of
lower than 300 m above sea-level. Based on detailed surveys, 10 populations of
A. c. var. bicolor were sampled (Figure 1). The estimated sizes of each popu-
lation ranged from 5 to about 1000 individuals (Table 1). Leaves were col-
lected randomly from individuals with a distance of at least 2 m between each
other in medium and large populations, avoiding collecting the same clones. In
small populations, as many as possible individuals were collected with a dis-
tance of at least 2 m between sampled individuals.
We isolated DNA with modified Doyle and Doyle’s (1987) procedure (Fan
et al. 2004). A set of random 10-mer primers was purchased from Sagon Inc.,
Shanghai. After screening more than 100 arbitrary primers, 10 primers that
consistently amplified clear banding patterns were chosen for further studies
(Table 1). RAPD assays were performed using the conditions described by Fan
et al. (2004). Samples were amplified at least two replicates and same pattern
was obtained by the primers used in this study. Five ll amplification product
was separated on 1.6% agarose gel in 0.5· TBE buffer and visualized by
[281]
1342
Figure 1. Locations of sampling sites of Ardisia crenata var. bicolor in Tiantong Forest Park and
adjacent areas.
[282]
1343
Table 1. RAPD primers used in the survey of Ardisia bicolor and number of scored bands.
staining with ethidium bromide and photographed under UV light with Bio-
RAD Gel Doc2000TM.
Data analysis
Each PCR product was assumed to represent a single locus and was scored for
presence and absence. The resulting data matrix was analyzed using Popgene
1.31 (Yeh et al. 1999). Gene diversities (H) at population and at the species
level were calculated based on Lynch and Milligan’s (1994) Taylor expansion
estimate using TFPGA (Tools For Population Genetic Analyses) v1.3 (Miller
1997).
Nei’s unbiased genetic identity (I) and genetic distance (D) between popu-
lations were also analysed using Popgene 1.31. Because the data were larger
than the up-limit of AMOVA, coefficient of gene differentiation (GST) was
calculated to estimate population
P differentiation. Shannon diversity index
(Lewontin 1972), SI = pi log2 pi, was calculated to provide a relative
estimate of the degree of variation at population and species levels using
Popgene 1.31 (Yeh et al. 1999). The proportion of diversity among populations
was estimated as (SIspSIpop)/SIsp, whereas SIsp and SIpop were SI at species
and population level, respectively. A Mantel type matrix randomization test
(Mantel 1967) was performed to evaluate the relationship between the matrix
of genetic distances and the matrix of geographic distances using TFPGA
(Miller 1997).
Relationships between measures of within-population genetic variation and
population size were analyzed using Regression methods in Microsoft Excel
program.
Results
The 10 primers used for analysis generated a total of 76 bands, among which
polymorphic bands were 48 (or 63.2%) and 51 (or 67.1%) based on 95 and
[283]
1344
100% criteria, respectively (Table 2). The number of scored bands ranged from
3 for primer S1221 to 11 for primer S1238. 308 of the 320 individuals from the
16 populations were found to have a unique multilocus genotype, and six
genotypes have two individuals. The individuals having the same multilocus
genotypes belonged to same populations. No population-specific band was
observed in the data set.
Genetic diversity
[284]
1345
Genetic differentiation
[285]
1346
Table 3. Nei’s unbiased genetic identity (below diagonal) and genetic distance (above diagonal) between populations of Ardisia crenata var. bicolor.
Population A B C D E F G H I J K L M N O P
A – 0.0372 0.0361 0.0212 0.0441 0.0254 0.1092 0.0429 0.0558 0.0603 0.0345 0.0383 0.0354 0.0382 0.0581 0.0445
B 0.9635 – 0.0397 0.0380 0.0442 0.0481 0.0893 0.0300 0.0552 0.0538 0.0300 0.0490 0.0296 0.0247 0.0676 0.0606
C 0.9646 0.9611 – 0.0324 0.0505 0.0452 0.0959 0.0372 0.0479 0.0655 0.0418 0.0626 0.0464 0.0296 0.0716 0.0631
D 0.9790 0.9627 0.9682 – 0.0426 0.0370 0.0922 0.0395 0.0532 0.0589 0.0398 0.0393 0.0356 0.0331 0.0462 0.0530
E 0.9569 0.9567 0.9507 0.9583 – 0.0389 0.0895 0.0338 0.0707 0.0574 0.0471 0.0725 0.0519 0.0281 0.0287 0.0457
F 0.9749 0.9530 0.9558 0.9636 0.9618 – 0.0953 0.0294 0.0718 0.0548 0.0499 0.0624 0.0414 0.0456 0.0660 0.0584
[286]
G 0.8965 0.9145 0.9085 0.9119 0.9144 0.9091 – 0.0892 0.1155 0.0858 0.0894 0.1216 0.0827 0.0608 0.0722 0.1487
H 0.9580 0.9705 0.9635 0.9613 0.9668 0.9711 0.9147 – 0.0411 0.0555 0.0397 0.0676 0.0385 0.0339 0.0526 0.0341
I 0.9458 0.9463 0.9532 0.9482 0.9317 0.9307 0.8909 0.9597 – 0.0732 0.0570 0.0771 0.0562 0.0526 0.0754 0.0447
J 0.9415 0.9477 0.9366 0.9428 0.9442 0.9467 0.9178 0.9460 0.9294 – 0.0184 0.0397 0.0387 0.0398 0.0713 0.1064
K 0.9661 0.9704 0.9590 0.9610 0.9540 0.9513 0.9145 0.9611 0.9446 0.9818 – 0.0304 0.0290 0.0296 0.0609 0.0789
L 0.9625 0.9522 0.9393 0.9614 0.9301 0.9395 0.8855 0.9346 0.9258 0.9610 0.9701 – 0.0257 0.0608 0.0926 0.1034
M 0.9652 0.9709 0.9547 0.9650 0.9494 0.9595 0.9206 0.9623 0.9454 0.9620 0.9714 0.9746 – 0.0413 0.0755 0.0876
N 0.9625 0.9756 0.9709 0.9675 0.9723 0.9555 0.9411 0.9667 0.9488 0.9609 0.9708 0.9410 0.9595 – 0.0518 0.0565
O 0.9435 0.9347 0.9309 0.9548 0.9717 0.9361 0.9303 0.9488 0.9274 0.9312 0.9409 0.9115 0.9273 0.9495 – 0.0536
P 0.9565 0.9412 0.9389 0.9484 0.9553 0.9433 0.8618 0.9665 0.9563 0.8991 0.9242 0.9018 0.9161 0.9451 0.9478 –
1347
Discussion
The present study reveals relative low genetic diversity in A. c. var. bicolor
compared to other species based on RAPD markers. Percentage of poly-
morphic bands, Nei’s gene diversity and Shannon index of the pooled data
were 67.1%, 0.192 and 0.246, respectively (Table 2). These values were even
lower than many endangered species, Metasequoia glyptostroboides
(P: 87.9%, H=0.318, SI = 0.476) (Li et al. 2005), Caesalpinia echinata
(P = 95.7%) (Cardoso et al. 1998), Leucadendron elimense (P = 98.8%)
(Tansley and Brown 2000), Boloria aquilonaris (H = 0.402) (Vandewoestijne
and Baguette 2002), but higher than Haplostachys haplostachya (H = 0.166)
(Morden and Loeffler 1999), Dryopteris cristata (P = 2.5%) (Landergott
et al. 2001).
Our results were not in accordance with the predictions based on the asso-
ciation of life history traits and genetic variation. According to the data of
RAPDs, there were strong associations between genetic diversity and breeding
system or successional status (Nybom and Bartish 2000; Nybom 2004). Out-
crossing species had significantly high genetic diversity than selfers. Higher
genetic diversity was found in late- than early - successional species. Species of
ingested seed dispersal also possessed relatively high genetic diversity (Nybom
2004). Given such considerations, high genetic diversity was expected based on
its high outcrossing rate (Chen et al. 2001), mid to late-successional status and
bird dispersal manner.
Relatively low genetic diversity in the studied A. c. var. bicolor populations
might be explained by their geographical positions. The studied populations
are located on the eastern margin of its distribution in mainland China (Figure
1). Due to effects of founder events, genetic drift and inbreeding, marginal
populations usually possess relative low genetic variation, which had been
confirmed in diverse species (Chen et al. 1997; Tyler 2002; Cassel and Tam-
maru 2003). Restricted geographical range in the present study might be
another explanation of the low genetic diversity. Though 10 populations were
sampled, their spatial distances were small. The largest distance of pairwise
populations is 4.6 km. If populations were sampled in a large range, more
genetic variation might be expected.
Though some studies failed to observe the distinct genetic consequences of
forest fragmentation on plant populations (Ellstrand and Elam 1993; Young
et al. 1996), our results indicated that habitat fragmentation had played a vital
role in genetic structure of A. c. var. bicolor populations. Fragmentation led to
the loss of genetic diversity. In small populations, significantly lower diversity
was observed than large and medium ones. Significant relationship was found
between estimated population size and within-population genetic variation as
measured by P, H and Shannon index.
Decreased genetic diversity in small populations was due to various reasons.
Firstly, the instantaneous effects of fragmentation (i.e., sampling effects) lead
to stochastic loss of rare alleles because only a small portion of the original
[287]
1348
gene pool remains after the decrease. Buchert et al. (1997) had compared the
genetic diversity in pre-harvest and post-harvest gene pools of two virgin
stands of eastern white pine (Pinus strobus). They found total and mean
number of alleles was reduced by 25% after tree density reductions of 75%.
About 40% of the low frequency alleles and 80% of the rare alleles were lost
because of harvesting (Buchert et al. 1997). Secondly, inbreeding and genetic
drift further decreased genetic variation in small populations (Young et al.
1996). In the present study, most small populations experienced bottleneck for
more than 10 generations, given a generation of 3 years and deforestation of at
least 50 years. It is enough for inbreeding and drift to virtually change the
genetic composition of small populations of less than 50 individuals. Thirdly,
founder effect might also contribute to low genetic diversity in some small
populations (Frankham et al. 2002). Population G, for instance, located in
dense high shrubs, had only two small individuals, indicating a recent founding
event.
High genetic differentiation was observed among populations of A. c. var.
bicolor. GST indicated that about 44.5% of the genetic variation occurred
among populations with short spatial distances. This value is higher than the
RAPD-based estimates of other widespread, or animal-dispersal species
(Nybom and Bartish 2000). High genetic differentiation among populations
was also in accordance with theoretical prediction of fragmentation, indicating
the effects of bottleneck and inbreeding.
No significant relationship between genetic distance and spatial distance was
found in the present study. This is usually interpreted that selection or drift
plays a more significant role than gene flow. In this study, no distinct difference
in habitats was found among populations, though the dominant species were
different. At local scale, populations from different communities usually
showed similar genetic composition in studied species, such as, Cyclobalanopsis
glauca (Chen and Song 1998). Therefore, selection plays a minor role in the
differentiation of A. c. var. bicolor populations, and drift led by fragmentation
contributed to the high differentiation.
Our findings in A. c. var. bicolor give a gloomy implication for forest
species because most species of forests are shrubs and herbs and among them
most are moderate- or low-density species. For example, there were about a
dozen of species in tree layer of EBLFs; among them, usually less than 3
species dominated the community. However, the number of species in shrub
and herb layers was about three to four folds of that in tree layer (Song and
Wang 1995). Among them, most are moderate or low density, like A. c. var.
bicolor, and are vulnerable to fragmentation. This situation is also common
in tropical, temperate or boreal forests. Thus, although some studies showed
no distinct effects on long-lived tree species which have survived hundreds of
years of fragmentation, our study indicated that distinct genetic consequences
of recent fragmentation may be expected for quite a number of plant species.
More attention should be paid to these species and conservation efforts are
needed.
[288]
1349
Acknowledgements
References
Bacles C.F.E., Lowe A.J. and Ennos R.A. 2004. Genetic effects of chronic habitat fragmentation on
tree species: the case of Sorbus aucuparia in a deforested Scottish landscape. Molecular Ecology
13: 573–584.
Buchert G.P., Rajora O.P., Hood J.V. and Dancik B.P. 1997. Effects of harvesting on genetic
diversity in old-growth eastern white pine in Ontario, Canada. Conservation Biology 11: 747–758.
Cardoso M.A., Provan J., Powell W., Ferreira P.C.G. and de Oliveira D.E. 1998. High genetic
differentiation among remnant populations of the endangered Caesalpinia echinata Lam.
(Leguminosae-Caesalpinioideae). Molecular Ecology 7: 601–608.
Cassel A. and Tammaru T. 2003. Allozyme variability in central, peripheral and isolated popu-
lations of the scarce heath (Coenonympha hero: Lepidoptera, Nymphalidae); implications for
conservation. Conservation Genetics 4: 83–93.
Chen X.Y. 2000. Effects of habitat fragmentation on genetic structure of plant populations and
implications for the biodiversity conservation. Acta Ecologica Sinica 20: 884–892.
Chen X.Y., Li N. and Shen L. 2001. The mating system of Ardisia crenata var. bicolor (Myrsin-
aceae), a subtropical understory shrub, in Tiantong National Forest Park, Zhejiang Province.
Acta Phytoecologica Sinica 25: 161–165.
Chen X.Y. and Song Y.C. 1998. Microgeographic differentiation in a Cyclobalanopsis glauca
poplation in western Huangshan, Anhui Province. Journal of Plant Resources Environment
7: 10–14.
Chen X.Y., Wang X.H. and Song Y.C. 1997. Genetic diversity and differentiation of Cycloba-
lanopsis glauca populations in East China. Acta Botanica Sinica 39: 149–155.
Cheon C.P., Chung M.Y. and Chung M.G. 2000. Allozyme and clonal diversity in Korean pop-
ulations of Ardisia japonica and Ardisia crenata (Myrsinaceae). Israel Journal of Plant Science
48: 239–245.
Cruzan M. 2001. Population size and fragmentation thresholds for the maintenance of genetic
diversity in the herbaceous endemic Scutellaria montana (Lamiaceae). Evolution 55: 1569–1580.
Doyle J.J. and Doyle J.L. 1987. A rapid DNA isolation procedure for small quantities of fresh leaf
tissue. Phytochemical Bulletin 19: 11–15.
Ellstrand N.C. and Elam D.R. 1993. Population genetic consequences of small population
size: implications for plant conservation. Annual Review of Ecology and Systematics 24:
217–242.
England P.R., Usher A.V., Whelan R.J. and Ayre D.J. 2002. Microsatellite diversity and genetic
structure of fragmented populations of the rare, fire-dependent shrub Grevillea macleayana.
Molecular Ecology 11: 967–977.
Fan X.X., Shen L., Zhang X., Chen X.Y. and Fu C.X. 2004. Assessing genetic diversity of Ginkgo
biloba L. (Ginkgoaceae) populations from China by RAPD markers. Biochemical Genetics 42:
269–278.
Frankham R., Ballou J.D. and Briscoe D.A. 2002. Introduction to Conservation Genetics.
Cambridge University Press, Cambridge.
[289]
1350
Hall P., Walker S. and Bawa K. 1996. Effect of forest fragmentation on genetic diversity and
mating system in a tropical tree, Pithecellobium elegans. Conservation Biology 10: 757–768.
Keller L.F. and Waller D.M. 2002. Inbreeding effects in wild populations. Trends in Ecology and
Evolution 17: 230–241.
Landergott U., Holderegger R., Kozlowski G. and Schneller J. 2001. Historical bottlenecks
decrease genetic diversity in natural populations of Dryopteris cristata. Heredity 87: 344–355.
Lewontin R.C. 1972. The apportionment of human diversity. Evolutionary Biology 6: 381–398.
Li Y.-Y., Chen X.-Y., Zhang X., Wu T.-Y., Lu H.-P. and Cai Y.-W. 2005. Genetic differences
between wild and artificial populations of Metasequoia glyptostroboides Hu et Cheng (Taxodi-
aceae): Implications for species recovery. Conservation Biology 19: 224–231.
Lienert J., Fischer M., Schneller J. and Diemer M. 2002. Isozyme variability of the wetland spe-
cialist Swertia perennis (Gentianaceae) in relation to habitat size, isolation, and plant fitness.
American Journal of Botany 89: 801–811.
Lynch M. and Milligan B.G. 1994. Analysis of population genetic structure with RAPD markers.
Molecular Ecology 3: 91–99.
Mantel N. 1967. The detection of disease clustering and a generalized regression approach. Cancer
Research 27: 209–220.
Miller M.P. 1997. Tools for population genetic analyses (TFPGA) v1.3: A windows program for
the analysis of allozyme and molecular genetic data. Department of Biological Sciences Northern
Arizona University, Flagstaff.
Morden C.W. and Loeffler W. 1999. Fragmentation and genetic differentiation among subpopu-
lations of the endangered Hawaiian mint Haplostachys haplostachya (Lamiaceae). Molecular
Ecology 8: 617–625.
Morris A.B., Baucom R.S. and Cruzan M.B. 2002. Stratified analysis of the soil seed bank in the
cedar glade endemic Astragalus bibullatus: Evidence for historical changes in genetic structure.
American Journal of Botany 89: 29–36.
Nybom H. 2004. Comparison of different nuclear DNA markers for estimating intraspecific genetic
diversity in plants. Molecular Ecology 13: 1143–1155.
Nybom H. and Bartish I.V. 2000. Effects of life history traits and sampling strategies on genetic
diversity estimates obtained with RAPD markers in plants. Perspectives in Plant Ecology,
Evolution and Systematics 3/2: 93–114.
Raijmann L.E.L., Van Leeuwen N.C., Kersten R., Oostermeijer J.R.B., Den Nijs H.C.M. and
Menken S.B.J. 1994. Alleleic variation and outcrossing rate in relation to population size in
Gentiana pneumonanthe L. Conservation Biology 8: 1014–1026.
Ritland K. 1990. A series of FORTRAN computer programs for estimating plant mating systems.
Journal of Heredity 81: 235–237.
Saunders D.A., Hobbs R.J. and Margules C.R. 1991. Biological consequences of ecosystem frag-
mentation: A review. Conservation Biology 5: 18–32.
Song Y.C. and Wang X.R. 1995. Vegetation and Flora of Tiantong National Forest Park, Zhejiang
Province. Shanghai Scientific Documentary Press, Shanghai.
Tansley S.A. and Brown C.R. 2000. RAPD variation in the rare and endangered Leucadendron
elimense (Proteaceae): implications for their conservation. Biological Conservation 95: 39–48.
Tyler T. 2002. Large-scale geographic patterns of genetic variation in Melica nutans, a widespread
Eurasian woodland grass. Plant Syst. Evol. 236: 73–87.
Van Rossum F., De Sousa S.C. and Triest L. 2004. Genetic consequences of habitat fragmentation
in an agricultural landscape on the common Primula veris, and comparison with its rare con-
gener, P. vulgaris. Conservation Genetics 5: 231–245.
Vandewoestijne S. and Baguette M. 2002. The genetic structure of endangered populations in the
Cranberry Fritillary, Boloria aquilonaris (Lepidoptera, Nymphalidae): RAPDs vs allozymes.
Heredity 89: 439–445.
Yeh F.C., Yang R.C. and Boyle T. 1999. Popgene Version 1.31: Microsoft Window-Based Free-
ware for Population Genetic Analysis. Department of Renewable Resources, University of
Alberta, Edmonton, AB Canada.
[290]
1351
Young A., Boyle T. and Brown T. 1996. The population genetic consequences of habitat frag-
mentation for plants. Trends in Ecology and Evolution 11: 413–418.
Young A., Merriam H.G. and Warwick S.I. 1993. The effects of forest fragmentation on genetic
variation in Acer saccharum Marsh. (sugar maple) populations. Heredity 71: 277–289.
[291]
Biodiversity and Conservation (2006) 15:1353–1374 Springer 2006
DOI 10.1007/s10531-005-5394-9
-1
Key words: Biodiversity, Cameroon, Campo-Ma’an, Central Africa, Conservation, Endemic spe-
cies, Forest refuge, Plant diversity, Tropical rain forest
Abstract. This study describes diversity patterns in the flora of the Campo-Ma’an rain forest, in
south Cameroon. In this area, the structure and composition of the forests change progressively
from the coastal forest on sandy shorelines through the lowland evergreen forest rich in Cae-
salpinioideae with Calpocalyx heitzii and Sacoglottis gabonensis, to the submontane forest at higher
elevations and the mixed evergreen and semi-deciduous forest in the drier Ma’an area. We tested
whether there is a correlation between tree species diversity and diversity of other growth forms
such as shrubs, herbs, and lianas in order to understand if, in the context of African tropical rain
forest, tree species diversity mirrors the diversity of other life forms or strata. Are forests that are
rich in tree species also rich in other life forms? To answer this question, we analysed the family and
species level floristic richness and diversity of the various growth forms and forest strata within 145
plots recorded in 6 main vegetation types. A comparison of the diversity within forest layers and
within growth forms was done using General Linear Models. The results showed that tree species
accounted for 46% of the total number of vascular plant species with DBH ‡1 cm, shrubs/small
trees 39%, climbers 14% and herbs less than 1%. Only 22% of the diversity of shrubs and lianas
could be explained by the diversity of large and medium sized trees, and less than 1% of herb
diversity was explained by tree diversity. The shrub layer was by far the most species rich, with both
a higher number of species per plot, and a higher Shannon diversity index, than the tree and the
herb layer. More than 82% of tree species, 90% of shrubs, 78% of lianas and 70% of herbaceous
species were recorded in the shrub layer. Moreover, shrubs contributed for 38% of the 114 strict
and narrow endemic plant species recorded in the area, herbs 29%, trees only 20% and climbers
11%. These results indicate that the diversity of trees might not always reflect the overall diversity
of the forest in the Campo-Ma’an area, and therefore it may not be a good indicator for the
diversity of shrubs and herbaceous species. Furthermore, this suggests that biodiversity surveys
based solely on large and medium sized tree species (DBH ‡10 cm) are not an adequate method for
the assessment of plant diversity because other growth form such as shrubs, climbers and herbs are
under-represented. Therefore, inventory design based on small plots of 0.1 ha, in which all vascular
plants with DBH ‡1 cm are recorded, is a more appropriate sampling method for biodiversity
assessments than surveys based solely on large and medium sized tree species.
Introduction
[293]
1354
Methods
Study area
The study was conducted in the Campo-Ma’an rain forest in south Camer-
oon. The site covers about 7700 km2 and is located between latitudes 210¢–
252¢ N and longitudes 950¢–1054¢ E. The Campo-Ma’an area is a Tech-
nical Operational Unit (TOU) that comprises a National Park, five forest
management units, two agro-industrial plantations, and a multi-uses zone.
Following the FAO classification system, soils in the Campo-Ma’an area are
generally classified as Ferrasols and Acrisols (Franqueville 1973; Muller 1979;
van Gemerden and Hazeu 1999). They are strongly weathered, deep to very
[294]
1355
deep and clayey in texture (except at the seashores and in river valleys where
they are mainly sandy), acid and low in nutrients with pH (H2O) values
generally around 4. The topography ranges from undulating to rolling in the
lowland area, to steeply dissect in the more mountainous areas. In the
Campo area, altitudes are mostly low, ranging from sea level to about 500 m.
In the eastern part, which is quite mountainous, the altitude varies between
400 and 1100 m and the rolling and steep terrain brings about a more var-
iable landscape.
The area has a typical equatorial climate with two distinct dry seasons
(November–March and July–mid-August) and two wet seasons (April–June
and mid-August–October). The average annual rainfall generally decreases
with an increasing distance from the coast, ranging from 2950 mm/year in
Kribi and 2800 mm in Campo to 1670 mm in Nyabissan in the Ma’an area.
The Ma’an region has significantly less rainfall than other areas. The average
annual temperature is about 25 C and there is little variation between years.
The hydrography of the area shows a dense pattern with many rivers, small
river basins, fast-flowing creeks and rivers in rocky beds containing many
rapids and small waterfalls. Generally, the area has a low population density of
about 10 inhabitants per km2 and is sparsely populated (ca. 61,000 inhabitants)
with most people living around Kribi, along the coast, and in agro-industrial
and logging camps (ERE Développement 2002; de Kam et al. 2002). Despite
the low population density, there are few employment opportunities. The local
people are very poor and so far rely solely on the forest resources to meet their
basic needs. As a result, local pressure on the Campo-Ma’an rain forest is
increasing and there are several activities that are carried out in the area with
varying ecological impacts on the forest ecosystem. These activities include
agriculture, logging, poaching and hunting.
Field sampling
[295]
1356
coverage. Most of the plots were located in the National Park and the forest
management units, which are less affected by human activities.
In each 0.1 ha plot, all trees, shrubs, herbs and lianas with DBH ‡1 cm were
measured, recorded and identified as far as possible. For unknown species, a
voucher specimen was collected. Herbaceous species and seedlings of trees,
shrubs and climbers were sampled in subplots of 5 m · 5 m each that were
established in the 0.1 ha plots. These subplots were not used for the analyses,
the output was only used to illustrate the contribution of the ground layer and
herbaceous species when all vascular plant species are included in the floristic
assessment of the forest.
Data analysis
The analysis focused on family and species level floristic richness within the
various life forms and forest strata recorded in the 145 plots. In this study tree
layer comprised all vascular plant species with DBH ‡10 cm, shrub layer
(1.5 cm £ DBH < 10 cm) and herbaceous layer (1 cm £ DBH < 1.5 cm).
Diversity was measured by recording the number of species and their relative
abundance in the different plots and vegetation types. This study focused on
the a diversity (species richness), which is defined as the number of species
within a chosen area, given equal weight to each species, and the b diversity,
which is the difference in species diversity between areas or communities
(Magurran 1988; Kent and Coker 1992; Bisby 1995). b diversity was quantified
with the Shannon diversity index (H¢) using all individuals above 1 cm DBH
and all species per plot. Phytosociological parameters (relative density and
relative frequency) and Shannon diversity index were calculated following
Whittaker (1975), Kent and Coker (1992) and Magurran (1988). The SPSS
package version 10.0 for Windows was used for statistical analyses. The
Spearman’s correlation test was used to correlate the species richness and
diversity between the various growth forms and forest layers. We compared
species diversity within forest layers and within growth forms using a General
Linear Model (GLM) followed by a Tukey Multiple Comparison test
(p<0.05).
Results
A total of 76360 trees, shrubs, climbers and other vascular plants with DBH
‡1 cm was recorded in 145 plots of 0.1 ha each in the various vegetation types.
They belonged to 1112 species, 420 genera and 97 families. In addition, 759
species of vascular plants (herbs, hemi-epiphytes, shrubs and seedlings of tree
species) belonging to 101 families and 327 genera were recorded in the subplots
[296]
1357
Table 1. Summary of the number of species, number of families, number of stems/ha and Shannon
diversity (H¢) recorded in each vegetation type for all vascular plants with DBH ‡ 1 cm.
The number of stems/ha and the number of vascular plant species per plot were
generally higher in the shrub layer compared to the herbaceous layer and the
tree layer. The number of stems/ha (Table 2) in the shrub layer varied from
3914 (mixed evergreen and semi-deciduous forest) to 4572 (coastal forest), in
the herbaceous layer from 905 (swamps) to 1963 (submontane forest), and in
the tree layer from 489 (coastal forest) to 785 stems/ha (submontane forest).
The number of species in the shrub layer varied from 231 species (swamps) to
413 (mixed evergreen and semi-deciduous forest), in the herbaceous layer from
99 (swamps) to 229 (Calpocalyx heitzii and Sacoglottis gabonensis forest) and in
the tree layer from 100 (swamps) to 183 species (submontane forest).
In terms of Shannon diversity (H¢), the shrub layer was the most diverse
followed by the tree and herbaceous layers (Table 2). The Shannon diversity
varied in the shrub layer from 4.39 (swamps) to 5.13 (forest rich in Cae-
salpinioideae), in the tree layer from 3.82 (swamps) to 4.83 (forest rich in
[297]
1358
Table 2. Summary of the number of species, number of families, number of stem/ha and Shannon diversity (H¢) recorded in the tree, shrub and herbaceous
layers for each vegetation types for all vascular plants with DBH ‡ 1 cm.
Table 3. Contribution to the total species richness from the various forest strata in 6 main veg-
etation types.
Tree species richness was relatively higher than that of shrubs, lianas and
herbs (Table 4). The total number of species varied from 172 species
[299]
1360
Figure 1. Correlation between the Shannon diversity (H¢) of all vascular plant species recorded in
the tree layer and that of the shrub/small tree layer within 145 plots of 0.1 ha each.
[300]
1361
Figure 2. Correlation between the Shannon diversity (H¢) of all vascular plant species recorded in
the tree layer and that of the herbaceous layer within 145 plots of 0.1 ha each.
There was a significant positive correlation between the number of large and
medium sized tree species and that of shrubs/small trees (F1, 143 = 112.033,
R2 = 0.439, p<0.001) and woody climbers (F1, 143 = 26.986, R2 = 0.159,
p<0.001). There was also a significant positive correlation between the
diversity of large and medium sized trees and that of shrubs/small trees and
woody climbers (Table 6; Figures 3 and 4). The correlations between the
diversity/species richness of large and medium sized trees and that of the
herbaceous species were not significant (F1, 143 = 0.001, R2 = 0.00002,
p = 0.975 for Shannon diversity and F1, 143 = 0.0387, R2 = 0.0003, p = 0.844
for species richness). The Shannon diversity (H¢) was significantly different
among the various growth forms (F3, 34.6 = 151.290, p<0.001) when correcting
for differences in vegetation types (GLM: F10, 30 = 2.727, p<0.01 for vege-
tation type included as a random factor). A Tukey multiple comparison test
showed that all growth forms were significantly different from each other
(p<0.05) with trees having the highest mean value (H¢ = 3.37) and the lowest
(H¢ = 0.24) for the herbaceous species. The species richness was also signifi-
cantly different between growth forms (GLM: F3, 33.2 = 221.889, p<0.001 for
layer and F10, 30 = 2.973, p<0.01 for vegetation type included as a random
factor), with similar relative difference.
[301]
1362
Table 4. Summary of the total number of species for the various growth forms recorded in 6 main vegetation types for all vascular plants with DBH ‡1 cm.
Total no. of tree species 207 172 237 211 222 216
Tree layer: DBH ‡ 10 cm 128 82 158 126 156 162
Shrub layer: 1.5 cm £ DBH <10 cm 177 132 198 169 192 173
Herb layer: 1 £ DBH <1.5 cm 94 44 97 100 86 102
Shannon diversity (H¢) 4.26 3.96 4.47 4.13 4.44 4.67
Total no. of shrubs/small tree species 127 71 164 155 161 152
Tree layer: DBH ‡10 cm 11 10 20 10 13 12
[302]
Shrub layer: 1.5 cm £ DBH <10 cm 113 58 153 143 152 137
Herb layer: 1 £ DBH <1.5 cm 76 37 105 104 100 101
Shannon diversity (H¢) 3.32 4.39 4.26 4.08 4.32 4.66
Total no. of herbaceous species 8 7 5 4 9 4
Tree layer: DBH ‡10 cm 0 0 0 0 0 0
Shrub layer: 1.5 cm £ DBH <10 cm 7 4 3 3 6 3
Herb layer: 1 £ DBH <1.5 cm 8 6 5 4 8 4
Shannon diversity (H¢) 0.53 0.54 0.78 1.03 0.85 0.90
Total no. of liana species 46 42 63 50 59 43
Tree layer: DBH ‡10 cm 8 7 7 8 12 6
Shrub layer: 1.5 cm £ DBH <10 cm 40 39 60 8 54 36
Herb layer: 1 £ DBH <1.5 cm 16 11 25 26 23 21
Shannon diversity (H¢) 3.13 3.29 3.34 3.27 3.65 3.19
Herbaceous species include herbs, herbaceous climbers and hemi-epiphytes. Lianas include small and large woody climbers. Note that species may overlap
within forest strata.
Table 5. Summary of the number of species of the various growth forms recorded in 136 vegetative subplots of 5 m · 5 m each covering 0.34 ha.
Coastal Swamps Forest rich in Forest rich in Calpocalyx Mixed evergreen and Submontane
forest Caesalpinioideae and Sacoglottis semi-deciduous forest forest
[303]
Trees 27 6 35 38 37 26
Shrubs/small trees 87 14 148 102 134 74
Herbs 88 14 142 98 105 77
Lianas 30 3 36 33 32 11
Total No of species 232 37 368 271 308 188
Total No of families 67 27 76 63 68 47
1363
1364
Table 6. Spearman correlation coefficients between the Shannon diversity (H¢) of the various
growth forms recorded for all vascular plants with DBH ‡1 cm in 145 plots of 0.1 ha each.
Large trees 1
Shrubs/small trees 0.29** 1
Herbs 0.13 0.20* 1
Lianas 0.24** 0.20* 0.18* 1
Spearman correlation is significant at the 0.05 level for * and at the 0.01 level for **.
Figure 3. Correlation between the Shannon diversity (H¢) of large and medium sized tree species
and that of the shrub/small tree species within 145 plots of 0.1 ha each.
Overall, shrubs/small trees have the highest number of families (75) followed
by medium trees (61), herbs and hemi-epiphytes (58), large trees (54) and
woody climbers (37). Rubiaceae was by far the most species rich family (204
species) followed by Euphorbiaceae (88), Leguminosae-Caesalpinioideae (85),
Annonaceae (63), Sterculiaceae (50), Apocynaceae (47) and Sapindaceae (40).
Leguminosae (especially the subfamily Caesalpinioideae) was the dominant
family for the large trees species (DBH ‡30 cm) in terms of relative density
and frequency in the Campo-Ma’an area (Table 7). Dominant large tree
species included Calpocalyx heitzii, Desbordesia glaucescens, Erythrophleum
ivorensis, Lophira alata, Lovoa trichilioides, Pycnanthus angolensis, Sacoglottis
[304]
1365
Figure 4. Correlation between the Shannon diversity (H¢) of large and medium sized tree species
and that of the climbers within 145 plots of 0.1 ha each.
gabonensis and Terminalia superba. Common shrubs and small tree species
included many species of the genera Cola, Diospyros, Drypetes, Psychotria,
and Rinorea. Common large woody climber species were from the genera
Agelaea, Dichapetalum, Combretum, Millettia, and Strychnos. The most
important herb species were of the genera Begonia, Culcasia, Dorstenia,
Geophila, Haumania, Hymenocoleus, Mapania, Marantochloa, Microcalamus
and Palisota.
Discussion
The Campo-Ma’an area is dominated by the lowland evergreen rain forest rich
in Caesalpinioideae and is characterized by a rich and diverse flora Letouzey
1968 and 1985; Thomas and Thomas 1993; Tchouto 2004). The number of
species/ha for all vascular plants ‡1 cm recorded varied from 293 in swamp
forest to 468 in the lowland evergreen forest rich in Caesalpinioideae. The
Shannon diversity (H¢) varied from 4.73 in coastal forests to 5.16 in forests rich
in Caesalpinioideae. The explanation for the high level of species richness and
diversity might stem partly from the fact that the area is part of a series of
postulated rain forest refugia in Central and West Africa (Hamilton 1982;
[305]
1366
Table 7. Five most important families/subfamilies recorded for the various growth forms in 145
forest plots (0.1 ha each) and the vegetative subplots. Note that species may overlap between size
classes.
White 1983; Maley 1987 and 1989; Sosef 1994). Furthermore, the Campo-
Ma’an forest falls within the Guineo-Congolian Centre of Endemism (White
1983; Gartlan 1989; Thomas and Thomas 1993; Davis et al. 1994). It is situated
in the middle of the rich Biafran forest type that extends from southeast
Nigeria to Gabon and the Mayombe area in Congo (White 1983; Letouzey
1968 and 1985) and shares with these sites the overall characteristic of lowland
evergreen rain forest with some semi-deciduous species.
In agreement with Hubbell and Foster (1983, 1986c), many species
showed no apparent distributional biases with respect to habitat boundaries.
Nevertheless, some species were strongly positively or negatively associated
with specific habitats such as swamps, hilltops, riverbank and disturbed
forests. Species composition was influenced by rainfall, altitude, soils, the
proximity of the sea, and the level of human disturbance that contributed to
[306]
1367
several environmental hostilities that may have limited the influx of plant
species and weaken competitive abilities of poorly adapted species (Tchouto
2004). The total number of vascular plant species was relatively influenced
by the proximity of the sea and rainfall, thus resulting in a gradual vari-
ation in dominant species and an increase in species richness with increasing
annual rainfall and distance from the sea. There was also an increase in
deciduous and semi-deciduous elements with decreasing annual rainfall.
Forest at higher altitudes between 800 and 1100 m above sea level appeared
to be relatively more species-rich than the disturbed lowland and the coastal
forests on sandy shorelines. In the undisturbed lowland evergreen forest rich
in Caesalpinioideae and the submontane forest, more than 93% of the
0.1 ha plots had above 100 species. Swamps and seasonal flooded forests on
hydric soils with poor drainage conditions and low nutrient concentrations
were species-poor. This result is in full agreement with several studies that
have shown that permanent water logging in soils is a main factor limiting
vascular plant species alpha diversity in Neotropical swamps (Duivenvoor-
den and Lips 1995; Gartlan et al. 1986; Newbery et al. 1996; Sheil et al.
2000).
[307]
1368
A total of 533 tree species with DBH ‡10 cm was recorded in 145 plots of
0.1 ha each, with an average of 12–50 species/0.1 ha. Most plots in the lowland
evergreen forest and the submontane forest had above 35 tree species/0.1 ha.
These results are comparable to that of 15–60 species/0.1 ha found by Gentry
(1988b) in the tree species rich Mishana and Yanamono plots near Iquitos.
Balslev et al. (1998) found 307 tree species with DBH ‡10 cm in 1 ha plots of
forest in Ecuador and Condit et al. (2000) between 673 and 996 tree species in
two 50-ha plots in Malaysia. In this study, shrubs and small trees were by far
the most species rich growth forms, more than 756 species were recorded
among treelets and shrubs with DBH £ 2.5 cm, and 793 species for treelets
between 2.5<DBH <10 cm in 145 0.1-ha plots, with an average of 13–101
species and 10–90 species/0.1 ha, respectively. Gentry and Dodson (1987)
found between 37 and 137 species/0.1 ha among treelets with DBH < 2.5 cm,
and 38–87 species among treelets between 2.5<DBH <10 cm in 3 plots in
western Ecuador. These results correspond to some extent to the general trend
of tree and shrub species densities recorded in other species-rich Neotropical
rain forests (Gentry 1988a, b; Valencia et al. 1994; Duivenvoorden and Lips
1995; Condit et al. 2000; ter Steege 2000). This suggests that the world’s highest
level of tree and shrub species richness are not only confined to the Amazonian
and south east Asian forests, but may be also found in the Atlantic Biafran
lowland evergreen rain forest in Central Africa.
The species richness of trees, treelets and shrubs was higher compared to that
of the climbers and herbs. Species richness among terrestrial herbs and climbers
were relatively low. A total of 161 species of vascular climbers, and 29 species
of herbs were recorded with an average of 2–23 species of climbers/0.1 ha and
0–5 herb species/0.1 ha with DBH ‡ 1 cm. Poulsen and Balslev (1991) found
25–41 obligate herb species in 0.1 ha in an upland forest in Ecuador, and
Duivenvoorden and Lips (1995), 21–25 (excluding palm) in Colombia. The
average of 9.7 individuals/ha (3–18 individuals/ha) recorded for large lianas
with DBH ‡10 cm in the Campo-Ma’an area, falls within the pan-tropical
range presented by Rollet (1974) for large lianas. Climbers represented 14% of
the total number of vascular plant species, indicating that climbers contribute
considerably to forest plant species richness in the area. Amazonian and
Australian moist forests are reported to be richer in climber species than
African rain forests since they contribute up to 25% of the total vascular plant
species richness (Putz and Chai 1987; Nabe-Nielsen 2001; Schnitzer and
Bongers 2002), however, the Campo-Ma’an area appeared to be one of the
richest sites for climbers in tropical Africa. The explanation for the high
number of climbers species might stem partly from the fact that the area falls
within the Congo Basin, which is known to be the richest in climbers in Central
and West African (Parren 2003).
The striking low number of herbs in the present study is partly due to the
fact that many herbaceous species were below 1 cm DBH and could not be
[308]
1369
sampled within the 0.1 ha plots. As a result, herbaceous species contributed less
than 1% of the total vascular plant species count. However, in the 136 subplots
of 5 · 5 m, there was a considerable increase in herb species, from 25 species in
145 plots of 0.1 ha to 257 species in the 5 · 5 m plots (Table 5). A stratified
sampling approach, with larger plots for larger individuals (e.g. 1 ha plot for
DBH ‡ 10 cm) divided into small subplots for smaller individuals (e.g. 0.1 ha
for DBH ‡ 1 cm and 0.01 ha for DBH <1 cm) seems therefore the best
approach to measure species richness in tropical rainforests.
Tree species appeared to be more diverse than shrubs, climbers and herbs.
Although there was a significant positive correlation between the diversity of
trees and that of the shrubs/small trees and woody climbers, the correlations
between the diversity of trees and that of the herbaceous species were not
significant. Moreover, there was a significant positive correlation between the
diversity of shrubs/small trees and that of herbaceous species. This is partly due
to the fact that most of the shrubs, small trees and herbaceous plants are
understorey species that live under the same physiological and biological
conditions. They are either shade bearers or non-pioneer light demanding
species that require little sunlight for survival. More than 40% of the herba-
ceous layer species contribution came from shade hemi-epiphytes that are often
restricted to the lower trunk of shrubs and small trees. Hemi-epiphytic species
of the genera Culcasia and Cercestis (Araceae) and several fern genera such as
Lomariopsis, Hymenophyllum and Trichomanes were very common (Table 7).
It is worth mentioning that species richness and diversity was higher in
undisturbed forest types than in the coastal forests where there was a pro-
nounced human disturbance. With the exception of swamps and mangroves,
the coastal forest types were floristically poorer with a lower diversity index
(H¢) than other forest types (Tables 1 and 2). The degree of human activities
seems to have an effect on species composition and decrease species diversity
within the various forest types. Plots located in secondary forests, past logging
concessions and plots near forest edges were more disturbed, and characterized
by a high number of herbaceous, climber and pioneer species. Furthermore,
undisturbed forests with open canopy, such as the mixed evergreen and semi-
deciduous forest in the drier Ma’an area, were also characterized by an in-
creased number of semi-deciduous, herbaceous and climber species.
Many botanical studies in tropical rain forest emphasise the structural aspect
of the forest, assuming that the diversity of large and medium sized trees (DBH
‡ 10 cm) reflect the overall diversity of the forest. When comparing the tree
diversity and floristic composition in 6 different forest types in the Campo-
Ma’an area, we noticed that tree species accounted for 46% of the total vas-
cular plant species with DBH ‡ 1 cm, shrubs/small trees 39%, climbers 14%
and herbs less than 1%. Only 22% of the diversity of shrubs and lianas could
[309]
1370
be explained by the diversity of large and medium sized trees, and less than 1%
of herb diversity was explained by the tree diversity (Figures 3 and 4). A higher
percentage of tree, shrub and climber species occurred in the shrub layer than
in the tree and herbaceous layers. Moreover, only 63% of the tree species were
recorded in the tree layer against 82% in the shrub layer. Less than 10% of the
total number of shrub/small tree species was found in the tree layer compared
to 90% in the shrub layer. Furthermore, shrubs contributed for 38% to the 114
strict and narrow endemic plant species recorded in the area, herbs 29%, trees
only 20% and climbers 11% (Tchouto 2004). It is worth mentioning that,
although there was a significant positive correlation between the diversity of
trees and that of shrubs and woody climbers, the correlation between tree and
herb diversity was not significant.
This study also demonstrated that the shrub layer was by far the most
species-rich in the different plots and vegetation types. It was significantly more
diverse and species-rich than the tree and herbaceous layers. More than 82% of
tree species, 90% of shrubs, 78% of lianas and 70% of herbaceous species were
recorded in this layer. The high number of species found in this layer can be
attributed to the fact that, in addition to immature large trees and woody
climbers, the shrub layer comprises shrubs, small trees, tall herbs, small
climbers and hemi-epiphytes which are not found in the upper tree layer. This
leads to the conclusion that tree diversity does not always reflect the overall
diversity of the forest. In addition, more than 75% of plant species of high
conservation value such as endemic species are shrub and herbaceous species
(Tchouto 2004). Similar studies carried out by Duivenvoorden and Lips (1995)
in Colombia, and by ter Steege (2000) and van Andel (2001) in Guyana, have
also shown that tree diversity is not a good indicator for the diversity of shrubs
and herbs. This suggests that sampling design, based on small plots of 0.1 ha, in
which all vascular plants with DBH ‡ 1 cm are recorded, is a more appropriate
sampling method for biodiversity conservation purposes, than assessments
based solely on large and medium sized trees. Although it requires additional
effort, time and financial involvement, it provides more information on other
growth forms such as shrubs, climbers and herbs that are under-represented
when the sampling design only includes large and medium sized trees (DBH ‡
10 cm).
In our study, large and medium sized tree species richness showed a strong
positive correlation with that of lianas, indicating that in undisturbed Central
African lowland evergreen rain forests, tree species richness may predict woody
climber species richness relatively well. Similar results were obtained by ter
Steege (2000) in Guyana. This is partly explained by the fact that woody
climbers are dependent on the presence of trees for their support. Although,
large trees are reported to have a negative impact on liana density through
direct competition with the lianas for light and nutrients, their composition and
physiognomy, as well as the forest structure, are important factors influencing
the species composition and diversity of liana (Schnitzer and Bongers 2002;
Parren 2003). Considering the fact that most of the Campo-Ma’an area is
[310]
1371
dominated by a lowland forest rich in large canopy and emergent tree species
(up to 60 m tall) that forms a fairly continuous canopy, only large non-pioneer
light-demanding climber species and small woody shade bearer climbers that
are adapted to low light understorey conditions can survive in such an envi-
ronment. The density and species richness of small light-demanding herbaceous
and woody climbers is therefore relatively low in undisturbed forests and high
at forest edges, in natural tree fall gaps, in secondary forests and in opened
forest with a discontinuous canopy.
Conclusion
There is a general perception among scientists that, in the tropical rain forest,
the diversity of large and medium sized trees (DBH ‡ 10 cm) can be used to
predict the diversity of other life forms, since, in most of these studies tree
species account for more than 50% of the overall species composition. This
study has demonstrated that the diversity of trees does not always reflect the
overall diversity of forest in the Campo-Ma’an area and, therefore, it may not
be a good indicator for the diversity of shrubs and herbaceous species. How-
ever it is a relatively good indicator for the diversity of lianas. In terms of
floristic composition, the number of tree species can be used to some extent to
predict the number of species of other growth forms. Furthermore, the shrub
layer (1.5 cm £ DBH <10 cm) was by far the most species rich in the different
vegetation types sampled and appeared to be more diverse and species-rich
than the tree and herbaceous layers. This suggests that sampling design, based
on small plots of 0.1 ha, in which all vascular plants with DBH ‡1 cm are
recorded, is a more appropriate sampling method for biodiversity conservation
purposes than assessments based solely on large and medium sized trees (DBH
‡ 10 cm). Moreover, if there are enough means, time and staff, a stratified
sampling approach, with larger plots for larger individuals (e.g. 1 ha plot for
DBH ‡ 10 cm) divided into small subplots for smaller individuals (e.g. 0.1 ha
for DBH ‡ 1 cm and 0.01 ha for DBH < 1 cm) is the best.
Acknowledgements
This study was carried out within the framework of the Campo-Ma’an Bio-
diversity Conservation and Management Project, Cameroon, and was finan-
cially supported by Tropenbos International, The Netherlands. We thank
G. Achoundong, J.M. Onana, B. Sonke, L. Zapfack and P. Mezili at the
National Herbarium, Cameroon, and F.J. Breteler, and C.C.H. Jongkind at
the Nationaal Herbarium Nederland, Wageningen University Branch, who
assisted in plant identification. The staff of Campo-Ma’an Project is also
acknowledged with gratitude for their assistance and support during the
fieldwork. Particular thanks are for my field assistants Elad Maurice and
[311]
1372
Ossele Mathilde for their enthusiastic support and cooperation. We also extend
our sincere thanks to all chiefs and village representatives, for their active
participation in the organization and collection of field data.
References
Achoundong G. 2000. Les Rinorea et l’étude des refuges forestiers en Afrique. In: Servant M. and
Servant-Vildary S. (eds), Dynamique à long terme des écosystèmes forestiers intertropicaux,
pp. 19–29.
Balslev H., Valencia R., Paz y Mino G., Christensen L.B. and Nielsen I. 1998. Species count of
vascular plants in one hectare of humid lowland forest in Amazonian Ecuador. In: Dallmeier F.
and Comiskey J.A. (eds), in Forest Biodiversity in North, Central and South America, and the
Caribbean. MAB series, 21: 585–594.
Bisby F.A. 1995. Characterization of biodiversity. In: Heywood V.H. and Watson R.T. (eds),
Global Biodiversity Assessment. UNEP, Cambridge University Press, pp. 192–294.
Cable S. and Cheek M. 1998. The plants of Mount Cameroon. A Conservation Checklist. Royal
Botanic Gardens, Kew.
Condit R., Ashton P.S., Baker P., Bunyavejchewin S., Gunatilleke S., Gunatilleke N., Hubbell S.P.,
Foster R.B., Itoh A., LaFrankie J.V., Lee S.H., Losos E., Manokaran N., Sukumar R. and
Yamakura T. 2000. Spatial patterns in the distribution of tropical tree species. Science 228: 1414–
1417.
Davis S.D., Heywood V.H. and Hamilton A.C. 1994. Centres of Plant Diversity. A Guide and
Strategy for Their Conservation 1.
de Kam M., Fines J.-P. and Akogo G. (eds) 2002. Schéma Directeur pour le developpement de
l’Unité Technique Opérationnelle de Campo-Ma’an. Cameroun. Campo-Ma’an Series 1.
Duivenvoorden J.F. and Lips J.M. 1995. A Land Ecological Study of Soils, Vegetation and Plant
Diversity in Colombia Amazonia. Tropenbos Series 12The Tropenbos Foundation, Wageningen.
ERE Développement 2002. Etude socio-économique dans l’UTO de Campo-Ma’an. Rapport final
phase 2: résultats d’enquètes auprès des ménages. SNV/Projet Campo-Ma’an.
Franqueville A. 1973. Atlas Regional Sud-Ouest 1. République du Cameroun. ORSTOM, Yao-
undé, Cameroun.
Gartlan J.S. 1989. La conservation des ecosystemes forestiers du Cameroun. IUCN.
Gentry A.H. 1988a. Change in plant community diversity and floristic composition on environ-
mental and geographical gradients. Annals of the Missouri Botanical Garden 75: 1–34.
Gentry A.H. 1988b. Trees species richness of upper Amazonian forests. Proceedings of the
National Academy of Science USA 85: 156–159.
Gentry A.H. and Dodson C. 1987. Contribution of non trees to species richness of a tropical rain
forest. Biotropica 19: 149–156.
Hamilton A.C. 1982. Environmental History of Africa: A Study of the Quaternary. Academic
Press, London.
Hart T.B., Hart J.A. and Murphy P.G. 1989. Monodominant and species-rich forests of the humid
tropics: causes for their co-occurrence. The American Naturalist 133: 613–633.
Hubbell S.P. and Foster R.B. 1983. Diversity of canopy trees in a neotropical forest and impli-
cations for conservation. In Sutton S.L., Whitemore T.C. and Chadwick A.C. (eds), in Forest:
Ecology and Management. Blackwell Science, Oxford, pp. 24–41.
Hubbell S.P. and Foster R.B. 1986c. Commonness and rarity in a neotropical forest: implications
for tropical tree conservation. In: Soule M.E. (ed.), Conservation Biology: The Science of
Scarcity and Diversity. Sinauer Associates, Massachusetts, pp. 205–231.
Kent M. and Coker P. 1992. Vegetation Description and Analysis. Belhaven Press, London.
Koubouana F. 1993. Les forêts de la vallée du Niari, Congo: Etudes floristiques et structurales.
Thèse de doctorat, Université de Paris 6.
[312]
1373
Lejoly J. 1995a. Biodiversité végétale dans le Parc National d’Odzala, Congo. Rapport technique.
Groupement Agreco-CTFT.
Lejoly J. 1995b. Utilisation de la méthode de transects en vue de l’étude de la biodiversité dans la
zone de conservation de la forêt de Ngotto, République Centrafricaine. Rapport technique.
Projet ECOFAC, Agreco-CTFT.
Letouzey R. 1968. Etude phytogéographique du Cameroun. Le Chevalier, Paris.
Letouzey R. 1985. Notice de la carte phytogéographique du Cameroun, vol. 1–5. Institut de la
Carte Internationale de la Végétation, Toulouse.
Magurran A.E. 1988. Ecological Diversity and its Measurement. Croom Helm, London.
Maley J. 1987. Fragmentation de la forêt dense humide Africaine et extention des biotopes mon-
tagnards du Quaternaire recent. Paleoecology of Africa 18: 307–309.
Maley J. 1989. Late Quaternary climatic changes in the African rainforest: the question of forest
refuges and the major role of sea surface temperature variations. In: Lewen M. and Sarthein M.
(eds), Paleoclimatology and Paleometeology: Modern and Past Patterns of Global Atmospheric
Transport. Kluwer, Londonmeteo, pp. 585–616.
Mosango M. 1990. Contribution à l’étude botanique et biogéochimique de l’écosystème des forêts
en région équatoriale (Ile Kongolo, Zaire). Thèse de doctorat, Université Libre de Bruxelles.
Muller J.P. 1979. Carte des sols du Cameroun. Atlas de la République Unie du Cameroun.
Nabe-Nielson J. 2001. Diversity and distribution of lianas in a neotropical rain forest, Yasuni
National Park, Ecuador. Journal of Tropical Ecology 17: 1–19.
Newbery D. Mc and Gartlan J.S. 1996. A Structural Analysis of Rainforest at Korup and Douala-
Edea, Cameroon. Proceedings of the Royal Society of Edinburg 104B: 177–224.
Parren M.P.E 2003. Lianas and logging in West Africa. Tropenbos-Cameroon Series 6. PhD
Thesis, Wageningen University, The Netherlands.
Poulsen A.D. and Balslev H. 1991. Abundance and cover of ground herbs in an Amazonian rain
forest. Journal of Vegetation Science 2: 315–322.
Putz F.E and Chai P. 1987. Ecological studies of lianas in Lambir National Park, Sarawak,
Malaysia. Journal of Ecology 75: 523–531.
Reitsma J.M. 1988. Forest vegetation in Gabon. Tropenbos Technical Series 1.
Robbrecht E. 1996. Geography of African Rubiaceae with reference to glacial rain forest refuges.
In: van der Maesen L.J.G., vander Burgt X.M. and van Medenbach de Rooy J.M. (eds), The
Biodiversity of African Plants. Kluwer Academic Publishers, Wageningen, pp. 564–581.
Rollet B. 1974. La régénération naturelle dans les trouées Un processus général de la dynamique
des forêts tropicales humides. Revue Bois et Forêts des Tropiques 201: 3–34.
Schnitzer S.A. and Bongers F. 2002. The ecology of lianas and their role in forests. Trend in
Ecology and Evolution 17: 223–230.
Sheil D., Jennings S. and Savill P. 2000. Long-term permanent plot observations of vegetation
dynamic in Budongo, a Ugandan rain forest. Journal of Tropical Ecology 16: 765–800.
Sonké B. 1998. Etudes floristiques et structurales des forêts de la Reserve de Faune de Dja,
Cameroun. Thèse présentée en vue de l’obtention du grade de Docteur en sciences, option
écologie végétale. Université Libre de Bruxelles.
Sonké B. and Lejoly J. 1998. Biodiversity study in the Dja Fauna Reserve (Cameroon): using
the transect method. In: Huxley C.R., Lock J.M. and Cutler D.F. (eds), Chorology, Taxonomy
and Ecology of the Floras of Africa and Madagascar. Royal Botanical Gardens, Kew,
pp. 171–179.
Sosef M.S.M. 1994. Refuge Begonias. Taxonomy, phylogeny and historical biogeography of
Begonia sect. Loasibegonia and sect. Scutobegonia in relation to glacial rain forest refuges in
Africa, PhD Thesis, Wageningen Agricultural University.
Sosef M.S.M. 1996. Begonias and African rain forest refuges: general aspects and recent progress.
In: van der Maesen, L.J.G., van der Burgt X.M. and van Medenbach de Rooy J.M. (eds), The
Biodiversity of African Plants. Kluwer Academic Publishers, Wageningen, pp. 602–611.
Tchouto M.G.P. 2004. Plant diversity in a Central African rain forest: implications for biodiversity
conservation in Cameroon. PhD Thesis, Wageningen University.
[313]
1374
ter Steege H. 2000. Plant diversity in Guyana, with recommendations for a National Protected area
strategy. Tropenbos Series 18.
Valencia R., Balslev H. and Pazy Mino G. 1994. High tree alpha-diversity in Amazonian Ecuador.
Biodiversity and Conservation 3: 21–28.
van Andel T. 2001. Floristic composition and diversity of mixed primary and secondary forests in
northwest Guyana. Biodiversity and Conservation 10: 1645–1682.
van Gemerden B.S. and Hazeu G.W. 1999. Landscape ecological survey (1:100,000) of the Bipindi-
Akom II-Lolodorf region, Southwest Cameroon. Tropenbos-Cameroon Document 1.
van Valkenburg J.L.C.H., Ketner P. and Wilks C.M. 1998. A floristic inventory and preliminary
vegetation classification of the mixed semi-evergreen rain forest in the Minkébé region, North
East Gabon. Adansonia 20: 139–162.
White F. 1983. The Vegetation of Africa. UNESCO, Paris.
White L.J.T. 1996. Determinants of vegetation composition in the Lopé Reserve. Report to
AGRECO/CTFT, Gabon.
Whittaker R.J. 1975. Communities and Ecosystems, 2nd ed.. Macmillan, London.
Wolter F. 1993. Etude des possibilités techniques, économiques et financières d’un aménagement
des forêts tropicales denses humides de la cuvette centrale du Zaı̈re, basée sur ses capacités
naturelles. Thèse de doctorat. Université de Louvain.
[314]
Biodiversity and Conservation (2006) 15:1375–1397 Springer 2006
DOI 10.1007/s10531-005-5397-6
-1
Key words: Forest composition, Ituri Forest, Secondary forest, Selective logging, Shifting culti-
vation, Timber tree regeneration, Tree diversity
Abstract. Mature tropical forests at agricultural frontiers are of global conservation concern as the
leading edge of global deforestation. In the Ituri Forest of DRC, as in other tropical forest areas,
road creation associated with selective logging results in spontaneous human colonization, leading
to the clearing of mature forest for agricultural purposes. Following 1–3 years of cultivation,
farmlands are left fallow for periods that may exceed 20 years, resulting in extensive secondary
forest areas impacted by both selective logging and swidden agriculture. In this study, we assessed
forest structure, tree species composition and diversity and the regeneration of timber trees in
secondary forest stands (5–10 and 40 years old), selectively logged forest stands, and undisturbed
forests at two sites in the Ituri region. Stem density was lower in old secondary forests ( 40 years
old) than in either young secondary or mature forests. Overall tree diversity did not significantly
differ between forest types, but the diversity of trees ‡10 cm dbh was substantially lower in young
secondary forest stands than in old secondary or mature forests. The species composition of
secondary forests differed from that of mature forests, with the dominant Caesalpinoid legume
species of mature forests poorly represented in secondary forests. However, in spite of prior log-
ging, the regeneration of high value timber trees such as African mahoganies (Khaya anthotheca
and Entandrophragma spp.) was at least 10 times greater in young secondary forests than in mature
forests. We argue that, if properly managed and protected, secondary forests, even those impacted
by both selective logging and small-scale shifting agriculture, may have high potential conservation
and economic value.
Introduction
The intrusion of human populations into primary forest areas that were
previously free of anthropogenic disturbance is becoming increasingly
[315]
1376
[316]
1377
[317]
1378
Study sites
The study was conducted at two sites in the Ituri Forest, in the northeastern part
of the Congo basin forest block (Democratic Republic of Congo, DRC). The
first site (Mandumbi) was a 17 years old logging concession located 25 km
northwest of the town of Beni (045¢ N latitude, 2915¢ E longitude), whereas a
second site was located at Epulu, in the 1,350,000 ha Okapi Wildlife Reserve.
Field investigation at this site was carried out within the 5 km2 Lenda Study
Area (LSA, 119¢ N and 2838¢ E) established by CEFRECOF. The elevation in
the region from varies 750 m to 950 m above sea level. Mean annual rainfall in
Beni is 1639 mm and 1725 mm in Epulu. A dry season occurs from December to
February, during which monthly average rainfall is less than 100 mm. May and
October are the wettest months of the year, with average precipitations of
186 mm and 200 mm, respectively. Annual average daily temperature at both
sites is 23–25.5 C and varies little through the year (Figure 1).
The vegetation in the region is a mixture of evergreen forest, including
extensive areas of ‘‘mbau forest’’ dominated by Gilbertiodendron dewevrei (De
Wild.) J. Léonard, and ‘‘mixed forests’’ in which no species is predominant, but
other Caesalpinoid legumes, such asJulbernardia seretii (De Wild.) Troupin
and Cynometra alexandri C.H. Wright, are abundant (Makana et al. 2004). At
the eastern edge of the region, evergreen forests grade into a semi-deciduous
forest whose canopy is dominated by light-demanding tree species that include
Entandrophragma spp. (Meliaceae), K. anthotheca (Meliaceae), Albizia spp.
(Mimosaceae) andCanarium schweinfurthii Engl. (Burseraceae). Evergreen
mixed forest is the main vegetation type in Epulu, while semi-deciduous forest
prevails at the Mandumbi site. Large-scale human activities at the Mandumbi
site have resulted in extensive areas of active crop fields and secondary vege-
tation of various ages. Secondary forests were generally young, less than
10 years old, and were dominated by the early pioneer tree Musanga cecro-
pioides R. Br. There was also an old secondary forest created by shifting
agriculture fields abandoned in the 1960s at the Epulu site. However, no
selective logging took place prior to forest clearing at this site.
Soils at both sites are derived from granitic or alluvial rocks and fall under
the order Oxisols, which dominates most of the Congo basin rain forest block
in central Africa. Their texture ranges from loamy sand to sandy clay. The soils
are very acidic, with mean pH values at 20 cm averaging 4 in Epulu, and low
in available nitrogen and phosphorus. Mean soil sand content at LSA is 70%
(Hart 1985).
Logging activities in the Ituri region are concentrated in the relatively drier
semi-deciduous forests near the transition between closed canopy forest and
eastern savanna woodlands, likely due to the proximity of export routes to the
Indian Ocean through Uganda and Rwanda. The forests at the savanna
margin are also richer in high-value timber trees such as African mahoganies
and Milicia excelsa than moist evergreen forests found in central and western
Ituri (Makana 2004).
[318]
1379
Figure 1. Climatic data of the two study sites, Epulu and Mandumbi. Data for Mandumbi came
from a weather station in the town of Beni. Mean annual rainfall and mean average daily
temperature are given at the top corners of each graph.
Methods
Vegetation sampling
[319]
1380
two forest types were parallel and 150 m apart. In Mandumbi, where small
farms are intermingled with patches of mature forest fragments, plot location
was dependent on the availability of appropriate secondary forest stands (e.g.
secondary forests of 5–10 years old). Plots were spatially interspersed across
forest patches, with a minimum distance of 50 m between any two adjacent
plots. At both sites, secondary forest plots were located at least 20 m away from
the adjacent mature forest edge. All free-standing trees ‡1 cm dbh (diameter at
breast height) were identified and measured for dbh in 5 m · 5 m plots. Trees
‡10 cm dbh were identified and measured in 10 m · 10 m plots extending from
each 5 m·5 m plot. For the most common tree species identifications were made
directly in the field. When definitive field identification was not possible leaf
samples were collected and compared to voucher specimens at CEFRECOF’s
herbarium in Epulu. Species names follow Lebrun and Stork (1997). In addition
to botanical data, environmental information was collected at each plot. This
information included soil texture, herbaceous cover, exposed mineral soil and
litter depth. Soil texture was assessed according to the finger assessment of soil
texture of the Ontario Institute of Pedology (1985). For the purposes of this
study, soil texture was classified only in three major categories: sandy, loamy
and clay soils. Herbaceous cover and the proportion of surface area made of
exposed mineral soil were visually estimated and recorded as a percent of total
area. Litter depth was measured by inserting a knife through the litter until it
reached mineral soil, then the thickness of the litter and organic matter layer
was determined to the nearest half centimeter by measuring the length of the
portion of the knife that was inserted into the litter. A total of 54 plots were
inventoried; 32 were at the Mandumbi site and 22 at the Epulu site.
Seedling demography
The regeneration of timber trees was assessed in each of the 54 vegetation plots
described above. Seedlings (30 cm height to 0.9 cm dbh) were identified, tagged
and measured for total aboveground height and collar diameter in 5 m · 5 m
plots. Saplings (1–9.9 cm dbh) were identified and measured for dbh in
10 m · 10 m. Diameter and height were taken at two different times to assess
growth. The first measurements were done in 2000–2001 and the second ones
took place in August–October 2002. Diameter was measured to the nearest
0.01 mm using an electronic caliper for seedlings and trees <4 cm dbh or to
the nearest millimeter for individuals ‡4 cm dbh, whereas height was measured
to the nearest cm using a graduated stick. Stem diameter of seedlings was
measured at 10 cm from the ground.
Data analysis
For each plot, which was the experimental unit of this study, the total number
of individuals, the number of species and Shannon-Wiener diversity index were
[320]
1381
calculated for seedlings, saplings, trees ‡1 cm dbh, and trees ‡10 cm dbh.
Seedling abundance was log-transformed before the analysis to homogenize
variance and produce approximately normal residuals (Sokal and Rolhf 1981).
ANOVA was used to assess the effects of site, forest type and their interaction
on tree and seedling abundance and diversity. If a significant interaction was
found, the means of site by forest type combinations were compared by a
Tukey-Kramer test. Analysis of covariance served to assess the effects of forest
type on the relationship between light availability and seedling relative growth
rates. Multivariate analysis was used to explore the variation of floristic
composition across sites and forest types. Detrended correspondence analysis
(DCA) was utilized because an arch effect was observed for correspondence
analysis (ter Braak 1995). Logistic regression was utilized to compare the
abundance of seedlings and saplings of individual timber tree species in sec-
ondary and primary forest plots. The number of seedlings/saplings per plot was
transformed into three categories: 0 for no seedling recorded, 1 for 1–4 seed-
lings, and 2 for >5 seedlings.
Results
Vegetation structure
The density of trees ‡1 cm dbh and ‡10 cm dbh differed significantly between
the two study sites, Mandumbi and Epulu (Table 1). On average, Mandumbi
had higher tree density than Epulu for both dbh cut-offs. There was a mar-
ginally significant effect of forest type on the density of trees ‡1 cm dbh
(F1,50 = 2.5, p = 0.094). In Epulu, primary forest had significantly more
stems than secondary forest, whereas the density of trees ‡1 cm dbh was only
slightly higher in primary forest compared to secondary forest in Mandumbi.
Basal area was significantly higher in secondary than in primary forests for
both dbh cut-offs (Table 1). Higher basal area in secondary forests was in the
most part due to the presence of fast growing pioneer species. The early suc-
cessional and short-lived Musanga cecropioides was very common in the sec-
ondary forest of Mandumbi and it represented 55% of total basal area. Some
of the basal area in secondary forest stands was also accounted for by remnant
trees. In Epulu, long-lived early colonizers such as Albizia sp., Petersianthus
macrocarpus (P. Beauv.) Liben, and Alstonia boonei De Wild. made up most of
basal area in secondary forest plots.
The number of stems decreased rapidly with increasing tree size (Figure 2).
Trees <10 cm dbh represented >80% of the total number of stems in each
forest type at both study sites. Although primary forest sites in Epulu had more
trees ‡10 cm dbh than secondary forest, the overall size distribution was similar
between the two forest types (v2 = 4.5, d.f. = 2, p = 0.210). In Mandumbi,
the size distribution was significantly different between the two forest types
(v2 = 16.3, d.f. = 2, p<0.001); small trees (<20 cm dbh) were more abun-
[321]
1382
Table 1. Structural characteristics and diversity of mature logged and undisturbed, and secondary forest stands in northeastern Congo basin. Values of stem
density are averages per hectare, and were calculated on the basis of 5 m · 5 m subplots. Species richness and Shannon’s diversity index were calculated for
5 m · 5 m quadrats for trees ‡1 cm dbh and for 10 m · 10 m quadrats for trees ‡10 cm dbh. Figures are least squares means and standard errors1.
Trees ‡1 cm dbh
Stem density (ha1) 6473(678)a 4509 (678)b 7571 (601)a 7343 (601)a 0.004 0.094 0.182
Basal area (m2 ha1) 26.82 (5.75)a 44.92 (5.75)b 28.17 (4.93)ab 39.16 (4.93)b 0.683 0.009 0.510
Richness 10.9 (0.89) 11.5 (0.96) 12.9 (0.86) 12.3 (0.76) 0.072 0.683 0.796
Shannon’s index 2.0 (0.14) ab 1.9 (0.14)a 2.3 (0.13)ab 2.3 (0.11)b 0.007 0.955 0.440
[322]
Figure 2. Whole-forest diameter distribution in 10 cm dbh intervals for (a) Epulu primary forest,
(b) Mandumbi primary forest, (c) Epulu secondary forest, and (d) Mandumbi secondary forest.
A total of 159 species were recorded in the census plots, of which 121 taxa were
identified to the species level, 17 to genus, 17 to family, and the remaining four
[323]
1384
were unidentified. Overall, Mandumbi plots had 122 species, while 82 species
were represented in Epulu plots. Species richness and Shannon’s diversity index
showed different patterns relative to tree sizes (Table 1). For trees ‡1 cm, mean
number of species per plot was similar in the two sites and forest types;
however, the value of Shannon’s index was significantly higher in Mandumbi
than in Epulu (F1,50 = 8.1, p = 0.007). In contrast, species richness of trees
‡10 cm dbh varied significantly according to both site and forest disturbance,
while Shannon’s index was only affected by forest disturbance. There were
significant interactions in both diversity measures between site and forest type.
In Epulu, both the mean number of species per plot and the mean value of
Shannon’s index were similar for primary and secondary forests. In contrast,
there was a significant difference for both parameters between the two forest
types in Mandumbi. The young secondary forest at the latter site was much less
diverse than primary forest for trees ‡10 cm dbh (Table 1). Among the four
combinations of site and forest type, richness and Shannon’s diversity index of
trees ‡10 cm dbh were highest for the primary forest of Mandumbi and lowest
for the young secondary forest of the same site.
The species accumulation curves in Figure 3 are far from asymptotic,
indicating that the area sampled was too small to estimate the total number
Figure 3. Species-area relationships for (a) all trees ‡1 cm dbh and (c) trees ‡10 cm dbh. Species–
individual curves for (b) all trees above 1 cm dbh and (d) trees ‡10 cm dbh. EP = Epulu and
MN = Mandumbi, PF = primary forest and SF = secondary forest.
[324]
1385
Figure 4. Axes 1 and 2 from a detrended correspondence analysis on tree species represented by at
least 8 individuals in the plots: (a) plot scores (triangles represent Mandumbi site and circles Epulu
site, filled symbols are for primary forest plots and open symbols for secondary forest sites) and (b)
species scores (symbols with circles represent commercial timber species). Species codes as in
Table 2.
[325]
1386
species accumulation for trees ‡10 cm dbh was much higher in primary
forest than in secondary forest.
The two sites and forest types shared most of their common species (Ta-
ble 2). Overall, Epulu and Mandumbi shared 14 of their 20 most abundant
species, while primary and secondary forests had 18 of their top 20 species
in common at each site. In spite of sharing the majority of the most
common species, there was no significant positive correlation of ranked
abundances of these species between sites or forest types. Epulu and
Mandumbi showed a significant negative correlation among abundance
ranks (Spearman’s rS = 0.549, p = 0.012), as did primary and secondary
forests (rS = 0.739, p<0.001). The most abundant species in Epulu,
Scaphopetalum dewevrei Wildem. and Th. Dur., was totally absent from
Mandumbi plots. Moreover, two of the most common species in the latter
site (Musanga cecropioides and Rinorea oblongifolia Marquand) were not
represented in Epulu plots. In addition, two species of African mahogany
(E. utile and K. anthotheca) were much more abundant in Mandumbi than
in Epulu (Table 2).
Pair-wise comparisons of species abundances for site-forest type combina-
tions showed that the young secondary forest of Mandumbi had significant
negative correlation with primary forests of both sites (rs = 0.743, p<0.001
and rs = 0.797, p<0.001 for Mandumbi and Epulu respectively). No sig-
nificant correlation was found between the old secondary forest of Epulu and
the primary forest of either site. Thus, the composition of the young secondary
forest in Mandumbi was significantly different from that of primary forests at
both sites, while that of the old secondary forest at Epulu was not.
The patterns of variation in floristic composition observed from correla-
tion analysis were corroborated by the results of multivariate analysis. The
first axis of a detrended correspondence analysis (DCA), which explained
12.9% of the total variation in the species data, appeared to be related to
the composition of canopy flora and it separated primary and secondary
forests at each site. For each site, secondary forests had higher scores of
axis 1 than primary forests. The separation between primary and secondary
forests on axis 1 was more distinct for Mandumbi than Epulu forests
(Figure 4a). The second axis (7.6% of the total variation) also distinguished
between primary and secondary forests of each site, but seemed to be more
related to the composition of understory vegetation. Primary forests had
higher scores than secondary forests on the second DCA axis. Primary
forests of both sites and the old secondary forest in Epulu overlapped on
both DCA axes. Mandumbi primary forest and Epulu secondary forest were
indistinguishable on the first axis, but the former had higher scores of axis 2
than the latter.
[326]
Table 2. Abundance of the 20 most common tree species for stems ‡1 cm dbh in primary forests (PF) and secondary forests (SF) at two study sites (Epulu
and Mandumbi) in northeastern Congo Basin. Figures are number of trees per hectare.
PF SF PF SF
1387
1388
Table 3. Abundance of seedlings and saplings of commercial timber tree species in primary forest (PF) and secondary forest (SF) in northeastern Congo
basin. Abundances are averages of the number of stems per hectare, and were calculated on the basis of 5 m · 5 m quadrats for seedlings and of 10 m · 10 m
quadrats for saplings. Significance levels for the difference in regeneration abundance between primary and secondary forests as follows: ns p>0.05, * p<0.05,
** p<0.01.
Mahogany regeneration
[329]
1390
Mandumbi. No seedlings of these species were found in any of the forest plots in
Epulu. The abundance of mahogany saplings (1–9.9 cm dbh) was also higher in
secondary forests than in primary forests at both sites, 331 (125) vs. 113 (39) stems
ha1 in Mandumbi and 27 (14) vs. 9 (9) stems ha1 in Epulu. To determine
environmental factors that affect the regeneration of African mahogany tree
species in the region, the abundance of mahogany regeneration was compared
between primary and secondary forests and the relationships between the
abundance of the regeneration and recorded environmental variables was tested
through regression analysis and analysis of variance. This analysis was limited to
data from Mandumbi because mahogany regeneration in Epulu was negligible.
The abundance of the regeneration of African mahogany species varied
significantly according to forest type (F1,44 = 17, p<0.001) and soil texture
(F1,44 = 2.43, p = 0.038). There were at least ten times more seedlings of these
species in secondary forests than in primary forests at Mandumbi (Table 3).
Loamy and sandy soils supported a higher density of seedlings and saplings of
mahogany than did clay-textured soils. Simple regression analysis showed that
two factors were significantly associated with mahogany regeneration. The
abundance of M. cecropioides, an early pioneer tree species, and litter depth
were positively associated with mahogany regeneration (r2 = 0.431, p<0.001
and r2 = 0.158, p = 0.022 respectively). Herb cover was marginally negatively
associated with the abundance of mahogany regeneration (r2 = 0.099,
p = 0.075).
A simultaneous test of all factors and their interactions through multiple
regression (stepwise selection) revealed that only the abundance of M. cecro-
pioides and the interaction between the latter factor and herb cover were sig-
nificant (p<0.001 for both variables). This model explained 61.4% of the
variation in the abundance of mahogany regeneration in Mandumbi. Litter
depth was not maintained in the model likely due to its strong positive cor-
relation with the abundance of M. cecropioides (r = 0.525, p = 0.002).
Discussion
[330]
1391
depending on land-use type and intensity. Forest clearing for traditional slash-
and-burn agriculture, such as practiced in our study sites, occupies the lower
end of severe forest disturbance as compared to clearing for large commercial
pastures or oil palm plantations (Lawrence et al. 1998; Mesquita et al. 2001).
This study shows that secondary forests growing after the initial clearing
of primary forests for shifting cultivation in the Ituri harbor surprisingly
high levels of tree species diversity for small stem sizes. Overall diversity
measures of trees ‡1 cm dbh were similar between secondary and primary
forest stands. Two factors that may account for the observed diversity
patterns in the secondary and primary forest stands include edge effects and
the dominance of primary forest stands by G. dewevrei. The proximity of
many secondary forest plots to primary forest stands will likely result in
increased diversity in these plots due to potentially high seed input from
mature forests (Mesquita et al. 2001; Kennard 2002). On the other hand,
most mature forest plots especially at Epulu were located in forest stands
dominated by G. dewevrei, which are known to have very low diversity at
small spatial scales (Hart et al. 1989; Makana et al. 2004). The diversity of
larger trees (i.e. ‡10 cm dbh) was, however, significantly lower in young
secondary forest stands (<10 years old). However, notwithstanding the
similarity in overall tree diversity, the floristic compositions of the two forest
types were very different. The flora of secondary forests was particularly
depauperate in common species characteristic of old-growth forests in the
region, particularly G. dewevrei and J. seretii, and understory specialists
such as S. dewevrei, Drypetes spp., Rinorea spp., and Pancovia harmsiana
Gilg. In this respect, our results corroborate those of other studies on
tropical forest succession (Brown and Lugo 1990; Guariguata et al. 1997;
Aide et al. 2000). For example, Aide et al. (2000) noted that a 40-year old
secondary forest derived from abandoned pasture had comparable tree
diversity to adjacent mature forest, whereas the floristic composition was
substantially different.
Stem density and basal area increased very rapidly after farm abandonment
with higher basal areas observed in 10-year old secondary forests than in
adjacent primary forest. However, most of the basal area in the young sec-
ondary forest of Mandumbi was accounted for by the presence of the early
pioneer tree M. cecropioides, which does not persist beyond the senescence of
the initial cohort, which may result in a reduction in the basal area of those
stands when the initial cohort of that species dies out. Similar trends have been
observed elsewhere in tropical forest succession (Aide et al. 2000). Our result
are consistent with other findings suggesting that many structural character-
istics of secondary forest stands in the tropics can reach levels encountered in
mature forest stands quite early during succession (Brown and Lugo 1990;
DeWalt et al. 2003). The rapid forest recovery observed in this study may be
the result of light land-use intensity and high seed fall into abandoned fields
due to the relatively small sizes of farms and the presence of remnant trees that
attract seed dispersers.
[331]
1392
[332]
1393
(rainfall), soil conditions (moisture and nutrients), historic events, and dis-
turbances (Bongers et al. 1999). Most African mahoganies have been described
as belonging to the deciduous or semi-deciduous forests (Hall and Swaine
1981). The relatively drier and semi-deciduous forests of the eastern fringe of
the Ituri basin probably offer more favorable conditions for the regeneration
and growth of African mahoganies than the evergreen forests of Epulu.
Densities of commercial size trees (‡ 80 cm dbh) were 5 times higher in
Mandumbi than in Epulu (J.-R. Makana, unpublished data). Low density of
adult trees may therefore partly explain the poor regeneration of K. anthotheca
and Entandrophragma spp in Epulu, due to limited seed availability (Plumptre
1995; Makana and Thomas 2004).
Natural forests in the Ituri region host an impressive diversity of tree and
animal species which has led to the designation of the Ituri forest as a
refugium during late Quaternary climate fluctuations (Grubb 1982; Hart
et al. 1996). The Ituri forest contains large mammal species including the
endemic Okapi, forest elephant, leopard, buffalo, as well as 13 species of
anthropoid primates, seven species of forest antelopes (duikers), and many
other species of mammals, birds and reptiles (Hart 1986; Thomas 1991;
Hart et al. 1986; Plumptre 1996). This abundant and diverse wildlife
assemblage has coexisted with small communities of horticulturalists in the
Ituri for centuries (Wilkie and Finn 1990). While the presence of small areas
of secondary vegetation derived from swidden agriculture may benefit some
mammal populations (Short 1983; Thomas 1991), the creation of large areas
of secondary forests poses a threat to most mature forest dwelling species
such as forest ungulates, okapi, leopard, and others. These species, espe-
cially forest ungulates, are the main source of dietary protein for local
populations (Hart and Hart 1986). Therefore maintaining large tracks of
undisturbed primary forest is essential to the conservation of animal
diversity and to the well-being of local hunter–gatherer communities.
Forest recovery on abandoned agricultural lands is possible through natural
regeneration. It is shown here that some aspects of forest structure and the
diversity of small trees can be rapidly restored to levels comparable to those of
mature forests. However, forest recovery is possible only if secondary forests
are protected from repeated clearing because the return of the species com-
position of secondary vegetation to assemblage similar to that of old-growth
forests may require over 100 years due to limited seed availability and dis-
persal, and slow growth of mature forest tree species. Small-sized clearings,
moderate land use intensity (long fallow periods), and a fine scaled landscape
mosaic may speed up the return of secondary forests to mature forest species
[333]
1394
Acknowledgements
Financial support for the research came through grants from ITTO, CTFS and
CEFRECOF, a University of Toronto fellowship, and an NSERC grant to
Sean Thomas. Logistical support at the field sites was provided by CEFRE-
COF and ENRA. We thank Sabuni Paluku, Simende, Nzambe, and Amisi for
assisting with fieldwork. Jay Malcolm, Justina Ray, Terry Carleton and two
anonymous reviewers provided constructive comments on an earlier version of
this manuscript.
References
Aide T.M., Zimmerman J.K., Herrera L., Rosario M. and Serrano M. 1995. Forest recovery in
abandoned tropical pastures in Puerto Rico. Forest Ecology and Management 77: 77–86.
Aide T.M., Zimmerman J.K., Pascarella J.B., Rivera L. and Marcano-Vega H. 2000. Forest
regeneration in a chronosequence of tropical abandoned pastures: implications for restoration
ecology. Restoration Ecology 8: 328–338.
[334]
1395
Bongers F., Poorter L., van Rompaey R.S.A.R. and Parren M.P.E. 1999. Distribution of twelve
moist forest canopy species in Liberia and Côte d’Ivoire: response curves to a climatic gradient.
Journal of Vegetation Science 10: 371–382.
Brown S. and Lugo A.E. 1990. Tropical secondary forests. Journal of Tropical Ecology 6: 1–32.
Cannon C.H., Peart D.R. and Leighton L. 1998. Tree diversity in commercially logged Bornean
rainforest. Science 281: 1366–1368.
Carrière S.M., Letourmy P. and McKey D.B. 2002. Effects of remnant trees in fallows on diversity
and structure of forest regrowth in a slash-and-burn agricultural system in southern Cameroon.
Journal of Tropical Ecology 18: 375–396.
Coomes O.T., Grimard F. and Burt G.J. 2000. Tropical forests and shifting cultivation: secondary
forest dynamics among traditional farmers of the Peruvian Amazon. Ecological Economics 32:
109–124.
DeWalt S.J., Malakal S.K. and Denslow J.S. 2003. Changes in vegetation structure and compo-
sition along a tropical forest chronosequence: implications for wildlife. Forest Ecology and
Management 182: 139–151.
Fredericksen T.S. 1998. Limitations of low-intensity selection and selective logging for sustainable
tropical forestry. Commonwealth Forestry Review 77: 262–266.
Fredericksen T.S. and Putz F.E. 2003. Silvicultural intensification for tropical forest conservation.
Biodiversity and Conservation 12: 1445–1453.
Garcia-Montiel D.C. and Scatena F.N. 1994. The effect of human activity on the structure and
composition of a tropical forest in Puerto Rico. Forest Ecology and Management 63: 57–78.
Grubb P. 1982. Refuges and dispersal in the speciation of African forest mammals. In: Prance G.T.
(ed.), Biological Diversification in the Tropics. Columbia University Press, NY, pp. 537–553.
Guariguata M.R. and Ostertag R. 2001. Neotropical secondary forest succession: changes in
structural and functional characteristics. Forest Ecology and Management 148: 185–206.
Guariguata M.R., Chazdon R.L., Denslow J.S., Dupuy J.M. and Anderson L. 1997. Structure and
floristics of secondary and old-growth forest stand in lowland Costa Rica. Plant Ecology 132:
107–120.
Gullison R.E., Panfil S.N., Strouse J.J. and Hubbell S.P. 1996. Ecology and management of
mahogany (Swietenia macrophylla King) in the Chimanes Forest, Beni, Bolovia. Botanical
Journal of Linnean Society 122: 9–34.
Hall J.B. and Swaine M.D. 1981. Distribution and ecology of vascular plants in a tropical rain
forest: forest vegetation in Ghana. Dr W. Junk, The Hague.
Hall J.S., Harris D.J., Medjibe V. and Ashton P.M.S. 2003. The effects of selective logging on forest
structure and tree species composition in a Central African forest: implications for management
of conservation areas. Forest Ecology and Management 183: 249–264.
Hart J.A. and Carrick P. 1996. Climate of the Réserve de Faune à Okapis: Rainfall and temper-
ature in the Epulu Sector, 1986–1995. CEFRECOF Working Paper n 2.
Hart J.A., Hart T.B. and Thomas S.C. 1986. The Ituri Forest of Zaire: primate diversity and
prospects for conservation. Primate Conservation 7: 42–44.
Hart T.B. 1985. The Ecology of a Single-Species Dominant Forest and a Mixed Forest in Zaire,
Equatorial Africa. PhD thesis, Michigan State University, East Lansing, MI.
Hart T.B., and Hart J.A. 1986. The ecological basis of hunter-gatherer subsistence in African rain
forests-the Mbuti of Eastern Zaı̈re. Human Ecology 14: 29–55.
Hart T.B., Hart J.A. and Murphy P.G. 1989. Monodominant and species-rich forests of the humid
tropics: causes for their co-occurrence. American Naturalist 133: 613–623.
Hart T.B., Hart J.A., Dechamps R., Fournier M. and Ataholo M. 1996. Changes in forest com-
position over the last 4000 years in the Ituri basin, Zaire. In: van der Maesen L.J.G. (ed.), The
Diversity of African Plants. Kluwer Academic Publishers, the Netherlands, pp. 545–563.
Holl K.D., Loik M.E., Lin E.H.V. and Samuels I.A. 2000. Tropical montane forest restoration in
Costa Rica: overcoming barriers to dispersal and establishment. Restoration Ecology 8: 339–349.
Johns A.G. 1997. Timber production and biodiversity conservation in tropical rain forests.
Cambridge University Press, Cambridge, UK.
[335]
1396
[336]
1397
Smith J., van de Kop P., Reategui K., Lombardi I., Sabogal C. and Diaz A. 1999. Dynamics of
secondary forests in slash-and-burn farming: interactions among land use types in the Peruvian
Amazon. Agriculture, Ecosystems and Environment 76: 85–98.
Snook L.K. 1996. Catastrophic disturbance, logging, and the ecology of mahogany: ground for
listing a major tropical timber species in CITES. Botanical Journal of the Linnean Society 122:
35–46.
Sokal R.R. and Rohlf F.J. 1981. Biometry: the Principles and Practice of Statistics in Biological
Research. W. H. Freeman and Company, NY.
Southgate D., Sierra R. and Brown L. 1991. The causes of tropical deforestation in Ecuador: a
statistical analysis. World Development 19: 1145–1151.
Struhsaker T.T. 1997. Ecology of an African Rain Forest: Logging in Kibale and the Conflict
Between Conservation and Exploitation. University Press of Florida, Gainesville, Florida.
Struhsaker T.T., Lwanga J.S. and Kasenene J.M. 1996. Elephants, selective logging, and forest
regeneration in the Kibale Forest, Uganda. Journal of Tropical Ecology 12: 45–64.
ter Braak C.J.F. 1987. Ordination. In: Jongman R.H.G., ter Braak C.J.F. and van Tongeren
O.F.R. (eds), Data analysis in community and landscape ecology. Cambridge University Press,
Cambgridge, UK.
Thomas S.C. 1991. Population densities and patterns of habitat use among anthropoid primates of
the Ituri Forest, Zaire. Biotropica 23: 68–83.
Verissimo A., Bareto P., Tarifa R. and Uhl C. 1995. Extraction of a high-value resource in
Amazonia: the case of mahogany. Forest Ecology and Management 72: 39–60.
White L.J.T. 1994. The effects of commercial mechanized selective logging on a transect in lowland
rainforest in Lopé Reserve, Gabon. Journal of Tropical Ecology 10: 313–322.
Whitmore T.C. 1999. Arguments on the forest frontier. Biodiversity and Conservation 8: 865–868.
Wijdeven S.M.J. and Kuzee M.E. 2000. Seed availability as a limiting factor in forest recovery
processes in Costa Rica. Restoration Ecology 8: 414–424.
Wilkie D.S. and Finn J.T. 1990. Slash-burn cultivation and mammal abundance in the Ituri forest,
Zaire. Biotropica 22: 90–99.
Wilkie D.S., Sidle J.G. and Boundzanga G.C. 1992. Mechanized logging, market hunting, and a
bank loan in Congo. Conservation Biology 6: 570–580.
Wilkie D.S., Shaw E., Rotberg F., Morelli G. and Auzel P. 2000. Roads, development, and con-
servation in the Congo basin. Conservation Biology 14: 1614–1622.
Witte J. 1992. Deforestation in Zaire: logging and landlessness. The Ecologist 22: 58–65.
Wolfire D.M., Bruner J. and Sizer N. 1998. Forests and the Democratic Republic of Congo:
opportunity in a time of crisis. World Resources Institute, Washington, D.C.
Zahawi R.A. and Augspurger C.K. 1999. Early plant succession in abandoned pastures in Ecuador.
Biotropica 31: 540–552.
[337]
Biodiversity and Conservation (2006) 15:1399–1415 Springer 2006
DOI 10.1007/s10531-005-5410-0
Key words: Black woodpecker, Dead wood, Dendrocopos leucotos ssp. lilfordi, Dryocopus martius,
Forest structure, Population dynamics, Spain, Sustainable forest management, White-backed
woodpecker
Abstract. The woodlands of Quinto Real (Quinto Real, Erreguerena and Legua Acotada) are a
3,000 hectare beech (Fagus sylvatica) forest managed by the shelterwood system applied to even-
aged (regular) stands. This study analyses how forest management determines the local distribution
of the white-backed woodpecker (Dendrocopos leucotos) and black woodpecker (Dryocopus mar-
tius) and its relationship with the type, structure and size of the stands used for nesting by both
species, as well as their dead wood requirements. The most suitable nesting habitat of both species
is the mature forest (stands of regular large final crop trees), but the size of the mature fragments
and a minimum quantity of dead wood is also important.
Introduction
[339]
1400
[340]
1401
Methods
Study area
The study was carried out in the Quinto Real group of woodlands, located in
northern Navarre in the Baztán and Erro valleys on the Spanish side of the
Pyrenees. The Quinto Real area consists of three woodlands: Quinto Real
[341]
1402
(1666 ha), Erreguerena (941 ha) and Legua Acotada (907 ha), listed under
numbers 2, 3 and 4 in the Navarre’s Public Utility Woodlands Catalogue
(Gobierno de Navarra 1998). These woodlands are some of the best-preserved
beech forests in the Pyrenees. As a result, the Regional Government of Navarre
recently proposed them as a Place of Community Interest within the frame-
work of the European Union’s Natura 2000 Network. Forest management in
this area is governed by a Management Project and its subsequent reviews
(Schwendtner and Larrañaga 2001).
The Quinto Real group of woodlands has been managed for timber pro-
duction since 1904. Extensive areas of regeneration, resulting from shelterwood
system harvesting, are present. This kind of exploitation was particularly
intensive in the period 1950–1970. There are also mature areas where thinning
has been carried out with varying degrees of intensity, while others have not
been harvested in the last 70 years. Other areas are characterised by their
heterogeneity and unevenness as a result of high-grading of the valuable
timber.
Forest characterization
The Quinto Real Natural Resources Management Plan carried out a detailed
inventory for the different forest stands and their classification according to
structural criteria (Schwendtner and Larrañaga 2001). Table 1 provides
information on the different stand types on which the three woodlands of
Quinto Real were divided. A total of 397 different homogenous forest patches
(stands) were distinguished in the three woodlands. The following variables
were measured to characterize the forest structure: dominant height (DH),
measured as the mean height of those trees with an average diameter, excluding
the stems under 20 cm of diameter, except for the young stands (RY), in which
only stems under 10 cm of diameter are excluded; basal area (BA) of all stems
over 10 cm of diameter; average diameter (AD) of the stems over 10 cm,
measured at breast height; and average age (AA) of the stand, based on several
individuals (approximately 1%) of each stand, by counting the growth rings on
wood samples. All these measures were made on all stands over all the stand
area for management purposes.
Site quality is a variable that is calculated from the relationship between the
mean growth rate and tree age. Site quality can be categorized on a scale
ranging from I to V: very good and good quality sites are I and II, while III and
IV would represent intermediate qualities, and V poor quality sites
(Schwendtner and Larrañaga 2001). In better quality sites, trees grow more
rapidly that on poorer quality sites, where the harvesting cycle is not as rapid.
Schwendtner and Larrañaga (2001) proposed to exploit site qualities I and II,
but advised against timber harvest on the lower quality sites.
Basal area of standing dead trees (DBA) and the amount of felled trunks
were also determined for each stand.
[342]
1403
Table 1. Stand classification by management objectives on the Quinto Real Natural Resources
Management Plan (Schwendtner and Larrañaga 2001).
1. Stand of regular large RLF Mature forest. Average tree diameter >45 cm.
final crop trees Suitable for final cutting.
2. Stand of regular medium RM Medium sized and aged stand. Average tree diameter
sized crop trees from 20 to 45 cm. For intermediate cuttings
with economic value.
3. Stand of regular RY Young stand. Average diameter from 10 to 20 cm.
young trees For thinning without economic value.
4. Heterogeneous and HI Mixed stand, It is heterogeneous when it has
irregular stand different species and irregular when it has
different age classes and structures.
Various diameters.
5. Low forest stand LF Low forest stand. Generally on sites
of poor quality.
6. Open large final crop trees OF Open zones in regeneration process.
Some residual large trees.
Basal Area <15 m2/ha.
Woodpeckers census
The distribution and density of the black woodpecker and the white-backed
woodpecker were established by determining their breeding territories during
spring 2001. The low density of both species makes sampling difficult.
However, since they are highly territorial animals, the location of breeding
territories was used to census these species (Svenssons 1979; Tellerı́a 1986;
Bibby et al. 1992). The method was the same as the one used in the previous
censuses by Fernández and Azcona (1996), so that the densities and distri-
bution of territories could be compared. Recordings of the birdcalls and
tapping patterns of both species were used as decoys for locating the breeding
territories. For the density estimates, the ‘open land’ or patches with no trees
were excluded of the total study area (3200 ha) and not sampled, though all
the other stands of the study area were sampled. To attract territorial birds
or to provoke their response during the search, the tapped calls were played
every one or two hundred meters, alternating with periods of 30 s of silence,
thus permitting to detect the bird response and to locate the individual. Once
an individual was located, it was followed to locate the nest and the partner,
registering the stands they defended and used for foraging. The wood boring
signs were only taken into account when they were extremely recent and very
abundant and they were only used to determine areas where investigation
should be intensified. Alone they were not considered as sufficient proof of
the existence of a nesting area. Sightings of non-territorial individuals were
excluded.
[343]
1404
Habitat selection
The effect of different forest variables on the breeding area selection of the two
bird species (presence or absence of bird territories) was investigated. There-
fore, the conclusions that may be drawn from this study concern the territories
used for breeding and not the habitats used during other seasons. For the study
on dead wood, the three woodlands of Quinto Real were subdivided into
‘quarters’ that represent smaller management units. Quinto Real was divided
into three quarters, and Erreguerena and Legua Acotada into two quarters
each (Schwendtner and Larrañaga 2001).
Statistical analysis
[344]
1405
Results
From the analysis of forest stands, it appears that approximately 40% of the
area consist of high-quality sites (I and II), 30% of intermediate quality sites
(III and IV), and 30% of low quality sites. In the low quality sites are included
the non-exploitable sites due to environmental constraints (known as protec-
tion patches). Extraction priority was given to areas that are more productive
or easily accessible (Schwendtner and Larrañaga, 2001), so there is a certain
imbalance in the age histogram according to site qualities (results not shown).
Stand type distribution in the studied area is explained in Table 2.
Census
Table 2. Area (hectares), number of stands and average stand size (hectares) of the tree woodlands
of Quinto Real: Quinto = Quinto Real; Erreg. = Erreguerena; Legua = Legua Acotada.
[345]
1406
2001 1993
WW BW WW BW
Number of territories
Total 11 14 12 13
Quinto 7.5 7.5 6 6
Erreguerena 1.5 4.5 4 4
Legua 2 2 2 3
Density (pairs/Km2)
Total 0.34 0.44 0.38 0.41
Quinto 0.49 0.49 0.48 0.48
Erreguerena 0.18 0.54 0.38 0.38
Legua 0.33 0.33 0.33 0.49
The territories of both species reveal certain mobility compared to the 1993
census carried out by Fernández and Azkona (1996). Many of the territories
are still located in exactly the same forest stands. Others clearly occupy the sites
situated between former territories, presumably using the areas that were less
used on the 1993 territory distribution. In territories where felling has been
carried out, the pairs affected have moved out, probably to other unoccupied
patches.
Frequencies of sightings of the two species in the different forest types show
significant differences (v2, p lt 0.005 for the black woodpecker and p < 0.00001
for the white-backed woodpecker). Both species show a clear preference for
regular large final crop stands (RLF). For both species there is also a distinct
negative selection against heterogeneous (mixed with conifers) and uneven
(mixed ages) stands (HI); this is less marked in the case of the black wood-
pecker. Sightings in regular medium sized crop stands (RM) and regular young
stands (RY) do not reveal any significant differences.
Although most of the territories cover various forest stand types, nearly all
cases – except in one black woodpecker and one white-backed woodpecker
territories – include an RLF stand. When nests were found, they were usually
located in this stand type, while the others – mainly RM and RY – are also
defended and used for feeding. In the two territories identified in a place
without RLF, there were RM stands of a considerable age (on the boundary
of stands regarded as RLF). In one of these cases, the territory may have
been moved from a recently exploited mature stand (RLF). Nevertheless, in
[346]
1407
Table 4. Means comparison (LSD) of dominant height (DH), average age, average diameter (AD)
and basal area (BA) from the Quinto Real Natural Resources Management Plan (Schwendtner and
Larrañaga 2001).
all cases in which a territory is included in only one big stand (3 white-backed
woodpecker territories and 4 black woodpecker territories), this is a RLF
stand.
No significant differences were found between the site quality of the stands
used by each species with those not used. Therefore, woodpeckers do not
appear to choose stands for their site quality, but rather for the physiognomic
characteristics of the forest which is best reflected in the stand classification (see
Table 1 and 4). This supports the option for only harvesting stands with high
site quality, and conserving the poorer quality sites with well-preserved mature
forest (RLF).
Stand size
One of the typical questions that arises from the management of these forests
concerns the minimum stand size that must be left as mature forest for these
species to establish their territory. To answer this question, a study was con-
ducted on the size differences between RLF stands where territories were
present and those where they were not. Figure 1 illustrates that the stands
where territories for both species were present were considerably larger than
those where there were not.
The RLF stands in which the black woodpecker appears have an average
size of 24 ha, while those of the white-backed woodpecker average 19 ha. The
[347]
1408
Figure 1. Analysis of variance on the regular large final crop trees (RLF) forest stands size
according to their occupation for each woodpecker species: BN, stands where the black wood-
pecker territories are studied; WN, stands where the white-backed woodpecker does not appear;
WT, stands where the white-backed woodpecker territories are situated. The rectangles represent
the standard error and the lines the standard.
Figure 2. Percentage of RLF stands of different sizes included in territories of black woodpecker
(triangles) and white backed woodpecker (circles).
actual surfaces required are probably somewhat greater, since some territories
occupy more than one mature stand. In fact, most RLF stands with an area
exceeding 30 ha (8 stands) are included in a territory, except three stands in
which recent cutting was carried out. In Figure 2 can be seen the percentages of
the different sizes of RLF stands that are included in the territories of each
species. The total percentage for each species is higher that 100 because one
territory usually extend over several stands. The occupation percentage in large
stands is much higher than that for small stands.
[348]
1409
Forest physiognomy
Dead wood
When stands of the three woodlands of Quinto Real were used, no significant
differences were found for total dead wood (number of trunks per ha) between
stands used by each species or those that were not used, probably because
many stands with considerable amounts of dead wood were not used by neither
of the two species. Nevertheless, there were notable differences in the amount
of thin dead wood (from 10 to 20 cm), i.e., the type of dead wood most
abundant and most representative (Table 5).
Comparing only stands from quarters that had comparatively little dead
wood, significant differences were found among stands where the white-backed
woodpecker appeared, particularly for the 20–30 cm range (p < 0.05). This
species did not choose stands with more dead wood in quarters where it was
abundant, but in those quarters with little dead wood, it did a positive selection
for stands with more dead wood.
For the dead wood in RLF, considerable differences might be identified
between stands where the white-backed woodpecker was found and stands
where it was absent for the two largest diameter classes of dead wood
[349]
1410
Figure 3. Principal components analysis of the stands described by the four variables that describe
the forest structure. The first two axis do account for the 85% of the variance. (a) scatter diagram
of all the stands; symbols indicate the stand classification according to Table 1. RLF, stands of
regular large final crop trees; RM, stands of regular medium sized crop trees; RY, stands of regular
young trees; HI, heterogeneous and irregular stands; LF, low forest stands. (b) scatter diagram of
the larger RLF stand of each woodpecker territory on the same axis that (a). Circles represent the
black woodpecker territories and triangles the white backed woodpecker territories. Also the de-
scriptors of the four variables that describe the forest structure are represented. BA, basal area;
DH, dominant height; AD, average diameter; AA, average age.
(20–30 cm and >30 cm; p < 0.001). This implies that the species clearly
chooses those RLF that contain a larger amount of dead wood. This
relationship remains significant comparing the differences between large-size
RLF stands (>15 ha) whether the white-backed woodpecker is present or not.
For the black woodpecker no differences were found for all the comparisons.
[350]
1411
Table 5. Standing dead wood on the seven quarters of Quinto Real Woodlands: three in Quinto
Real (Q1,Q2, Q3), two in Erreguerena (E1, E2) and two in Legua Acotada (L1, L2).
Trunk diameter
Discussion
Although the overall densities of the two woodpecker species remain rather
stable for each of the woodlands of Quinto Real, according to the 1993 census
(Fernández and Azkona 1996) and the one obtained in 2001, it appears that the
white-backed woodpecker is declining in Erreguerena, and the black wood-
pecker is declining in Legua Acotada and increasing in Erreguerena. These
differences are probably due to forestry management. Nevertheless, the Quinto
Real populations have increased, which seems to indicate that management has
been more appropriate than in Erreguerena and Legua Acotada. It does not
seem that this difference can be explained by other ecological variables, because
topography, climate, and other non-anthropic factors are quite similar.
Over the years both species remained faithful to their breeding territories (as
also observed by McClelland and McClelland 1999). By comparing both
censuses (1993–2001), it appears that there have been some shifts of territories,
possibly due to forestry activities. A possible explanation could be that changes
on the forest structure due to felling in a woodpecker territory, (a RLF stand is
converted on a RY stand) may cause a territorial movement towards another
place with more mature forest, thus ‘pushing’ adjacent territories.
One of the most obvious conclusions is that the most suitable habitat for
both species is the mature forest stands (RLF). These are also referred to in the
Natural Resources Management Plan (Schwendtner and Larrañaga 2001) as
stands where final cutting is most likely to be carried out because these sites
have the largest amount of timber trees for felling. If the exploitation of these
woodlands by town councils is intensified, these stands will soon become
extremely scarce.
A negative selection is observed, in both species, against heterogeneous and
irregular stands. The same occurs within the stands in which other species
[351]
1412
rather than beech are dominant (Larix, Pinus, Quercus). Although the black
woodpecker occasionally feeds in these forests, the white-backed woodpecker
has been found exclusively in monospecific beech forests.
The dominant height (26 m), age (149 years), mean diameter (34 cm) and
basal area (27 m2/ha) of stands where both the black and white-backed
woodpecker territories coincide may help to determine the characteristics that
the stands left for conservation should have (5% of the total). A sufficient
amount of this type of stand should be left in the rest of the woodlands to
ensure that the population of these species do not decrease.
It has been confirmed that the size of the RLF stand is another decisive
factor for both species in establishing breeding territories, with minimum sizes
close to 20–30 ha. As these are territorial birds, it does not appear to be a good
idea to leave all the RLF stands grouped in one area. It seems much more
appropriate to keep sufficiently large (>30 ha) RLF patches separated from
one other (the number of patches depends on the desired size of the popula-
tion). Moreover, given that there is no correlation between the site quality and
the distribution of territories, it is advisable to concentrate exploitation in the
best sites with short felling cycles. The worst sites should be left unexploited in
order to fulfill the above objectives.
The amount of dead wood does not appear to be the main factor for
choosing breeding territories for these birds at the scale of the whole study
area, probably because it is very abundant in most of the area and therefore it
is not a limiting factor. In fact, when the analysis is concentrated in the areas
where dead wood is scarcer, this variables becomes an important factor for the
distribution of the white-backed woodpecker territories, but the black wood-
pecker territories distribution do not seem to be affected by this variable. This
can be explained by the diet of the black woodpecker, which feeds mainly on
ants, and is not so dependent on dead wood.
The white-backed woodpecker territories distribution shows a relationship
with the amount of thicker standing dead wood in the quarters where total
dead wood is less abundant. Also the amount of dead wood seems important
when comparing the territories occupancy frequencies between all RLF stands
and also only with large RLF stands. But not all dead wood classes are of the
same importance: thick standing dead wood seems to be more important than
other classes of dead wood. This reinforce the importance of well conserved,
large enough RLF stands, also with sufficient amount of standing thick dead
trees, that could be increased by ringing some trees if necessary. It is more
important in the places where the surroundings have less dead wood. There are
other authors that have also found important the amount of dead wood for the
woodpeckers, in particular with the specialist species (Angelstam et al. 2003;
Butler et al. 2004).
Although the number of felling activities is insufficient to analyse their effect
from a statistical point of view, their effects on the territories of both species
seem to be very clear. When a RLF stand included in a 1993 territory disap-
peared, the territory has ‘moved’ to include another RLF stand in it. In this
[352]
1413
study, the movement of territories have not been analysed, and we do not know
if the new territories (or even the old ones) are done by the same individuals or
different ones. As example, the black woodpecker territory that has disap-
peared in Legua Acotada, correspond to a place where the RLF stand included
in it has been fallen down, and no mature stand can be found nearby.
At the moment in Legua Acotada it seems unlikely that the population of
both species can be increased to levels similar to those in Quinto Real unless
management is changed, for example by using smaller stand sizes and leaving
the small number of mature RLF stands (there were 3 in 1993, but one has
already been harvested). There is a territory of each species in each of these
stands, but the nearby forests are regular young stands (RY) or final cuttings
that are extremely homogenous and have been exhaustively ‘cleaned out’. As a
result, they are of no use to the picidae. If the RLF stands that remain are cut
down, the territories in them will surely disappear.
Nevertheless, the situation in Erreguerena is slightly more encouraging.
There are sufficiently large RLF stands in this area. Although it appears that
several white-backed woodpecker territories have disappeared due to recent
cuttings, they may have established in other RLF stands that were unoccupied.
Black woodpecker seems to be attracted by these felling activities, possibly due
to an increase of felled dead wood and therefore an increase of the amount of
ants. Appropriate management would mean cutting the unoccupied stands in
order not to disturb existing pairs. As the territories may vary in location, it is
necessary to conduct yearly censuses in order to determine the situation before
planning felling activities.
The patchwork situation in Quinto Real – stands that are relatively small,
with a relatively high abundance of large enough RLF stands sufficiently
separated ones from the others – has allowed the creation of a large number of
breeding territories. Nevertheless, it is advisable to exploit only the stands that
remain between territories and leave those that currently contain breeding
territories.
Another general recommendation from the results is to leave enough dead
wood in all stands and to leave dead trees standing because this is where the
white-backed woodpecker mainly feeds. Girdling can be carried out instead of
harvesting in some cases, since this technique leaves standing dead wood. Some
management plans which aimed at protecting yew trees (Taxus baccata) by
girdling the beech trees that overshadow them, may be also beneficial to spe-
cialist species (see Carlson 2000). White-backed woodpeckers are also benefi-
cial for the health of the beech forest as it eats a lot of the forest plagues, and
keeps them under control (Butler and Schlaepfer 2003).
Acknowledgements
[353]
1414
provided about Picidae. Our thanks also to the ‘‘Sección de Montes del Servicio
de Conservación de la Biodiversidad del Gobierno de Navarra’’ for their support
and to the Foreign Language Co-ordination Office at the Polytechnic Uni-
versity of Valencia for their help in translating this paper.
References
Angelstam P. 1990. Factors determining the composition and persistence of local woodpecker
assemblages in taiga forests in Sweden – a case for landscape ecological studies. In: Carlson A.
and Aulen G. (eds), Conservation and Management of Wo‘odpecker Populations. Swedish
University of Agricultural Sciences, Department of Wildlife Ecology, Uppsala, pp. 147 – 191.
Angelstam P. and Mikusinski G. 1994. Woodpecker assemblages in natural and managed boreal
and hemiboreal forest – a review. Annales Zoologici Fennici 31: 157 – 172.
Angelstam P., Butler R., Lazdinis M., Mikusinski G. and Roberge J.M. 2003. Habitat thresholds
for focal species at multiple scales and forest biodiversity conservation – dead wood as an
example. Annales Zoologici Fennici 40(6): 473 – 482.
Aulén G. 1988. Ecology and distribution history of the White-backed Woodpecker Dendrocopos
leucotos in Sweden. Ph.D. Thesis, Report 14, Swedish University of Agricultural Sciences
Department of Wildlife Ecology, Uppsala.
Aulén G. and Lundberg A. 1991. Sexual dimorphism and patterns of territory use by the White-
backed Woodpecker Dendrocopos leucotos. Ornis Scandinavica 22: 60 – 64.
Avery M. and Leslie R. 1990. Birds and Forestry. T. & A.D. Poyser, London.
Bibby C.J., Burgess N.D. and Hill D.A. 1992. Bird Census Techniques. Academic Press, London.
Blanco J.C. and González J.L. 1992. Libro Rojo de los Vertebrados de España. ICONA, ed.
Ministerio de Agricultura Pesca y Alimentación, Madrid.
Brooks J. 1985. Handbook of the Birds of Europe, the Middle East and North Africa. Vol IV.
Oxford University Press.
Butler R., Algelstam P., Ekelund P. and Schlaepfer R. 2004. Dead wood threshold values for the
three-toed woodpecker presence in boreal and subalpine forest. Biological Conservation 119:
305 – 318.
Butler R. and Schlaepfer R. 2003. Three-toed woodpecker as an alternative to bark beetle control
by traps?. In: Pechacek P and d’Oleire-Oltmanns W. (eds), International Woodpeckers Sym-
posium. Berchtesgaden, pp. 13 – 26.
Carlson A. 2000. The effect of habitat loss on a deciduous forest specialist species: the White-
backed Woodpecker (Dendrocopos leucotos). Forest Ecology and Management 131: 215 – 221.
Conner R. 1979. Seasonal changes in woodpecker foraging methods: Strategies for winter survival
Pages 95 – 105. In: Dickson J.G., Conner R.N., Fleet R.R., Kroll J.C. and Jackson J.A. (eds),
The role of insectivorous birds in forest ecosystems. Academic Press, NY, pp. 381.
Cramp S. (ed.) 1985. Handbook of the Birds of Europe, the Middle East and North Africa. Vol. 4.
Oxford, Oxford University Press.
Cuisin M. 1967. Essai d’une monographie du Pic noire (Dryocopus martius). Oiseau Rev Fr
Ornithol. 37:163 – 192, 37:285 – 315, 38:20 – 52, 38:103 – 126, 38:209 – 224.
Fernández C., Azkona P. and Lorente L. 1994. Corologı́a y caracterización del hábitat del pico
dorsiblanco (Dendrocopos leucotos) en el Pirineo occidental español. Ardeola 41(2): 135 –
140.SEO.
Fernández C. and Azkona P. 1996. Influence of forest structure on the density and distribution of
the White-backed Woodpecker Dendrocopos leucotos and black Woodpecker Dryocopus martius
in Quinto Real (Spanish western Pyrenees). Bird Study 43: 305 – 313. British Trust for
Ornithology.
Fleishman E., Blair R.B. and Murphy D.D. 2001. Empirical validation of a method for umbrella
species selection. Ecological Applications 11: 1489 – 1501.
[354]
1415
Fleishman E., Jonsson B.G. and Sjagren-Gulve P. 2000. Focal species modeling for biodiversity
conservation. Ecological Bulletins 48: 85 – 99.
Gobierno De Navarra 1990. Ley Foral 13/1990, de 31 de diciembre. Protección Y Desarrollo del
Patrimonio Forestal de Navarra. Boletı́n Oficial de Navarra núm. 6, de 14 de enero de 1991.
Gobierno De Navarra 1992. Decreto Foral 59/1992, de 17 de febrero. Reglamento de Montes en
Desarrollo de la Ley Foral 13/1990, de 31 de Diciembre, de Protección y Desarrollo del Patri-
monio Forestal de Navarra. Boletı́n Oficial de Navarra núm. 48, de 20 de abril de 1992.
Gobierno De Navarra 1998. Plan Forestal de Navarra. www.cfnavarra.es, .
Gobierno De Navarra 1999. Estrategia Navarra para la Conservación de la Biodiversidad.
www.cfnavarra.es.
Hogstad O. 1970. On the ecology of the three-toed woodpecker Picoides tridactylus (L.) outside the
breeding season. Nytt Magasin for Zoologi 18: 221 – 227.
Imbeau L. and Desrochers A. 2002. Foraging ecology and use of drumming trees by three-toed
woodpeckers. Journal of Wildlife Management 66: 222 – 231.
Martinez-Vidal R. 1999. Hábitat de crı́a del pito negro (Dryocopus martius) en las Sierras de Cadı́ y
Moixeró: caracterización, tipologı́a y pérdidas de árboles nido. Gestión y Conservación de la
biodiversidad en ecosistemas forestales. C.T.F.C, Solsona.
Mcclelland B. and Mcclelland P. 1999. Pileated woodpecker nest and roost trees in Montana: links
with old-growth and forest ‘‘health’’. Wildlife Society Bulletin 27(3): 846 – 857.
Mikusinski G. and Angelstam P. 1997. European woodpeckers and anthropogenic habitat change:
a review. Vogelwelt 118: 277 – 283.
Murphy E.C. and Lehnhausen W.A. 1998. Density and foraging ecology of woodpeckers following
a stand-replacement fire. Journal of Wildlife Management 62: 1359 – 1372.
Nilsson S.G. 1992. Population trends and fluctuations in Swedish woodpeckers. Ornis Svecica 2:
13 – 21.
Purroy F.J. 1972. El pico dorsiblanco Dendrocopos leucotos del Pirineo. Ardeola 20: 145 – 158.
Purroy F.J., Alvarez A. and Pettersson B. 1990. Bosque y fauna de vertebrados terrestres en
España. Ecologı́a 1: 349 – 363.
Roberge J.M. and Angelstam P. 2004. Usefulness of the umbrella species concept as a conservation
tool. Conservation Biology 18(1): 1 – 10.
Schwendtner O. and Larrañaga A. (coords.) 2001. Sexta Revisión de la Ordenación del grupo de
Montes de Quinto Real. BASOA/Gobierno de Navarra.
Simberloff D. 1999. The role of science in the preservation of forest biodiversity. Forest Ecology
and Management 115: 101 – 111.
Svenssons S.E. 1979. Census Efficiency and number of visits to a study plot when estimating bird
densities by the territory mapping method. Journal of Applied Ecology 16: 61 – 68.
Tellerı́a J.L. 1986. Manual para el Censo de los Vertebrados Terrestres. Ed. Raı́ces, Madrid.
Tellerı́a J.L. 1992. Gestión forestal y conservación de las aves en España peninsular. Ardeola 39:
99 – 114.
Tucker G.M. and Heath M.F. (eds) 1994. Birds in Europe – their conservation status. BirdLife
International, Conservation Series No. 3, Cambridge.
Voous K.H. 1947. On the history of the distribution of the genus Dendrocopos. Limosa 20: 1 – 142.
[355]
Biodiversity and Conservation (2006) 15:1417–1424 Springer 2006
DOI 10.1007/s10531-005-0308-4
STEVEN M. VAMOSI
Department of Biological Sciences, University of Calgary, 2500 University Drive NW, Calgary,
Alberta, Canada, T2N 1N4; E-mail: [email protected]
Key words: Brazil, Dioecy, Floristic survey, Reproductive biology, Seed shadow effect
Abstract. Published species lists that include breeding system designations of vascular plants are
rare in the primary literature and, thus, can be potentially valuable sources of information for
comparative studies. The published list for vascular plants in the Volta Velha Reserve suffered from
a number of errors, notably applying the designation of monoecious to all species with imperfect
flowers. Here, I reconsider the breeding systems for 97 woody vascular plant species. The majority
of species initially categorized as monoecious are found to be hermaphroditic. I then examine the
relationship between breeding system and numbers of individuals in a 1 ha plot. The mean number
of individuals was marginally higher in dioecious than hermaphroditic and monoecious species
combined. Furthermore, although only 28% of the species were characterized as possessing a
dioecious breeding system, 42% of the individuals encountered belonged to a dioecious species.
These results suggest that dioecious species can, at least under certain circumstances, overcome the
reductions in the number of seed-bearing individuals and mate assurance that accompany pos-
sessing spatially segregated sexes.
Introduction
[357]
1418
Methods
[358]
1419
The main effect of re-scoring breeding systems was, perhaps not surprisingly, to
drastically reduce the apparent prevalence of monoecious species (Table 1). In
agreement with other studies, hermaphroditic species were most common
(68%), followed by dioecious and androdioecious (28%) species combined,
and monoecious and polygamous (4%) species combined. These values are
remarkably similar to those observed for a sample of 139 tree species from a
tropical rainforest in Mexico (Ibarra-Manrı́quez and Oyama 1992). In Los
Tuxtlas, the frequency of species with the different sexual systems was reported
as 63, 27, and 9%, respectively. The apparently high incidence of dioecy in the
current survey likely reflects the fact that only individuals with DBH ‡10 cm
were sampled. Because of the correlation between dioecy and a woody growth
form, especially in the tropics (Vamosi and Vamosi 2004), the incidence of
dioecy in the Volta Velha Reserve will almost certainly be lower with the
inclusion of herbaceous species (see also Ibarra-Manrı́quez and Oyama 1992).
There was a marginally higher number of individuals in dioecious
(mean = 7.92) than hermaphroditic + monoecious (mean = 4.81) species
[359]
1420
Table 1. Breeding system of 97 plant species surveyed in a 1 ha plot of the Volta Velha Reserve.
Podocarpus sellowii (Podocarpaceae) is a gymnosperm, whereas the remaining species are angio-
sperms.
[360]
1421
Table 1. (Continued).
[361]
1422
Table 1. (Continued).
10
9
8
Square root number
7
of individuals
6
5
4
3
2
1
0
Dioecious Nondioecious
Breeding system
Figure 1. Square root transformed number of individuals with a dioecious (diamonds) or non-
dioecious (hermaphroditic and monoecious; triangles) breeding system in a 1 ha plot of the Volta
Velha Reserve. N = 27 dioecious species, 70 nondioecious species.
[362]
1423
Acknowledgements
References
Arroyo M.T.K. and Squeo F. 1990. Relationship between plant breeding systems and pollination.
In: Kawano S. (ed.), in Biological Approaches and Evolutionary Trends in Plants. Academic
Press, London, pp. 205–227.
Bush M.B. 1995. Neotropical plant reproductive strategies and fossil pollen representation. Am.
Nat. 145: 594–609.
Charlesworth D. 1985. Distribution of dioecy and self-incompatibility in angiosperms. In:
Greenwood J., Harvey P.H. and Slatkin M. (eds), Evolution: Essays in Honour of John May-
nard Smith. Cambridge University Press, Cambridge, pp. 237–268.
Chazdon R.L., Careaga S., Webb C. and Vargas O. 2003. Community and phylogenetic structure
of reproductive traits of woody species in wet tropical forests. Ecol. Monogr. 73: 331–348.
Darwin C. 1877. The Different Forms of Flowers on Plants of the Same Species. John Murray,
London.
Darwin C. 1878. The Effects of Cross and Self Fertilisation in the Vegetable Kingdom. John
Murray, London.
Flores S. and Schemske D.W. 1984. Dioecy and monecy in the flora of Puerto Rico and the Virgin
Islands: ecological correlates. Biotropica 16: 132–139.
[363]
1424
Geber M.A., Dawson T.E. and Delph L.F. (eds) 1999. Gender and Sexual Dimorphism in
Flowering Plants. Springer-Verlag, New York.
Heilbuth J.C. 2000. Lower species richness in dioecious clades. Am. Nat. 156: 221–241.
Heilbuth J.C., Ilves K. and Otto S.P. 2001. The consequences of dioecy on seed dispersal: modeling
the seed-shadow handicap. Evolution 55: 880–888.
Hutchinson J. 1959. The Families of Flowering Plants, 2nd edn. Clarendon Press, Oxford.
Ibarra-Manrı́quez G. and Oyama K. 1992. Ecological correlates of reproductive traits of Mexican
rain forest trees. Am. J. Bot. 79: 383–394.
Kessler P.J.A. 1993. Annonaceae. In: Kubitzki K., Rohwer J.G. and Bittrich V. (eds), The Families
and Genera of Vascular Plants. Volume II. Flowering Plants. Dicotyledons: Magnoliid,
Hamamelid and Caryophyllid Families. Springer-Verlag, Berlin, pp. 93–129.
Kubitzki K., Rohwer J.G. and Bittrich V. (eds) 1993. The Families and Genera of Vascular Plants.
Volume II. Flowering Plants. Dicotyledons: Magnoliid, Hamamelid and Caryophyllid Families.
Springer-Verlag, Berlin.
Kubitzki K. (ed.) 2004. The Families and Genera of Vascular Plants. Volume VI. Flowering Plants
Dicotyledons: Celastrales, Oxalidales, Rosales, Cornales, Ericales. Springer-Verlag, Berlin.
Lieberman M. and Lieberman D. 1994. Patterns of density and dispersion of forest trees. In:
McDade L.A., Bawa K.S., Hespenheide H.A. and Hartshorn G.S. (eds), in La Selva: Ecology
and Natural History of a Neotropical Rain Forest. University of Chicago Press, Chicago,
pp. 106–119.
Nanami S., Kawaguchi H. and Yamakura T. 1999. Dioecy-induced spatial patterns of two
codominant tree species, Podocarpus nagi and Neolitsea aciculata. J. Ecol. 87: 678–687.
Negrelle R.R.B. 2002. The Atlantic forest in the Volta Velha Reserve: a tropical rain forest site
outside the tropics. Biodiv. Conserv. 11: 887–919.
Pannell J.R. and Barrett S.C.H. 1998. Baker’s Law revisited: reproductive assurance in a meta-
population. Evolution 52: 657–668.
Pitman N.C.A., Terborgh J.W., Silman M.R., Núñez V.P., Neill D.A., Cerón C.E., Palacios W.A.
and Aulestia M. 2001. Dominance and distribution of tree species in upper Amazonian terra
firme forests. Ecology 82: 2101–2117.
Renner S.S. and Ricklefs R.E. 1995. Dioecy and its correlates in the flowering plants. Am. J. Bot.
82: 596–606.
Rohwer J.G. 1986. Produmus einer monographie der gattung Ocotea Aubl. (Lauraceae), sensu lato.
Mitteilungen aus dem Institut für Allgemeine Botanik Hamburg 20: 3–278.
Vamosi J.C. and Vamosi S.M. 2004. The role of diversification in causing the correlates of dioecy.
Evolution 58: 723–731.
Watson L. and Dallwitz M.J. 1992 onwards. The Families of Flowering Plants: Descriptions,
Illustrations, Identification, and Information Retrieval. Version: 14th December 2000. http://
biodiversity.uno.edu/delta/.
Yampolsky C. and Yampolsky H. 1922. Distribution of sex forms in the phanerogamic flora. Bib.
Genet. 3: 1–62.
[364]
Biodiversity and Conservation (2006) 15:1425–1440 Springer 2006
DOI 10.1007/s10531-005-0310-x
-1
Abstract. Patterns of rodent species abundance and diversity were examined over a 5 months
period in two areas of a Kenyan relict tropical rainforest. The two areas are subjected to different
administrations which lead to various levels of anthropogenic disturbance: one can be considered
relatively disturbed and one relatively undisturbed. Anthropogenic disturbance causes a reduction
in woody stem density between 0 and 1.5 m and reduced understory tree canopy cover. Rodent
abundance was estimated using the program CAPTURE and compared with the number of
individuals actually captured. Density was estimated with three different methods, two of these
utilised a boundary strip to estimate effective size of the area trapped. Density resulted in being
relatively high in both areas, so population might have been at a peak. Species richness was higher
in the disturbed forest, while species diversity and evenness was higher in the undisturbed forest.
We suggest that in the disturbed forest the increase in number of species might be due to sporadical
entrance in the forest by non-forest species, while the decrease in diversity might be due to the
decrease of lower strata vegetation that occurs in the disturbed forest, hence this factor might affect
species equitability. Bibliographic data supports this hypothesis as rodent species diversity and
ground vegetation cover have been found to be correlated.
Introduction
[365]
1426
[366]
1427
The study area was located in the Kakamega Forest, Kenya (latitude: 0010¢ N–
0021¢ N, longitude 3447¢ E–3458¢ E; 1500–1700 m a.s.l). According to Lucas
(1968) the forest is the only reasonably large patch of Central African type
lowland rainforest in Kenya. Mean annual precipitation is 2000 mm (Cords
1990). Some of commonest trees are Celtis africana, Prunus africana, Albizia
gummifera, Antiaris toxicaria (Cords 1990; KIFCON 1994).
Research was carried out in the field from November 2002 to April 2003,
during the dry season.
Rodents were live-trapped using large aluminium Sherman traps. Diced fried
coconut mixed with peanut butter was used as bait. Traps were inspected once
a day. Animals were marked by toe clipping and standard data was taken from
each animal before releasing it: specimen number, trap location, sex, repro-
ductive condition, body weight.
The trapping pattern was determined by the number of traps temporarily
available and by logistical factors. The main objective was to obtain two
specular sets of grids in each type of forest in order to allow comparisons. All
grids covered an area of 0.81-ha. Each forest area had:
(1) one 7 · 7 grid with 15 m of trap spacing, two traps at each station, one at
ground level and one arboreal trap (1–3 m), a trapping period of 4 days;
(2) one 7 · 7 grid with 15 m of trap spacing, one ground level trap at each
station, a trapping period of 4 days;
(3) one 10 · 10 grid with 10 m of trap spacing, two traps at each station, one at
ground level and one arboreal trap (1–3 m), a trapping period of 3 days
(Table 1).
Due to the higher interspersion between cultivated fields and forest in the
disturbed forest area, the grids resulted in being closer to the forest edge
(nearest trap of each grid approximately 50, 80 and 150 m) while the grids of
the undisturbed forest were relatively distant (nearest trap approximately 300,
600 and 700 m).
Three trapping sessions (December, February, March) were performed in
the undisturbed forest whereas, due to unpredictable logistic problems, two
trapping sessions (February, March) in the disturbed Forest. The 10 · 10 grid
Table 1. Grid label and grid collocation, trap numbers and trap collocation.
[367]
1428
was used only in February and March sessions of each forest. A removal grid
with variable number of traps was occasionally added in each forest in order to
gather 10 skulls of each sex of each species for accurate species determination.
We followed the classification and diagnostic characters of Lecompte et al.
(2001, 2002) and Delany (1975).
Rodent abundance in each grid was estimated using the program
CAPTURE which selects the most appropriate model (Otis et a1. 1978).
Density was estimated in three different ways: the first method consisted in
dividing the number of individuals/grid area; the other two methods applied a
correction to the grid area by adding a boundary strip in order to estimate the
effective size of the area trapped (Krebs 1999). The simplest procedure is to add
a strip one-half the movement radius of the animals under study (Krebs 1999).
In the first case, we used movement radius of animals of the same species in the
same grid. As this first method may be biased by a low number of recaptures, in
the second case we used movement radius of all individuals of the same species
captured in the whole trapping period. Abundance comparisons were per-
formed for each species using a capture index (number of unique individuals/
trap nights for each trapping session, Nichols and Dickman 1996) with a t-test.
Species richness was estimated by the number of species captured. Rodent
species diversity was estimated using the Shannon–Wiener and Simpson indices
(Krebs 1999); the importance of species was measured by the number of
individuals actually captured (Krebs 1999). Values for Shannon–Wiener Index
are espressed as N1 that is number of equally common species that would
produce the same diversity as H¢, while values for Simpson Index are espressed
as 1/D (Krebs 1999). Community evenness was measured using the Shannon–
Wiener measurement (H/Hmax) and Simpson measurement. (1/D/number of
species) (Hair 1980; Krebs 1999).
Results
The study covered a total of 5340 trap nights: 2376 in the undisturbed forest
and 2964 in the disturbed forest. In the whole forest a total of eight species was
captured. Species and number of individuals caught were: 274 Praomys jack-
soni, 82 Hylomyscus stella, 13 Lophuromys flavopunctatus, 4 Mus (Nannomys)
minutoides, 1 Graphiurus sp., 1 Otomys sp., 1 Lemniscomis sp., 1 Mastomys sp.).
Less abundant species (one single capture) were identified only to the genus
level as no skull was available for accurate species determination. All species
except for Graphiurus sp. were caught in the disturbed forest, instead only the
first five species were caught in the undisturbed forest. The number of indi-
viduals of both Praomys and Hylomyscus varied, changing both in grids and in
time; no significant correlation was found between rodent abundance and trap
numbers (Spearman correlation: r = 0.153, p = 0.43, n = 28). Comparisons
of the abundance of each species between macrohabitats was performed with a
t-test on the capture index (unique individuals captured/trap nights, Table 2)
[368]
1429
Table 2. Capture index (unique individuals captured/trap nights) for Praomys jacksoni and
Hylomyscus stella.
g6 g6 g6 g3 g3 g1 g1 g4 g4 g4 g5 g5 g2 g2
s1 s2 s3 s1 s2 s1 s2 s1 s2 s3 s1 s2 s1 s2
Praomys 0.06 0.09 0.07 0.08 0.11 0.14 0.1 0.07 0.11 0.11 0.03 0.02 0.05 0.06
jacksoni
Hylomyscus 0.04 0.02 0.01 0.03 0.04 0 0.01 0.04 0.04 0.02 0.01 0.01 0.02 0.01
stella
Results are shown for each session of each trapping grid; g – grid; s –session.
Figure 1. Models used by program CAPTURE. In black Praomys jacksoni, in white Hylomyscus
stella. X axis – model name; Y axis – absolute frequencies. See Otis et al. (1978) for models
description.
[369]
1430
Figure 2. Comparison between CAPTURE estimate and actual number of individuals captured.
Praomys jacksoni (above) and Hylomyscus stella (below): in white CAPTURE estimate, in black
actual number individuals captured. X axis – number of rodents; Y axis – grid number and session
number.
Figure 3. Praomys jacksoni CAPTURE estimate (with selected model) and confidence interval.
Leg: p – Praomys jacksoni; g6-1 – grid 6 session 1.
[370]
1431
Figure 4. Hylomyscus stella. CAPTURE estimate (with selected model) and confidence interval.
Leg: h – Hylomyscus stella; g6-1 – grid 6 session 1.
(undisturbed forest) and grid 1 and 2 (disturbed forest); Graphiurus sp. was
caught only in grid 6 (undisturbed forest), Lemniscomys was caught only in
grid 3 (disturbed forest); Otomys sp. and Mastomys sp. were caught only in grid
2 (disturbed forest). Grid 1 and 2 of the disturbed forest supported a higher
number of species than their corresponding grids in the undisturbed forest, but
in grid 6 of the undisturbed forest a higher number of species was caught in
comparison to its analogous in the disturbed forest (Table 5).
In all the comparisons with the Shannon–Wiener Index between analogous
grids, species diversity resulted higher in the undisturbed forest, while with the
Simpson Index in grid 6 (undisturbed forest) we found a lower diversity than its
equivalent grid in the disturbed forest. As one more trapping session was
performed in grid 4 and 6 of the undisturbed forest, the Diversity Indices were
also calculated without considering individuals caught in the last session, re-
sults partially confirm the higher values for the undisturbed forest: in this case
in grid 6 (undisturbed forest) we found a higher value of the Simpson Index in
comparison to the analogous grid (grid 3) in the disturbed forest thus, all the
undisturbed forest grids were characterised by a higher diversity (Table 5).
In grid 4 and 5 of the undisturbed forest there was a higher Evenness than
their correspondant grids in the disturbed forest, but this pattern was inverted
in grid 6 of the undisturbed forest.
If we pool data from all the grids of the same forest type we obtain a higher
community diversity and higher evenness in the undisturbed forest (Table 5).
In the double-trap grids Hylomyscus stella was captured above ground (1–3
m) 55.7% of times (n = 159), Praomys jacksoni 25.3 % of times (n = 247).
The number of above ground captures for Hylomyscus stella resulted sig-
nificantly higher than the number of above ground captures for Praomys
[371]
1432
Table 3. Praomys jacksini: density estimates.
1433
1434
Table 5.
Grid 6 undisturbed 5 (P. jacksoni, H. stella, 2.03 (2.11) 1.66 (2.08) 0.44 (0.46) 0.33 (0.41)
L. flavopunctatus, M. minutoides,
Graphiurus sp.)
Grid 3 (disturbed) 4 (P. jacksoni, H. stella, 1.91 1.75 0.59 0.58
L. flavopunctatus, Lemniscomys sp.)
Grid 4 undisturbed 3 (P. jacksoni, H. stella, 1.93 (2.05) 1.74 (1.83) 0.60 (0.37) 0.58 (0.61)
L. flavopunctatus)
Grid 1 (Disturbed) 4 (P. jacksoni, H. stella, 1.5 1.29 0.29 0.32
L. flavopunctatus,
M. minutoides)
Grid 5 undisturbed 3 (P. jacksoni, H. stella, 2.24 2.14 0.74 0.71
[374]
L. flavopunctatus)
Grid 2 (disturbed) 6 (P. jacksoni, 2.1 1.78 0.43 0.29
H. stella, L. flavopunctatus,
M. minutoides, Otomys sp.,
Mastomys sp.)
Undisturbed forest 5 (P. jacksoni, H. stella, 2.24 2.03 0.5 0.4
(all data pooled) L. flavopunctatus, M. minutoides,
Graphiurus sp.)
Disturbed forest 7 (P. jacksoni, H. stella, 2.03 1.69 0.36 0.24
(all data pooled) L. flavopunctatus, M. minutoides,
Otomys sp.,
Mastomys sp., Lemniscomys sp.)
a
The value is the number of equally common species that would produce the same diversity as H¢ (Krebs 1999).
b
The value is 1/D (Krebs 1999); in brackets, value calculated excluding last session.
1435
Discussion
Our results show that anthropogenic disturbance does not necessarily lead to
an increase in rodent species diversity. We found that damage to the vegetation
of the lower strata and increase in the forest – open area boundaries lead to a
variation in the diversity, evenness and richness of the rodent community.
More specifically it appears that the proximity to forest edge increases the
species richness while the decrease in ground vegetation cover leads to a
decrease in the community diversity.
The population study results show that there is a high similarity between the
CAPTURE estimate and the actual numbers of individuals trapped. This might
suggest that the estimate was accurate and that most of the animals present in
the trapping area were caught. At the same time from Figures 3 and 4 we can see
that the estimates were often precise as they have a very small confidence
interval. The comparison of the capture index of each species between macro-
habitats did not give significant results, in fact most variations occur between
grids and between trapping sessions and not between macrohabitats: this might
depend on actual differences in population sizes, or sampling error as well as
differences in trap numbers (but no significant correlation was found between
rodent abundance and trap numbers). Struhsaker (1997), reports that in some
cases these two species showed higher abundance in logged forest, but in other
cases differences were not significant. Chapman and Chapman (1999) reported
that their capture success doubled in the disturbed forest, while Isabirye-Basuta
and Kasenene (1987) found significant differences for Hylomyscus stella but not
for Praomys jacksoni. Waweru and Odanga (2004) in the Kakamega Forest
found higher abundance in a portion of mature forest when compared to a
fragment of regenerating forest (clear felled 15 years before). In all these cases
anthropogenic disturbance coincided with logging; as previously stated this is
not the case for our study areas, so the fact that we did not observe significant
differences could be due the fact that this kind of anthropogenic disturbance
does not affect this demographic parametre. However, since populations of
African rainforest rodents fluctuate to some extent (Struhsaker 1997) long-term
studies are needed to confirm this hypothesis.
The density estimate varies with the method used. For example in grid 5
(disturbed forest), with the first method, that might overestimate density, we
obtain a value of 46.91 individuals/ha, with one of the corrections the estimate
drops down to 27.9 individuals/ha. Even if we consider the lowest values still we
obtain very high values of density. The boundary strip correction is a very useful
tool to estimate the effective size of Area trapped and is often used, also in
[375]
1436
[376]
1437
higher in the disturbed forest. If we pool data from all the grids of the same
forest type, species richness is higher in the disturbed forest (7 species) than in
the undisturbed forest (5 species). With the Shannon–Wiener Function we
found a higher community diversity in all the grids of the undisturbed forest,
while with the Simpson’s Index we obtained the same result in two of the three
comparisons. The decrease of rodent species diversity together with the de-
crease in lower strata vegetation seems to be coherent with the considerations
of Isabyrie-Basuta and Kasenene (1987) on the positive correlation between
rodent species diversity and ground vegetation cover. The only difference is
that in their case anthropogenic disturbance (logging) leads to an increase of
ground vegetation cover, which is the opposite of our case. On the other hand,
our data reveals an increase in species richness in the disturbed forest which is
characterised by lower ground vegetation density. Isabyrie-Basuta and
Kasenene (1987) instead, found a positive correlation between ground vege-
tation cover and rodent species richness. This particular pattern of inversion
between rodent species richness and diversity in the two areas of the Kakamega
Forest may be due to two different aspects of anthropogenic disturbance: (1)
proximity to the forest edge might favour the occasional entrance of non-forest
species, thus increasing rodent species richness; (2) alteration of lower strata
vegetation might increase the other component of diversity: equitability, that is
the relationships of dominance between species.
First, as was anticipated in the Introduction, the disturbed forest grids
were near to the forest edge, so very near to the cultivated fields and pastures,
that sustain a different rodent fauna (Delany 1975; Delany 1986). Although
identification of single capture species stops at the genus level, Mastomys,
Lemniscomys and Otomys are three genus with only non-forest species and
they were captured only in the disturbed forest, Lophuromys flavopunctatus
and Mus (Nannomys) minutoides are habitat generalists and are often cap-
tured in forests (Kingdon 1974; Delany 1975; Delany 1986), these species
were captured in both forest types. Graphiurus, Praomys jacksoni and Hylo-
myscus stella are the only forest-specialist species (Kingdon 1974; Delany
1975; Delany 1986; Lecompte et al. 2002). Graphiurus sp. was captured only
in grid 6 of the undisturbed forest, which is the only grid of the undisturbed
forest characterised by higher species richness than its corresponding in the
disturbed forest. Thus this data supports the hypothesis that the proximity to
forest edge is the main factor responsible for increase in species richness. A
short-term study such as this is not able to assess whether the 3 non-forest
species captured in the disturbed forest are permanent components of the
rodent forest community, however, these same open area species are known
to invade forest, particularly secondary forest (Delany 1975; Isabyrie-Basuta
and Kasenene 1987; Struhsaker 1997).
As regards point two, in two out of three comparisons evenness is higher in
the undisturbed forest, if we pool data we confirm that evenness is higher in
the undisturbed forest. As structural complexity or heterogeneity of a habitat
increases, the number of microhabitats potentially available increases (Hair
[377]
1438
1980). In the other part of our research (Mortelliti and Boitani, submitted)
we found that in the Kakamega Forest microhabitat heterogeneity decreases
in the disturbed forest as a result of the reduction of low strata vegetation.
The higher variability of microhabitats in the undisturbed forest might hence
lead to a more even distribution of species increasing and/or decreasing the
proportion of individuals of a certain species in the community. The pattern
observed in the corresponding grids 6 and 3 does not fit in this hypothesis, so
one can only postulate that some other factor might be affecting community
diversity and evenness. In synthesis our results support the hypothesis that
proximity to forest edge increases species richness while degradation of
ground vegetation cover reduces the evenness parametre, thus decreasing
community diversity.
A comprehensive formulation of this hypothesis will need further long-term
studies with the utilisation of various types of traps at various heights in order
to reduce bias of the species checklist. Furthermore it is important to highlight
that woody stem density or low-strata vegetation in general might not be the
actual feature responsible for this change, it might be correlated to the actual
factors effectively influencing rodent species diversity.
Data on vertical stratification appears to be consistent with bibliographic data
(Kingdon 1974; Delany 1975; Delany 1986; Struhsaker 1997): Praomys jacksoni
and Hylomyscus stella are two scansorial species, the latter being relatively more
arboreal. Our results show that in the Kakamega forest anthropogenic distur-
bance does not seem to influence vertical stratification since no significant
difference in above-ground captures was found between macrohabitats.
Struhsaker (1997) suggested that anthropogenic disturbance leads to an in-
crease in the degree of segregation between arboreal and terrestrial niches
(more specifically Hylomyscus stella appeared to be strictly arboreal rather
than scansorial). However, in his case anthropogenic disturbance coincided
with logging and thus with an increase of ground vegetation cover, which is the
opposite of our case. Our data refers only to the first 3 m of height, a detailed
analysis of vertical stratification will require traps located right up to the
overstory canopy (Delany 1986).
Our short-term study shows that during the period of study the Kakamega
forest supported an extremely high abundance of Praomys jacksoni, and in
some cases of Hylomyscus stella, this highlights the importance of studies
focusing on possible effects of this high density of rodents on seed ecology. This
study also supports the hyphothesis that ground vegetation might be (directly
or indirectly) the factor responsible for variation in community diversity
(through a variation in evenness), while proximity to the forest edge might be
responsible for variation in species richness. This study further demonstrates
the ability of man to modify the structure of animal communities, even in the
form of a ‘light’ impact such as wood collection and cattle passage by local
communities inhabiting forest surroundings and thus pones evidence on the
importance of these studies to forest management.
[378]
1439
Acknowledgements
References
Barnett A. and Dutton J. 1995. Expedition field techniques: small mammals (excluding bats).
Expedition Advisory Centre, London, England.
Chapman C.A. and Chapman L.A. 1999. Forest restoration in abandoned agricultural land: a case
study from east Africa. Conserv. Biol. 13(6): 1301–1311.
Cords M. 1990. Mixed-species association of East-African Guenons: general pattern or specific
examples? Am. J. Primat 21: 101–114.
Delany M.J. 1975. The Rodents of Uganda. British Museum of Natural History.
Delany M.J. 1986. Ecology of Small Rodents in Africa. Mammal review Vol. 16.
Fleming T.H. 1975. The role of small mammals in tropical ecosystems. In: Golley F.B., Petrusewicz
K. and Ryszkowski L. (eds), Small Mammals: Their Productivity and Population Dynamics.
Cambridge University Press.
Genest-Villard H. 1980. Regime alimentaire des rongeurs Myomorphes de foret equatoriale (region
de M’Baiki Republique Centrafricaine). Mammalia 44: 432–484.
Gurnell J. and Flowerdew J.R. 1994. Live Trapping Small Mammals –A Practical Guide –
Occasional Publication No3. Mammal Society, London.
Hair J.D. 1980. Measurements of ecological diversity. In Schennitz (ed.), Wildlife Management
Techniques Manual.
Isabirye-Basuta G. and Kasenene J.M. 1987. Small rodent population in selectively felled and
mature tracts of Kibale forest, Uganda. Biotropica 19(3): 260–266.
Jefferey S.M. 1977. Rodent ecology and land use in western Ghana. J. Appl. Ecol. 14: 741–755.
Jordan C.F. 1986 Local effects of tropical deforestation. In: Soulè E. (ed.)1986 Conservation
Biology: The Science of Scarcity and Diversity. Sinauer Associates.
KIFCON (Kenya Indigenous Forest Conservation) 1994. Kakamega Forest: The Official Guide.
KIFCON, Nairobi.
Kingdon and J. 1974. East African Mammals (Rodentia, Insectivora, Macroscelida). Academic Press.
Krebs Charles J. 1999. Ecological Methodology 2nd edn. Addison Wesley Longman Inc.
Lecompte E., Denys C. and Granjon L. 2001. An identification key of the Praomys species (Rodentia:
Muridae). In: 1’IRD (ed.), African Small Mammals. colloques et séminaires, paris, pp. 127–139.
Lecompte E., Granjon L. and Denys C. 2002. The phylogeny of the Praomys complex (Rodentia:
Muridae) and its phylogeographic implications. J. Zool. Syst. Evol. Res. 40(2002): 8–25.
Lucas G.L. 1968. Kenya. In: Hedberg I. and Hedberg O. (eds), Conservation of Vegetation in
Africa South of the Sahara. Acta phytogeographica suecica 54, pp.152–166.
Mares M.A. and Enest K.A. 1995. Population community ecology of small mammals in a gallery
forest of central Brazil. J. Mammal. 76(3): 750–768.
Malcolm J.R. 1995. Forest structure and the abundance and diversity of neotropical small mammals.
In: Lowman M.D. and Nadkarni N.M. (eds), Forest Canopies. Academic Press, New York, 624 pp.
Nichols and Dickmanin Wilson D.E., Cole R.T., Nichols J.D., Rudran R. and Foster M.S. 1996.
Measuring and Monitoring Biological Diversity: Standard Methods for Mammals. Smithsonian
Institution Press.
Primack R.B. 2000. A Primer in Conservation Biology. Sinauer Associates.
[379]
1440
Otis D.L., Burnham K.P., White G.C. and Anderson D.R. 1978. Statistical Inference from Capture
Data on Closed Animal Populations. Wildlife Monographs – A publication of the Wildlife
Society No62.
Rogo L., Lwande W., Miller S., Herren H. and Chapya A. 1999. Kakamega Forest: an integrated
conservation project. Bull. East African Nat. History Soc. 29(3): 9–13.
Struhsaker T.T. 1997. Ecology of an African Rainforest: Logging in Kibale and the Conflict
between Conservation and Exploitation. University Presses of Florida, Gainesville.
Waweru C. and Odanga J.C. 2004. Demographic aspects of sympatric Praomysjacksoni and
P. Stella in a tropical lowland forest in Kekamega, Kenya. Afr. J. Ecol. 42: 93–99.
Whitmore T.C. 1998. An Introduction to Tropical Rainforests. Oxford University press.
[380]
Biodiversity and Conservation (2006) 15:1441–1457 Springer 2006
DOI 10.1007/s10531-005-0598-6
-1
J. LUIS HERNANDEZ-STEFANONI1,2,*
1
Watershed Ecosystems Graduate Program, Trent University, Peterborough, Ontario, Canada;
2
Present address: Servicio de Información y Estadı´stica Agroalimentaria y Pesquera, Av. Benjamı´n
Franklin 146, Col. Escandon, C.P. 11800, Me´xico, USA; *Author for correspondence (e-mail:
[email protected]; phone: +1-52-55-5271-7111, ext. 134)
Key words: Landscape fragmentation, Landscape patterns, Plant diversity, Shannon diversity
index, Simpson diversity index, Tropical forest
Abstract. The relationships among landscape characteristics and plant diversity in tropical forests
may be used to predict biodiversity. To identify and characterize them, the number of species, as
well as Shannon and Simpson diversity indices were calculated from 157 sampling quadrats (17,941
individuals sampled) while the vegetation classes were obtained from multi-spectral satellite image
classification in four landscapes located in the southeast of Quintana Roo, Mexico. The mean
number of species of trees, shrubs and vines as well as the mean value of the total number of species
and the other two diversity indices were calculated for four vegetation classes in every one of the
four landscapes. In addition, the relationships between landscape patterns metrics of patch types
and diversity indices were explored. The multiple statistical analyses revealed significant predictor
variables for the three diversity indices. Moreover, the shape, similarity and edge contrast metrics
of patch types might serve as useful indicators for the number of species and the other two diversity
variables at the landscape scale. Although the association between the three diversity indices and
patch types metrics showed similar behavior, some differences were appreciated. The Shannon
diversity index, with its greater sensitivity to rare species, should be considered as having a greater
importance in interpretation analysis than Simpson index.
Introduction
Tropical forests of the world are being destroyed by degradation and con-
version to other forms of land use, induced by increasing human needs or
simply by economic gain. The loss of biodiversity is considered to be one of the
most important of all negative effects on these forests. High diversity implies
that there is a source of new species executing functions or ecosystem services
for human needs (Bengtsson 1998). Therefore, a reduction of biological
diversity means less environmental functions and ecological processes that
generate and maintain soils, convert solar energy into plant tissue, absorb
pollutants, supply clean air and water, store essential nutrients, regulate
weather, and climate and so on (Myers 1995).
The Yucatan peninsula has been recognized as one of the world’s biodi-
versity ‘‘hotspots’’ areas with high levels of biological diversity (Myers et al.
[381]
1442
[382]
1443
Methods
The study was conducted in a tropical forest over four contiguous landscapes
of 4 km · 4 km, located in the southeastern portion of the Yucatan peninsula,
Mexico. Tropical sub-deciduous forests in different stages of succession, which
are characterized by the age, as well as secondary associations that prosper
mainly in flood areas, cover the majority of the four landscapes (Cabrera et al.
1982). The forest consisting of 2 or 3 canopy layers with trees, shrubs and vines
between 3 and 25 m of height. Mayan farmers identified the stages of succes-
sion with indigenous local names. ‘‘kanah kax’’ refers to a forest from 20 to
60 years old; ‘‘kelenche’’ is used for vegetation between 11 and 19 years of age;
‘‘juche’’ used for plant species between 4 and 10 years of age and ‘‘saakab’’
with plants species of 3 or less years of age. The secondary plant associations in
the area are ‘‘savanna’’, which have few sparse tree species between 3 and 10 m of
height and ‘‘akalche’’ (in local Mayan language) consisting of a shrub stratum.
A plant survey based on a stratified random sampling design was performed
in the study area during the summer of 2000, 2001 and 2003. This survey had a
total of 157 sampling quadrats, which were located on the ground using a GPS
unit in the six vegetation types. Of the total number of quadrats 42 fell within
the class ‘‘kanah kak’’, 27 in ‘‘kelenche’’, 22 in ‘‘juche’’, 25 in ‘‘saakab’’, 22 in
‘‘akalche’’ and 19 in the ‘‘savanna’’ vegetation class. The sampling quadrat
[383]
1444
consisted of two nested sites, one of them of 10 · 10 m used to sample trees and
vines that have 3 or more meters of height, the other, a nested sub-site of
5 · 5 m used for sampling all the shrubs taller than 1.0 m. In every one of the
quadrats three diversity indices were computed, those are species richness (i.e.,
the number of species present in an area) and two measures based on species
frequencies or abundance, including exponent Shannon and reciprocal Simp-
son indices (Magurran 1988; Krebs 1989). A total of 17,941 sampled individ-
uals were identified to species and enumerated.
A land cover map for the entire area was obtained from Landsat 7 Thematic
Mapper (TM) imagery acquired on April 2000, after applying a supervised
classification on bands 5 (short-wave infrared: 1.55–1.75 gm), 4 (near infrared:
0.76–0.90 gm) and 3 (red: 0.63–0.69 gm). Each band was geo-referenced and
radiometrically corrected. The ‘‘Maximum Likelihood Algorithm’’ imple-
mented by the image analysis software ER MapperTM 6.1 (Earth Resource
Mapping Ltd. 1998) was used as the classification method. The sampling
quadrats were used for assessing the accuracy of the classified land cover maps,
which resulted in an overall accuracy of 82.3%. The final land cover map of the
four landscapes is shown in Figure 1. Details of the classification and the
accuracy assessment procedures of the resulting land cover maps are found in
Hernandez-Stefanoni (2004) and Hernandez-Stefanoni and Ponce-Hernandez
(2004).
The ER MapperTM raster files of the four landscape mosaics were exported to
the GIS program IDRISI (Eastman 1999), in order to calculate the landscape-
pattern metrics using the program software FRAGSTATS 3.0 (McGarigal
et al. 2002). The six vegetation types identified during the classification and the
remaining of the land cover classes grouped as ‘‘background’’ were considered
for the calculations. The individual patches were classified as clusters of ver-
tical, horizontal or diagonal pixels as in other studies (Gustafson et al. 1994),
while a patch type considers all the individual patches of the same vegetation
class. Most of the metrics applied to vegetation classes can be interpreted as
fragmentation indices, because they measure the configuration of a particular
patch type (McGarigal et al. 2002). The four landscapes in the study area were
considered for the computation of the indices per class (patch type).
The division of the study area in four landscapes was done to obtain different
replicas of landscape configurations. The four landscapes, defined for esti-
mating landscape metrics of patch types, still retain a sufficient size as to allow
for the occurrence of several patches of plant diversity. Gustafson (1998) used
[384]
1445
Figure 1. Land cover maps obtained from supervised classification of the four landscapes.
[385]
1446
[386]
1447
Table 1. Values used to give a similarity weight between the different patch types.
Statistical analysis
[387]
1448
of the problem. This was done while trying to avoid the well-known multico-
linearity problem that often emerges among explanatory variables when the
correlations between them are high. The variables need to be transformed with
1/x, log10(x), log10(x + 1) and sqrt(x) as necessary to meet the assumptions of
normality and linearity (Tabachnick and Fidell 1996). The final components at
each level were found by a rotation procedure using the varimax method,
yielding the final principal components. Then, these were interpreted using
component loading (correlation between the principal component and the
original variable). In order to create a predictive model of plant diversity as a
function of the computed ‘‘compound’’ variables represented by the principal
components, the factor scores and the plant diversity indices were related using
multiple regression analysis, producing models to predict plant diversity indices
based on components created from landscape metrics of patch types. Finally, to
avoid that the correlations between landscape pattern metrics and species
diversity were confounded by habitat type, the analysis was conducted just for
the four stages of succession in the tropical sub-deciduous forest class.
Results
The mean values of the tree plant diversity indices in each patch type for the
four landscapes are presented in Table 2. The mean values of the diversity
indices for the six vegetation types in every one of the four landscapes showed a
similar pattern. Thus, kanah kax class (i.e., the oldest stage of succession in the
forest) has more species and less dominance than kelenche, juche and saakab,
which are early successional stages of the forest.
To evaluate the degree of association between metrics of patch types and plant
diversity indices, correlation coefficients between them were computed (see
Table 3). The total edge contrast index (TECI), percentage of land (PLAND),
edge density (ED) and mean area-weighted shape index (SHAPE_AM) showed
the highest correlation coefficients with most of the plant diversity variables.
These coefficients varied from 0.351 to 0.869 in absolute values. PLAND, ED
and SHAPE_AM are positively correlated while TECI is negatively corre-
lated with the diversity indices. This means that the diversity of a patch type
increases with the augment in its area, irregular shape and perimeter, and when
the contrast with other patch types decreases. So, classes that occupy larger
proportion of the area of the landscape and show less contrast with neigh-
boring patch types favor diversity of plants. Moreover, another variable of
patch type metrics (SIMI_AM) was found moderately correlated with plant
diversity indices varying from 0.351 and 0.395 in absolute values.
[388]
1449
Table 2. Mean plant diversity values for the four landscapes of the study area.
Landscape 1
Kanah Kax 34.45* 26.70* 5.09* 2.63* 23.28* 16.95*
Kelenche 33.00* 27.57* 6.00* 1.92* 20.13* 14.12*
Juche 27.67* 19.00* 6.00* 2.66* 12.52* 7.31*
Saakab 19.67 12.00 4.67* 2.00* 13.22* 9.32*
Landscape 2
Kanah Kax 35.36* 27.57* 5.78* 1.92* 22.74* 16.10*
Kelenche 34.67* 25.66* 7.00* 3.00* 20.06* 13.38*
Juche 29.75* 22.50* 5.37* 1.87* 16.78* 10.86*
Saakab 16.20 10.90 3.90 1.40* 9.49 6.67
Landscape 3
Kanah Kax 34.08* 28.08* 4.33* 1.66* 20.63* 13.65*
Kelenche 30.27* 23.45* 5.00* 1.88* 18.50* 13.18*
Juche 28.71* 21.57* 5.28* 2.00* 15.90* 10.96*
Saakab 15.11 9.77 3.33* 1.88* 8.57 6.24
Landscape 4
Kanah Kax 36.60* 28.00* 4.80* 4.00* 19.66* 12.51*
Kelenche 30.71* 24.85* 4.42* 1.42* 18.04* 12.06*
Juche 27.00* 20.25* 3.75* 3.00* 15.32* 10.45*
Saakab 17.66 10.67 4.33* 1.66* 9.85* 6.95*
*A Turkey HSD test was performed to compare the mean diversity values among vegetation classes
by landscape, no significant differences between these groups.
Table 3. Pearson correlation coefficients between landscape spatial patterns of patch and between
plant diversity indices and landscape patterns of patch types.
PLAND
PD 0.340*
ED 0.513 0.411
SHAPE_AM 0.651 0.178* 0.590
SIMI_AM 0.470 0.305* 0.151* 0.736
TECI 0.740 0.514 0.091* 0.333* 0.283*
Number of species
Total 0.791 0.317* 0.596 0.561 0.353 0.490
Trees 0.829 0.289* 0.616 0.652 0.357 0.520
Shrubs 0.351 0.093* 0.593 0.464 0.395 0.125*
Vines 0.186* 0.354* 0.051* 0.016* 0.074* 0.046*
Exp Shannon 0.865 0.323* 0.591 0.583 0.351 0.603
Rec Simpson 0.869 0.314* 0.575 0.575 0.369 0.619
*Correlations are not significant at p < 0.05.
Values in bold are for pairs highly correlated (r > |0.5|).
[389]
1450
tree-species, the correlation with PLAND is very high (r = 0.829, Table 3),
alternatively in the case of shrubs, there is a moderate correlation with the
percentage of land (r = 0.351, Table 3) and there was not significant corre-
lation between PLAND and number of species of vines. This gives some idea of
the importance of the percentage of land in the distribution of number of tree-
species. In the case of richness of trees and shrubs it can be observed that edge
density and shape have an important weight as explanatory factors, as opposed
to the richness of vines, which is not associated with any of landscape metrics.
On the other hand, no significant correlation was found between the number
of patches by hectare or patch density (PD) and plant diversity variables
(p > 0.05, Table 3). The correlations between the 6 metrics of landscape pat-
terns for patch types are also shown in Table 3. Most of the paired combi-
nations of the metrics of patch types were significantly correlated between them
and 6 of the 15 possible paired combinations are highly correlated (r > |0.5|,
Table 3), which may indicate a degree of redundancy in terms of the infor-
mation that they provide about the structure of the landscape.
Principal component analysis was performed on 6 landscape metrics of patch
types for 16 observations. Two components in the studied area were selected as
meaningful factors with eigenvalues greater than one, which explained 77.24%
of the variation. After applying a varimax rotation, the components were
interpreted as a gradient in percentage of land corresponding to a class
(PLAND), the shape (SHAPE_AM), similarity (SIMI_AM) and the total edge
contrast of the patch type (TECI). This first principal component (PC1) ex-
plained 50.50% of the total variation in the original data. The second principal
component (PC2) explained the additional 26.74% of the variation, and it is
essentially a measure of number of patches found in a hectare (PD) and the
edge density (ED). The percentage of variance explained by the two compo-
nents and the correlation between the principal component and the original
metrics of patch types are shown in Table 4.
To find a model for predicting plant diversity indices from uncorrelated
variables derived from landscape metrics of patch types, a regression analysis
was performed. The regression uses number of species (total, trees, shrubs and
vines), exponent Shannon and reciprocal Simpson indices as dependent vari-
ables and the two components retained from the Principal Component Anal-
ysis (PCA) as independent variables. The results are shown in Table 5. The
variability in total number of species and in the other two indices is explained
by a moderately correlated variable related to percentage of land, similarity,
shape and total edge contrast index (PC1) with Sr2 ranged from 0.453 to 0.479.
Also the number of patches and edge density (PC2) explained the variability of
the diversity indices, with Sr2 ranged from 0.113 to 0.154. The positive cor-
relation with axis 1 of the PCA indicates that a patch types is more diverse
when it is larger in area, more similar to its neighbors, has irregular shape and
it is less contrasted with other patch types.
The results of the regression considering the number of species of the three
groups of plants and the two factors of landscape structure of patch types show
[390]
1451
Table 4. Variance explained by two principal components derived from metrics of patch types and
the weights of the variables in each component after rotation.
PC1 PC2
Explained variance
Observed Eigenvalue 3.03 1.60
% Variance 50.50 26.74
Cum. % Variance 50.50 77.24
Variable
PLAND 0.90 0.11
PD 0.50 0.79
ED 0.42 0.85
SHAPE_AM 0.83 0.36
SIMI_AM 0.75 0.05
TECI 0.73 0.33
*Marked loadings are >0.70.
different responses and reveal that the factor associated with area, shape,
similarity and contrast is more related to the richness of trees, and shrubs
(Sr2 = 0.528 and 0.313, Table 5). On the other hand the factor (PC2) associated
with number of patches and edge density is of variable importance depending on
the group of species considered. Axis 2 of PCA appears to be only correlated
with the richness of trees (Sr2 = 0.115, Table 5).
Discussion
The analysis of the relationships between plant diversity variables and metrics
of habitat types yield the following main results. The percentage of land
(PLAND), a measure of landscape composition, showed a strong correlation
Table 5. Summary of the regression procedures for predicting plant diversity indices from prin-
cipal components derived from metrics of patch types.
Number of species
Total 28.18 PC1* 4.84 0.453 0.500 0.566
PC2** 2.41 0.113
Trees 21.15 PC1* 4.90 0.528 0.588 0.643
PC2** 2.28 0.115
Shrubs 4.94 PC1* 0.54 0.313 0.264 0.313
Vines – – – – – –
Exponent Shannon 16.54 PC1* 3.26 0.485 0.584 0.640
PC2* 1.84 0.154
Reciprocal Simpson 11.29 PC1* 2.29 0.479 0.577 0.632
PC2* 1.34 0.153
*Variables included in the model with p < 0.05.
**Variables included in the model with p < 0.09.
[391]
1452
with species richness (total, trees, and shrubs) and the other two additional
diversity indices. This result together with the weak relationship between plant
diversity and area of a fragment (Hernandez-Stefanoni 2005) may indicate the
importance that fragmentation and diversity of habitats have on plant species
diversity. Several small spread habitat patches usually contain more plant
species than a few large habitat patches (Margules et al. 1994; Honnay et al.
1999), indicating that the availability of different resources may be important
for the establishment of plant species. Canopy openings events such as treefall,
and ‘‘slash and burn’’ agriculture promote spatial variation of patch types that
create several physical environments for plants that offer an increasing in the
availability of resources (Martinez-Ramos et al. 1988; Denslow 1995). This can
improve species diversity for a specific patch type. However, forests with both
very high or low frequency of disturbances, may both lead to low diversity, due
to the fact that either pioneer species (fast-growing) or non-pioneer species
(highly competitive and slow-growing) are respectively selected in each situa-
tion (Martinez-Ramos et al. 1988). Thus, this allows for the regeneration of
similar group of species.
The different response to percentage of land by various groups of species
(trees, shrub and vines) shows the importance of the frequency of disturbances
and the different life history of the species within each group. Given the fact
that newly formed patches cover a lower proportion of the forest in the studied
area (Hernandez-Stefanoni 2004), and that the community of oldest patches
contains more individuals and includes the shade-intolerant species, the pre-
dominant class life history in the tropical forest (Whitmore 1989), it would be
expected that a higher association between tree-species and percentage of land
existed. The reason for this expectation is that the development of a similar
group of species established under the canopy (shade-intolerant species) is
favored. In contrast, in the shrub community the pioneer, shade-intolerant
species are the predominant group (Denslow et al. 1990), making this group
more dependant of gaps.
In the case of vines, they depend on large plants for support and living,
where vines maturation takes place. The availability of light however, has been
proposed as other factor that promotes liana-species distribution, particularly
near the edges where lianas can grow faster (Ibarra-Martinez and Martinez-
Ramos 2002). Not all liana species however, are light demanding (Putz and
Chai 1987). Therefore, vines-species presence could be favored by two main
factors. First, vines can be established in oldest patches, where trees can pro-
vide them support. Second, several liana species are light demanding and grow
well in natural or man-made disturbances (Putz 1984). Consequently, it is
difficult to establish an association between richness of vines and the different
metrics if there is not a division of these two main groups of vines.
The degree of contrast between a patch type and its neighbors classes,
measured as total edge contrast index, was a metric highly related to species
richness (total, tree and shrub) and the other two plant diversity indices. These
results may be explained by the fact that resource availability of a class is given
[392]
1453
[393]
1454
[394]
1455
On the other hand the use of Shannon and Simpson diversity indices to
assess plant diversity and its relationships to landscape patterns should be
considered with caution. However, the question remains as to which diversity
index should de used, and under what situations. Occurrences of rare species
are among the most frequently used criteria for selecting and prioritizing
habitat sites for preservation (Prendergast et al. 1993; Rossi and Kuitunen
1996). Moreover, the relative species abundance distribution of a tropical
forest often follows a J-inverted shape (Magnussen and Boyle 1995), in par-
ticular in the studied area (Hernandez-Stefanoni 2004), which implies the
probability of finding many species with low abundances. Thus, the Shannon
diversity index, with its greater sensitivity to rare species, should be considered
as having a greater importance in interpretation of the analysis. In fact, several
published studies of plant diversity in tropical forests have elected to use this
index (Fanliang et al. 1996; Nangendo et al. 2002). However, in particular
cases where a single dominant species is of interest for management or con-
servation proposes, the Simpson diversity index should be preferred.
References
Alvarez-Buylla E. and Garcia-Barros R. 1991. Seed and forest dynamics: a theoretical framework
and an example from the neotropics. Am. Nat. 137(2): 133–154.
Alvarez-Buylla E. and Martinez-Ramos M. 1992. Demography and allometry of Cecropia
obtusifolia, a neotropical pioneer tree and evaluation of the climax-pioneer paradigm for tropical
rain forest. J. Ecol. 80: 275–290.
Bengtsson J. 1998. Which species? What kind of diversity? Which ecosystem function? Some
problems in studies of relations between biodiversity and ecosystem function. Appl. Soil. Ecol.
10: 191–199.
Brokaw N.V.L. 1985. Treefalls, regrowth and community structure in tropical forest. In: Pickett
S.T.A. and White P.S. (eds), The Ecology of Natural Disturbances and Patch Dynamics. Aca-
demic Press, Orlando, Florida, USA, pp. 53–69.
Brokaw N.V.L. 1987. Gap-phase regeneration of three tree pioneer species in a tropical forest.
J. Ecol. 75: 9–19.
Cabrera C.E., Souza S.M. and Tellez V.O. 1982. Imagenes de la flora Quintanaroense. Centro de
Investigaciones de Quintana Roo, Mexico.
Carroll S.S. 1998. Modelling abiotic indicators when obtaining spatial predictions of species
richness. Environ. Ecol. Stat. 5: 257–276.
Debinski D. and Holt R.D. 2000. A survey and overview of habitat fragmentation experiments.
Conserv. Biol. 14: 63–76.
Denslow J.S., Schultz J.C., Vitousek P.M. and Strain B.R. 1990. Growth responses of tropical
shrubs to treefall gap environments. Ecology 71(1): 165–179.
Denslow J.S. 1995. Disturbance and diversity in tropical rain forest: the density effect. Ecol. Appl.
5(4): 962–968.
Eastman J.R. 1999. IDRISI 32: User’s Guide. Clark Labs for Cartography, Technology and
Geographic Analysis. Clark University, Worcester, Massachusetts, USA.
Earth Resource Mapping Ltd. 1998. ER Mapper 6.1. User Guide. San Diego CA.
Fanliang H., Legendre P. and LaFrankie J.V. 1996. Spatial patterns of diversity in a tropical rain
forest in Malaysia. J. Biogeogr. 23: 57–74.
Grashof-Bokdam C.J. 1997. Forest plants in an agricultural landscape in the Netherlands: effect of
habitat fragmentation. J. Veg. Sci. 8: 21–28.
[395]
1456
Gustafson E.J. 1998. Quantifying landscape spatial patterns: what is the state of the art? Ecosys-
tems 1: 143–156.
Gustafson E.J., Parker G.R. and Backs S.E. 1994. Evaluating spatial patterns of wildlife habitat: a
case study of the wild turkey (Meleagris gallopavo). Am. Midl. Nat. 131: 24–33.
Hartshorn G.S. 1989. Application of gap theory to tropical forest management: natural re-
generation on strip clear-cuts in the Peruvian Amazon. Ecology. 70(3): 567–569.
Hargis C.D., Bossonette J.A. and David J.L. 1998. The behavior of landscape metrics commonly
used in the study of habitat fragmentation. Landscape Ecol. 13: 167–186.
Hernandez-Stefanoni J.L. 2004. A Framework for Evaluating the Diversity Status of Trees, Shrubs
and Vines in a Tropical Forest in Mexico based on Landscape Patterns Metrics and Spatial
Modelling with GIS. Ph. D. dissertation, Trent University, Peterborough, Ontario, Canada.
Hernandez-Stefanoni J.L. 2005. Relationships between landscape patterns and species richness of
trees, shrubs and vines in a tropical forest. Plant Ecol. 179: 53–65.
Hernandez-Stefanoni J.L. and Ponce-Hernandez R. 2004. Mapping the spatial distribution of plant
diversity indices in a tropical forest using multi-spectral satellite image classification and field
measurements. Biodiv. Conserv. 13: 2599–2621.
Hernandez-Xolocotzi E., Baltaza E. and Tache S. 1995. La milpa en Yucatan, Un sistema agricola
tradicional. Colegiode Postgraduados, Mexico.
Honnay O., Hermy M. and Coppin P. 1999. Effects of area, age and diversity of forest patches in
Belgium on plant species richness, and implications for conservation and reforestation. Biol.
Conserv. 87: 73–84.
Honnay O., Piessens K., Van Landuyt W., Hermy M. and Guilinck H. 2003. Satellite based land
use and landscape complexity indices as predictors for regional plant species diversity. Landscape
Urban Planning 63(4): 241–250.
Howe H.F. 1990. Habitat implications of gap geometry in tropical forest. Oikos. 59: 141–143.
Hubbell S.P. and Foster R.A. 1986. Biology, change and history and the structure of tropical rain
forest communities. In: Diamond J. and Case T.J. (eds), Community Ecology. Harper and Row,
New York, pp. 314–329.
Ibarra-Martinez G. and Martinez-Ramos M. 2002. Landscape variation of liana communities in a
neotropical rain forest. Plant ecology. 160: 91–112.
Kollman J. and Schneider B. 1999. Landscape structure and diversity of flashy-fruited species at
forest edges. Plant Ecol. 144: 37–48.
Krebs C.J. 1989. Ecological Methodology. Harper Collins, New York.
Krummel J.R., Gardner R.H., Sugihara G., O’Neill R.V. and Coleman P.R. 1987. Landscape
patterns in a disturbed environment. Oikos. 48: 421–424.
Magnussen S. and Boyle T.J.B. 1995. Estimating sample size for inference about Shannon-Weaver
and the Simpson indices of species diversity. For. Ecol. Manage. 78: 71–84.
Magurran A.E. 1988. Ecological Diversity and its Measurement. Princeton University Press,
Princeton, NJ.
Margules C.R., Nicholls A.O. and Usher M.B. 1994. Apparent species turnover, probability of
extinction and the selection of natural reserves: a case of study of the Ingleborough limestone
pavements. Conserv. Biol. 8: 398–409.
Martinez-Ramos M., Alvarez-Buylla E., Sarukhan J. and Piñero D. 1988. Treefall age determi-
nation and gap dynamics in a tropical forest. J. Ecol. 76: 700–716.
Mazerolle M.J. and Villard M.A. 1999. Patch characteristics and landscape context as predictor of
species presence and abundance: a review. Ecoscience 6(1): 177–124.
McGarigal K. and Cushman S.A. 2002. Comparative evaluation of experimental approaches to the
study of habitat fragmentation studies. Ecol. Appl. 12(2): 335–345.
McGarigal K., Cushman S.A., Neel M.C. and Ene E. 2002. FRAGSTATS: Spatial Pattern
Analysis for Categorical Maps. University of Massachusetts.
Myers N. 1995. Environmental services of biodiversity. Proc. Natl. Acad. Sci. USA. 3: 2764–2763.
Myers N., Mittermeier R.A, Mittermeier C.G., da Fonseca G.A.B. and Kent J. 2000. Biodiversity
hotspots for conservation priorities. Nature 43(24): 853–858.
[396]
1457
Nangendo G., Stein A., Gelens M., Gier A. and Albricht R. 2002. Quantifying differences in
biodiversity between a tropical forest area and a grassland area subject to traditional burning.
For. Ecol. Manage. 164: 109–120.
Peet R.K. 1974. The measurement of species diversity. Annu. Rev. Ecol. Syst. 5: 285–307.
Prendergast J.R., Quinn R.M., Lawton J.H., Eversham B.C. and Gibbons D.W. 1993. Rare spe-
cies, the coincidennce of diversity hotspots and conservation strategies. Nature 365: 335–337.
Putz F.E. 1984. The natural history of lianas on Barro Coloraco Island, Panama. Ecology 65(6):
1713–1724.
Putz F.E. and Chai P. 1987. Ecological studies in lianas in Lambir National Park, Sarawak. J. Ecol.
75: 523–531.
Riitters K.H., O’Neill, R.V., Hunsaker C.T., Wickham J.D., Yankee D.H., Timmins S.P., Jones
K.B. and Jackson B.L. 1995. A factor analysis of landscape pattern and structure metrics.
Landscape Ecol. 10: 23–39.
Roberts M.R. and Gilliam F.S. 1995. Patterns and mechanisms of plant diversity in forested
ecosystems: implications for forest management. Ecol. Appl. 5(4): 969–977.
Rossi E. and Kuitunen M. 1996. Ranking of habitats for the assessment of ecological impact in
land use planning. Biol. Conserv. 77: 227–234.
Schupp E.W., Howe H.F., Augspurger C.K. and Levey D.J. 1989. Arrival and survival in tropical
treefall gaps. Ecology 70(3): 562–564.
Tabachnick B.G. and Fidell L.S. 1996. Using Multivariate Statistics. Harper Collins College
Publishers, New York.
Uhl C., Clark K., Dezzeo N. and Maquirino P. 1988. Vegetation dynamics in Amazonian treefall
gaps. Ecology 69: 751–763.
Whigham D.F., Olmsted I., Cabrera Cano E. and Harmon M.E. 1991. The impact of hurricane
Gilbert on trees, literfall and woody debris in a dry tropical forest in the northeastern Yucatan
peninsula. Biotropica 23: 434–441.
Whitmore T.C. 1989. Canopy gaps and the mayor groups of forest trees. Ecology 70(3): 536–538.
[397]
Biodiversity and Conservation (2006) 15:1459–1466 Springer 2006
DOI 10.1007/s10531-005-0599-5
-1
JANE HERBERT
School of Biology, Sir Harold Mitchell Building, University of St Andrews, KY16 9TH, UK; Present
address: School of Integrative Biology, University of Queensland, Brisbane QLD 4072, Australia;
(e-mail: [email protected]; phone: +01334-463372; fax: +01334-463366)
Key words: Bush fire, Endangered, In situ conservation, Mining, Primary forest, Serpentine,
Ultramafic
Abstract. The monotypic genus Canacomyrica Guillaumin is a small tree endemic to the rare
remaining fragments of primary forest growing on ultramafic geology in New Caledonia. In the
rich flora of this island it is one of many endemics to be threatened by habitat loss due to a variety
of factors, most significantly open-cast mining for nickel. Using field observations and data from
herbarium specimens the extent of occurrence of Canacomyrica monticola is established to be
approximately 1420 km2. Within this area the distribution of C. monticola is very fragmented and
limited to just 11 known localities. Six localities are outside protected areas; two of these may be
imminently threatened by mining activity and another may be threatened by bush fires. It is
recommended that the IUCN Red List status of Endangered (EN B1ab (i,ii,iv,v)) is assigned to this
species.
Introduction
[399]
1460
A list of all known localities for Canacomyrica was compiled from herbarium
specimens held in three collections: Royal Botanic Garden, Edinburgh (E) (16
[400]
1461
Results
A survey of the herbarium specimens held in Edinburgh (E), Paris (P) and
Noumea (NOU) revealed 10 localities for Canacomyrica, and another was
reported by T. Jaffré (personal communication). The exact positions of col-
lection sites given on herbarium specimens were determined by consulting the
H.S. MacKee gazetteer (Muséum National D’Histoire Naturelle, Paris; http://
phanero.novcal.free.fr/site.html). The eleven localities, all of which were in the
south of the island on ultramafic substrate, are shown in Figure 1 and details
are given in Table 1. Manual measurement of the area between the 11 known
localities gave an extent of occurrence of 1420 km2.
Canacomyrica was examined in the field at Mont Bouo and Mont Mamié
(Table 1, Figure 1). Plants at both sites were highly localised, occurring in
almost monospecific stands with few or no outlying individuals. The local
distribution of the populations appeared to be limited by water availability.
At Mont Bouo plants of Canacomyrica were found only between altitudes of
1050 m and approximately 1150 m, growing in a rainforest community on
ultramafic substrate. Mature individuals were 3–4 m in height. The habitat at
Mont Bouo appeared to be undisturbed and access was difficult.
At Mont Mamié plants of Canacomyrica were found at 500 m, the lowest
recorded altitude for the plant (most collections have been made above
800 m). Canacomyrica was growing in a low scrub community co-dominated
by Cyperaceae species on ultramafic substrate, inundated with water from
abundant natural springs. Mature individuals were up to 1 m in height; many
[401]
1462
Figure 1. Map of New Caledonia showing the extent of ultramafic geology and the position of the
11 known localities for Canacomyrica (extent of ultramafic follows Jaffré et al. 1987).
seedlings were observed in this population. The habitat at Mont Mamié was
disturbed and access was relatively easy due to the presence of mining pros-
pecting tracks.
At both sites there were more than 30 mature individuals but, it was not
possible to estimate the total number of mature plants at either site due to the
time constraints of the expedition. The total area occupied by the population at
Mont Bouo was estimated to be approximately 100 m2. Time constraints also
prevented estimation of the total area occupied by the population at Mont
Mamié.
It was expected that further populations would be found at Mont Mou and
Rivière Bleue but plants of Canacomyrica were not found at these localities
during the expedition. All suitable habitat for Canacomyrica was searched at
Mont Mou. At Rivière Bleue it was not possible to search all suitable habitat
due to the difficulty of the terrain. On the basis of these observations, it is
considered that a single collection from Mont Mou (Baumann-Bodenheim
15679, P) is a doubtful locality for Canacomyrica. It is thought likely that the
specimen was collected elsewhere and incorrectly labelled, alternatively (but
less likely) this collection may represent an extinct population. It was not
possible to gain access to populations within the strictly protected Nature
Reserve of Montagne des Sources. Other locations were not visited due to the
time constraints of the field study.
[402]
Table 1. Known localities for Canacomyrica, with coordinates and protected status details.
Mont Bouo (Koghis range)b 2210¢ S 16630¢ E Protected – amenity protected area, adjacent to the Strict Nature Reserve of Montagne des Sources
Mont Mamiéb 2206¢ S 16653¢ E Unprotected – mining activity observed
N’Goic 2149¢ S 16630¢ E Unprotectede – adjacent to existing Special Botanical Reserve of Mt Humboldt (5 km)
Montagnes des Sourcesc 2207¢ S 16633¢ E Protected – Strict Nature Reserve of Montagne des Sources
Pourinac 2201¢ S 16644¢ E Unprotected – adjacent to Special Botanical Reserve of Haute Pourina (3 km) and Natural park
of Riviére Bleue (5 km)
Ouinnéc 2157¢ S 16642¢ E Unprotectede
[403]
Kouakouéc 2157¢ S 16632¢ E Protectedf – Special Fauna and Flora Reserve of Mt Kouakoué
Humboldtc 2153¢ S 16625¢ E Protectedf – Special Botanical Reserve of Mt Humboldt
Nembrouc 2145¢ S 16613¢ E Unprotected – adjacent to Special Botanical Reserve of Forêt de Saille (5 km)
Nékandoc 2150¢ S 16620¢ E Unprotected – adjacent to Special Botanical Reserve of Mt Humboldt
Forêt de Sailled 2140¢ S 16613¢ E Protected – Special Fauna and Flora Reserve of Forêt de Saille
a
Sources of information on protected status: Pintaud et al. (1999), J. Manaute (personal communication).
b
Locality visited by the author.
c
Locality determined from herbarium specimens.
d
Locality according to T. Jaffré (personal communication).
e
Site adjacent to proposed ‘Ni-Kouakoué-Ouinné’ reserve.
f
Site expected to be included in the proposed ‘Ni-Kouakoué-Ouinné’ reserve. Map coordinates are intended as a guide only.
1463
1464
Discussion
[404]
1465
of genetic diversity, both within and among populations, should be carried out
to act as a guide for the prioritisation of populations in future conservation
management.
Protection of the above mentioned populations would raise the number of
protected localities for Canacomyrica from five to nine, or 82% of all known
sites. These measures represent the first steps towards ensuring the continued
survival of Canacomyrica. Furthermore, conservation measures targeted at
Canacomyrica will help to raise the profile and survival prospects of some of
the last remaining fragments of New Caledonia’s primary forest.
Acknowledgements
References
Bouchet P., Jaffré T. and Veillon J.M. 1995. Plant extinction in New-Caledonia – Protection of
sclerophyll forests urgently needed. Biodivers. Conserv. 4: 415–428.
Brooks R.R. 1987. Serpentine and its Vegetation. A Multidisciplinary Approach. Croom Helm,
London, UK, pp. 330–353.
Farjon A. and Page C.N. (compilers) 1999. Status Survey and Conservation Action Plan. IUCN/
SSC Conifer Specialist Group, IUCN, Gland, Switzerland, pp. 41–50.
Guillaumin A. 1940. Matériaux pour la flore de la Nouvelle-Calédonie. LVII. La présence d’une
Myricacée. Bulletin de la Société Botanique de France 87: 299–300.
Herbert J. 2005 Systematics and biogeography of Myricaceae. Unpublished Ph.D. Thesis, Uni-
versity of St Andrews, St Andrews.
Herbert J., Hollingsworth P.M., Gardner M.F., Mill R.R., Thomas P.J. and Jaffré T. 2002.
Conservation genetics and phylogenetics of New Caledonian Retrophyllum (Podocarpaceae)
species. New Zeal. J. Bot. 40: 175–188.
IUCN 1997. Red List of Threatened Plants. IUCN Species Survival Commission, http://iucn.org/
themes/ssc/97plrl/table5.htm [accessed 26 April 2004].
IUCN 2001. IUCN Red List Categories and Criteria: Version 3.1. IUCN Species Survival Com-
mission, Gland, Switzerland.
[405]
1466
Jaffré T., Bouchet P. and Veillon J.M. 1998. Threatened plants of New Caledonia: Is the system of
protected areas adequate? Biodivers. Conserv. 7: 109–135.
Jaffré T., McCoy S., Rigault F. and Navarro E. 2001a. A comparative study of flora and symbiotic
microflora diversity in two Gymnostoma formations on ultramafic rocks in New Caledonia.
S. Afr. J. Sci. 97: 599–603.
Jaffré T., Morat Ph., Veillon J.M. and MacKee H.S. 1987. Changements dans la végétation de la
Nouvelle-Calédonie au cours du Tertiaire: la végétation et la flore des roches ultrabasiques.
Adansonia 4: 356–391.
Jaffré T., Morat Ph., Veillon J.M., Rigault F. and Dagostini G. 2001b. Composition and Char-
acteristics of the Native Flora of New Caledonia. IRD, Nouméa, pp. 28–29.
Mittermeier R.A., Werner T.B. and Lees A. 1996. New Caledonia – A conservation imperative for
an ancient land. Oryx 30: 104–112.
Myers N., Mittermeier R.A., Mittermeier C.G., da Fonseca G.A.B. and Kent J. 2000. Biodiversity
hotspots for conservation priorities. Nature 403: 853–858.
Olson D.M., Dinerstein E., Abell R., Allnutt T., Carpenter C., McClenachan L., D’Amico J.,
Hurley P., Kassem K., Strand H., Taye M. and Thieme M. 2000. The Global 200: A Repre-
sentation Approach to Conserving the Earth’s Distinctive Ecoregions. World Wildlife Fund-US,
Conservation Science Program, Washington, DC. http://www.panda.org/about_wwf/where_
we_work/ecoregions/global200/downloads [accessed 26 April 2004].
Pintaud J.C., Jaffré T. and Veillon J.M. 1999. Conservation status of New Caledonia palms. Pac.
Conserv. Biol. 5: 9–15.
Primack R.B. 2000. A Primer of Conservation Biology. Sinauer Associates Inc., Sunderland,
Massachusetts, pp. 183.
Proctor J. 2003. Vegetation and soil and plant chemistry on ultramafic rocks in the tropical Far
East. Perspect. Plant Ecol. Evol. Syst. 6: 105–124.
Takhtajan A. 1986. Floristic Regions of the World. University of California Press, Berkeley,
pp. 248–253.
Whitlock B.A., Karol K.G. and Alverson W.S. 2003. Chloroplast DNA sequences confirm the
placement of the enigmatic Oceanopapaver within Corchorus (Grewioideae : Malvaceae S. L.,
formerly Tiliaceae). Int. J. Plant Sci. 164: 35–41.
Willis F., Moat J. and Paton A. 2003. Defining a role for herbarium data in Red List assessments: a
case study of Plectranthus from eastern and southern tropical Africa. Biodivers. Conserv.
12: 1537–1552.
WWF and IUCN 1995. Centres of Plant Diversity. A Guide and Strategy for their Conservation
Vol. 2. IUCN Publications Unit, Cambridge, UK, pp. 529–541.
Zanis M.J., Soltis D.E., Soltis P.S., Mathews S. and Donoghue M.J. 2002. The root of the
angiosperms revisited. Proc. Natl. Acad. Sci. USA 99: 6848–6853.
[406]
Biodiversity and Conservation (2006) 15:1467–1495 Springer 2006
DOI 10.1007/s10531-005-1876-z
-1
[407]
1468
Introduction
[408]
1469
with some areas burnt more often than others, forest succession into the
woodland became possible. Hence, an added component of our study is the
extension of a historically important succession gradient, exploring its range
into the woodland areas.
In our study, special emphasis was placed on how the existing vegetation
types can be characterized in terms of the woody plants and the implication of
the observed species patterns to conservation of woody plants in such land-
scapes. The hypothesis made is that all the vegetation types that exist within the
study area, and the species they support, are an integral part of a composi-
tional/successional gradient that stretches across the FWS mosaic. We asked
the following questions: Is it possible to quantify the gradient? What species are
specific for certain areas? How does the species composition vary along the
succession gradient? Can the gradient be explained in relation to environmental
variables?
A further question we address is whether a satellite image classification of the
area can be used to adequately map the vegetation and its composition in the
area. For this we made use of discrete vegetation cover classes, obtained from a
classification carried out using a combination of spectral information and
environmental variables’ information (Nangendo et al., submitted). The veg-
etation classes are considered a proxy of the vegetation types found in the area.
Standard vegetation indices (NDVI and Tasseled Cap vegetation index) based
on the same image were also compared in their ability to explain the observed
gradient. Finally, we discuss the conservation and management implications of
our results.
Study area
The work was carried out in the northern part of Budongo Forest Reserve in
north-western Uganda. The area is located between 135¢ and 155¢ N and
3118¢ and 3142¢ E. It receives between 1397 and 1500 mm of rain annually on
100 to 150 days. There are two main forest blocks: the main Budongo Forest
block and the Kaniyo-Pabidi Forest block (Figure 1). A woodland area,
interspersed with forest patches, commonly referred to as Kaniyo-Pabidi
woodland, separates these two blocks.
The underlying geology of the Budongo Forest is Precambrian origin con-
sisting of high-grade metamorphic rocks of the 2.9 billion-year-old granulite
group (van Straaten 1976). The soils over 90% of the study area are orthic
Ferralsols: highly weathered, deep, well drained soils with low pH. The
remaining 10% of the area has typically shallow soils, called Lithosols. These
soils are mainly found on hilltop regions and are predominantly underlain by
rocks. In river valleys, eutric Fluvisols are present.
[409]
1470
In the woodlands, fire has been prevalent for hundreds of years (Paterson
1991). The woodland burning was initially carried out by the local people for
purposes of hunting and refreshing grass for both domestic and wild ungulates
(Buechner and Dawkins 1961). With the transfer of the control of the wood-
lands from the local people (Bunyoro Kingdom) to the central government
(Forest Department) in 1968, measures to control burning were put in place
(Forest Department Uganda 1997). These were not very effective, however,
until the establishment of the joint management between Forest Department
and Uganda Wildlife Authority in the mid 1980s. Fewer, and smaller, areas are
now burnt and the burning is also less frequent. The woodland is therefore
heterogeneous and made up of vegetation patches at varying stages of recovery
since they were last burnt.
Data collection
Data was collected from 591 plots, 266 of which had an area of 400 m2 and
326 with an area of 500 m2. All data were collected during the same period
(August–October 2002). Along a transect, perpendicular lines were laid every
300 m. Along each perpendicular line, data were collected at every 75 m. For
sites 1–5, a plot size of 400 m2 was used (Figure 2), while for sites a–e, it was
500 m2. Based on a 2002 satellite image of the study area, sites 1–5 were
located in areas that showed a similar spectral reflectance, whereas sites a–e
were located in areas that showed varying spectral reflectance. The variation
[410]
1471
Figure 2. The location of the data collection points. 1, 2, 3, 4 and 5 are locations where the plots
were 400 m2 and a, b, c, d, e and f are locations where the plot size was 500 m2.
[411]
1472
of the site locations was to ensure that we capture as much as possible of the
species variation within the area. In each plot, the following data were col-
lected:
• Plot coordinates
• Species names, diameter at breast height (DBH) for all woody plants ‡10 cm
DBH, measured at 130 cm. If the tree was buttressed and abnormal at
130 cm, the diameter was measured just above the buttress where the stem
assumes a near cylindrical shape.
• Canopy cover percentage, using a canopy densiometer (Robert E: Lemmon,
Forest Densiometers, Oklahoma, USA), following the provided guidelines.
Four measurements were taken in each plot and an average of these mea-
surements was calculated to determine the final canopy cover of the plot.
• A fire indicator value. The fire indicator value was based on several factors
(1) the degree of scorching on the woody stems i.e. if it was fresh or old, (2) if
there existed remains of burnt grass in the undergrowth and (3) whether fresh
ash was found in the area. The last two factors were used to confirm areas
with recent fire. Plots with fresh fire scorching on the woody stems, remains
of burnt grass or ash were recorded as ‘recent burns’ and labelled class 2.
Plots with old signs of fire were labelled class 1 (old fires) and plots with no
sign of fire were labelled class 0 (no fire).
Species identification was based on Eggeling and Dale (1952) and Hamil-
ton (1991). Samples of the species that could not be clearly identified in the
field by the botanists on the team (Israel Tinka and Hezekias Ddumba) were
sent to the Uganda National Herbarium, Makerere University, where they
were identified.
Data preparation
[412]
1473
Remote sensing
Accessibility
We used distance from the southern forest boundary to each plot as a surrogate
for accessibility, by the local people, to the sampled areas. The conservation
area gate marks the southern boundary between the conservation area and the
local people’s settlements. From here on, distance will be referred to as ‘dis-
tance from gate.’ During fieldwork, it was observed that because of the gate
control, the local people entered the protected area at other points along the
boundary of the protected area, instead of using the road. Having recorded the
coordinate of the gate location, an east–west line was established at this point
[413]
1474
and distance for each plot was calculated based on this line. This provided the
plot distance relative to the conservation area gate.
Two approaches were used in analysing remote sensing outputs. First, plot
values obtained from vegetation indices (such as TC), which are continuous
classifiers, were compared to DCA plot scores. To identify the vegetation index
that best explained the gradient, a non-linear regression method was used since
the scatter plot of the DCA vs. the index values showed a non-linear rela-
tionship. Second, discrete classes obtained from an earlier classification
(Nangendo et al., submitted) were analyzed for differences in terms of species
composition and diversity and, in basal area. Although the same satellite image
was used for the classification and for the creation of the index maps, the plots
used for the classification are not the same as those used in the analysis.
Differences in composition
[414]
1475
particular group should always be present and should also be exclusive to that
group (not occurring in other groups). From the analysis, an indicator value is
obtained for each species in each group (Dufrêne and Legendre 1997; McCune
and Mefford 1999; McCune et al. 2002). The indicator values are tested for
statistical significance using a Monte Carlo randomization. Species diversity
was expressed as species dominance, which was calculated using the Simpson
Index (SI) (Magurran 1988), and Fisher’s a (Fa), (Fisher et al. 1943). These
indices have low sensitivity to plot size differences (Magurran 1988). Differ-
ences between plots in different fire and cover classes with respect to SI, Fa and
BA were tested with ANOVA using SPSS (SPSS 10, SPSS Inc. USA).
To check for variation in species abundance and diversity in relation to
disturbance, graphs of number of species per 100 m2 and Fisher’s a per plot
were made. Having the assumption, which was also backed by field observa-
tion, that disturbance was lowest in the forest class and highest in the wooded
grassland class, plots were arranged according to vegetation cover classes. The
order of plot arrangement was; forest (1–147), closed woodland (148–310),
open woodland (311–459), very open woodland (460–555) and wooded
grassland (556–592). Within each vegetation cover class the plots are randomly
ordered.
Results
Species distribution
A total of 26,076 individuals from 121 species, 89 genera, and 38 families were
recorded on the 591 plots. The most species-rich family was Moraceae with
11% of all species found (13), followed by Euphorbiaceae and Mimosaceae
with 8% each (10). The most species-rich genus was Ficus with 5% of all
species (6), followed by Acacia, Albizia, Celtis and Combretum with 3% each
(4). Nine species or 7% of all species could not be identified to genus level. A
full species list with abundances is given in Appendix 3. The most abundant
genus, in terms of total individuals encountered, was Combretum, with close to
16% of all individuals, followed by Terminalia (14%), Grewia (13), Stereo-
spermum (6%), and Uvariopsis (6%).
The DCA analysis on combined and trimmed data (491 plots and 45 species)
ordered the plots mainly along 1 axis (Figure 3a). This axis had a relatively
high eigenvalue (0.465) suggesting significant woody species variation along
this axis. The eigenvalue for the second axis was 0.172. With 491 plots in-
cluded, axis 1 explained 11.8% of the variation. There was, however, one
outlier plot strongly influencing the second axis. This outlier plot was domi-
nated by Sapium elipticum, a species that rarely occurred in the study area.
After removing this plot, axis 1 explained 12.5% of the variation and axis 2 an
additional 4.6%. Plots with a low axis score (close to 0) are found in the forest
area, plots with a high score (>7) are found in the most open areas. As most of
[415]
1476
[416]
1477
the discussion here on will pertain to axis 1, the main gradient, we will
abbreviate ‘DCA axis 1 plot scores’ to ‘DCA scores.’
The species plot (Figure 3b) also shows most of the variation along the
first axis. The effect of the second axis is only evident close to zero along
axis 1, the forest side, where there appear to be two groups (the same can be
said for the plot scores). Based on this interpretation the species can be
divided into three groups; A, B and C (Figure 3b). Groups A and B occur
within the forest area and group C, probably starting at the forest edge,
stretches through to the woodland area. Species found in group A include
Cynometra alexandri, Diospyros abyssinica and Khaya anthotheca. Group B
species include Uvariopsis congensis, Celtis wightii, Holoptelea grandis
and Funtumia elastica. And species found in group C include Albizia
grandibracteata, Terminalia velutina, Grewia mollis, Combretum molle and
Lonchocarpus laxiflorus.
Fire indicator best explained the gradient in species composition followed by
slope and then distance from gate. Using the stepwise regression analysis, fire
alone had r2 of 0.324 with a standard error of 1.395. Including slope in the
model the r2 was raised to 0.354 and the standard error reduced to 0.365. When
distance from gate was included, the r2 increased to 0.359 and the standard
error was reduced to 1.361. Vegetation cover type was not significant and so it
does not appear in the results table. Relating the site variables individually to
DCA (results not shown) showed that while all the other variables had a
positive correlation with the DCA, distance from gate had a negative corre-
lation.
All the vegetation-indices explained well the DCA variation. TC-wetness and
TC-greenness showed the best relationship with DCA scores with r2 of 0.73
and 0.70, respectively. TC-brightness had the lowest value (r2 = 0.46). NDVI
had an r2 of 0.64.
The classes derived from the analyses of the satellite image differed consider-
ably in their DCA scores (Figure 4a and b). Plots of the ‘No-fire’ class had
consistently low DCA scores, whereas the plots from the class ‘Recent-fire’
have high DCA scores. Plots from the class ‘Old-fire’ were intermediate. The
Fire classes also differed considerably in their TC-greenness values. Conse-
quently a combination of DCA scores and TC-greenness value segregated the
fire classes well.
A similar result was found for the cover classes. These classes are segregated
both by their DCA scores and TC-greenness values (Figure 4b).
[417]
1478
Figure 4. DCA axis 1-Tasseled Cap relationship as subdivided by (a) fire regimes and (b)
vegetation cover classes.
[418]
1479
forest area. In addition, Pterygota mildbreadii was also exclusively found in the
forest (Appendix 2). Funtumia elastica, Uvariopsis congensis and Celtis wightii
had the highest relative frequency in the forest class; 55, 54 and 50%,
respectively. Species with the highest relative frequency in the closed woodland
are Terminalia velutina and Grewia mollis with 97 and 70%, respectively. In the
open woodland plots, Terminalia velutina and Grewia mollis still had the
highest relative frequency of 85 and 88%, respectively. In the very open
woodland, Grewia mollis occurred in 99% of the plots while in the wooded
grassland, Stereospermum kunthianum had the highest relative frequency of
67%. Overall, Grewia mollis in the very open woodland had the highest relative
frequency i.e. it occurred in 99% of the closed woodland plots.
Whereas in the forest some of the species that had the highest relative fre-
quency are part of those that had the highest relative abundance, it is different
for the other cover classes. In the closed woodland, the species with the highest
relative abundance were Bridelia michrantha (70%), Albizia grandibracteata
(55%) and Maesopsis eminii (55%). In the open woodland there were no
species with relative abundance above 50%. The highest was Ficus exasperata
with 48%. In the very open woodland, Combretum molle, Securinega virosa and
Dombeya rotundifolia had the highest relative abundance with 74, 71 and 75%,
respectively. Combretum guenzi exclusively occurred in the wooded grassland.
Other species with high relative abundance in the wooded grassland
were Combretum binderanun, Grewia bicolor, Lonchocarpus laxiflorus and
Hymenocardia acida with 78, 66, 58 and 50%, respectively.
Most of the species identified as belonging to groups A and B e.g. Cynometra
alexandri, Khaya anthotheca, Diospyros abyssinica, Uvariopsis congensis and
Holoptelea grandis (Figure 3b) were also identified through indicator species
analysis as good indicators for the No-fire class. Of these, Diospyros abyssinica,
Uvariopsis congensis, Holoptelea grandis and all Celtis species were also good
indicators of the forest class (Appendix 2). The species in group C belonged
both to Old-fire and Recent-fire classes. Considering the cover classes,
Terminalia velutina and Albizia grandbracteata were good indicators for closed
woodland, Grewia mollis and Combretum mole for very open woodland and
Lonchocarpus laxiflorus, Grewia bicolor and Combretum guenzi were good
indicators for the wooded grassland class. Several of these species e.g.
Uvariopsis congensis, Terminalia velutina and Grewia mollis have distinctively
high abundance in specific areas along the gradient (Figure 5).
Although the closed woodland had the largest area sampled followed by the
open woodland, the forest had the highest number of species and genera
identified (Appendix 3). The lowest number of species and genera was found in
the wooded grassland. The highest ratio of species to genera was in very open
woodland (1.4) and the lowest in wooded grassland (1.2). Eleven species occur
in all classes and most species occur in more than one cover class but their
abundance varies greatly between classes. Forest and closed woodland classes
had an equal number of families and wooded grassland class had the lowest
number of families.
[419]
1480
Figure 5. Relationship between DCA axis 1 and some of the most abundant species whose
maximum abundance occur in different areas along the gradient. The selected species also display a
variation in their distribution range.
The Simpson index of all vegetation classes differed only slightly except that
of wooded grassland (Figure 6a). The wooded grassland had the highest value
and the highest standard error. The forest class had the highest mean Fisher’s a
(Figure 6b) followed by the closed woodland class. These two classes were
significantly different from all other classes but not from each other. The open
woodland was also significantly different from the wooded grassland. The
wooded grassland had the lowest Fisher’s a.
The basal area (Figure 6c) decreased from the forest, which had the highest
value, to the wooded grassland, which had the lowest. The forest also showed
the highest variation. All cover types were significantly different from each
other. The mean stem density values for the forest, closed woodland and open
woodland were very close (Figure 6d) and there was no significant difference
between them. The very open woodland also had a high mean value although
slightly lower than the other 3. The wooded grassland is much lower than all
others. The very open woodland and the wooded grassland are each signifi-
cantly different from all others. So while many individual trees may be found in
each cover type, they vary in size with the forest having larger trees than any of
the other cover types. Details of the species occurring in each cover type and
their abundance are indicated in Appendix 3.
Discussion
The species composition along the gradient gradually changes from species
that attain maximum abundance in areas of minimum disturbance e.g.
[420]
1481
Figure 6. comparison of cover class mean and standard deviation for (a) Simpson index, (b)
Fisher’s a, (c) basal area and (d) stem density. The class numbers consistently represent 1, forest; 2,
closed woodland; 3, open woodland; 4, very open woodland; and 5, wooded grassland. The letters
beside each bar indicate significance differences. Bars, for a specific variable, which have the same
letter mean that they are not significantly different (ANOVA: p = 0.05).
[421]
1482
Of the environmental variables recorded, fire best explained the gradient. This
is evidenced by the high correlation between DCA and fire (Table 1) and the
fact that the compositional gradient could be divided using the fire regime
(Figure 4a). Areas that had recent fires, and are probably most frequently
burnt, had species that characteristically display fire resistant traits e.g. a thick
bark, pealing off of the old bark and good sprouting ability after a fire
(Gashaw et al. 2002; Saha and Howe 2003; Vesk and Westoby 2004). The
occurrence of some species is thus influenced by their fire-tolerance level
(Cauldwell and Zieger 2000) with increasingly more of the less fire resistant
species in the Old-fire class. Here, seed dispersal (a factor not explored in this
study) may have an important role. A number of the species that occurred in
the Old-fire class were most abundant in the No-fire class. Their seeds were
probably dispersed into the Old-fire class areas e.g. by wind and, when con-
ditions became favorable, they got established. Hence we suggest that the
existent fire regime influences their low occurrence (Huston 1994).
Although water is often a limiting factor for plant survival, in humid FWS
mosaics, water distribution is not a critical controlling factor (Favier et al.
2004). Despite the variation in rainfall over Budongo Forest Reserve, with the
northern part receiving less rain than the south (Plumptre 1996), the north still
receives over 1200 mm a year (Forest Department Uganda 1997) which is
sufficient for forest maintenance. Also elephants that previously restricted
forest expansion (Laws et al. 1975) are no longer present. The species turnover
could possibly be explained by an additive effect of the environmental variables
considered in this study, the historical impact by elephants and probably other
factors that were not considered in this study e.g. seed dispersal mechanisms,
which have been shown to favor establishment of species with higher dispersal
ability in the post disturbance period (Hovestadt et al. 1999; Ohsawa et al.
2002). However, just like in other studies where FWS occur (Elliott et al. 1999;
Hovestadt et al. 1999), fire plays a major role in controlling species distribution
pattern but it does not explain all the variation (Weiher 2003). Accessibility to
the protected areas, where local people mainly utilize areas closest to them
(Acharya 1999; Obiri et al. 2002), also showed a significant relationship with
the species composition gradient.
[422]
1483
shown (Cousins and Lindborg 2004) to correspond well with the succession
gradient. Classification of mosaic areas using remotely sensed data could
therefore be a good start for identification of the vegetation types that exist
within them. This would require less time (Schmidt et al. 2004) as compared to
when only field surveys would have been used.
Our study has shown that although the forest significantly differed in species
diversity and vegetation structure, especially basal area, there was a systematic
decrease in variation from forest to wooded grassland (Figure 5). A major
gradient stretching from the forest to the wooded grassland is evident (Dezzeo
et al. 2004) and species composition and forest structure vary along this gra-
dient. Most of the areas sampled by Eggeling (1947) and followed up in Sheil
et al. (2000) had not had disturbance for a long time. Areas sampled in this
study, however, cover both areas with ranging times since last disturbance and
areas that are still experiencing frequent disturbance. Thus, in this study we
observe a wider range of vegetation variation.
Although subtle variations in vegetation structure may be evident in some
landscapes, the species composition variation is often more complex (Muh-
lenberg et al. 1990). In our study, the observed gradual change in species
composition along the gradient and the compositional interrelationship
between the vegetation cover classes indicate that the FWS mosaic is a single,
interacting, integrated unit.
[423]
1484
Figure 7. DCA graph obtained after combining a resampled set of Eggeling’s data to the data
used in this paper. Axis 1 had an eigenvalue of 0.38 and explained 9.8% of the variation. The
second axis had an eigenvalue of 0.19 and explained 4.9%.
[424]
1485
ought to be preserved (Sheil and Burslem 2003), the woodland areas should
not all be allowed to become forest since that would mean loosing the
woodland dependant species. And the highest number of species can only be
conserved when complementary areas are included in the conservation plan
(Howard et al. 1998). The maintenance of the high diversity of Budongo,
being an isolated forest with no immediate source of additional forest species,
may be more attributed to the existence of all stages of the succession gra-
dient (Richardson-Kageler 2004; Shea et al. 2004) than acquisition of more
forest species from elsewhere, which, additionally, often takes a long time
(Chapman et al. 1997). Hence, if reforestation of Budongo Forest Reserve
would continue to the extent that the woodland areas would be lost, the
biodiversity of the reserve would probably decrease. For purposes of con-
serving woody plants in a dynamic landscape, it is thus important that each
vegetation type represented is included and maintained within the conserva-
tion area (Bengtsson et al. 2003). In the area under study, fire disturbance is a
requirement for species coexistence (Shea et al. 2004).
In areas where fire may be applied, the vegetation type and its develop-
ment stage may affect the potential for ignition and spread of the fire
(Everett et al. 2000). Although no evidence exists of fires having destroyed
tropical rain forests in Uganda, it has been observed elsewhere that tropical
forests can burn (Cochrane and Schuize 1999; Cochrane and Laurance 2002;
Laurence 2003). This, however, mainly occurs in the presence of very dry
conditions, in fragmented forest landscapes and when fire is carelessly
applied in or adjacent to logged over areas. Fire also remains a highly de-
bated conservation management tool (Mentis and Bailey 1990; Trollope et al.
1995; van Wilgen et al. 1998). It is therefore important that fire be used
cautiously and, probably learning and using burning methods that have been
used in the past (Goma et al. 2001) will be a prerequisite. In this respect,
conservationists need to focus more attention on the current vegetation
management practices of local people surrounding conservation areas (Leone
and Lovreglio 2004) since they have been noted to use fire destructively
(Condit et al. 1998; Wheater 1971).
In Africa FWS mosaics are prevalent in areas surrounding the Congo basin
forests, including Uganda. These areas have been defined as transitional zones
between the moist tropical forest and the drier savanna landscape typical of
much of Africa. On the northern side, the transition occurs at about 8 N with
the exception of Togo and Benin and part of Ivory Coast (Gautier and Spi-
chiger 2004). Many FWS mosaics occur in Uganda because of its location in a
zone of overlap between the ecological communities characteristic of the dry
East African savannas and the West African rainforests (Howard 1991). The
observations made in this study and their management implications are,
therefore, relevant to many areas in Africa and in much of the tropical world
where such landscapes occur.
[425]
1486
Acknowledgements
We thank Mr Hezekias Ddumba for being there for us to sort out our species
identification problems and for all the logistical support he provided while we
stayed at Kaniyo-Pabidi ecotourism camp. We express our gratitude to Mr Oli-
ver van Straaten who contributed part of the data used in this study. We also
express our gratitude to Professor Dr Alfred De Gier for his continued support
and for his valuable comments during the preparation of this manuscript.
Appendix 1. The Indicator Species Analysis output based on fire indicator classes.
Species names RA-0 RF-0 RA-1 RF-1 RA-2 RF-2 IV p Fire class
[426]
1487
Appendix 1. Continued.
Species names RA-0 RF-0 RA-1 RF-1 RA-2 RF-2 IV p Fire class
Albizia grandibracteata 64 41 31 22 5 6 26.2 0.001 0
Stereospermum kunthianum 22 28 44 50 34 47 21.9 0.003 1
Combretum molle 1 5 31 24 68 49 33.7 0.001 2
Lonchocarpus laxiflorus 10 12 19 22 71 45 31.7 0.001 2
Vitex doniana 38 33 36 36 26 25 12.8 0.585
Funtumia elastica 94 31 6 3 0 0 28.9 0.001 0
Lanea barteri 27 25 41 32 32 29 13.1 0.174
Celtis wightii 100 24 0 0 0 0 24.2 0.001 0
Acacia hockii 22 19 46 29 32 23 13.4 0.033
Piliostgma thonningii 25 18 53 31 23 15 16.2 0.004 1
Caloncoba schweinfurthii 86 23 14 6 0 0 20.2 0.001 0
Holoptelea grandis 99 22 1 1 0 0 21.6 0.001 0
Maesopsis eminii 62 21 38 12 0 0 12.8 0.003 0
Diospyros abyssinica 100 10 0 0 0 0 9.6 0.001 0
Ficus sur 43 15 48 15 9 3 7.1 0.245
Grewia bicolor 8 5 23 9 69 18 12.8 0.001 2
Khaya anthotheca 86 15 14 2 0 0 12.6 0.001 0
Dombeya mukole 79 8 17 3 5 1 6.4 0.013 0
Bridelia micrantha 87 12 9 1 4 1 10.7 0.001 0
Celtis durandii 86 12 14 2 0 0 9.9 0.003 0
Combretum binderanum 2 0 37 6 61 9 5.6 0.012 2
Margaritaria discoidea 71 11 29 2 0 0 7.6 0.004 0
Phyllanthus discoideus 100 10 0 0 0 0 9.6 0.001 0
Albizia zygia 56 2 20 2 24 3 1.3 0.829
Celtis zenkeri 100 10 0 0 0 0 10.4 0.001 0
Pterygota mildbreadii 70 5 30 1 0 0 3.8 0.059
Hymenocardia acida 11 3 27 5 62 9 5.7 0.008 2
Olea welwitschii 95 9 5 1 0 0 8.8 0.001 0
Oncoba spinosa 37 5 14 1 49 6 2.8 0.326
Tapura fisheri 80 9 20 2 0 0 7.1 0.009 0
Securinega virosa 0 0 38 5 62 10 6.4 0.004 2
Dichrostachys cinerea 73 5 27 1 0 0 3.9 0.064
Alstonia boonei 100 9 0 0 0 0 8.8 0.001 0
Cynometra alexandri 100 5 0 0 0 0 5 0.007 0
Ficus exasperata 37 3 63 3 0 0 2.2 0.272
Combretum gueinzii 0 0 73 1 27 1 0.5 0.749
Sapium ellipticum 81 2 19 1 0 0 1.2 0.403
Carpololobia alba 0 0 21 3 79 6 4.5 0.006 2
Dombeya rotundifolia 8 2 25 5 67 8 5.4 0.012 2
Trichilia prieuriana 100 5 0 0 0 0 5.4 0.007 0
It indicates the concentration of each species in each class (Relative abundance, RA), the faith-
fulness of occurrence of the species in that class (Relative frequency, RF), the highest species
indicator value across the classes (IV) the statistical significance of the indicator value (p) and the
class in which a particular species had the highest indicator value (Fire class). For species that were
not significant indicators for any class, fire class was left blank. RA is expressed as a proportion of a
particular species in a particular class relative to its abundance in other classes. RF is expressed as
the percentage of sample units in a class that contain that species.
p is significant at 0.01.
0, No fire; 1, old fire; and 2, recent fire.
[427]
1488
Appendix 2. The Indicator Species Analysis output based on vegetation cover classes.
Species RA-1 RF-1 RA-2 RF-2 RA-3 RF-3 RA-4 RF-4 RA-5 RF-5 IV P Cover
cord
[428]
1489
Appendix 2. Continued.
Species RA-1 RF-1 RA-2 RF-2 RA-3 RF-3 RA-4 RF-4 RA-5 RF-5 IV P Cover
cord
Celtis zenkeri 100 22 0 0 0 0 0 0 0 0 21.6 0.006 1
Pterygota 100 13 0 0 0 0 0 0 0 0 12.8 0.025
mildbreadii
Hymenocardia 0 0 3 2 20 9 27 7 50 22 11.2 0.045
acida
Olea 92 17 8 3 0 0 0 0 0 0 15.4 0.016
welwitschii
Oncoba 48 9 6 1 12 3 33 6 0 0 4.2 0.166
spinosa
Tapura fisheri 90 18 10 2 0 0 0 0 0 0 16.6 0.014 1
Securinega 0 0 8 1 21 3 71 14 0 0 10.1 0.03
virosa
Dichrostachys 63 6 18 4 12 1 8 1 0 0 4 0.138
cinerea
Alstonia 86 14 14 3 0 0 0 0 0 0 12.4 0.033
boonei
Cynometra 100 10 0 0 0 0 0 0 0 0 10.4 0.022
alexandri
Ficus 6 1 46 5 48 4 0 0 0 0 2.2 0.454
exasperata
Combretum 0 0 0 0 0 0 0 0 100 22 22.2 0.001 5
gueinzii
Sapium 82 2 18 1 0 0 0 0 0 0 2 0.174
ellipticum
Carpololobia 0 0 32 1 38 4 30 4 0 0 1.4 0.607
alba
Dombeya 0 0 8 2 17 4 75 13 0 0 9.7 0.04
rotundifolia
Trichilia 100 11 0 0 0 0 0 0 0 0 11.2 0.03
prieuriana
It indicates the concentration of each species in each class (Relative abundance, RA), the faithfulness of
occurrence of the species in that class (Relative frequency, RF), the highest species indicator value across the
classes (IV) the statistical significance of the indicator value (p) and the class in which a particular species had the
highest indicator value (Cover cord). For species that were not significant indicators for any class, cover cord was
left blank. RA is expressed as a proportion of a particular species in a particular class relative to its abundance in
other classes. RF is expressed as the percentage of sample units in a class that contain that species.
p is significant at 0.01.
1, Forest; 2, closed woodland; 3, open woodland; 4, very open woodland; and 5, wooded grassland.
Appendix 3. The 121woody species identified in the field, their abundance per class.
Number of individuals
[429]
1490
Appendix 3. Continued.
Number of individuals
Mimosaceae Acacia sieberiana Dc. Var. 1 6 . 3 .
woodii (Burtt Davy)
Keay & Brenan
Mimosaceae Acacia spp. 2 . . . .
Euphorbiaceae Acalypha neptunica Müll. 9 . . . .
Arg. Var.
Mimosaceae Albizia coriaria Oliver 9 8 2 . .
Mimosaceae Albizia grandibracteata Taub. 120 257 57 7 .
Mimosaceae Albizia spp. . . 1 . .
Mimosaceae Albizia zygia (DC.) Macbr. 33 5 3 3 .
Apocynaceae Alstonia boonei de Wild 24 5 . . .
Sapotaceae Aningeria altissima (A. Chev.) 15 2 1 2 .
Aubr. & Pellegr.
Annonaceae Annona senegalensis Pers. 35 157 194 129 1
Balemetea gramofolia 1 . . . .
Rubiaceae Belonophora glomerata 2 1 . . .
Sapindaceae Blighia unijugata Baker 2 6 . . .
Euphorbiaceae Bridelia micrantha (Hochst.) Baill. 15 49 1 1 .
Euphorbiaceae Bridelia scleroneuroides Pax. . 2 5 2 .
Elacourtiacea Caloncoba schweinfurthii Glig. 107 42 3 . .
Polygalaceae Carpololobia alba G. Don . 9 12 4 .
Caesalpinioideae Cassia siamea Lam. 2 13 . . .
Caesalpinioideae Cassia spp. . 1 . . .
Ulmaceae Celtis durandii Engl. 59 6 . . .
Ulmaceae Celtis mildbraedii Engl. 7 . . . .
Ulmaceae Celtis wightii Planch. 204 1 . . .
Ulmaceae Celtis zenkeri Engl. 43 . . . .
Moraceae Chlorophora excelsa (Welw.) Benth . 3 . . .
Sapotaceae Chrysophyllum albidum G. Don 6 . . . .
Rutaceae Citropsis articulata (Wild. Ex Spreng) 1 . . . .
Swingle & M. Kellerm
Annonaceae Cleistopholis patens (Beth.) Engl. & Diels 1 . . . .
Closophila magida 1 . . . .
Rubiaceae Coffea canephora Pierre ex Froechner. 5 . . . .
Rubiaceae Coffea euginiodes 5 . . . .
Sterculiaceae Cola gigantea A. Chev. 13 . . . .
Combretaceae Combretum binderanum Kotschy . 3 16 14 17
Combretaceae Combretum collinum Fresen. 107 240 431 137 12
Combretaceae Combretum gueinzii Sond. . . 2 . 21
Combretaceae Combretum molle R. Br. Ex G. Don 4 18 92 243 6
Boraginaceae Cordia millenii Baker 7 7 . . .
Aralliaceae Cussonia arborea Hochst. Ex A. Rich. . 1 4 12 .
Caesalpiniaceae Cynometra alexandri CH Wright 29 . . . .
Mimosaceae Dichrostachys cinerea (L.) Wright & Arn 19 8 3 1 .
Ebenaceae Diospyros abyssinica (Hiern) F. White 90 3 1 . .
Sterculiaceae Dombeya mukole Sprague 56 3 6 1 .
Sterculiaceae Dombeya rotundifolia (Hochst.) Planch. 2 3 6 12 .
Mimosaceae Entada abyssinica Steud. Ex A. Rich . . 1 . .
Meliaceae Entandrophragma angolense (Welw.) C. DC. 3 . . . .
[430]
1491
Appendix 3. Continued.
Number of individuals
Meliaceae Entandrophragma cylindricum 1 . . . .
(Sprague) Sprague
Papilionaceae Erythrina abyssinica Lam. Ex DC 1 4 1 3 .
Leguminosae Erythrophleum suaveolens 8 . . . .
(Guill. & Perr.)Brenan
Rutaceae Fagaropsis angolensis (Engl.) 5 1 . . .
HM. Gardner
Moraceae Ficus capensis Thunb 1 1 . . .
Moraceae Ficus casuarina . 1 . . .
Moraceae Ficus exasperata Vahl 1 10 8 . 5
Moraceae Ficus mucuso Welw ex Ficalho 1 5 1 . .
Moraceae Ficus polita Vahl 9 5 1 . .
Moraceae Ficus saussureana DC. . 1 . . .
Moraceae Ficus spp. 1 . . . .
Moraceae Ficus sansibarica Warb. 1 . . . .
Moraceae Ficus sur Forssk 17 41 26 3 .
Apocynaceae Funtumia elastica (Preuss) Stapf 227 32 . . .
Rubiaceae Gardenia Jovis-tonantis (Welw.) Hiern. . 1 1 1 .
Tiliaceae Grewia bicolor Juss. . 10 27 26 17
Tiliaceae Grewia mollis Juss. 58 424 761 531 6
Simaroubaceae Harrisonia abyssinica Oliv. . 1 . . .
Ulmaceae Holoptelea grandis (Hutch.) Mildbr. 99 10 . . .
Euphorbiaceae Hymenocardia acida Tul. 3 3 22 15 3
Meliaceae Khaya anthotheca (Welw.) C. DC. 50 16 4 . .
Meliaceae Khaya grandifolia C. DC. 1 . . . .
Bignoniaceae Kigeria africana (Lam.) Benth . . 5 1 .
Anacardiaceae Lanea barteri (Oliv.) Engl. 5 78 107 48 2
Anacardiaceae Lannea welwitschii (Hiern.) Engl. . 3 . 1 .
Rhamnaceae Lasiodiscus mildbraedii Engl. 1 . . . .
Sapindaceae Lepisanthes senegalensis (Juss. Ex Poir.) 6 5 . . .
Papilionaceae Lonchocarpus laxiflorus Guill. & Perr. . 20 116 108 34
Capparidaceae Maerua duchensii 12 . . . .
Rhamnaceae Maesopsis eminii Engl. 47 61 . . .
Meliaceae Mahogany spp. 8 . . . .
Euphorbiaceae Margaritaria discoidea (Baill.) Webster 29 21 . . .
Rignoniaceae Markhamia platycalyx (Baker) Sprague 2 1 . . .
Celastraceae Maytenus undata (Thunb.) Blakelock . 1 4 7 4
Papilionaceae Mildbraediodendron excelsum (Harms) 3 . . . .
Moraceae Milicia excelsa (Welw.) CC Berg 2 2 . . .
Rubiaceae Mitragyna stipulosa (DC.) O. Ktze 1 . . . .
Moraceae Morus lactea (Sim) Mildbr. . 1 . . .
Moraceae Myrianthus holstii Engl. 4 . . . .
Oleaceae Olea welwitschii (Knobl.) Gilg & Schellenb. 36 4 . . .
Flacourtiaceae Oncoba spinosa Forsk. 24 3 5 8 .
Palmae Phoenix reclinata Jacq. 12 . . . .
Euphorbiaceae Phyllanthus discoideus Muell. 46 2 . . .
Caesalpiniaceae Piliostgma thonningii (Schum.) 21 66 75 18 12
Verbenaceae Premna angolensis Guerke 18 2 . . .
Proteaceae Protea madiensis Oliv. . . . 9 .
Anacardiaceae Pseudospondias microcarpa (A. Rich.) Engl. 4 2 . . .
[431]
1492
Appendix 3. Continued.
Number of individuals
Sterculiaceae Pterygota mildbraedii Engl. 40 2 . . .
Euphorbiaceae Ricinodendron excelsum 8 2 . . .
Violaceae Rinorea dentata (P. Beauv.) Kuntze 1 . . . .
Violaceae Rinorea ilicifolia (Welw. Ex Oliv.) 13 . . . .
Capparidaceae Ritchiea albersii Gilg 1 . . . .
Rubiaceae Rothmannia urcelliformis (Hiern) . . 4 5 6
Bullock exRobyns
Celestraceae Salacia elegans Welw. Ex Oliv. . . 3 2 5
Euphorbiaceae Sapium ellipticum Pax. 20 1 . . .
Oleaceae Schrebera arborea A. Chev. 9 . . . .
Polygalaceae Securidaca spp. . 4 10 19 .
Euphorbiaceae Securinega virosa (Roxb. Ex Willd.) Baill 2 3 5 . .
Bignoniaceae Spathodea campanulata P. Beauv. 4 12 . . .
Umbelliferae Steganotaenia araliacea Hochst. 2 2 2 9 .
Bignoniaceae Stereospermum kunthianum Cham. 41 162 126 55 18
Apocynaceae Tabernaemontana holstii K. Schum 13 1 . . .
Chailletiaceae Tapura fisheri 32 6 . . .
Rutaceae Teclea nobilis Del. 11 10 2 . .
Combretaceae Terminalia velutina Rolfe 308 1288 824 103 14
Euphorbiaceae Thecacoris lucida . . . 4 .
Ulmaceae Trema orientalis (L.) Blume 1 2 . . .
Meliaceae Trichilia prieuriana A. Juss 23 . . . .
Meliaceae Trichilia spp. 3 . . . .
Meliaceae Turrae floribunda 1 . 1 . .
Annonaceae Uvariopsis congensis Robyns & Ghesq. 663 . . . .
Rubiaceae Vangueria apiculata K. Schum . 2 . . .
Compositae Vernonia amygdalina Delile 1 2 . . .
Verbenaceae Vitex doniana Sweet. 20 143 92 29 6
Rhamnaceae Zizyphus abyssinica Hochst. Ex A. Rich 1 . 1 . .
Total individuals 3042 3394 3172 1602 196
Total species 95 77 48 39 18
Total genera 73 60 35 28 15
Total families 33 33 23 19 13
Total area (sq. m) 65300 71000 65500 42100 16500
References
Acharya B. 1999. Forest biodiversity assessment: a spatial analysis of tree species diversity in
Nepal. Ph.D. Thesis, Leiden University, Enschede, 199 pp.
Alados C.L. et al. 2004. Variations in landscape patterns and vegetation cover between 1957 and
1994 in a semiarid mediterranean ecosystem. Landscape Ecol. 19: 543–559.
Balvanera P., Lott E., Segura G., Siebe C. and Islas A. 2002. Patterns of B-diversity in a Mexican
tropical dry forest. J. Vege. Sci. 13(2): 145–158.
Bengtsson J. et al. 2003. Reserves, resilience and dynamic landscapes. Ambio 32(6): 389–396.
Buechner H.K. and Dawkins H.C. 1961. Vegetation change induced by elephants and fire in the
Murchison falls national park, Uganda. J. Ecol. 42(4): 752–766.
[432]
1493
Cauldwell A.E. and Zieger U. 2000. A reassessment of the fire-tolerance of some miombo woody
species in the Central Province, Zambia. Afr. J. Ecol. 38: 138–146.
Chapman C.A., Chapman L.J., Wrangham R., Isabirye-Basuta G. and Ben-David K. 1997. Spatial
and temporal variability in the structure of a tropical forest. Afr. J. Ecol. 35: 287–302.
Cochrane M.A. and Laurance W.F. 2002. Fire as a large-scale edge effect in Amazonian forests.
J. Trop. Ecol. 18: 311–325.
Cochrane M.A. and Schuize M.D. 1999. Fire as a recurrent event in tropical forests of eastern
Amszon: effects on forest structure, biomass, and species composition. Biotropica 31: 2–16.
Collins B.J. and Woodcock E.C. 1996. An assessment of several linear change detection techniques
for mapping forest mortality using multitemporal Landsat TM data. Remote Sens. Environ.
56: 66–77.
Condit R., Sukumar R., Hubbell S.P. and Foster R.B. 1998. Predicting population trends from size
distributions: a direst test in a tropical tree community. Am. Natural. 152(4): 495–509.
Connell J.H. 1978. Diversity in tropical rain forests and coral reefs. Science 199: 1302–1310.
Cousins S.A.O. and Lindborg R. 2004. Assessing changes inplant distribution patterns-indicator
species versus plant functional types. Ecol. Indicators 4: 17–27.
Crist E.P. and Cicone R.C. 1984. Application of the Tasseled Cap concept to simulated Thematic
Mapper data. Photogramm. Eng. Remote Sens. 50: 343–352.
Crist E.P., Laurin R. and Cicone R.C. 1986. Vegetation and soils information contained in
transformed Thematic Mapper data. Proceedings of the IGARSS ’86 Symposium, Zurich,
Switzerland, ESA, Paris, pp. 1465–1470.
Crow T.R. and Perera A.H. 2004. Emulating natural landscape disturbance in forest management –
an introduction. Landscape Ecol. 19: 231–233.
Dezzeo N., Chacón N., Sanoja E. and Picón G. 2004. Changes in soil properties and vegetation
characteristics along a forest–savanna gradient in southern Venezuela. Forest Ecol. Manag.
200: 183–193.
Dufrêne M. and Legendre P. 1997. Species assemblages and indicator species: the need for a flexible
asymmetrical approach. Ecol. Monogr. 67(3): 345–366.
Eggeling W.J. 1947. Observations on the ecology of the Budongo rain forest, Uganda. J. Ecol.
34(1): 20–87.
Eggeling W.J. and Dale R.I. (eds) 1952. The Indigenous Trees of the Uganda Protectorate. Uganda
Government printer, Entebbe.
Eilu G., Hafashimana D.L.N. and Kasenene J.M. 2004. Tree species distribution in forests of the
Albertine Rift, Western Uganda. Afr. J. Ecol. 42: 100–110.
Elliott K.J., Hendrick R.L., Major A.E., Vose J.M. and Swank W.T. 1999. Vegetation dynamics
after a prescribed fire in the southern Appalachians. Forest Ecol. Manag. 114: 199–213.
Everett R.L., Schellhaas R., Keenum D., Spurbeck D. and Ohlson P. 2000. Fire history in the
ponderosa pine/Douglas-fir forests on the east slope of the Washington Cascades. Forest Ecol.
Manag. 2000: 207–225.
Favier C., Chave J., AFabing A., Schwartz D. and Dubois M.A. 2004. Modelling forest–savanna
mosaic dynamics in man-influenced environments: effects of fire, climate and soil heterogeneity.
Ecol. Model. 171: 85–102.
Fisher R.A., Corbet A.S. and Williams C.B. 1943. The relation between the number of species and
the number of individuals in a random sample of an animal population. J. Anim. Ecol. 12: 42–58.
Forest Department Uganda 1997. Forest management plan for Budongo Forest Reserve for 1997
to 2007. Forest Department, Kampala, 107 pp.
Forest Department Uganda 1999. Forestry Nature Conservation Master Plan, Ministry of Water.
Lands and Environment, Kampala.
Gashaw M. et al. 2002. Post fire regeneration strategies and tree bark resistance to heating in
frequently burning tropical savanna woodlands in Ethiopia. Nord. J. Bot. 22(1): 19–33.
Gautier L. and Spichiger R. 2004. The forest–savanna transition in West Africa. In: Poorter L.,
Bongers F., Kouamé F.N. and Hawthorne W.D. (eds), Biodiversity of West African Forests: An
Ecological Atlas of Woody Plant Species. CAB International, Wallingford, pp. 33–40.
[433]
1494
Goma H.C., Rahim K., Nangendo G., Riley J. and Stein A. 2001. Participatory studies for agro-
ecosystem evaluation. Agric. Ecosyst. Environ. 87: 179–190.
Hamilton A. 1991. A Field Guide to Ugandan Forest Trees. Makerere University, Kampala.
Hovestadt T., Yao P. and Linsenmair E.K. 1999. Seed dispersal mechanisms and the vegetation of
forest islands in a West African forest savanna mosaic (Comoé National Park, Ivory Coast).
Plant Ecol. 144: 1–25.
Howard P.C. 1991. Nature Conservation in Uganda Tropical Forest Reserves. IUCN, Gland.
Howard P.C. et al. 1998. Complementarity and the use of indicator groups for reserve selection in
Uganda. Nature 394: 472–475.
Huston M.A. 1994. Biological Diversity. The Coexistence of Species on Changing Landscapes.
Cambridge University Press, London.
Laurence W.F. 2003. Slow burn: the insidious effects of surface fires on tropical forests. Trends
Ecol. Evol. 18(5): 209–212.
Laws R.M., Parker I.S.C. and Johnstone R.C.B. 1975. Elephants and Their Habitats. The Ecology
of Elephants in North Bunyoro, Uganda. Oxford University Press, London.
Leone V. and Lovreglio R. 2004. Conservation of Mediterranean pine woodlands: scenarios and
legislative tools. Plant Ecol. 171: 221–235.
Li J., Loneragan W.A., Duggin J.A. and Grant C.D. 2004. Issues affecting the measurement of
disturbance response patterns in herbaceous vegetation – a test of the intermediate disturbance
hypothesis. Plant Ecol. 172: 11–26.
Magurran A.E. 1988. Ecological Diversity and its Measurement. Princeton University Press,
Princeton, New Jersey, 179 pp.
McCune B., Grace J.B. and Urban D.L. 2002. Analysis of the Ecological Communities. MjM
Software, Gleneden Beach, Oregon, US.
McCune B. and Mefford M.J. 1999. Multivariate Analysis of Ecological Data. Version 4.25. MjM
Software, Gleneden Beach, Oregon, US.
Mentis M.T. and Bailey A.W. 1990. Changing perceptions of fire management in savanna parks.
J. Grassland Soc., Southern Africa 7: 81–85.
Muhlenberg M., Galat-Luong A., Poilecot P., Steinhauer-Burkart B. and Kuhn I. 1990. The
importance of forest islands within wet savannas for the conversation of rainforest animals in the
Ivory Coast. Revue d’Ecologie (Terre et la Vie) 45(3): 197–214.
Mwami P.M. and McNeilage A. 2003. Natural regeneration and ecological recovery in Bwindi
Impenetrable National Park, Uganda. Afr. J. Ecol. 41: 93–98.
Nangendo G. 2005. Changing forest–woodland–savanna mosaics in Uganda: with implications for
conservation. Ph.D. Thesis, Wageningen University, Wageningen, 139 pp.
Nangendo G., Skidmore A.K. and van Oosten H. submitted. Mapping East African tropical forests
and woodlands: a comparison of classifiers. ISPRS J. Photogramm. Remote Sensing.
Nangendo G., van Straaten O.ter Steege H., de Gier A. and Bongers F. submitted. Vegetation
cover change and its relationship with species distribution dynamics. J. Biogeogr..
Obiri J., Lawes J. and Mukolwe M. 2002. The dynamics and sustainable use of high-value tree
species of the coastal Pondoland forests of the Eastern Cape Province, South Africa. Forest Ecol.
Manag. 166: 131–148.
Ohsawa K., Kawasaki K., Takasu F. and Shigesada N. 2002. Recurrent habitat distribution and
species diversity in a multiple-competitive species system. J. Theor. Biol. 216: 123–138.
Paterson D.J. 1991. The ecology and history of Uganda’s Budongo forest. Forest Conserv. Hist.
35: 179–186.
Petraitis P.S., Latham R.E. and Niesenbaum R.A. 1989. The maintenance of species diversity by
disturbance. Quarter. Rev. Biol. 64: 393–418.
Plumptre A.J. 1996. Changes following 60 years of selective timber harvesting in the Budongo
Forest Reserve, Uganda. Forest Ecol. Manag. 89: 101–113.
Richardson-Kageler S.J. 2004. Effects of large herbivore browsing on the functional groups of
woody plants in a southern African savanna. Biodivers. Conserv. 13: 2145–2163.
[434]
1495
Saha S. and Howe H.F. 2003. Species composition and fire in a dry deciduous forest. Ecology
84(12): 3118–3123.
Schmidt S.K. et al. 2004. Mapping costal vegetation using an Expert System and Hyperspectral
imagery. Photogramm. Eng. Remote Sens. (in press).
Schwartz M.W. and Caro T.M. 2003. Effect of selective logging on tree and understory regener-
ation in Miombo woodland in western Tanzania. Afr. J. Ecol. 41: 75–82.
Schwilk D.W., Keeley J.E. and Bond W.J. 1997. The intermediate disturbance hypothesis does not
explain fire and diversity pattern in fynbos. Plant Ecol. 132: 77–84.
Shea K., Roxburgh S.H. and Rauschert E.S.J. 2004. Moving form pattern to process: coexistence
mechanisms under intermediate disturbance regimes. Ecol. Lett. 7: 491–508.
Sheil D. 1999. Developing tests of successional hypotheses with size-structured populations, and an
assessment using long-term data from a Ugandan rain forest. Plant Ecol. 140: 117–127.
Sheil D. and Burslem F.R.P.D. 2003. Disturbing hypothesis in tropical forests. Trends Ecol. Evol.
18(1): 18–26.
Sheil D., Jennings S. and Savill P. 2000. Long-term permanent plot observations of vegetation
dynamics in Budongo, a Ugandan rain forest. J. Trop. Ecol. 16: 765–800.
Smart N.O.E., Hatton J.C. and Spence D.H.N. 1985. The effect of long-term exclusion of large
herbivores on vegetation in Murchison Falls National Park, Uganda. Biol. Conserv. 33: 229–245.
Swaine M.D., Hawthorne W.D. and Orgle T.K. 1992. The effects of fire exclusion on savana
vegetation at Kpong, Ghana. Biotropica 24: 166–172.
Trapnell C.G. 1959. Ecological results of woodland burning experiments in Northern Rhodesia.
J. Ecol. 47: 129–168.
Trollope W.S.W., Biggs H.C., Potgieter A.L.F. and Zambatis N. 1995. A structured versus a
wilderness approach to burning in the Kruger National Park in South Africa. In: West N.E.
(ed.), Rangelands in a Sustainable Biosphere. Society for Rangeland Management, Denver.
Tucker J.C. 1979. Red and Photographic Infrared linear combinations for monitoring vegetation.
Remote Sens. Environ. 8: 127–150.
van Straaten H.P. 1976. Präkambrium und junges Western Rift im Bunyoro District NW-Uganda.
Unpublished Ph.D., Universität Göttingen, Hanover.
van Straaten O. 2003. Changing woodland systems: post-disturbance woody species succession
dynamics and spatial trends. M.Sc. Thesis, International Institute for Geoinformation Science
and Earth Observation, Enschede, 70 pp.
van Wilgen B.W., Biggs H.C. and Potgieter A.L.F. 1998. Fire management and research in the
Kruger National Park, with suggestion on the detection of thresholds of potential concern.
Koedoe 41: 69–87.
Vesk P.A. and Westoby M. 2004. Sprouting ability across diverse disturbances and vegetation types
worldwide. J. Ecol. 92(2): 310–320.
Walter H. 1985. Vegetation of the Earth and Ecological Systems of the Geo-Biospere. Springer-
Verlag, 318 pp.
Weiher E. 2003. Species richness along multiple gradients: testing a general multivariate model in
oak savanna. OIKOS 101: 311–316.
Wheater R.J. 1971. Problems of controlling fires in Uganda National parks, Annual Tall Timbers
Fire Ecology Conference.
[435]
Biodiversity and Conservation (2006) 15:1497–1508 Springer 2006
DOI 10.1007/s10531-005-2356-1
Key words: Blattodea, Dermaptera, Habitat size, Road corridors, Herbivores, Marsh corridors,
Orthoptera
Abstract. We studied Orthoptera, Dermaptera, and Blattodea of the Białowie_za Forest (Poland) in
order to assess (1) the minimum patch size of open habitat necessary for each species, (2) the role of
linear corridors as habitat, and (3) the impact of herbivores on diversity by comparing the fauna at
periods of different ungulate densities. Many species occurred in the farthest clearings from the
forest edge to arable land. Two third of species occurred in clearings smaller than 10,000 m2. Dry
linear corridors of 10–40 m width and wet linear corridors of 100–200 m width had a species
richness that corresponded to that of clearings of about 10,000 m2. Four species disappeared from
the Białowie_za Forest when ungulate density decreased from 20 individuals/km2 (3000 kg/km2
biomass) at the beginning of the 20th century to 10 individuals/km2 (1000 kg/km2) at the end of the
20th century. We conclude that most Orthoptera, Dermaptera, and Blattodea species could survive
in Central Europe if human land use was replaced by intensive grazing and browsing by wild
herbivores.
Introduction
Closed forest has long been regarded as the natural vegetation of most parts
of Central Europe (Birks 2005). As a result, it is widely believed that there
was little space available to species associated with open landscape until
humans cleared the forest and created meadows and heathland. This image of
closed forest as the natural vegetation has been discussed over the last decade
(e.g. Svenning 2002). Vera (2000) argued that natural forests in the lowlands
of Central Europe were rather park-like landscapes that were shaped and
maintained by herbivores. This ‘‘wood-pasture’’ hypothesis is in opposition
to the ‘‘high-forest" hypothesis (Bradshaw et al. 2003), which has been
favoured throughout the 20th century and that considers closed forest as the
climax for Central Europe but does not acknowledge the major influence of
herbivores. Maybe because the concept of forest has long been that of a
closed canopy without larger clearings, forests are generally not considered
an important habitat for Orthoptera, which are considered indicators of
[437]
1498
The study area lies in the Polish lowlands on the border to Belarus and includes
the Polish side of the Białowie_za Forest (600 km2) and its surroundings (Figure
1). The Białowie_za Forest is a forest complex of 1450 km2 (5230¢–5300¢ N,
2330¢–2415¢ E) that straddles the Polish–Belarussian border. The forest is a
mosaic of deciduous, coniferous, and mixed tree stands where large ungulates
such as the European bison (Bison bonasus), moose (Alces alces), red deer
(Cervus elaphus), roe deer (Capreolus capreolus), and wild boar (Sus scrofa)
occur. The Polish side of the Białowie_za Forest consists of the Białowie_za
National Park and a commercial forest (480 km2), in which timber harvest,
[438]
1499
Figure 1. Sample sites (open circles) in the Białowie_za Forest (light grey) and the near sur-
roundings (sample sites at the Bug river not within the ranges of the figure) in 1997–2000. Open
land (white), strict reserve of the Białowie_za National Park (dark grey), rivers and lakes (black lines
and area), state border (dashed and dotted line).
reforestation, and hunting take place. Fifty km2 of the Białowie_za National
Park have been protected as a strict reserve since 1921. No hunting or forestry
is permitted in the strict reserve. The vegetation structure of the strict reserve
and of some places in the commercial forest is little influenced by humans.
Most of the study area is covered by closed forest (Figure 1). Open habitats
that occur within the forest are clearings of natural origin (usually few square
meters but sometimes up to several thousand square meters), young pine re-
growth areas (either plantations or natural, usually dry and sunny), sandpits
(mainly small-scale with bare sandy parts and older parts with vegetation),
mesophilic forest meadows (usually covered by high grasses), and sandy
meadows (mostly dry and with low vegetation). Linear corridors in the forest
consist in roads and railway lanes (10–40 m wide) and open marshes along
rivers, which are semi-natural as the use of meadows and reed in the marsh
almost completely stopped in the 1950s, so the marsh is now often covered by
reed and willow shrubs. The open land around the village Białowie_za is
[439]
1500
connected with the open land outside the Białowie_za Forest by a large river
marsh (Narewka) and a railway lane 30–40 m wide. The open land outside the
Białowie_za Forest is mainly non-intensive arable land, but also includes
sandpits, fallow land, dry and wet meadows.
Koźmiński (1925) recorded 32 species of Orthoptera, 1 species of Dermap-
tera and 1 species of Blattodea (see Table 1) in the Białowie_za Forest at a
period just after the total density of ungulates (including cattle) was of 20
individuals/km2 and of a crude biomass of 3000 kg/km2 (Je˛drzejewska et al.
1997). The forest still bears the signs of the almost medieval use that persisted
until recently (mid 20th century) as clearings were used for hay making, sand
and stone quarries were created throughout the forest, cattle grazed extensively
in the forest whilst some game species were protected. The impact of herbivores
was therefore much higher. Human use of the forest has changed, forestry
exploitation is now being promoted and ungulate densities are being kept low
by hunting. In addition, the European bison are being fed in winter, which
reduces their impact on forest re-growth, As a result, the total ungulate density
was around 10 individuals/km2 for a biomass of 1000 kg/km2 at the time of
this study (Je˛drzejewska et al. 1997).
On 150 days from April to October in 1997–1999 and in September 2000, we
recorded Orthoptera, Blattodea, and Dermaptera at 187 sites in the Białowie_za
Forest and its surroundings (Figure 1). The furthest 2 sites in the surroundings
were 50 km south-west from the Białowie_za Forest: dunes in the Kozki Nature
Reserve 5 km south of the town Siemiatycze and sandy meadows along the
Bug river east of the town Drohiczyn. As we were primarily interested in the
minimum patch size and the total numbers of species in each patch size class, a
standardised sampling effort or a complete species list of each site were not
necessary. However, as we recorded species everywhere where we encountered
them, the sample size for each patch size class was finally comparable con-
sidering that larger plots need a larger sample size (Table 1), except for the size
class from 0.1–1 km2. The small sample size of this size class was related to the
little diversity of habitats in these clearings (mainly fields and meadows for
cattle grazing). All size classes included sites in dry and wet habitats and from
all months to avoid any influence of habitat or season on the patch size
analyses. We visited most sites only once and searched for animals until we
could no longer find new species. However, if we expected species to exist in a
site but could not find them during our first visit, we usually returned at least
once to the site (287 samples on the 187 sites).
Because of earlier experience in identifying Orthoptera in the field, we were
able to identify species in the field acoustically (Bellmann 1985; Bellmann
2004) or by their morphology (Harz 1957) using magnifying glasses. Identi-
fication of Orthoptera using combinations of stridulation and morphological
characteristics are not only ethically preferable but also more reliable than
retrospective identifications of dead individuals. When we were not sure about
a field identification of Blattodea or Dermaptera, we collected the animals and
identified them later with a stereoscope and a key (Harz 1957). We used a
[440]
Table 1. Frequency of occurrence (in % of sites) of Orthoptera, Dermaptera, and Blattodea at 187 sites in 6 size classes of clearings in the Białowie_za Forest, on linear corridors
(roads of 10–40 m width, river marshes of 100–200 m width), in the glade of Białowie_za (106–107 m2) and in the open land within 50 km around the Białowie_za Forest ( > 108 m2).
The maximal distance to forest edge was measured from sites to the forest edge with agricultural land outside the Białowie_za Forest or the glade of Białowie_za (max. possible: 10 km).
Orthoptera
Podisma pedestris *
Stenobothrus lineatus* +
Psophus stridulus* +
Stenobothrus stigmaticus* 51
Aiolopus thalassinus 51
Chorthippus vagans 51
Omocestus rufipes 111
[441]
Meconema thalassinum 5 5
Decticus verrucivorus* 10 21
Tettigonia viridissima* 4 29 53 6.2
Gryllotalpa gryllotalpa* 11 9 19 5 2.0
Phaneroptera falcata 11 5 11 3.1
Conocephalus dorsalis* 22 9 33 19 11 8.7
Chorthippus montanus* 3 22 9 33 14 16 5.8
Stethophyma grossum* 4 22 13 50 14 16 4.6
Conocephalus discolor 4 10 4.1
Barbitistes constrictus* 4 5 2.8
Gryllus campestris* 7 3 17 17 5 16 6.4
Chrysochraon dispar* 7 9 44 9 17 43 11 8.7
Tetrix tenuicornis* 11 9 17 10 5 6.3
Chorthippus mollis 14 11 13 33 19 47 9.8
Metrioptera bicolor 14 13 14 84 9.8
1501
Omocestus haemorrhoidalis* 18 6 17 50 10 37 9.8
Euthystira brachyptera 7 9 11 9 9.8
Table 1. (Continued).
1502
Species Size of open habitat patch (m2) Max distance
2 2 3 3 4 4 5 5 6 6 7 8 to forest
0–10 10–10 10 –10 10 –10 Roads Marsh 10 –10 10 –10 10 –10 > 10
edge (km)
Number of sites 12 (12) 23 (23) 12 (12) 28 (39) 34 (43) 9 (18) 23 (44) 6 (7) 21 (55) 19 (34)
(number of samples)
*species found by Koźmiński (1925) in the Białowie_za Forest, +found by Bönsel and Runze (2000) 100 km north-west of the Białowie_za Forest, 1found only on dunes in the Kozki
Nature Reserve and on sandy meadows along the Bug river 50 km south-west of the Białowie_za Forest. Scientific names of Orthoptera follow Heller et al. (1998) and of Blattodea
and Dermaptera follow Harz (1957).
1503
Results
Discussion
[443]
1504
Figure 2. Best fit models (power regression) of mean numbers of species (regression line for single
sites and 95% confidence intervals for sites of each size class) and of total numbers of species (solid
regression line with closed circles for each size class, dotted regression line with open circles for
cumulative number of species) of clearings in the Białowie_za Forest and the open land around the
forest.
[444]
1505
in land use such as the disappearance of forest pasture and the abandonment of
heathland. In contrast, Liana (1981) found that dry habitats in forested areas
were richer in Orthoptera species than those in the open land. Five spe-
cies (Barbitistes constrictus, Chorthippus pullus, Euthystira brachyptera,
Gomphocerippus rufus, Metrioptera brachyptera) were even restricted to
forested areas. Psophus stridulus occurred in the Biebrza area (north-east
Poland) only on a dune surrounded by trees, which had a warmer micro
climate than the surroundings (Bönsel and Runze 2000). It is likely that
clearings in forested areas are better habitats for thermophilous species due to
thermal advantages. Dragonflies also exhibited a similar dependence to forest
in the Białowie_za Forest where the most thermophilous species reproduced
only in ponds of forest clearings but not in the surroundings (Theuerkauf and
Rouys 2001). Forests might therefore play an important role in maintaining
Orthoptera diversity in Central Europe.
Although forest can improve the habitat quality for thermophilous species, it
can also reduce the survival chances of species that depend on relatively large
open habitat patches. Bieringer and Zulka (2003) found that shading affects the
occurrence of Orthoptera up to 30 m from the forest edge. Clearings under
1000 m2 are therefore almost entirely affected by shading from the forest edge,
and indeed we found only about one third of species on these clearings. Two
third of species already occurred in the next size class (0.01–0.1 km2), which is
probably linked to a reduced shading effect. Those species that disappeared
from the Białowie_za Forest, however, probably needed a larger surface of
habitat. High density of cattle and wild ungulates might have provided the
necessary extension of habitat for these thermophilous Orthoptera species at
the beginning of the 20th century, but the current grazing intensity did not
allow these species to persist. We think that these species probably disappear in
Central Europe where human land use is discontinued and not taken over by
intensive ungulate grazing. In the Bieszczady National Park (Southeast
Poland), Psophus stridulus and Aiolopus thalassinus probably disappeared after
human land use (cattle grazing) was discontinued 60 years ago (Theuerkauf
et al. 2005).
Habitat corridors are known to be important for the dispersal of Orthoptera
(Collinge 1998; Berggren et al. 2002; Jordán et al. 2003). However, our study
indicated that linear corridors in forests are also an important habitat for
Orthoptera. Both corridors in river marshes and along forest roads had
numbers of species that corresponded to clearings of about 10,000 m2. Whilst
the number of species on the wet marsh corridors can be explained by their
width of 100–200 m, the number of species on the road corridors of 10–40 m
was much larger than might be expected by the patch size. Obviously, the
shading effect was not an important limitation in species numbers on these dry
corridors. From an evolutionary perspective, it is possible that species of drier
habitats must be able to exist on smaller habitat patches than those of wet
habitats. The reason might be that river marshes are open on a larger scale than
dry clearings due to regular flooding or activity of beavers (Castor fiber). Dry
[445]
1506
clearings on the other hand can be maintained by ungulates but first need to be
created (for example by wind), which occurs probably on a smaller area.
We conclude that Central Europe might loose a few Orthoptera species
(e.g. the first 7 species in Table 1) if human land use was replaced by grazing by
wild animals. However, under the current ecological situation with low her-
bivore densities and only a few species in most regions, it would take consid-
erable management measures to re-establish natural grazing communities.
Many species of Orthoptera and even some Dermaptera species would prob-
ably disappear from regions where human land use is discontinued and not
replaced by natural grazing. Rather than keeping forest clearings open for the
conservation of Orthoptera (Kati et al. 2003), we argue that the concept of
species conservation in open habitats needs to be reconsidered to include a full
array of herbivores at densities that might appear high but that were common
in historical times (Beutler 1996).
Acknowledgements
We thank C. Okołow, the director of the Białowie_za National Park, for per-
mission to work in the strict reserve, B. Jaroszewicz for providing us with
information during the study, P.-M. A. Dettinger-Klemm and two anonymous
reviewers for useful comments.
References
[446]
1507
Birks H.J.B. 2005. Mind the gap: how open were European primeval forests? Trends Ecol. Evol. 20:
154–156.
Bönsel A. and Runze M. 2000. Ein Habitat der Rotflügeligen Schnarrschrecke (Psophus stridulus
L., 1758) im nordöstlichen Polen [A habitat for Psophus stridulus L., 1758 in northeastern
Poland]. Articulata 15: 49–62. [In German with English abstract]
Bouget C. and Duelli P. 2004. The effects of windthrow on forest insect communities: a literature
review. Biol. Conserv. 118: 281–299.
Bradshaw R.H.W., Hannon G.E. and Lister A.M. 2003. A long-term perspective on ungulate-
vegetation interactions. Forest Ecol. Manage. 181: 267–280.
Clayton J.C. 2002. The effects of clearcutting and wildfire on grasshoppers and crickets (Orthop-
tera) in an intermountain forest ecosystem. J. Orthoptera Res. 11: 163–167.
Collinge S.K. 1998. Spatial arrangement of habitat patches and corridors: clues from ecological
field experiments. Landscape and Urban Planning 42: 157–168.
Firbank L.G., Telfer M.G., Eversham B.C. and Arnold H.R. 1994. The use of species-decline sta-
tistics to help target conservation policy for set-aside arable land. J. Environ. Manage. 42: 415–422.
Harz K. 1957. Die Geradflügler Mitteleuropas [The Orthoptera of Central Europe]. Gustav
Fischer, Jena. [In German]
Heller K.G., Korsunovskaya O., Ragge D.R., Vedenina V., Willemse F., Zhantiev R.D. and
Frantsevich L. 1998. Check-list of European Orthoptera. Articulata Beiheft 7: 1–61.
Ingrisch S. and Köhler G. 1998. Die Heuschrecken Mitteleuropas [The Orthoptera of Central
Europe]. Neue Brehm-Bücherei Bd. 629. Westarp Wissenschaften, Magdeburg Germany. [In
German]
Je˛drzejewska B., Je˛drzejewski W., Bunevich A., Miłkowski L. and Krasiński Z. 1997. Factors
shaping population densities and increase rates of ungulates in Białowie_za Primeval Forest
(Poland and Belarus) in the 19th and 20th centuries. Acta Theriologica 42: 399–451.
Jordán F, Báldi A, Orci K.-M., Rácz I. and Varga Z. 2003. Characterizing the importance of
habitat patches and corridors in maintaining the landscape connectivity of a Pholidoptera
transsylvanica (Orthoptera) metapopulation. Landscape Ecol. 18: 83–92.
Kati V., Dufrêne M., Legakis A., Grill A. and Lebrun P. 2003. Conservation management for
Orthoptera in the Dadia reserve, Greece. Biol. Conserv. 115: 33–44.
Koźmiński Z. 1925. Ökologische Untersuchungen an Orthopteren des Urwalds von Białowie_za
[Ecological studies of Orthoptera in the Białowie_za Forest], Bulletin de l’Académie Polonaise des
Sciences et des Lettres – Classe des Sciences Mathématique et Naturelles – Série B: Sciences.
Naturelles 1925: 447–475. [In German]
Laußmann H. 1993. Die Besiedlung neu entstandener Windwurfflächen durch Heuschrecken
[Colonisation of newly created wind-fall sites by Orthoptera]. Articulata 8: 53–59. [In German
with English abstract]
Liana A. 1981. Prostoskrzydłe (Orthoptera) w siedliskach kserotermicznych Pojezierza Mazur-
skiego [Orthoptera in xerothermic habitats of the Mazurian lakeland]. Fragmenta Faunistica 25:
479–510. [In Polish with Russian and French abstracts]
Liana A. 1992. Owady prostoskrzydłe Orthoptera [Orthopteroid insects]. In: Głowaciński Z. (ed),
Czerwona lista zwierza˛t gina˛cych i zagro_zonych w Polsce [Polish Red List of Threatened Ani-
mals]. Polish Academy of Sciences, Nature Protection Research Centre, Kraków, pp. 85–91. [In
Polish with English abstract]
Liana A. 2001. Orthoptera-Blattodea. In: Gutowski, J.M. and Jaroszewicz B. (eds), Katalog Fauny
Puszczy Białowieskiej. [Faunal catalogue of the Białowieza Forest]. Instytut Badawczy
Leśnictwa, Warszawa, pp. 92–93.
Shure D.J., Phillips D.L. 1991. Patch size of forest openings and arthropod populations. Oecologia
86: 325–334.
Svenning J.-C. 2002. A review of natural vegetation openness in north-western Europe. Biol.
Conserv. 104: 133–148.
Theuerkauf J., Rouys S. 2001. Habitats of Odonata in the Białowie_za Forest and its surroundings
(Poland). Fragmenta Faunistica 44: 33–39.
[447]
1508
Theuerkauf J., Rouys S. Grein G. Becker A. 2005. New records of Orthoptera in the Bieszczady
Mountains (Southeast Poland) with special regard to the genus Isophya. Fragmenta Faunistica
48: 9–14.
Vera F.W.M. 2000. Grazing Ecology and Forest History. CABI Publishing, Wallingford.
[448]
Biodiversity and Conservation (2006) 15:1509–1527 Springer 2006
DOI 10.1007/s10531-005-2632-0
-1
Abstract. The floristic composition and diversity of tropical dry deciduous and gallery forests were
studied in Chacocente Wildlife Refuge, located on the Pacific coast in Nicaragua during 1994 and
2000. Density, dominance and frequency as well as species and family important values were
computed to characterize the floristic composition. A variety of diversity measures were also
calculated to examine heterogeneity in each forest community. A total of 29 families, 49 genera and
59 species were represented in 2 ha dry deciduous forest. In the gallery forest, the number of
families, genera and species recorded in 2000 inventory was 33, 48 and 58, respectively and slightly
higher than the 1994 inventory. The number of stems ‡10 cm dbh varied from 451 to 489 per
hectare in the deciduous forest, and from 283 to 298 per hectare in the gallery forest. The basal area
was much larger for species in the gallery than dry deciduous forest. Fabaceae, sub family
Papilionoideae, was the most specious family in the deciduous forest while Meliaceae was the
dominant family in the gallery forest. Similarity in species composition and abundance between
deciduous and gallery forests was low. In terms of species diversity, the gallery forest was found
more diverse than the deciduous forest using Fisher’s diversity index. Both forest communities were
characterized by a typical inverse J shape. Therefore, emphasis should be given to the protection of
rare species, i.e. as the forests are still under continued human pressure, an immediate action should
be taken to conserve the remaining flora.
Introduction
Dry forests once covered more than 40% of the total area of tropical forests
(Murphy and Lugo 1986). They are considered to be one of the most threa-
tened of all the major tropical forest habitats and are argued to deserve a high
priority for conservation (Janzen 1988; Gillespie et al. 2000). According to the
Holdridge system of life zone classification, dry tropical and subtropical forests
and woodlands occur in frost-free areas with a mean annual temperature
[449]
1510
higher than 17 C, a mean annual rainfall between 250 and 2000 mm, and an
annual ratio of potential evapotranspiration to precipitation exceeding unity
(Murphy and Lugo 1995).
The area of natural forests in Central America is estimated to be 190,000 km2,
representing ca. 15% of the total land cover. In addition, some 130,000 km2
deforested land is considered suitable only for forestry, adding to 24% of the total
area (Segura et al. 1997). The deforestation rate in Central America is estimated
as 0.5 km2 per year (Roldan 2001). The tendency of human populations to
concentrate in drier climates is hastening the rate of dry forest degradation
(Murphy and Lugo 1995) and deforestation has increased dramatically with
population growth during the last century. Large areas are cleared for grazing
and agriculture and only fragments of dry forests remain (Gerhardt 1994).
Nicaragua has 2500 km2 of tropical dry forests, representing ca. 2% of the
total forest cover (Harcourt and Sayer 1996). The dry forests are found mainly
on the Pacific coast where ca. 50% of the population also lives. Nicaraguan dry
forests have been intensively exploited for commercial timber production. The
major commercial timber species are Swietenia humilis, Cedrela odorata, Bom-
bacopsis quinata, Dalbergia retusa and Guaiacum sanctum (Sabogal 1992). The
extraction of valuable commercial trees for export started in early 1900 (Tercero
and Urrutia 1994), and continued for decades, resulting in considerable
reduction of commercially important species. The Nicaraguan Pacific railway
was constructed in the 1950s and most of the railway sleepers used was extracted
from the dry forest in Chacocente (Tercero and Urrutia 1994). The dry forests
are still major sources of wood for fire, poles and timber, and provide oppor-
tunities for hunting and collection of other important non-timber forests
products (NTFP). In addition to cutting of trees for wood and related products,
the major causes of deforestation have been conversion of dry forests into coffee
plantations, crop fields and ranches (Roldan 2001).
Chacocente National Wildlife Refuge was established in 1983 to protect the
nesting beach of marine turtles and the last area of the tropical dry forest due to
the social, economic, ecological and scientific relevance of this type of ecosys-
tem. During the Sandinista Revolution big ranches were expropriated and be-
came property of the state. In 1990 this land was given to peasant cooperatives.
By 1998, the land tenure changed very rapidly since land was being sold and
cooperative land was converted into private land (Anonymous 2002). Today,
the Chacocente National Wildlife Refuge consists mainly of private farms (84
owners), although some are quite small. The only state land in the refuge is a
small property donated to Ministry of Natural Resources and Environment
(MARENA) by the International Fauna and Flora Organization. The refuge is
not fully protected against human impact and is utilized both legally and ille-
gally by the local people living inside as well as outside the refuge. Anthropo-
genic disturbances such as burning, grazing, wood collection and illegal cutting
are factors affecting plant population density (Gillespie et al. 2000).
The effect of this land use dynamics and forest fragmentation on biological
diversity in Chacocente is not well documented (Sabogal and Valerio 1998). An
[450]
1511
Study area
Figure 1. Distribution of tropical dry forests in Nicaragua and location of the study site,
Chacocente.
[451]
1512
(1962 mm) and the minimum in 1991 (991 mm). During October 1998,
hurricane Mitch passed over the area and the precipitation that month was as
high as 775 mm. The average annual temperature is 26 C (Anonymous 2002).
The gallery forest, defined as narrow patches along the fringes of semi-
permanent watercourses (Lamprecht 1989), occurs along the main water
course, the Rı́o Escalante. It has a different species composition, structure and
stand density than the more common deciduous forest. The vegetation is
mostly evergreen, the trees are tall and the majority of the trees have a diameter
exceeding 35 cm at breast height. The deciduous forest trees totally or partially
shed their leaves during the dry season.
[452]
1513
Results
Floristic composition
A total of 29 families, 49 genera and 59 species were found in the dry deciduous
forest during both inventories (Table 1). While the stem density slightly in-
creased in 2000 inventory, the basal area was relatively less compared to the
inventory made in 1994. Fabaceae, sub-family Papilionoideae was the most
specious family with higher FIV (Table 2). Other families (sub-families) with
‡4 species were Caesalpinioideae and Boraginaceae. Hernandiaceae, though
represented by one species (Gyrocarpus americanus), had the second and third
higher FIV in 1994 and 2000 inventories, respectively owing to the large stem
density per hectare (62 individuals/ha in 1994 and 37 individuals/ha in 2000).
Gyrocarpus americanus stood out as the most abundant species during both
inventories in terms of basal area, relative dominance, relative frequency and
IVI (Table 3). While Tabebuia ochracea was the second most abundant species
during both inventories, Lonchocarpus minimiflorus and Myrospermum
frutescens were more abundant in 1994 and 2000 inventories, respectively. The
rarest species during both inventories were Celtis caudata and Zanthoxylum
caribaeum (Table 4). Four species, Acacia costaricensis, Ficus obtusifolia,
Pithecellobium saman and Trichilia hirta, recorded in 1994 inventory were not
encountered in 2000, but four other species, Adelia barbinervis, Casearia
[453]
1514
Table 1. Summary of floristic composition and structure of trees ‡10 cm dbh in dry deciduous and
gallery forests inventoried in 1994 and 2000.
Forest types-Inventory time Families Genera Species Stem densitya Basal areab
Table 2. The ten most important families (sub-families) in the dry deciduous and gallery forests of
Chacocente in 2000 inventory according to decreasing order of family importance value (FIV).
corymbosa, Cordia dentata and Trema micrantha, were found in 2000 inventory
(Appendix).
In the gallery forest, the number of families, genera and species encountered
in 1994 inventory were 33, 47 and 55, respectively while 48 genera and 58 species
were recorded in 2000 inventory (Table 1). The total stem density (298 indi-
viduals/ha) was comparatively higher in 1994 inventory than in 2000 inventory
while the basal area was relatively larger in 2000 than in 1994 inventory. In both
inventories, Meliaceae was the most specious family with higher IFV (Table 2).
Most of the important families were represented by 1 or 2 species. The most
abundant species, in terms of basal area, relative dominance, and IVI was
Pithecellobium saman, followed by Trichilia hirta (Table 5). In both inventories,
Cordia alliodora was the rarest species, followed by Hymenaea courbaril in the
[454]
1515
Table 3. The ten most abundant species in the dry deciduous forest of Chacocente in 1994 and
2000 inventories according to decreasing order of importance value index (IVI) together with
structural characteristics.
2000 Inventory
Gyrocarpus americanus 3.548 12.23 9.41 8.28 29.9
Tabebuia ochracea 2.1671 7.47 10.04 7.63 25.1
Myrospermum frutescens 1.4233 4.91 6.52 6.10 17.5
Caesalpinia exostemma 1.8082 6.23 7.28 3.92 17.4
Stemmadenia obovata 1.0028 3.46 7.53 5.88 16.9
Achatocarpus nigricans 2.0956 7.23 5.90 3.05 16.2
Lonchocarpus minimiflorus 0.9979 3.44 7.90 4.79 16.1
Gliricidia sepium 1.8574 6.40 2.38 2.83 11.6
Luehea candida 0.6862 2.37 3.14 3.92 9.4
Allophylus psilospermus 0.5432 1.87 2.76 2.40 7.0
1994 inventory, and Triplaris melaenodendronin the 2000 inventory (Table 6).
Four species, Casearia tremula, Pithecellobium dulce, Piper aduncum and Randia
nicaraguensis recorded in 1994 inventory were missing in 2000 inventory while
seven other species were encountered in 2000 inventory; namely, Acacia cos-
taricensis, Bursera simaruba, Caesalpinia exostemma, Caesalpinia violacea,
Coccoloba sp., Licania arborea and Tabebuia rosea (Appendix).
As a whole, the number of species recorded in dry deciduous and gallery
forests was nearly the same. However, the stem density in the dry deciduous
forest was twice higher than the gallery forest while the basal area was much
bigger in the latter. It was found that the similarity in species composition
between the two forest communities in both inventories was very low, as shown
by low Jaccard’s (0.27) and Morisita’s (0.35) similarity indices.
Species diversity
[455]
1516
Table 4. The ten rarest species in the dry deciduous forest of Chacocente in 1994 and 2000
inventories according to increasing order of IVI together with structural characteristics.
[456]
1517
Table 5. The ten most abundant species in the gallery forest of Chacocente in 1994 and 2000
inventories according to decreasing order of importance value index together with structural
characteristics.
2000 Inventory
Pithecellobium saman 12.7787 25.89 1.22 2.07 29.2
Trichilia hirta 2.4232 4.91 15.29 8.55 28.8
Capparis pachaca 1.3614 2.76 13.15 7.51 23.4
Thouinidium decandrum 2.9474 5.97 9.63 6.74 22.3
Gyrocarpus americanus 4.6233 9.37 4.43 3.63 17.4
Simarouba glauca 2.01 4.08 4.89 8.29 17.3
Annona reticulata 1.3844 2.80 4.13 6.74 13.7
Astronium graveolens 2.0686 4.19 3.21 5.44 12.8
Stemmadenia obovata 0.8589 1.74 6.57 4.15 12.5
Guarea glabra 1.2341 2.50 5.35 4.40 12.3
evenness into a single value, declined over time within each forest community,
and identified the dry deciduous forest as more diverse than the gallery forest.
The complement of Simpson’s index, which attaches more weight to the
abundance of the most common species, also identified the dry deciduous
forest as more diverse than the gallery forest. Fisher’s diversity index, the most
widely recommended measure of diversity, revealed that the galley forest is
more diverse than the dry deciduous forest.
Local uses
Although we did not make a systematic ethno-botanical study, the local uses
of the tree species in both forest communities were identified based on
information gathered from the local people, expert consultation and existing
literature. Accordingly, ten major use categories were identified (Figure 3). It
was found that the largest number of species in both forest communities (53%
of the total species) was used for firewood, followed by timber extraction
(35%), rural construction (27%) and charcoal production (23%). Interest-
[457]
1518
Table 6. The ten rarest species in the gallery forest of Chacocente in 1994 and 2000 inventories
according to increasing order of IVI together with structural characteristics.
ingly, 24% of the species are not currently under any kind of use. The
abundance of species by use group was also examined for each forest com-
munity (Figure 3). Given the large number of species used for firewood, the
overall abundance was also high. The most interesting part of this result is
that the abundance of species used for firewood and timber declined from
1994 to 2000 in both forest communities. Although the abundance of the ‘‘not
used’’ species in the gallery forest showed an increasing tendency, the reverse
held true in the deciduous forest.
Most of the species in our plots were represented by few individuals (Figure 2).
Some of the rarest species were already short-listed in IUCN red list directory
as threatened species. Among these threatened species, five species were cate-
gorized as vulnerable and six species as endangered (Table 8). Bombacopsis
quinata, considered as vulnerable, was not encountered in our plots. We also
identified seven candidate species that could be included in IUCN red list
directory in the future (Table 8).
[458]
1519
125 125
Deciduous-94 Deciduous-00
100 100
Abundance
Abundance
75 75
50 50
25 25
0 0
0 10 20 30 40 50 0 10 20 30 40 50
Species rank Species rank
100 100
Gallery-94 Gallery-00
75 75
Abundance
Abundance
50 50
25 25
0 0
0 10 20 30 40 0 10 20 30 40
Species rank Species rank
Figure 2. Species abundance plots for dry deciduous and gallery forests inventoried in 1994 and
2000.
Discussion
The number of families, genera and species reported in the present study lies
within the range reported earlier in most Neotropical dry forests. For example,
Gentry (1988) reported 35 families and 55 species per hectare in a gallery forest
in Guanacaste, Costa Rica and Sabogal and Valerio (1998) reported on
average 44 species per hectare in Chacocente dry deciduous forest in Nicara-
gua. The most common family in the deciduous forest was Fabaceae/Papilio-
noideae with 10 species, a pattern common in most Neotropical dry forests
(Gentry 1988). This result also coincides with a study carried out in Central
America where Fabaceae was found to be the dominant tree and shrub family
in six of seven sites studied (Gillespie et al. 2000). Gillespie et al. (2000) made
an inventory in Chacocente and found the same common species as in the
present study. However, the present study found L. minimiflorus and C. exo-
stemma as common species. In tropical dry forest across the north central
Yucatan, the following important natural forest species were reported: Bursera
simaruba, Caesalpinia gaumeri, Gymnopodium floribundum and Piscidia piscip-
ula (White and Hood 2004), which are also encountered in our study. It was
observed that some species recorded in the first inventory (1994) were missing
in the subsequent inventory (2000) while new species were encountered in the
second inventory. Given the large number of species with 1 or 2 individuals,
[459]
1520
unknown
carvings
ornamental
timber
fruits
construction
fodder
firewood
live fence
Charcoal
0 10 20 30 40 50
No. species
1994 unknown
2000 carvings
ornamental
timber
fruits
construction
fodder
firewood
live fence
Deciduous Charcoal
1994 unknown
2000 carvings
ornamental
timber
fruits
construction
fodder
firewood
live fence
Gallery Charcoal
[460]
1521
Table 7. Diversity measures for trees ‡10 cm dbh in the dry deciduous and gallery forests
inventoried in 1994 and 2000 on 2 ha plots.
Table 8. List of threatened species and suggested candidate species for future IUCN listing.
Species Status
illegal cutting might have caused the disappearance of this species. However,
the plausible explanation for the appearance of species in the second inventory
could be ascribed to the transition from seedling class in the 1994 inventory to
higher class (trees ‡ 10 cm dbh) in the subsequent inventory.
Tree species richness is difficult to compare for different sample sizes and
geographical variation (Murphy and Lugo 1995). Dry forest at Palo Verde on
the Pacific side of Costa Rica had approximately 52 tree species per hectare
(Murphy and Lugo 1995). Lower values have been found in the drier areas and
particularly in insular forests, such as in Southwestern Puerto Rico near
Guanica where 30–50 tree species per hectare were found. Gentry (1995) re-
ported an average of 65 tree species per ha in 23 Neotropical dry forests, which
[461]
1522
is considerably higher than the number of species found in the deciduous forest
of Chacocente. However, Gentry’s data set included individuals with dbh
‡2.5 cm. In addition, dry forests of Chacocente have a history of severe
selective logging which may be the main factor causing the low number of
species in this forest. Gillespie and Jaffré (2003) compared species richness in
seven different countries using 1000 m2 area and found that species richness is
high in Chamela-Mexico (89), Quiapaca (86) and Chaquimayo-Bolivia (79).
The lower species richness was found in Mudumalai (India) with 15 species.
The number of species for trees ‡10 cm dbh ranged from 3 to 28 species with a
mean value of 16 species per hectare in the Vindhyan dry tropical forest of
India (Sagar and Singh 2005). As a whole, the total number of species recorded
in the present is comparable with other tropical dry forests.
With regard to stem density and basal area, our result lies with the range of
values reported earlier for other tropical dry forests, and in some cases com-
parably higher. For example, Sabogal and Valerio (1998) reported 389 trees/ha
with a basal area of 14.48 m2/ha in the Chacocente dry deciduous forest,
Rundel and Boonpragob (1995) reported 20–88 trees/ha and a basal area
ranging from 7 to 42 m2/ha for tropical dry forest in Thailand. For tropical dry
forest at the north central Yucatan, White and Hood (2004) documented the
basal area in two sites as 20.7 m2/ha and 28.4 m2/ha. Gillespie and Jaffré (2003)
inventoried two tropical dry forests of New Caledonia and found the following
basal area per hectare for each site: Ouen-Toro 32.7 m2/ha and Pindai 32.3 m2/
ha. Gillespie and Jaffré (2003) also pointed out that tropical dry forests in the
Neotropics have greater structural similarity. In the present study, the similarity
in species composition and abundance between dry deciduous and gallery for-
ests was low. The stem density was much higher in the dry deciduous forest
while the basal area was much greater for the gallery forest. This could be
related to better soil moisture condition in the latter than the former, as
moisture is the major environmental factor limiting tree growth in dry areas.
A variety of diversity measures were computed to describe the heterogeneity
of the two forest communities, and it was Simpson’s index and Fisher’s a that
consistently differentiated the two communities. According to Simpson’s
dominance index the dry deciduous forest is more diverse than the gallery
forest. This could be related to the relatively large number of abundant species
in the deciduous forest than the gallery forest (cf. 7 species in dry deciduous
and 4 species in gallery forests with abundance >24 individuals/ha). In 1994
and 2000 inventories of the dry deciduous forest, Gyrocarpus americanus,
Lonchocarpus minimiflorus, Stemmadenia obovata, Tabebuia ochracea and
Caesalpinia exostemma represented 49% and 42% of the total individuals per
hectare, respectively. While Trichilia hirta, Capparis pachaca, Thouinidium
decandrum and Stemmadenia obovata represented 44% and 45% of the total
individuals per hectare found in the gallery forest during 1994 and 2000
inventories, respectively.
Fisher’s diversity index showed that the gallery forest is more diverse than the
dry deciduous forest. One implication of this finding would be the majority of the
[462]
1523
Acknowledgements
We thank Claudio Calero for his support during the fieldwork. Also he made a
great contribution with his knowledge about species uses. Thanks also to Ali
Water and Alvaro Noguera for their help during the fieldwork. The Ministry of
Natural Resources and Environment kindly allowed us to carry out this
research at Chacocente National Wildlife Refuge. The study was financed by
[463]
1524
Appendix
Appendix. List of tree species ‡10 cm dbh recorded in 1994 and 2000 inventories in dry deciduous
and gallery forests in Chacocente Wildlife Refuge, Nicaragua together with their uses (CH –
charcoal; LF – live fence; FW – fire wood; FOD – fodder; RC – rural construction; FRU – fruit; W
– timber; P – pole; O – ornamental; HC – handicrafts and carvings).
[464]
1525
Appendix. Continued
[465]
1526
Appendix. Continued
References
Anonymous 2002. Plan de manejo del refugio de vida silvestre Rı́o Escalante- Chacocente. Min-
isterio de Recursos Naturales y del Ambiente. Ramboll/Posaf. Managua, Nicaragua.
Duque A. and Cavelier J. 2003. Strategies of tree occupation at a local scale in terra firme forests in
the Colombian Amazon. Biotropica 35: 20–27.
Gentry A.H. 1988. Changes in plant community diversity and floristic composition on environ-
mental and geographical gradients. Ann. Missouri Bot. Garden 75: 1–34.
Gentry A.H. 1995. Diversity and floristic composition of Neotropical dry forest. In: Bullock S.H.,
Mooney H.A. and Medina E (eds), Seasonally Dry Tropical Forests. Cambridge University
Press, Cambridge, UK, pp. 9–34.
Gerhardt K. 1994. Seedling Development of Four Tree Species in Secondary Tropical Dry Forest
in Guanacaste, Costa Rica. PhD thesis, Uppsala, Sweden.
Gillespie T.W., Grijalva A. and Farris C.N. 2000. Diversity, composition, and structure of tropical
dry forests in Central America. Plant Ecol. 147: 37–47.
Gillespie T.W. and Jaffré T. 2003. Tropical dry forests in New Caledonia. Biodiv. Conserv. 12:
1687–1697.
Harcourt C.S. and Sayer J.A. 1996. The Conservation Atlas of Tropical Forest. The Americas,
Simon & Schuster.
IUCN. 2004. 2004 IUCN Red List of Threatened Species. www.iucnredlist.org. Downloaded on 14
December 2004. International Union for Conservation of Nature and Natural Resources,
Cambridge, UK.
Janzen D. 1988. Management of habitat fragments in a tropical dry forest: growth. Ann. Missouri
Bot. Garden 75: 105–116.
Krebs C.J. 1999. Ecology Methodology. 2nd ed. Addison-Wesley Educational Publishers. Inc.
Lamprecht H. 1989. Silviculture in the Tropics: Tropical Forest Ecosystems and their Tree Species.
GTZ, Eschborn.
Magurran A.E. 2004. Measuring Biological Diversity. Blackwell Publishing, Malden, Oxford and
Victoria.
Murphy P.G. and Lugo A.E. 1986. Ecology of tropical dry forest. Ann. Rev. Ecol. Syst. 17: 67–88.
[466]
1527
Murphy P.G and Lugo A.E. 1995. Dry forest of Central America and the Caribbean. In: Bullock
S.H., Mooney H.A and Medina E (eds), Seasonally Dry Tropical Forests. Cambridge University
Press, Cambridge, England, pp. 9–34.
Roldan H. 2001. Recursos forestales y cambio en el uso de la tierra, Republica de Nicaragua.
Santiago, Chile.
Rundel P.W. and Boonpragob K. 1995. Dry forest of Central America and the Caribbean. In:
Bullock S.H., Mooney H.A. and Medina E (eds), Seasonally Dry Tropical Forests. Cambridge
University Press, Cambridge, England, pp. 93–119.
Sabogal C. 1992. Regeneración de bosques secos naturales en Centro América, con ejemplos de
Nicaragua. J. Veget. Sci. 3: 407–416.
Sabogal C. and Valerio L. 1998. Forest Composition, Structure and Regeneration in a dry forest of
the Nicaraguan Pacific Coast. In: Dallmeier F and Comiskey J.A. (eds), Forest Biodiversity in
North Central and South America, and the Caribbean: Research and Monitoring. Man and The
Biosphere Series, Vol. 21. UNESCO, New York pp.187–212.
Sagar R. and Singh J.S. 2005. Structure, diversity, and regeneration of tropical dry deciduous forest
of northern India. Biodiv. Conserv. 14: 935–959.
Salas J.B. 1993. Arboles de Nicaragua. Instituto Nicaragüense de Recursos Naturales y del Am-
biente. Servicio Forestal Nacional., Managua, Nicaragua.
Segura O., Kaimowitz D. and Rodrı́guez J. 1997. Polı́ticas forestales en Centro América: Análisis
de las restricciones para el desarrollo del sector forestal. IICA-Holanda/LADERAS.
Stevens W.D., Ulloa C.U., Pool A. and Montiel O.M. 2001. Flora de Nicaragua. Missouri
Botanical Garden Press.
Tercero M.G. and Urrutia G.S. 1994. Caracterización florı́stica y estructural del bosque de galerı́a
en Chacocente, Carazo, Nicaragua. Trabajo de diploma. Universidad Nacional Agraria,
Managua, Nicaragua.
Thiollay J.M. 2002. Forest ecosystems: threats, sustainable use and biodiversity conservation.
Biodiv. Conserv. 11: 943–946.
White D.A. and Hood C.S. 2004. Vegetation patterns and environmental gradients in tropical dry
forests of the northern Yucatán Peninsula. J. Veget. Sci. 15: 151–160.
[467]
Biodiversity and Conservation (2006) 15:1529–1543 Springer 2006
DOI 10.1007/s10531-004-6678-1
Key words: Conservation priority zones, Eastern Ghats, Endemism, Geographical information
system, Red listed plants, Shervarayan hills, Tropical forest
Abstract. There are thousands of protected forest areas existing on earth, yet the deforestation rate
continues unabated both inside and outside the protected areas especially in the tropical forests. It
identifies the less effectiveness of the current conservation strategies, which is normally oriented
around the forest area cover rather than the quality of the protected areas. This calls for realistic
and effective management strategies for forests. Based on the drawbacks the present study aims at
identifying conservation priority sites within the protected areas (Reserved Forests) of Shervarayan
hills, Eastern Ghats of Tamil Nadu, India. The remnant forest patches having less effective man-
agement/protection is identified and analysed for its qualitative contribution to the ecosystem.
Quadrats of 20 · 20 m were laid in different vegetation based on the percentage of forest cover and
assess the species diversity pertaining the richness, Endemism and Red list categories. Thematic
layers (maps) such as vegetation type, floristic species richness, floristic endemism, and red list flora
are created and categorized according to their weightage classes and overlaid in GIS domain to
demarcate the Conservation Priority Zones (CPZ). The CPZ are classified according to the priority
status i.e., high, moderate and low based on the contributing species richness, levels of endemism
and concentration of Red listed plants.
Introduction
[469]
1530
conservation policy (Robertson and Hull 2001). Within the realm of conser-
vation measures, the forest strategists have identified and conserved large tracts
of vegetation as protected areas (Gaston et al. 2002; Margules et al. 2002). Still
the deforestation rate has markedly increased (Downton 1995) and has spread
to the protected areas of tropical region too (Hamilton 1984; Howard 1991;
Redford 1992) rendering ineffectiveness in arresting it. Ecologists nowadays are
on the consensus that biological diversity is not effectively conserved by re-
serves alone (Wilcove 1989). Various quantitative methods that allow relatively
expeditious identification of conservation-priority areas have been proposed in
recent years and these approaches include identification of hotspots of biodi-
versity (Myers 1988, 1990; Dobson et al. 1997), rapid biodiversity assessment
(Oliver and Beattie, 1993 and 1996), identification of indicator and surrogate
species (Curnutt et al. 1994), development or rarity and complementary sets
(Williams et al. 1996), identification of key eco-region (Olson and Dinerstein
1998), and cost-minimizing or land-values analyses (Ando et al. 1998).
This may be due to the very size of the forest tract whereby the porosity of
the protected forest will lead to its ineffective management. Now, it is better to
identify the quality of the vegetation in the protected and non protected areas,
rather than the area size for effective conservation management (Sheil 2001).
Most often we had adopted the conventional approach to maintain biological
diversity by following a protocol based on species by species and threat-by-
threat approach, but it too has its own detriments i.e., the financial drawbacks,
inaccurate complicated database of the forest community (Hutto et al. 1987;
Scott et al. 1987, 1991; Margules 1989, Noss 1991) etc.
In recent years the focus for conservation has shifted from single species
management approach to protection of biodiversity in the aggregate i.e., to
maintain the native plant species in extensive natural landscapes (habitats)
restricting to a minimal size factor, that are sufficiently linked (i.e. corridors) to
allow interaction and genetic interchange among distinct populations (Noss
1983). This approach requires a cohesive and representative system of areas to
be managed for the maintenance of biodiversity. Hence there is a need to
prioritize only those areas, which are considered most essential for conserva-
tion, which are termed as biodiversity priority areas (Olson and Dinerstein
1998). The procedures involve scoring and ranking, which make priority setting
more systematic and explicit (Margules et al. 2002). Prioritization of strategies
is essential to ensure that efforts at conservation yield best possible results and
undesirable side effects, such as the alienation and impoverishment of local
communities can be avoided (Singh and Taneja 2000). Prioritization of sites for
conservation also needs to be done with reference to the (often least studied)
vegetation type (Williams et al. 2002), species richness (Terborgh and Winter
1983; Scott et al. 1987), endemism based on Kier and Barthlott (2001) and
concentration of red listed plants (Ahmedullah 2000; Kumar et al. 2000).
The methods for identifying priority areas vary with the entity selected for
the overall biological conservation planning and management (Margules et al.
2002; Gaston et al. 2002) and for example Ramesh et al. (1997) have suggested
[470]
1531
Study area
The Shervarayan hills (a part of Eastern Ghats) are located in the northern
part of Salem city, Tamil Nadu, South India and with an area of 469.9 km2.
The study area lies between latitudes 1143¢00¢¢ to 1200¢00¢¢ N and longitudes
of 7800¢00¢¢ to 7822¢30¢¢ E (Map 1) and falls in the Survey of India toposheets
(SOI) 581/1, 2, 5 and 6 (i.e., 1: 50,000 scales). The mean annual rainfall at the
upper hill reaches is 1638 mm and 850 mm at the foothills. The temperature
[471]
1532
ranges from 13–29C on the hill plateau to 25C and 40C at the foothills. The
soil is red loamy and lateritic. The area is made up of Archaean crystalline rock
like amphibolites, leptinites, garnetiferous granites and charnockites. Bauxite
and Magnesite are the chief mineral resources in the Shervarayan hills. There
are 71 villages, which are administrated by two taluks (political unit equivalent
of an English county) i.e., Yercaud and Omalur. Most of the hill plateau is in
private ownership, which includes coffee estates, villages and their agricultural
lands. Colonial planters had been maintaining and harvesting the coffee estates
till the time of independence of the country and later, the ownership has been
entrusted to the natives. There are 45 reserved forests, which are administered
by the Salem Forest Division. Almost all the reserved forest area is on the outer
slopes of the hill tract facing the human habitats on the fringing foothills
thereby enhancing the proneness to deforestation and very much is the evident
fact.
Methodology
Vegetation type map of Shervarayan hills (Balaguru et al. 2003) is used which
covers nearly half (49.50%) of the hill area (23260.76 ha) under reserved forests
comprising about six major forest types - evergreen (111.33 ha), semi evergreen
(1057.67 ha), riparian (1145.15 ha), dry mixed deciduous (10179.10 ha),
southern thorn scrub (10735.70 ha), and evergreen scrub (31.81 ha), respec-
tively (Map 1). To evolve potential conservation priority elements, the virgin
and primary forest patches comprising the evergreen, semi evergreen, riparian,
and dry mixed deciduous forests are used as the base, while the evergreen scrub
and southern thorn scrub forests are excluded due to their highly degraded
nature. The scores for each forest type are attributed according to the species
concentration (Figure 1) and substituted to all the representing polygons
accordingly.
Representative polygons for each forest type are analysed for assessing species
richness contribution by adopting quadrat method (20 · 20 m) (CES 1998;
Ferreira and Prance 1998). This study has taken optimum sampling quadrats
to cover all variations within each type of the vegetation and the number of
quadrats for each forest type is based on the area percentage of the forest cover
(>1000 ha area 0.5%; 1000–2000 (0.5%) and >2000 (0.01%). All living plant
species within the quadrat are identified and the number of species in each
forest type is summed and represented by species richness values and these
[472]
1533
The plant species thus collected in the quadrats are identified with the endemic
flora of Peninsular India as enlisted by Ahmedullah and Nayer (1986). Pro-
cedures for deciding on CPZ need more systematic and explicit approach for
priority setting wherein multiple criteria are given scores. These scores are then
combined and ranked accordingly and priority (high, moderate or low) is given
to those areas (Margules et al. 2002). The number of endemic species are
allocated to the respective scores/classes based on their significant status in
[473]
1534
Indian context (Roy 1999; Ajith Kumar et al. 2000) i.e. individual species
endemic to India is considered as ‘Indian Endemic’ in distribution, hence they
received low score (1), similarly individual species endemic to peninsular India
is considered as ‘Regional Endemic’ (2) and species endemic to Eastern Ghats
(including Shervarayan hills) is ‘Local Endemic’ and received the highest score
(3). The number of species and their scores in each of the polygon is then
summed up and values attributed as described for species richness. To produce
endemic species map, the polygons are finally regrouped/reclassified into low,
medium and high degree of endemism according to the summed values
attributed to each polygon (see Figure 1), the polygon with the highest score
had the high degree of endemism and likewise.
The methodology to map red listed plants is the same as for the endemic plants
map or species richness map. The ensuing plant species in the quadrats are
evaluated based on version 3.1; IUCN/SSC (1999) criteria and cross checked
with Indian Red Data books (Nayer and Sastry 1987–1990) and other relevant
literature Kumaravelu and Chaudhuri 1999). The red listed categories and their
scores are classified into (a) Critically Endangered (CE) 5; (b) Endangered
(En) 4; (c) Vulnerable (VU) 3; (d) Lower risk (LR)/Least Concern (Lc) 2;
(e) Data Deficient (DD) 1 (Table 2). To produce the red listed plant species
map, the polygons are finally grouped into low, medium and high wherein the
scoring is similar to the one adopted for the endemic classification.
The components of various units (classes) from the thematic maps like the
vegetation type, floristic richness, endemism and red lists with their respective
weightages (Figure 1) are essential to develop conservation priority zones for
this study. Considering the conservational importance and status for each class
(unit) of the respective thematic maps, the classes are given weightages to
designate and identify the CPZ. Overlay or superimposition creates a com-
posite output GIS file by combining a number of input GIS files based on the
minimum or maximum values of the input files (Murthy 2000). To prepare the
CPZ map, the respective thematic maps (species richness map, red list map and
finally endemism map) are overlaid on the vegetation type map, which com-
prised the lowermost tier (the base map) using a remote sensing and GIS
software (ERDAS imagine). The model maker (a tool in the ERDAS software)
is used to analyse the overlays, wherein the different features of the thematic
layers are intersected/extracted and new class values are attributed to the
resulting polygons. The polygons are classified according to the conservation
priority status and finally integrated (union criteria in model maker) to generate
[474]
1535
the CPZ. The authenticity of the areas/zones proposed for conservation pri-
ority is confirmed with ground truthing.
Results
Totally 322 species are recorded from the Shervarayan hills (based on the
quadrats studied in the study area), of which 24 species are endemic (Table 1)
and 23 species are red listed (Table 2). The floristic richness (Map 2) are re-
grouped/reclassed into high (>80 species), medium (40–80 species) and low
(>40 species) rich areas respectively. The endemism and red listed species are
grouped into three zones based on the number of contributing species. The
CPZ map (Map 3) is generated with three classes according to the criteria
described before, based on the combination of scores – high, moderate and low
priority zones.
High priority zone is distributed in five sites with moderate to high species
richness. This zone accounts for 1582.53 ha (6.80%) of the total hill forest area.
The priority sites are authenticated with the presence of select/target species
(under different criteria) like Rubia cordifolia, Crotalaria shevaroyensis, Litsea
oleoides, Smilax zeylanica, Ixora notoniana, Neolitsea scrobiculata, Psychotria
octosulcata, Randia candolleana var. candolleana, Peperomia dindigulensis,
Celastrus paniculatus and Nothopegia colebrookiana in the evergreen forests.
The riparian forests comprise Terminalia arjuna, Mangifera indica, Ficus mi-
crocarpa and Syzygium cumini and on the other hand the semi evergreen forests
is represented mainly by Nothopegia colebrookiana, Celastrus paniculatus, De-
calepis hamiltonii, Santalum album, Naravelia zeylanica, Gymnema sylvestre,
[475]
1536
The zone occupies an area of about 6282.4 ha (27%) enclosing parts of ever-
green forests and dry mixed deciduous forests with species richness ranging
from moderate to low. The evergreen forest under this class includes the en-
demic and red listed species like Symplocos cochinchinensis, Vaccinium neil-
gherrense, Gnetum edule, Rubia cordifolia, Peperomia dindigulensis, Elaeagnus
indica and Curcuma neilgherrensis. The endemic and IUCN red listed plant
species in the dry mixed deciduous forests has both moderate and high richness
and the representing species are Withania somnifera, Naravelia zeylanica,
Dolichandrone arcuata, Hemidesmus indicus, Sapindus emarginatus,
[476]
1537
This zone with moderate to low species richness occupies an area of about
4524.92 ha (19.45%) of the total forest area and the zone comprises mostly of
the dry mixed deciduous types and to a lesser extent the riparian forests. The
red listed plant species in dry mixed deciduous forests species are Celastrus
paniculatus, Nothopegia colebrookiana, Pseudarthria visida and Hemidesmus
indicus and the select endemic species like Mallotus stenanthus, Pamburus
missionis, Shorea roxburghii and Pavetta blanda The riparian forest has only
one endemic and endangered plant species i.e. Cycas circinalis,
[477]
1538
Discussion
Most of the forests on the outer slopes and plateau of Shervarayan hills are still
facing the wrath of deforestation in spite of its protected status. There are
multidimensional reasons to it and the size of the protected area is the first
detriment rendering the very base of protection as ineffective. Secondly it is
followed by easy accessibility to the forest patches by the illegal loggers wherein
the dense network of the footpaths crisscrossing the forest patches confirm the
same. Thirdly the ineffectiveness of the protection status is the poor knowledge
of conservation prior sites within the protection realms of the forest.
This study also identifies a similarity in species contribution between the
evergreen forests of Shervarayan hills (a part of Eastern Ghats) with that of the
evergreen forests of Western Ghats. Species like Chionanthus ramiflorus,
[478]
1539
[479]
1540
The potential utility of remote sensing and GIS to identify the CPZ in this
study and culmination of all aspects dealing with the sole purpose of conser-
vation has been effective and reliable (based on the ground truth and field
checks). The resultant maps gives a picture of the CPZ providing a birds eye
view of the areas thus identified. The accessibility to the zones thus identified
can be deciphered and planned, finally paving way for better and effective
conservation.
Conclusion
Acknowledgements
References
Ahmedullah M. and Nayer M.P. 1986. Endemic plants of the Indian Region (Vol. I) – Peninsular
India. Deep printers, New Delhi.
Ahmedullah M. 2000. Prioritisation of endangered plants of India. In: Singh S., Sastry A.R.K.,
Mehta R. and Uppal V. (eds), Setting Biodiversity Priorities for India, vol. 2. World Wide Fund
for Nature-India, New Delhi, pp. 442–459.
[480]
1541
Ajith Kumar, Walker S. and Molur S. 2000. Prioritization of Endangered species. In: Singh S.,
Sastry A.R.K., Mehta R. and Uppal V. (eds), Setting Biodiversity Priorities for India. World
Wide Fund for Nature-India, New Delhi, pp. 341–425.
Amarnath G., Murthy M.S.R., Britto S.J. and Dutt C.B.S. 2003. Diagnostic analysis of conser-
vation zones using remote sensing and GIS techniques in Wet evergreen forests of the Western
Ghats- An ecological hotspot, Tamil Nadu, India. Biodiv. Conserv. 12(12): 2331–2359.
Ando A., Camm J., Polasky S. and Solow A. 1998. Species distribution, land values and efficient
conservation. Science 279: 2126–2128.
Balaguru B., John Britto S., Nagamurugan N., Natarajan D., Soosairaj S., Ravipaul S. and
Arockiasamy D.I. 2003. Vegetation mapping and slope characteristics in Shervarayan hills,
Eastern Ghats using Remote Sensing and GIS. Current Science 85(5): 645–653.
Balaguru B. 2002. Plant Diversity Assessment and its Conservation in Shervarayan Hills Eastern
Ghats of Tamil Nadu Using Remote Sensing and Geographical Information System. Ph.D
thesis, Bharathidasan University, Tiruchirappalli, Tamil Nadu, India.
Bawa K.S. and Seidler R. 1998. Natural forest management and conservation of biodiversity in
tropical forests. Conserv. Biol. 12(1): 46–53.
Bowles I.A., Rice R.E., Mittermeier R.A. and da Fonseca G.A.B. 1998. Logging and tropical forest
conservation. Science 280: 1899–1900.
Calridge G. and O’Callaghan B. (eds) 1997. Community Involvement in Wetland Management:
Lessons from the Field. Incorporating the proceeding of workshop 3: wetlands, local people, and
development, of the international conference on wetlands, Kuala Lumpur, Malaysia, pp. 19–24.
CES (Centre for Ecological Sciences) 1998. Monitoring Biodiversity – A manual of methods. IISC,
Bangalore.
Curnutt J., Lockwood J., Luh H., Nott P. and Russell G. 1994. Hotspots and species diversity.
Nature 367: 326–327.
Dawson K.J. 1996. Interior and exterior land use as factors in landscape biodiversity from case
studies. In: Szaro R.C. and Johnston D.W. (eds), Biodiversity and managed landscapes-Theory
and Practice. Oxford University Press, Oxford, pp. 514–530.
Dobson A.P., Rodriguez J.P., Roberts W.A. and Wilcove D.S. 1997. Geographic distribution of
endangered species in the United States. Science 275: 550–553.
Downton M.W. 1995. Measuring tropical deforestation: Development of the methods. Environ.
Conserv. 22(3): 229–240.
Ferreira L.V. and Prance G.T. 1998. Species richness and floristic composition in four hectares in
the Jau National Park in upland forests in Central Amazonia. Biodiv. Conserv. 7: 1349–1364.
Gaston K.J., Pressey R.L. and Margules C.R. 2002. Persistence and vulnerability: retaining bio-
diversity in the landscape and in protected areas. J. Biosci. 27(4): 361–384.
Hamilton A.C. 1984. Deforestation in Uganda. Oxford University Press and The African Wildlife
Society, Nairobi, Kenya.
Hartshorn G.S. 1995. Ecological basis for sustainable development in tropical forests. Ann. Rev.
Ecol. Systemat. 26: 155–175.
Howard P.C. 1991. Nature Conservation in Uganda’s Tropical Forest Reserves. Publication 3.
Forest Conservation Programme, World Conservation Union, Gland, Switzerland.
Hutto R.L., Heel S. and Landres P.B. 1987. A critical evaluation of the species approach to
biological conservation. Endangered species Updata 4(11): 1–4.
IUCN/SSC criteria Review working group. 1999. IUCN Red list criteria review provisional report;
draft of the proposed changes and recommendations. Species 31 & 32: 43–57.
Kier G. and Barthlott W. 2001. Measuring and mapping endemism and species richness: a new
methodological approach and its application on the flora of Africa. Biodiv. Conserv. 10: 1513–
1529.
Kumar A., Walker S. and Molur S. 2000. Prioritisation of endangered species. In: Singh S., Sastry
A.R.K., Mehta R. and Uppal V. (eds), Setting Biodiversity Priorities for India. World Wide
Fund for Nature – India, New Delhi, pp. 341–386.
[481]
1542
Kumaravelu G. and Chaudhuri K.K. 1999. Proceedings of a National Seminar on Endemic and
Endangered Plants and Animal Species of Eastern and Western Ghats. Tamil Nadu Forest
Department, Chennai.
Malcolm J.R. and Ray J.C. 2000. Influence of timber extraction routes on central African small-
mammal communities, forest structure and tree diversity. Conserv. Biol. 14(6): 1623–1628.
Margules C.R. 1989. Introduction to some Australian developments in Conservation evaluation.
Biol. Conserv. 50: 1–11.
Margules C.R., Pressey P.L. and Williams P.H. 2002. Representing biodiversity: data and proce-
dures for identifying priority areas for conservation. J. Biosci. 27(4): 309–326.
Menon S., Gil Pontius J.R.R., Rose J., Khan M.L. and Bawa K.S. 2001. Identifying conservation –
Priority areas in the tropics: A land - use change modeling approach. Conserv. Biol. 15(2): 501–
512.
Mooney H.A. and Chapin F.S.III 1994. Future directions of global change research in terrestrial
ecosystems. Trends Ecol. Evol. 9: 371–372.
Murthy K.S.R. 2000. Ground water potential in a semi-arid region of Andhra Pradesh - a geo-
graphical information system approach. Int. J. Remote Sensing 21(9): 1867–1884.
Myers N. 1980. Conversion of Tropical Moist Forests. National Academy of Science, Washington,
DC, USA, pp. 250 (illus).
Myers N. 1988. Threatened biotas: ‘‘hotspots’’ in tropical forests. The Environmentalist 8: 1–20.
Myers N. 1990. The biodiversity challenge: expanded hotspots analysis. The Environmentalist 10:
243–256.
Nayer M.P. and Shastry A.R.K. (eds) 1987–1990. Red Data Book on Indian Plants 1,2,3. BSI,
Calcutta.
Noss R.F. 1983. A regional landscape approach to maintain diversity. Bioscience 33: 700–706.
Noss R.F. 1991. From endangered species to biodiversity. In: Kohm K. (ed), Balancing on the
Brink of Extinction the Endangered Species Act and Lessons for the Future. Island Press,
Washington, DC, pp. 227–246.
O’Neill R.V., Hunsaker C.T., Jones K.B., Ritters K.H., Wickham J.D., Schwartz P.M., Goodman
I.A., Jackson B.L. and Bailargeon W.S. 1997. Monitoring environmental quality at the landscape
scale. Bioscience 47: 513–519.
Oliver I. and Beattie A.J. 1993. A possible method for the rapid assessment of biodiversity.
Conserv. Biol. 7: 562–568.
Oliver I. and Beattie A.J. 1996. Designing a cost-effective invertebrate survey: a test of methods for
rapid assessment of biodiversity. Ecol. Appl. 6: 594–607.
Olson D.M. and Dinerstein E. 1998. The global 200: a representation approach to conserving the
Earth’s most biodiversity valuable ecoregions. Conserv. Biol. 12: 502–515.
Peters C.M., Gentry A.H. and Mendelson R.O. 1989. Valuation of an Amazonian rain forest.
Nature 339: 655–656.
Ramesh B.R., Menon S. and Bawa K.S. 1997. A vegetation based approach to biodiversity gap
analysis in the Agasthyamalai region, Western Ghats, India. Ambio 26(8): 529–536.
Redford K.H. 1992. The empty forest. Bioscience 22: 412–422.
Robertson D.P. and Hull R.B. 2001. Beyond biology: toward a more public ecology for conser-
vation. Conserv. Biol. 15(4): 970–979.
Roy P.S. 1999. Remote sensing application in forestry and environment: retrospective and per-
spective. In: Proceedings of ISRS National Symposium on Remote Sensing Applications for
Natural Resources Retrospective and Perspective. Bangalore, pp. 169–185.
Scott J.M., Csuti B., Jacobi J.D. and Estes J.E. 1987. Species richness: a geographic approach to
protection future biological diversity. Bioscience 37: 782–788.
Scott J.M., Csuti B., Smith K., Estes J.E. and Caicco S. 1991. Gap analysis of species richness and
vegetation cover; an integrated biodiversity conservation strategy. In: Kohm K. (ed), Balancing
on the Brink of Extinction the Endangered Species Act and Lessons for the Future. Island Press,
Washington, DC, pp. 282–297.
[482]
1543
Serrao E.A.S. and Homma A.K.D. 1993. Sustainable Agriculture and the Environment in the
Humid Tropics. National Academy Press, Washington, DC, pp. 265–351.
Sheil D. 2001. Conservation and Biodiversity monitoring in the Tropics: Realities, Priorities, and
distractions. Conserv. Biol. 15(4): 1179–1182.
Singh S. and Taneja B. 2000. Prioritisation for biodiversity conservation: values and criteria. In:
Singh S., Sastry A.R.K., Mehta R. and Uppal V. (eds), Setting Biodiversity Priorities for India.
World Wide Fund for Nature – India, New Delhi, pp. 28–33.
Terborgh J. and Winter B. 1983. A method of sitting parks and reserved with species reference to
Colombia and Ecuador. Biol. Conserv. 27: 45–58.
Upreti G. 1994. Environmental conservation and sustainable development require a new devel-
opmental approach. Environ. Conserv. 21(1): 18–29.
Western D. and Wright R.M. 1994. Natural Connections: Perspective in Community Based
Conservation. Island press, Washington, DC, pp. 581.
Wilcove D.S. 1989. Protecting biodiversity in multiple-use lands: Lessons from the U.S. Forest
Service. Trends Ecol. Evol. 4: 385–388.
Williams P.H., Margules C.R. and Hilbert D.W. 2002. Data requirements and data sources for
biodiversity priority area selection. J. Biosci. 27(4): 327–338.
Williams P.H., Gibbons D., Margules C.R., Rebelo A., Humphires C. and Pressey R. 1996. A
comparison of richness hotspots, rarity hotspots, and complementary areas for conserving
diversity of British birds. Conserv. Biol. 10: 155–174.
Wilson E.O. (ed) 1988. Biodiversity. National Academy press, Washington, DC.
[483]
Biodiversity and Conservation (2006) 15:1545–1575 Springer 2006
DOI 10.1007/s10531-005-2930-6
Key words: Avian species richness, Conservation, Deforestation, Diversity, Guatemala, Land-use
Abstract. Cloud forests in central Guatemala are fragmented and decreasing in area due to slash-
and-burn agricultural activities. We studied bird species composition, abundance, guild composi-
tion, and site tenacity of a 102 ha plot located in a cloud forest region of the Sierra Yalijux in
Guatemala, half of which was primary forest and half young secondary forest (<7-years-old). Of
the 100 species present 14 were restricted to the Endemic Bird Area ‘Northern Central American
highlands’ (i.e. 66% of a total of 21 endemics). Five of the 100 analysed species, including one of
the restricted-range species (Troglodytes rufociliatus), had a significantly different abundance in
primary and secondary forests. Theoretical analysis suggests that seven species out of a community
comprised of 141 bird species are already extirpated and only three out of the 14 present restricted-
range species might survive the current state of deforestation. Insectivores were the dominant guild
on the plot in terms of numbers of species, followed by omnivores, frugivores and granivores.
However, in terms of individuals, omnivores made up nearly half of the bird individuals in primary
forest, but declined by 44% in secondary forest, whereas granivores more than doubled in this
habitat type. Numbers of species per guild were not significantly different between habitats, while
numbers of individuals per guild were significantly different. In general, individuals per species are
significantly different in the two habitats. Results suggest that most of the species that are currently
surviving in the remnant forests of the Sierra Yalijux might be fairly well adapted to a range of
forest conditions, but that populations of a number of restricted-range species might be small. Even
generalists species like the Common Bush Tanager (Chlorospingus ophthalmicus) are less abundant
in secondary vegetation than in primary forest of the study plot.
Introduction
[485]
1546
forest remnants world-wide (e.g., Hughes et al. 2002; Schulz et al. in press) and
especially in Latin America (Markussen 2004). In central Guatemala, natural
habitats, such as primary cloud forests, are often converted and used as part of
the milpa system (slash-and-burn agriculture) (Renner 2003; Markussen 2004;
Markussen and Renner 2005). It is questionable whether secondary forest re-
growth after farming can preserve numbers of species comparable to natural
habitats in tropical landscapes over the long term (Hughes et al. 2002; Renner
2003). In addition, the time required for secondary forest biodiversity to ap-
proach that of primary forest is not known (e.g., Shankar-Raman et al. 1998;
Terborgh 1999; Shankar-Raman 2001; Shankar-Raman and Sukumar 2002).
Cloud forest areas in many parts of the tropics are increasingly deforested,
and many forest remnants have likely already lost a large proportion of their
original avifauna (Kappelle and Brown 2001). Some sites are presumed to have
lost forest bird populations, e.g., the Brazilian Atlantic forests. Other sites may
still retain endemic forest species because of persisting habitat heterogeneity
(Marsden et al. 2004). Nevertheless, fragmentation and loss of forest habitats
are important causes of regional species extinction (Marsden et al. 2004). The
province Alta Verapaz in central Guatemala is located within a Biodiversity
Hotspot (Myers et al. 2000) and an Endemic Bird Area (Stattersfield et al.
1998), and is therefore of high importance for conservation (Veblen 1976;
Islebe 1995; Stattersfield et al. 1998; Markussen and Renner 2005). This region
contains primary forests subject to decreasing area and increasing fragmen-
tation that currently contain endemic and specialised bird species.
We examined species richness, abundance, site tenacity and body mass dis-
tribution of birds in primary cloud forest compared to neighbouring young
secondary forest in the highlands of central Guatemala near the village of
Chelemhá. We used within-one-habitat recaptures as a measure of habitat
quality. Special emphasis was placed on restricted-range species and Central
American cloud forest specialists.
Study area
The Sierra Yalijux where the study site was located, belongs to the north-
easternmost slopes of the northern Cordillera of Guatemala (Municipio Tu-
curú, Alta Verapaz). The total forest cover above 1,800 m (ca. the altitudinal
border of cloud forests in Guatemala) is estimated at 5,500 ha (Markussen
2004). In this region, the climate is tropical to sub-tropical (MAGA 2001) and
precipitation is high, up to 4000 mm per year (Markussen 2004).
We chose a 102 ha study plot in the Sierra Yalijux near the settlement of
Chelemhá (Figure 1). The geographic plot co-ordinates are 9004¢ W and
1523¢ N. It is situated at an elevation between 1,980 m and 2,550 m and
contains approximately 51 ha of primary forest and 51 ha of secondary
forest. The primary forest in the region is mainly mixed oak and pine forest
with Pinus maximinoi as the dominant species (Veblen 1976; Moziño 1996).
[486]
1547
Figure 1. (a) Location of the study plot (n) in Guatemala and Endemic Bird Area ‘Central
American highlands’ (Stattersfield et al. 1998) (- - - -). Grey indicates areas in Guatemala with
elevations above 1000 m. (b) The study plot (bold line — ) with main trails for transect counts (thin
line —), mist net sites (black dots •) and vegetation classification (n grey: primary forest, white:
secondary habitats of all kinds). Dotted line indicates the Rı́o Chelemhá; arrow on map resembles
approximately 400 m in nature.
The canopy exceeds 35 m and there are generally three vegetation strata, i.e.
understorey <7 m, mid-story 7–20 m, and overstorey >20 m. Stem diameter
at breast height for overstorey trees exceeds 1 m. Young secondary forest
(<10 years-old), is – aside from primary forest – the most abundant vege-
tation in the study region and probably also within the whole of Alta Ve-
rapaz. Secondary forests >10 years cover <5 ha within the community of
Chelemhá. Young secondary vegetation is the result of slash-and-burn agri-
culture (milpa system), and the secondary forest we studied had an age of less
than 7 years. After corn and bean cultivation, shrubs can grow up to 10 m in
height during a fallow period of 7 years. However, a fallow period of this
duration is exceptional occurring only in the study plot. Generally in the
community of Chelemhá there is a fallow period of <2 years. The shrubby
habitat is composed of many different tree species, including oaks (Quercus
sp.); but in contrast to primary forest the pines (Pinus sp.) are rare and the
shrubs form only a single vegetation layer.
For both the understorey and overstorey, the estimated tree height, diameter
at breast height and density were measured for each 25 m section of the bird
count transects.
Methods
[487]
1548
In primary and secondary forest, an existing trail system was used to count
birds by means of regular transect counts in order to assess population den-
sities for different habitats of the plot (Bibby et al. 1995; Krebs 1999; Gilbert
et al. 2000). A total of 3,300 m of counts were conducted along the trails (no
trail intersections allowed, all trails were separated at least by 50 m), 50% in
each vegetation type. A series of three visits per trail were conducted in 2002
between 0530 h and 0900 h. A maximum trail length of 450 m was visited per
day. All sight records and singing individuals were recorded while slowly
walking along the trail. The trails of the entire plot were divided into 22 trail
sections of 150 m each. Eleven sections were situated in primary forest, and 11
in secondary forest. All displaying birds within 100 m distance to both sides of
the trail were recorded, so that each trail section represented a 3-ha-sized
rectangular strip (150 · 2 · 100 m = 3 ha). Each record contained the per-
pendicular observation distance, as well as information on species, sex, age,
and number of individuals. Density estimates were derived from the maximum
number of birds recorded on each 3-ha strip. For some cryptic species, (e.g.
Grallaria guatimalensis) or species with low voices (Buarremon brunneinucha,
Atlapetes gularis), the strip width was reduced to 50 m and the corresponding
observations within 50 m from the transect for the density calculations. Cya-
nocorax melanocyaneus and Cyanolyca pumilo were estimated conservatively by
directly counted individuals. Most flocks (especially the frequent Chlorospingus
ophthalmicus flocks) disbanded early in the breeding season and could be noted
as pairs. Singing and territorial display behaviour of males was counted as a
breeding unit for all other species, except for species where females also display
territorial behaviour. Trochilidae were treated in the same way, except for trap-
liners where only females (if distinctive) were recorded as a breeding unit.
Identification was made using Land (1970), Howell and Webb (1995), and
Edwards (1998). Taxonomy and systematic order follows the checklist of the
American Ornithologists’ Union (1998) with the relevant supplements
(American Ornithologists’ Union 2000, 2002).
Transect and mist netting procedures are often combined in tropical forest
bird community studies since the two methods select different portions of the
community, thus achieving a more accurate assessment of community com-
position (Poulsen 1994; Remsen 1994; Remsen and Good 1996). Twelve mist
net lines were established, six in natural and six in secondary forest. The net
lines were distributed randomly at existing tracks to avoid pseudo-replication
(Hurlbert 1984). Each net line consisted of eight nets of 12 m each and was
opened for 8.5 h per capture day. Capturing was conducted for 2 days at each
line in 2001, and for 4 days in 2002 with a total of 4,896 net hours (12 m net).
Each captured bird was marked and body mass determined with spring bal-
ances.
Effects of land-use were measured as differences in bird community struc-
ture, including changes in species richness, species composition, numbers of
individuals (trail data) as well as in individual turnover, body size and body
mass (mist net data) between primary forest and secondary forest.
[488]
1549
Statistics
Species richness is normally not detected completely, that is neither all species,
nor all individuals are detected in natural environments (e.g., Rosenzweig 1995;
Begon et al. 1996; Krebs 1999). For the purpose of estimation of total species
richness, several indices were established with different purposes (e.g., Chao
and Lee 1992; Colwell and Coddington 1994; Rosenzweig 1995). These esti-
mators and indices give an approximate number of species that might be de-
tected when making repeated observations (Magurran 1988; Rosenzweig 1995;
Krebs 1999). Different estimators are of different value according to their
purpose. Here we used the Bootstrap estimator due to the large sample (Ma-
gurran 1988). Otherwise the Jackknife estimator must to be used.
Based on the trail data, we first described species accumulation from the
pooled data of the whole study plot (Sobs in Colwell 2000). Then differences
between primary forest and secondary forest were established from the sepa-
rate samples. In addition, primary and secondary forest were compared with
regard to species composition. To describe the community, observed species
numbers (Sobs), observed individuals numbers (N), and indices were calculated,
and Bootstrap (SBoot), Evenness (E), and the Abundance-based Estimator of
Species Richness (ACE) were used. For calculation and formulas of the esti-
mators see Magurran (1988) and Colwell (2000). See Krebs (1999) and Renner
(2003) for a complete summary.
Body mass analyses required splitting of sexes for certain species. A group is
either a species or one sex of a species. For Diglossa baritula, Lampornis
amethystinus, L. viridipallens, Lamprolaima rhami, and Turdus infuscatus, a
[489]
1550
sexual separation for analysis is essential due to significant body mass differ-
ences between the sexes.
Recapture data
Results
Vegetation
Mean distance between trees with diameter at breast height of <0.20 m of the
Chelemhá Plot was 2.94 m±1.77 in primary forest and 1.16 m±0.82 in sec-
ondary forest. Mean height of overstorey trees was 25.9 m±6.6 in primary
forest, 4.9 m±1.9 for understorey trees in primary forest, and 2.9 m±1.9 for
trees in secondary forest (see Table 1).
Overstorey Understorey
dbh h d dbh h d
[490]
1551
On the 102-ha plot, a total of 100 avian species was found to be present of
which 75 are presumably residents. Sixty-six of the 100 avian species recorded
were detected by mist netting, 75 from regular counts along trails, and eight by
incidental records. Forty-seven species were recorded by both mist netting and
transect counts, 17 by mist netting alone and 28 by transect counts alone. Out
of the 100 species, 75 species were presumed resident breeding species, hereafter
treated as ‘residents’. This number is an estimate and includes incidental re-
cords of species which are believed to breed but were only rarely recorded (1–2
observations) by regular transect counts (Coragyps atratus, Lophostrix cristata,
Glaucidium gnoma, Streptoprocne zonaris, Sclerurus guatemalensis, Tachycineta
bicolor, Saltator caerolescens, Dives dives). The total species numbers for the
entire plot was estimated (Colwell 2000) at 61 (Sobs) and 62.75±0.00 (Boot-
strap±s.d.).
Further results on the bird community structure and comparisons to other
tropical and temperate studies are drawn in Renner (2003).
Based on the maximum abundance of each species at each trail section, a total
of 1405 individuals was estimated to be present on the plot. While, in total,
more species were observed in secondary forest (51 and 55 observed species in
primary forest and secondary forest, respectively), more individuals were ob-
served in primary forest (753) than in secondary forest (652). The difference in
overall population size is significant (v2 = 7.41, pdf1<0.01), representing a
decrease of 14% between primary and secondary forest. The higher observed
species numbers in secondary forest is also reflected in higher estimated total
species richness: the Bootstrap estimator in primary forest amounted to 52.91
(±0.77 s.d.) and in secondary forest to 57.80 (±0.00). This suggests that only a
few more species remained undetected (Figures 1 and 2). Diversity statistics of
birds are summarised in Table 2.
Table 2. Summary of diversity statistics (calculated with Colwell 2000) depending on habitat
compiled from transect count data.
Sobs 51 – 55 –
Individuals 1563 – 1766 –
Singletons 0.0 0.00 0.0 0.00
Doubletons 0.0 0.00 0.0 0.00
ACE 51.0 0.00 55.0 0.00
Bootstrap 52.9 0.77 57.1 0.00
[491]
1552
Figure 2. Expected species E (transect counts data from 2002) in primary forest and secondary
forest. Error bars were left out for better illustration. Data were calculated with Rarefaction 1.3
(Holland 2003).
Figure 3. Observed species Sobs (transect count data from 2002) in primary forest and secondary
forest in the Chelemhá Plot. Error bars were left out for better illustration. Calculated with Colwell
(2000).
[492]
1553
using the bird data from the twelve net lines, there are two distinct groups
(primary, secondary forest) forming the ordination plot (Figure 5), with
sample scores of trapped bird assemblages showing significant differences
between habitats (Rao’s R2,9 = 8.27, p<0.01).
[493]
1554
Forty-three species had individuals in both habitats; three species had similar
mean individual numbers, 17 had higher mean individual numbers in primary
forest and 23 had higher mean individual numbers in secondary forest. Five
species showed significant differences in abundance in the two habitats
(Table 6). Zimmerius villisimus, Troglodytes rufociliatus, Atthis ellioti,
Myadestes unicolor, and Atlapetes gutteralis had significantly different indi-
vidual numbers in primary forest and secondary forest. C. ophthalmicus (p =
0.08), Henicorhina leucophrys (p = 0.06), and Sclerurus mexicanus (p = 0.06)
had almost significant differences.
The most common species, C. ophthalmicus, was comparatively less abun-
dant in secondary forest than in primary forest (Table 6). Three species
(Z. vilissimus, T. rufociliatus, and M. unicolor; Tables 3 and 7) were signifi-
cantly less abundant in secondary forest compared to primary forest. While 23
species are less abundant in primary forest than in secondary forest, at least 22
have 50% more individuals in primary forest.
At the species level, five species were found in significantly different numbers
in the two habitats (Table 3). This represents a minor part of the 100 species
observed, and should be treated with caution. Given that from an accepted
significance level of 5% (p = 0.05) and 61 analysed species with any N to
perform the Mann–Whitney-U test out of the 100 residents (see Table 6), al-
ready 3.05 species (61 · 0.05) could be expected to show significant differences
in abundance between habitat types by chance alone.
Within the Chelemhá Plot there are three species influenced significantly
negatively and two positively in accordance to individual numbers. Several pri-
mary forest birds (Table 7) and especially endemic primary forest birds of the
Central American highlands are still present in the respective parts of this plot.
Guild composition
Table 3. Species showing significant differences in abundance between habitat types. Abundance
given as mean numbers of individuals calculated from the maximums recorded on 3-ha transect
strips. p: p-level (corrected) based on the Mann-Whitney-U test. Exclusively species with significant
difference (p<0.05) are listed.
[494]
Table 4. Numbers of species and individuals per guild (transect count data). Primary forest represents expected and secondary forest observed frequency for
v2-test. The first number is total N, the second the relative portion. Differences to Table 2 occur to the fact that here counted and there estimated individuals
are given. As indicated in Table 6, several species and individuals were accounted for more than one guild when clear nutrition preferences were not given.
Exclusiona Insectivores Omnivores Frugivores Nectarivores Granivores Carnivores Total (N) v2 (p)
1555
1556
Table 5. Body mass changes in recaptured individuals. BM 1: body mass at first capture, BM 2:
body mass at final recapture in the same breeding season. PF: primary forest, SF: secondary forest.
[496]
Table 6. Abundance for forest species, given as maximum records on 51 ha primary forest (PF) and 51 ha secondary forest (SF) and population estimates
(for 102 ha and 5500 ha, the latter represents the natural pine–oak forests of the Sierra Yalijux). All species potentially abundant in the Sierra Yalijux are
listed. However, guild, habitat, etc. is only listed for the relevant species definitely observed. Taxonomy and systematics follows American Ornithologist’s
Union (1998, 2000, 2002).
Family Species Exp.a Guildb Habitatc Methodd Abundancee Estimate for the Sierra Yalijux
(PF)
PF SF Ind./102 ha Ind./5,500 ha
Cracidae
Ortalis vetula H,C F TC 2 0 4 216
Penelope purpurascens H
Oreophasis derbianus H F –
Penelopina nigra H,C F PF TC 16 16 31 1725
Phasianidae
Dendrortyx leucophrys H,C G TC 1 4 2 108
[497]
Odontophorus guttatus H
Dactylortyx thoracicus H
Cyrtonyx ocellatus H G TC
Ardeidae
Ardea herodias C
Bubulcus ibis C
Cathartidae
Coragyps atratus H,C Ca I
Accipitridae
Chondrohierax uncinatus H,C C TC
Elanoides forficatus C
Accipiter chinogaster H,C
Asturina nitida C
Buteo platypterus C
Buteo solitarius H
Buteo jamaicensis H,C
1557
Table 6. (Continued).
1558
Family Species Exp.a Guildb Habitatc Methodd Abundancee Estimate for the Sierra Yalijux
(PF)
PF SF Ind./102 ha Ind./5,500 ha
Falconidae
Micrastur ruficollis H,C
Falco sparverius H,C
Scolopacidae
Bartramia longicauda C
Columbidae
Columba livbia H
Patagioenas fasciata H,C F,G TC 9 3 18 971
Zenaida asiatica H
Zenaida macroura C
Columbina inca H
[498]
Columbina passerina H
Claravis mondetoura H F,G N
Leptotila verreauxi C
Geotrygon albifacies H,C F,G N
Psittacidae
Bolborhynchus lineola H,C G,F TC 2 1 4 216
Cuculidae
Piaya cayana H
Tytonidae
Tyto alba H
Strigidae
Megascops trichopsis H
Megascops barbarus H
Lophostrix cristata H C I
Glaucidium gnoma H,C C I
Ciccaba virgata H,C C TC
Strix fulvescens superspecies varia H,C C TC 0 1 0 0
Asio stygius H
Aegolius ridgway H
Caprimulgidae
Chordeiles acutipennis H
Caprimulgus vociferus) arizonae H,C I TC 1 0 2 108
Apodidae
Cypseloides niger H,C I TC
Streptoprocne rutila H,C
Streptoprocne zonaris H,C I I
Chaetura vauxi H,C
Aeronautes saxatalis H,C I TC
Trochilidae
Campylopterus hemileucurus H,C N N
Colibri thalassinus H,C N N,TC 0 1 0 0
Abeillia abeillei H N N,TC 0 2 0 0
Hylocharis leucotis H,C N N
[499]
Amazilia cyanocephala H N N
Lampornis viridipallens H,C N N,TC 0 2 0 0
Lampornis amethystinus H,C N N,TC 46 45 90 4961
Lamprolaima rhami H,C N N,TC 11 7 22 1186
Eugenes fulgens H,C N N,TC
Doricha enicura H N N
Tilmatura dupontii H,C N TC
Atthis ellioti H,C N N,TC 0 6 0 0
Trogonidae
Trogon collaris H I,F TC 2 0 4 216
Trogon mexicanus H,C I,F TC 11 7 22 1186
Pharomachrus mocinno H,C F N,TC 17 13 33 1833
Momotidae
Aspatha gularis H,C F,I N,TC 3 7 6 324
Ramphastidae
Aulacorhynchus prasinus H,C F N,TC 7 5 14 755
1559
Table 6. (Continued).
1560
Family Species Exp.a Guildb Habitatc Methodd Abundancee Estimate for the Sierra Yalijux
(PF)
PF SF Ind./102 ha Ind./5,500 ha
Picidae
Melanerpes formicivorus H
Melanerpes aurifrons H,C
Sphyrapicus varius C
Colaptes auratus H,C I N,TC 1 3 2 108
Picoides villosus H,C I N,TC 10 8 20 1078
Piculus rubiginosus H
Furnaridae
Anabacerthia variegaticeps H,C
Automolus rubiginosus H,C I N,TC 5 2 10 539
Sclerurus mexicanus H,C I TC 5 2 10 539
[500]
Sclerurus guatemalensis - I I
Dendrocolaptidae
Xiphorhynchus promeropirhynchus H
Xiphorhynchus erythropygius H,C I N,TC 1 1 2 108
Lepidocolaptes affinis H,C I N,TC
Formicaridae
Grallaria guatimalensis H I N
Tyrannidae
Camptostoma imberbe H I N
Elaenia frantzii C
Zimmerius vilissimus H,C I SF N,TC 26 13 51 2804
Mitrephanes phaeocercus H,C I TC 12 12 24 1294
Cantopus pertinax H I TC
Cantopus sordidulus C
Contopus virens C
Cantopus cinereus C
Contopus borealis C
Empidonax flaviventris C
Empidonax virescens C
Empidonax minimus C
Empidonax hammondii C
Empidonax oberholseri C
Empidonax affinis H I N
Empidonax flavescens H,C I SF N,TC 0 1 0 0
Empidonax fulvifrons H,C
Sayornis nigricans H,C
Pachyramphus major H
Pachyramphus aglaiae H,C
Vireonidae
Vireo plumbeus H,C I N,TC 7 0 14 755
Vireo huttoni H,C
Vireo gilvus C
Vireo leucophrys C
[501]
Vireo philadelphicus C
Cyclarhis gujanensis C
Corvidae
Cyanocitta stelleri H
Cyanocorax melanocyaneus H,C O N,TC 1 3 2 108
Cyanolyca pumilo H,C O N,TC 16 8 31 1725
Aphelocoma unicolor H
Corvus corax H O TC
Hirudinidae
Tachycineta bicolor H,C I I
Tachycineta thalassina H I TC 0 1 0 0
Notiochelidon pileata H,C I TC
Petrochelidon pyrrhonota C
Hirundo rustica C
Certhididae
Certhia americana H
1561
Table 6. (Continued).
1562
Family Species Exp.a Guildb Habitatc Methodd Abundancee Estimate for the Sierra Yalijux
(PF)
PF SF Ind./102 ha Ind./5,500 ha
Troglodytidae
Thryothorus modestus H,C I PF N,TC 13 13 25 1402
Troglodytes musculus H,C I N,TC 7 10 14 755
Troglodytes rufociliatus H,C I N,TC 14 2 27 1510
Henicorhina leucophrys H,C I N,TC 35 29 69 3775
Turdidae
Sialia sialis H,C
Myadestes occidentalis H,C O N,TC 15 16 29 1618
Myadestes unicolor H,C O N,TC 30 15 59 3235
Catharus aurantiirostris H,C O N,TC 4 5 8 431
Catharus frantzii H,C O N,TC 35 31 69 3775
[502]
Catharus mexicanus H
Catharus dryas H
Catharus ustulatus C
Catharus guttatus C
Turdus infuscatus H,C O N,TC 15 8 29 1618
Turdus plebejus H,C O N,TC 5 8 10 539
Turdus grayi H,C O N,TC 35 33 69 3775
Turdus rufitorques H,C O TC 1 4 2 108
Mimidae
Dumatella carolineus C
Melanotis hypoleucus H,C I,F SF N,TC 11 13 22 1186
Ptilogonatidae
Ptilogonys cinereus H,C
Peucedramidae
Peucedramus taeniatus H,C
Parulidae
Vermivora chrysoptera H I PF N
Vermivora peregrina C
Parula superciliosa H,C I N,TC 3 7 6 324
Dendroica pensylvanica C
Dendroica coronata C
Dendroica virens C
Dendroica townsendi C
Dendroica occidentalis C
Dendroica fusca H,C I N,TC
Dendroica graciae H
Mniotilta varia C
Seiurus noveboracensis C
Seiurus motacilla C
Oporornis tolmiei H,C I N,TC 5 1 10 539
Geothlypis poliocephala H,C I N
Wilsonia pusilla H,C I N,TC 4 0 8 431
Wilsonia canadensis C
[503]
1563
Table 6. (Continued).
1564
Family Species Exp.a Guildb Habitatc Methodd Abundancee Estimate for the Sierra Yalijux
(PF)
PF SF Ind./102 ha Ind./5,500 ha
Aimophila rufescens H,C G N
Zonotrichia capensis H,C G N,TC 3 5 6 324
Spizella passerina H G N
Cardinalidae
Saltator caerolescens C G I
Saltator atriceps H G TC 3 2 6 324
Pheucticus ludovicianus C
Icteridae
Dives dives H,C G,I I
Quiscalus mexicanus H,C O TC 1 4 2 108
Molothrus aeneus H
[504]
# Species Global statusa Endemicb Mean individuals 102 ha (primary Sierra Yalijux (primary
in primary forest forest) ind./102 ha forest) ind./5500 ha
ind./51 ha*
1565
1566
Table 7. (Continued).
# Species Global statusa Endemicb Mean individuals 102 ha (primary Sierra Yalijux (primary
in primary forest forest)ind./102 ha forest) ind./5500 ha
ind./51 ha*
Additional species of conservation concernc
22 Pharomachrus mocinno Near threatened 17 34 1870
23 Penelopina nigra Near threatened 16 32 1760
24 Chlorospingus ophthalmicus 163 336 18,480
[506]
both habitats, the largest group in terms of numbers of species was insectivores
(24 sp. in primary vs. 19 sp. in secondary forest), followed by omnivores (9 vs.
11 sp.). However, in primary forest omnivores were much more abundant than
insectivores. This was mainly a consequence of the large numbers of the
tanager C. ophthalmicus in primary forest which is responsible for roughly half
(45% drop in C. ophthalmicus from primary to secondary forest) of the 56%
drop in individual numbers of omnivores in secondary forest. M. unicolor and
Cyanolyca pumilo are the second and third species with the largest drop of
individuals between primary and secondary forest for omnivores with 8% and
4%, respectively. If these three species (C. ophthalmicus, M. unicolor, and
C. pumilo) are excluded from the analyses, omnivores have a percentage of
25.2% individuals in primary forest and 15.4% in secondary forest. When
excluding the three species, insectivores stay the largest group but granivores
double their relative proportions (10.5% vs. 22.8%) (Table 4).
The only other guild that showed remarkable differences in abundance
between habitats was granivores. Granivores were more than twice as abun-
dant in secondary forest (219% of primary forest) than in primary forest, a
result largely caused by high numbers of the emberizids Atlapetes gutteralis
(responsible for 34% of the increase) and Carduelis notata (11%), but also of
the dove Columba fasciata (2% decrease). When excluding the three species,
granivores are represented with 10.9% in secondary forest and 0.6% in pri-
mary forest.
Numbers of individuals and species of frugivores, nectarivores and carni-
vores are each represented in similar proportions in both habitats. When
excluding these six species as mentioned before the differences in individuals
per guild are still significantly different with p<0.01 (Table 7).
Recaptures
There were 25 species with at least one recapture, of which seven had recap-
tures in both habitats. Twelve species were recaptured exclusively in primary
forest and six exclusively in secondary forest. However, using different species
as sampling units, the numbers of recaptures for all 25 species were not sig-
nificantly different between the two habitats (Wilcoxon test, T=125.0,
Z=0.714; p = 0.48).
Members of five species endemic to Central American highland forests (not
to be confused here with the Endemic Bird Area; Stattersfield et al. 1998) were
recaptured in primary forest (Abeillia abeillei, D. baritula, Empidonax flaves-
cens, Asphata gularis, Melanotis hypoleucus). The recapture rate Rt is £ 0.01,
except for E. flavescens with a recapture rate of Rt = 0.20 in primary forest.
Leaving species level and going one step beyond to the individual level, of all
140 recaptures (plus 28 excluded same-day recaptures) 51 individuals were
recaptured at the same net line (Table 5). Except for three individuals recap-
tured at the same net line, all marked individuals were recaptured at different
[507]
1568
localities within the same habitat, i.e. at one of the six net lines in the same
habitat but differing from the original first-capture net line. Individual recap-
tures show that just three out of 180 were recaptured in a habitat different from
previously captures during the study period in 2001 and 2002. Two of the three
were first captured in primary forest and one moved into primary forest after
being marked in secondary forest.
Two habitat-switching individuals had a higher body mass in secondary
forest than in primary forest. A female L. amethystinus first captured in pri-
mary forest on 20 March 2002 was recaptured in secondary forest on 12 April
2002 (5.5 g fi 6.5 g). C. frantzii first captured in primary forest on 17 March
2002 was recaptured in secondary forest on 17 May 2002 (26.0 g fi 28.3 g).
C. ophthalmicus banded in secondary forest on 17 July 2001 was recaptured in
primary forest on 17 April 2002 with dramatically decreased body mass
(16.5 g fi 8.5 g).
Of all recaptures, 28 individuals were recaptured twice, one individual of
B. belli three times within three months in primary forest at neighbouring mist
nets.
Body mass
Individual recaptures (within same habitat and within the same season, i.e.
within the 2001 or 2002 netting period) indicate that 10 individuals lost body
mass in secondary forest and 12 in primary forest. While eight species were
heavier in primary forest, 14 had higher body mass in secondary forest. For all
species with both more than five captures in each habitat
(Nprimary forest 5 \ Nsecondary forest 5), the species’ body mass distribution was
not significantly different between habitats (v2-test, p>0.05). Forty-eight
individuals were recaptured and body mass was determined (Table 5).
Thirteen of the 23 groups (for definition of ‘group’ see methods) with at least
two individuals in each habitat had a higher mean body mass in secondary
forest. Eight groups had higher body mass in primary forest. The body mass
differences are not significant (MANOVA, variable: ‘habitat’, p = 0.21, Post-
hoc: Newman-Keuls test).
Body mass distribution did not indicate that primary forest represented a
different habitat quality compared to secondary forest in the Sierra Yalijux
(Table 5). Also, the total biomass was generally similar.
Restricted-range species
[508]
1569
extinct in the recent past (Table 6). Eight of the remaining 13 restricted-range
species were exclusively observed in secondary forest in the Sierra Yalijux (see
also Table 7). Of the remaining five species, four did not reveal important
differences in abundance between primary and secondary forest. One re-
stricted-range species, T. rufociliatus, was significantly more abundant in pri-
mary than secondary forest (Table 3).
Discussion
[509]
1570
(see above) and (ii) most species are patchily distributed and/or rare (Thiollay
1994a and b). Even in 100 ha of tropical landscapes ca. 95% of species are
recorded with the methods we used according to Terborgh et al. (1990). The
second aspect would suggest 133 recorded species at the Sierra Yalijux (95% of
141 species), the missing 33 species may be the result of the area–species-
relationship and extinction. The area–species-relationship suggests the presence
of 107 species at our study site, assuming here Aoriginal = 5500 ha, Asurviving =
102 ha, Soriginal = 133, and a mean empirically proven z-value = 0.26 for
tropical bird communities (Brooks et al. 1997, 1999a, b and c). Therefore seven
species are most likely already extinct. Due to the fact that this calculation is
not a species-specific analysis, the nomination of the seven species is impossi-
ble. Nevertheless, based on the previous analysis, it appears that at least seven
species are extirpated, and more species are likely to vanish from the study plot.
Differences at the community level between primary and secondary forest were
generally small. Species richness and species similarity were comparable be-
tween habitat types. Only the numbers of individuals revealed a significantly
higher bird abundance in secondary than in primary forest. While multi-
dimensional scaling of the community data yielded significant differences when
using only mist net captures, the differences for transect counts were not sig-
nificant. This might be the result of a collapsed stratification of the bird
community trapped in secondary forest, where even canopy species occur at
levels that make them vulnerable to capture.
Accurate comparisons of the importance of one habitat vs. another requires
measurement of individual fitness. The question of whether species of con-
servation interest (e.g., endemics and specialists, see Table 7) are able to find
suitable conditions to reproduce in secondary forest as well as they can in
primary forest is of considerable interest. Reproductive success for P. mocinno
is most likely directly linked to primary forest (Renner 2005). The species
breeds in primary forest and finds stable breeding sites only there (Renner
2003; Renner 2005; compare also Gillespie 2001). The same potentially applies
for other forest species and endemics listed in Table 7 (cf. Renner 2003).
Some authors affirm that young secondary forest and old secondary forest
are suitable to preserve a degree of diversity and species richness comparable to
primary forests (Shankar-Raman et al. 1998; Shankar-Raman 2001; Shankar-
Raman and Sukumar 2002). Older secondary forest (10+ years) might be
suitable for preservation of biodiversity and a high degree of species richness,
but unfortunately, at least in the study region of central Guatemala, secondary
forests older than 7 years are rare, even more rare than primary forest.
Therefore, primary forest must be preserved as long as data are lacking to
indicate that young secondary forest might also be suitable for conservation of
biodiversity. Currently and most likely in the future also, young secondary
[510]
1571
forest is more frequent than old secondary forest and primary forest in the
Sierra Yalijux. The relative proportion of the different types of forest areas is a
consequence of expanding slash-and-burn agriculture and ongoing agricultural
activities in secondary re-growth areas (Markussen 2004; Markussen and
Renner 2005).
Guild composition
Guild structure is different for the two habitats, indicating that the resources
are not equal. While the percentage of insectivore bird individuals did not differ
significantly between primary forest and secondary forest, there is a higher
portion of insectivore species in primary forest than in secondary forest,
probably an indication of a more complex vegetation structure and absolute
numbers show significant differences (see results). Nectarivore species numbers
differ slightly and granivore species numbers are almost doubled in secondary
forest. This indicates that the food availability for species and for individuals
per species is higher in secondary forests.
One striking difference between primary and secondary forest is the large
amount of granivores in the latter, twice that of primary forest. This is
attributed to the larger supply of granivore food. Omnivores are more abun-
dant in primary forest, which superficially implies that there are less specialised
resources and species present. However, most omnivores in primary forest are
ground dwellers, which is obvious because the vegetation structure near ground
in secondary forest is much denser than in primary forest (see ‘Vegetation’ and
compare Renner 2003).
Recaptures
[511]
1572
habitat for birds and could contain more territories per area (cf. Renner 2003;
Matthysen et al. 1995; Reitsma et al. 2002).
Conclusions
Results indicate that, in the cloud forests of the Sierra Yalijux, specialised
forest species are extinct or at least diminished in remaining primary forest and
mostly habitat generalists have survived which also inhabit secondary forest.
Even abundant species that are recorded in both habitats (C. ophthalmicus)
show lower populations in secondary forest (Table 7). Based on our calcula-
tions of population sizes, the forest-dependent species P. mocinno, P. nigra and
Z. vilissimus are threatened since they already have small populations.
P. mocinno depends on nesting sites (nest holes in rotting trees), only found in
primary forest. This species will not survive complete deforestation, even when
few breeding holes remain in old secondary forest for approximately 25 years
(Renner 2005). P. nigra is feeding in old secondary forest also, but not in slash-
and-burn agriculture or young secondary forest. Both species are still abun-
dant, but they will disappear when primary forests vanish. Z. vilissimus has
twice as many individuals in primary forest than in secondary forest and will
suffer from deforestation and further fragmentation. All forest endemics from
the Central American highlands (Table 7) will be affected by definition. Con-
trastingly, generalists and non-forest birds and non-forest breeders will gain
from deforestation or at least will not be influenced.
Acknowledgements
This study was supported by the German Research Foundation (DFG) and
was part of the DFG GK 642/1 ‘‘Valuation and Conservation of Biodiversity
in Guatemala’’ and partly by the Gesellschaft für Tropenornithologie. John
Rappole and two unknown reviewers gave valuable comments and Aerin Jacob
was helpful with English. This study was performed with the current laws of
Guatemala and CONAP authorised the study (No. 139–2001). We also would
like to thank the local people of all nationalities and ethnic groups in the
community of Chelemhá supporting our fieldwork and discussing with us
interesting aspects of conservation and ‘development’.
References
American Ornithologists’ Union 1998. The Checklist of North American birds. American Orni-
thologists’ Union, Washington DC.
American Ornithologists’ Union 2000. 42nd supplement to the American Ornithologists’ Union
check-list of North American birds. Auk 117: 847–858.
American Ornithologists’ Union 2002. 43rd supplement to the American Ornithologists’ Union
check-list of North American birds. Auk 119: 897–906.
[512]
1573
Begon M., Harper J. and Townsend C. 1996. Ecology. Individuals, Populations, and Communities.
Blackwell, London.
Bibby C.J., Burgess N.D. and Hill D.A. 1995. Bird Census Techniques. Neumann, Radebeul.
Bierregaard R. 1990. Avian communities in the understorey of Amazonian forest fragments. In:
Keast A. (ed.), Biogeography and Ecology of Forest Bird Communities. SPB Academic Pub-
lishing, Den Haag, pp. 333–343.
Bierregaard R. and Stouffer P. 1997. Understorey birds and dynamic habitat mosaics in Amazo-
nian rainforests. In: Laurence W. and Bierregaard R. (eds), Tropical Forest Remnants. Uni-
versity of Chicago Press, Chicago, pp. 139–155.
Brooks T.M., Pimm S.L. and Collar N.J. 1997. The extent of deforestation predicts the number of
threatened birds in insular south-east Asia. Conservation Biology 11: 382–394.
Brooks T.M., Tobias J. and Balmford A. 1999a. Deforestation and bird extinctions in the Atlantic
forest. Animal Conservation 2: 211–222.
Brooks T.M., Pimm S.L., Kapos V. and Ravilious C. 1999b. Threat from deforestation to montane
and lowland forest birds and mammals in insular south-east Asia. Journal of Animal Ecology 68:
1061–1078.
Brooks T.M., Pimm S.L. and Oyugi J.O. 1999c. Time lag between deforestation and bird extinction
in tropical forest fragments. Conservation Biology 13: 1140–1150.
Chao A. and Lee S. 1992. Estimating the number of classes via sample coverage. Journal of the
American Statistical Association 87: 210–217.
Colwell R. 2000. EstimateS 6.0: Statistical estimation of species richness and shared species from
samples. Version 6.0. http://viceroy.eeb.uconn.edu/estimates.
Colwell R. and Coddington J. 1994. Estimating terrestrial biodiversity through extrapolation. Philo-
sophical Transactions of the Royal Society of London. Series B: Biological Sciences 345: 101–118.
Dranzoa C. 1998. The avifauna 23 years after logging in Kibale National Park, Uganda. Biodi-
versity and Conservation 7: 777–797.
Edwards E. 1998. A Field Guide to the Birds of Mexico and Adjacent Areas. Belize, Guatemala,
and El Salvador. University of Texas Press, Austin.
Eisermann K. 2000. Avifaunistisch-ökologische Untersuchungen in der Sierra Caquipec, Coban,
Alta Verapaz, Guatemala. Fachhochschule Eberswalde, Eberswalde.
FAO 2001. State of the Worlds Forests. FAO, Roma.
Gaston K.J. and Blackburn T. 2000. Pattern and Process in Macroecology. Blackwell, London.
Gilbert G., Gibbons D.W. and Evans J. 2000. Bird Monitoring Methods. A Manual of Techniques
for Key UK Species. A & C Black Publishers.
Gillespie T.W. 2001. Application of Extinction and Conservation Theories for Forest Birds in
Nicaragua. Conservation Biology 15: 699–709.
Holbech L. 1996. Faunistic diversity and game production contra human activities in the Ghana
high forest zone with references to the western region. Dissertation, University of Copenhagen,
Copenhagen.
Holland S. 2003. Analytic rarefaction 1.3. http://www.uga.edu/ strata/software/.
Howell S.N.G. and Webb S. 1995. A Guide to the Birds of Mexico and Northern Central America.
Oxford University Press, Oxford.
Hughes J., Daily G. and Ehrlich P. 2002. Conservation of tropical forest birds in countryside
habitats. Ecology Letters 5: 121–129.
Hurlbert S.H. 1984. Pseudoreplication and the design of ecological field experiments. Ecological
Monographs 54: 187–211.
Islebe G.A. 1995. High elevation coniferous vegetation of Guatemala: A phytosociological
approach. Vegetatio 116: 7–24.
Johns A. 1992. Vertebrate responses to selective logging: implications for the design of logging
systems. Philosophical Transactions of the Royal Society of London. Series B: Biological Sci-
ences 335: 437–442.
Kappelle M. and Brown A.D. 2001. Bosques Nublados Del Neotrópico. Instituto Nacional de
Biodiversidad, San José.
[513]
1574
[514]
1575
Schulze C.H., Waltert M., Kessler P., Pitopang R., Shahabudin, Veddeler D., Steffan-Dewenter I.,
Mühlenberg M., Gradstein S., Leuschner C. and Tscharntke T. 2004. Biodiversity indicator
groups of tropical land-use systems: comparing plants, birds and insects. Ecological Applica-
tions. Vol. 14(5): 1321–1333.
Shankar-Raman T. 2001. Effect of slash-and-burn shifting cultivation on rainforest birds in Mi-
ozoram, north-east India. Conservation Biology 15: 685–698.
Shankar-Raman T., Rawat G. and Johnsingh A. 1998. Recovery of tropical rainforest avifauna in
relation to vegetation succession following shifting cultivation in Mizoram, north-east India.
Journal of Applied Ecology 35: 214–231.
Shankar-Raman T. and Sukumar R. 2002. Responses of tropical rainforest birds to abandoned
plantations, edges and logged forest in the Western Ghats, India. Animal Conservation 5: 201–
216.
Stattersfield A. and Capper D. 2000. Threatened Birds of the World. BirdLife International, Lynx
Ediciones, London, Barcelona.
Stattersfield A., Crosby M., Long A. and Wege D. 1998. Endemic Bird Areas of the World.
Priorities for Biodiversity and Conservation. Bird Life International, London.
Terborgh J., Robinson S.K., Parker T.A., Munn C. and Pierpont N. 1990. Structure and orga-
nization of an Amazon forest bird community. Ecological Monographs 60: 213–238.
Terborgh J. 1999. Requiem for Nature. Island Press, Washington DC.
Thiollay J.M. 1994a. Disturbance, selective logging and bird diversity: a Neotropical forest study.
Biodiversity and Conservation 6: 1155–1173.
Thiollay J.M. 1994b. Structure, density and rarity in an Amazonian rainforest bird community.
Journal of Tropical Ecology 10: 449–481.
Veblen T. 1976. The urgent need for forest conservation in highland Guatemala. Conservation
Biology 4: 141–154.
Waltert M., Langkau M., Maertens M., Härtel M., Erasmi S. and Mühlenberg M. 2004. Predicting
losses of bird species from deforestation in Central Sulawesi. In: Gerold G., Guhardja E. and
Fremerey M. (eds), Land Use, Nature Conservation and the Stability of Rainforest Margins in
Southeast Asia. Springer, Berlin, Heidelber, pp. 327–338.
Waltert M. and Mühlenberg M. 2001. Zur Beziehung zwischen Abundanz und Habitatqualität:
Fangraten territorialer Gelbbartbülbüls Andropadus latirostris. In: Gottschalk E., Barkow A.,
Mühlenberg M. and Settele J. (eds), Naturschutz und Verhalten. Umweltforschungszentrum
Halle Leipzig, Leipzig, pp. 103–110.
Winker K., Rappole J.H. and Ramos M. 1995. The use of movement data as an assay of habitat
quality. Oecologia 101: 211–216.
World Bank. 2001. World development indicators 2001. Worldbank, Washington, DC.
World Conservation Monitoring Centre. 1992. Global Biodiversity. Status. Chapman & Hall,
London.
[515]
Biodiversity and Conservation (2006) 15:1577–1607 Springer 2006
DOI 10.1007/s10531-005-2352-5
-1
Key words: Agriculture, Anamalai hills, Bird communities, Cardamom, Coffee, Countryside bio-
geography, Fragmentation, Habitat structure and floristics, Landscape matrix, Tropical wet
evergreen forest
Abstract. As large nature reserves occupy only a fraction of the earth’s land surface, conservation
biologists are critically examining the role of private lands, habitat fragments, and plantations for
conservation. This study in a biodiversity hotspot and endemic bird area, the Western Ghats
mountains of India, examined the effects of habitat structure, floristics, and adjacent habitats on
bird communities in shade-coffee and cardamom plantations and tropical rainforest fragments.
Habitat and birds were sampled in 13 sites: six fragments (three relatively isolated and three with
canopy connectivity with adjoining shade-coffee plantations and forests), six plantations differing
in canopy tree species composition (five coffee and one cardamom), and one undisturbed primary
rainforest control site in the Anamalai hills. Around 3300 detections of 6000 individual birds
belonging to 106 species were obtained. The coffee plantations were poorer than rainforest in
rainforest bird species, particularly endemic species, but the rustic cardamom plantation with
diverse, native rainforest shade trees, had bird species richness and abundance comparable to
primary rainforest. Plantations and fragments that adjoined habitats providing greater tree canopy
connectivity supported more rainforest and fewer open-forest bird species and individuals than sites
that lacked such connectivity. These effects were mediated by strong positive effects of vegetation
structure, particularly woody plant variables, cane, and bamboo, on bird community structure.
Bird community composition was however positively correlated only to floristic (tree species)
composition of sites. The maintenance or restoration of habitat structure and (shade) tree species
composition in shade-coffee and cardamom plantations and rainforest fragments can aid in
rainforest bird conservation in the regional landscape.
Introduction
[517]
1578
[518]
1579
Study area
The Western Ghats is a 1600 km long chain of hills running along the west coast
of the Indian Peninsula (8–21 N) and is recognised as a unique biogeographic
province (Mani 1974). Moist forests, including tropical wet evergreen rainforest,
are found largely south of 16 N, particularly south of the Palghat Gap at 11 N,
a region often called the southern Western Ghats (Pascal 1988). The Anamalai
hill ranges are a major conservation area in the southern Western Ghats con-
taining mid-elevation rainforest in the Indira Gandhi Wildlife Sanctuary
(958 km2, 1012¢ N to 1035¢ N and 7649¢ E to 7724¢ E) and in private-owned
fragments on the Valparai plateau (Figure 1). The natural vegetation of this
region, receiving around 3500 mm of rainfall annually, particularly during the
southwest monsoon (June–September), is classified as mid-elevation tropical
wet evergreen forest of the Cullenia–Mesua–Palaquium type (Pascal 1988).
The Valparai plateau contains around 220 km2 of tea, coffee, and cardamom
plantations surrounded by Wildlife Sanctuaries, National Parks, and Reserved
Forest. Clearing of primary rainforest for plantations began in 1896 and was
mostly complete by the 1930s, although some clearing and conversion of coffee
and cardamom to tea plantations continues to the present day (Congreve 1942,
Raman and Mudappa 2003a). The plateau has one town (Valparai) and a
population of over 106,000 people (1991 Census), mostly estate labourers,
scattered across the town and estates. At least 25 rainforest fragments have
been identified so far in and around the Valparai plateau (Umapathy and
Kumar 2000) and additional sites do exist. Besides two large fragments (2000
and 2600 ha) within the Indira Gandhi Wildlife Sanctuary, the remaining
rainforests all occur as fragments of 0.3–650 ha in size, much of which is on
private land. These fragments are vital for conservation as they contain sig-
nificant proportions of the native fauna (Umapathy and Kumar 2000) and
provide landscape-level connectivity between patches critical for wide-ranging
species such as the Great Hornbill (Buceros bicornis, Raman and Mudappa
2003b), Asian elephants Elephas maximus, tigers Panthera tigris, and wild dogs
Cuon alpinus (Kumar et al. 2002).
[519]
1580
[520]
Figure 1. Map of the Indira Gandhi Wildlife Sanctuary showing location of the plantation areas in the Valparai plateau (within dashed lines), some
rainforest fragments (light grey), and reservoirs (stippled) on the Valparai plateau.
1581
Avifauna
Thirteen sites were selected for vegetation and bird sampling: six rainforest
fragments, six plantation sites, and a ‘control’ or reference site containing a
large and relatively undisturbed tract of tropical rainforest in the same ele-
vation range (Table 1). Of the six fragments, three were relatively isolated as
they occurred within tea plantations that had a very sparse canopy of pruned,
non-indigenous silver oak (Grevillea robusta) trees planted at 12 m · 12 m
spacing. The remaining three fragments adjoined shade-coffee estates with
extensive canopy cover. The five shade-coffee plantations differed in the
canopy tree species composition and the cardamom plantation, maintained
by a local tribal settlement, contained a canopy entirely of native rainforest
tree species (rustic cardamom, see Table 1). Three plantation sites adjoined
continuous forest tracts within the Indira Gandhi Wildlife Sanctuary,
whereas the remaining three sites only adjoined smaller fragments that were
in private lands. The control site was an approximately 2600 ha tract of
tropical rainforest (Iyerpadi–Akkamalai complex) within the Indira Gandhi
Wildlife Sanctuary adjoining the plantations on the Valparai plateau. All sites
were within a restricted elevation range of 900–1400 m containing mid-ele-
vation tropical wet evergreen forest vegetation (Pascal 1988). Sites (except
largest fragments and plantations) were mapped by walking around them
with a hand-held GPS (Garmin 12 XL) with the track option activated. The
larger fragments and plantation sites were digitised using a combination of
Survey of India 1:50,000 topographic sheets, GPS tracking, and satellite
imagery. Maps were prepared and areas of sites estimated using MapInfo
Professional software (version 7.0). Areas of plantation sites are approximate
as exact areas and boundaries were not indicated by all the private
companies.
[521]
1582
Table 1. Study sites selected for vegetation and bird sampling in the Valparai plateau and Indira
Gandhi Wildlife Sanctuary (IGWLS), Anamalai hills; RF = rainforest fragment, P = plantation.
Methods
Vegetation sampling
[522]
1583
trees) were laid and an additional 28 random trees were identified to species.
Tree species were identified using available guides (Gamble and Fischer 1935;
Pascal and Ramesh 1997). Shrubs (or in plantations: coffee bushes, cardamom
plants) were counted in 25 plots of 2 m radius and the presence of cane
(Calamus sp.), bamboos, and lianas was recorded within 5 m radius of the
centres of these 25 plots. Elevation was noted at these points using an altimeter.
Canopy variables (height, cover, stratification) and leaf litter depth were
measured at 25 points, evenly spaced 25 m apart as described elsewhere
(Raman and Sukumar 2002).
Bird sampling
I attempted to sample all sites in a relatively uniform and efficient manner over
the winter and breeding season (December to May) when both migrants and
residents were present in the study area. Point counts (Verner 1985; Bibby et al.
1992; Ralph et al. 1995) were used for bird surveys. Point count surveys of
5 min duration were carried out during the first three hours after sunrise when
bird activity was highest (see Raman 2003 for further details). Densities were
estimated using a fixed radius (50 m) approach as they are known to be highly
correlated to variable-radius point count estimates across species (Raman
2003). As some degraded fragment and plantation sites contained relatively
more open vegetation some bias due to detectability differences may have
existed and the results can only be taken as a conservative assessment of the
effects of fragmentation and plantations.
In each site, 30 point count surveys (25 in Sankarankudi cardamom and 26
in Manamboli) were carried out yielding 173–308 detections and an estimated
321–633 individual birds per site. I attempted to ensure independence of data
points to the extent possible by spacing out points and survey days. Successive
points sampled in any day were at least 100 m apart to avoid overlap and
intermediate points were sampled on different days. Although points sampled
on different days overlapped to some extent, the procedure followed ensured
uniform coverage of the site.
Data analysis
For each site, tree density and basal area were calculated using the PCQ
method (Krebs 1989). Average values across replicate samples in each site were
calculated for other habitat variables. Vertical stratification (average number
of strata with foliage) and its coefficient of variation (indexing horizontal
heterogeneity) were calculated following Raman et al. (1998). Tree species
richness was indexed by the number of tree species recorded in the PCQ plots.
The vegetation data was summarised by principal components analysis to
determine fewer uncorrelated components. The factor matrix was rotated by
[523]
1584
[524]
1585
Results
[525]
1586
cleared for planting cardamom (Table 2). Tree density, canopy cover, and litter
depth were highest in primary rainforest, intermediate in fragments, and least
in plantations. Canopy height and basal area were highest in the cardamom
plantation, and shrub density was highest in fragments (Table 3). Notably,
bamboos, canes, and lianas were absent from the plantation sites due to their
elimination during plantation maintenance and weeding.
Principal components analysis of the eleven vegetation variables extracted
two components (PC1 and PC2) accounting for 76.8% of the total variation in
the data-set (Table 4). PC1 was significantly positively correlated to variables
reflecting the density of woody plants, cane, and bamboo. PC2 was positively
correlated to vertical structure and canopy closure and negatively to hori-
zontal patchiness of sites (Table 4). The ordination of the sites using PC scores
shows the primary rainforest and two large fragments (Manamboli and
Andiparai) forming a cluster at the top right indicating a well developed
woody plant community, vertical structure, and canopy closure (Figure 2).
The plantation sites aggregate on the left indicating poorer development of
[526]
1587
Figure 2. Ordination of primary rainforest control (n), fragment (m), and plantation (d) sites in
and around the Valparai plateau using principal components analysis of vegetation variables. Site
codes as in Table 1.
[527]
1588
Point count sampling yielded 3299 bird detections with an estimated 5987
individuals belonging to 106 species across the 13 sites. Of the 106 bird species,
70 (66%) were rainforest birds and 36 (34%) were open-forest (non-rainforest)
birds. Birds detected at least thrice during sampling comprised 78 species,
including 57 (73%) rainforest and 21 (27%) open-forest species. The total
numbers of bird species seen in the four main habitat strata were: 43 (primary
rainforest control – one site), 95 (rainforest fragments – five sites), 76 (shade-
coffee plantations – five sites), and 49 (cardamom plantation – one site). The
percentage of rainforest bird species was highest in the primary rainforest
control (95.3%) and the cardamom plantation under natural shade (89.8%).
More open-forest birds had infiltrated into rainforest fragments and shade-
coffee plantations and the percentage of rainforest bird species in the com-
munity was lower at 70.5% and 59.2%, respectively, in these strata. These
differences in the number of rainforest vs. open-forest species across the four
habitat strata were statistically significant v2 = 34.1, df = 3, p < 0.001).
Bird species richness and abundance: effects of stratum and adjacent habitat
Both stratum (control vs. rainforest fragment vs. plantation) and adjacent
habitat (with low vs. high canopy connectivity) had significant effects on bird
species richness and abundance (MANOVA, p < 0.001, Table 5). The repeated
measure (point), which represented the replicate samples taken within each site,
had no significant direct effect or 2-way or 3-way interactions with the other
main effects (stratum and connectivity) in the multivariate analysis (Table 5).
However, there was a significant 2-way interaction between stratum and
connectivity (p < 0.001, Table 5).
Table 5. Results of multivariate analysis of variance (MANOVA) on the effects of habitat stratum
and connectivity on total, rainforest, and open-forest bird species richness and abundance in point
count samples in the Anamalai hills.
[528]
1589
The effects of habitat stratum and adjacent habitat were tested using analysis of
similarities (ANOSIM, using the Bray–Curtis similarity index between sites
and 1000 random permutations). As sufficient replicate sites were unavailable
for a simultaneous two-way analysis of both factors, each factor was analysed
separately. Bird community composition differed significantly between the two
[529]
1590
Figure 3. Effects of habitat stratum (rainforest fragment, primary rainforest control, and plan-
tations) and adjacent habitat (with low vs. high canopy connectivity) on bird community variables
in the Anamalai hills. Figures illustrate estimated means per point and their standard errors for bird
species richness (panels on the left) and bird abundance (panels on the right).
[530]
1591
The species richness and abundance of bird species categories (all, rainforest,
open-forest, endemics, priority species, rainforest migrants, and all migrants)
was compared between rainforest and plantation sites. Except for open-forest
birds and all migrants, values were lower in plantations than rainforests
(Table 6). Species richness and density of rainforest birds was 43–47% lower in
plantations than in rainforest sites. Although the richness of priority species
and rainforest migrants was not significantly different between the strata, the
abundance of these categories was significantly lower in plantations (Table 6).
Multiple regression analysis indicated a significant positive influence of PC1 on
rainforest and rarefaction bird species richness as well as on the abundance of
all birds, rainforest birds, endemic birds, and priority species (Table 6). The
results indicated that woody plant variables, cane, liana, and bamboo, repre-
sented in PC1 had a generally positive effect on rainforest birds, endemic and
priority species and a negative effect on open-forest bird species richness.
Canopy structural variables represented on PC2 had significant positive effects
on species richness of all birds taken together and for the subset of rainforest
birds as well as on the abundance of priority species and rainforest migrants.
PC2 had a negative effect only on open-forest bird abundance (Table 6). The
level of canopy connectivity in adjacent habitat was significantly positively
related only to richness of priority species. Site area did not appear to have a
significant influence on any of these variables barring a weak positive effect on
migrant species richness (Table 6).
Mantel tests were used to examine the effects of dissimilarity in vegetation
structure and floristics (tree species composition) on bird community dissimi-
larity between sites. Bird community dissimilarity was strongly positively
[531]
1592
Analysis of species habitat use with the deviation index showed that most
open-forest species (residents and migrants) occurred more often than expected
in shade-coffee plantations, and less often than expected in rainforest control,
[532]
1593
Discussion
[533]
1594
Bird community change: effects of habitat structure and tree species composition
[534]
1595
[535]
1596
in these plantations benefited two species of migrant wagtails (Grey and Forest
Wagtails).
The process of incursion of a large number of open-forest species into
rainforest fragments (‘biological infiltration’, Raman 2001) is explained by
alteration of habitat structure as well as the influence of the surrounding
landscape. Fragments, particularly the more-disturbed ones with their patchier
canopy and denser shrubbery, allow the persistence of species that thrive in
open areas and weedy vegetation (e.g., Common Tailorbird, Red-whiskered
Bulbul) due to changes in microhabitat and microclimate. Such open-forest
species derive from the deciduous and thorn forests of the region and such
infiltration into disturbed rainforests is known from many other rainforest
regions (Leck 1979; Daniels et al. 1990; Raman 2001). Dense thickets of
invasive weeds such as Lantana camara within these fragments, with more light
and warmth due to canopy openness, allows the persistence of such species that
do not occur in the cool, dark, evergreen vegetation of undisturbed rainforest
understorey.
Another aspect of conservation relevance to consider is whether these plan-
tations provide resources for birds throughout the year. The data from this
study over the main breeding season only indicate usage as it was not possible to
establish breeding. Plantations that consist mainly of exotic trees may offer few
resources for frugivorous and nectarivorous birds (Greenberg et al. 1997a).
Such plantations may only temporarily support frugivores such as the hornbills
and Pompadour Pigeon that visit the scattered fruiting trees (e.g., Ficus sp.,
Litsea glabrata, and Actinodaphne angustifolia). Nectar-seeking birds such as
sunbirds, drongos, orioles, and Vernal Hanging-Parrot visited flowers of exotics
such as Eucalyptus sp., Grevillea robusta, Erythrina sp., and even coffee bushes
(Crimson-backed Sunbird) in season, besides flowers of native plants. In the
more open coffee plantations, however, an open-forest species (Purple Sunbird)
was often more abundant than the endemic Crimson-backed Sunbird.
[536]
1597
Conservation implications
This study shows that the nature of adjoining habitats affects rainforest bird
communities in tropical rainforest fragments. Specifically, having shade-coffee
rather than tea plantations adjoin fragments has beneficial effects on rainforest
birds, and these effects are probably mediated through the influence of habitat
structure and canopy tree species composition on bird community structure.
Such plantations can thus promote the persistence in the entire landscape of
larger populations of rainforest birds (Beehler et al. 1987; Renjifo 2001).
Individual rainforest birds resident or dispersing from such sites can reduce
the likelihood of chance extinction in fragments through recolonization
(‘rescue effect’, Brown and Kodric-Brown 1977). Increased canopy connec-
tivity in adjoining habitats also has value for shade-coffee plantations. Again,
benefits accrue mainly to rainforest birds: more species and individuals were
supported per unit area in plantations that adjoined continuous forest.
Although there were no completely isolated (surrounded by tea estates) shade-
coffee plantations in the study area, it seems likely that such plantations will
be more depauperate in rainforest birds than those adjoining fragments or
continuous forest. If increased canopy connectivity in adjoining forest habitats
benefits rainforest birds it may also benefit coffee plantation owners. The bird
populations may enable ‘eco-friendly’ or ‘bird-friendly’ coffee certification
(Sherry 2000; Rappole et al. 2003) and reduce insect attack on coffee leaves
(Guatemala: Greenberg et al. 2000), while bees from adjoining forest may
enhance pollination of coffee plants (Costa Rica: Ricketts 2004, Ricketts et al.
2004).
In the study region, ongoing conversion of shade-coffee to tea plantations
driven largely by market forces, is therefore of conservation concern because
tea plantations represent a poorer habitat for rainforest birds and because of
the fallouts for fragments in the landscape. Efforts should be made to halt such
changes while encouraging landowners through tax and other incentives to
promote the relatively more benign form of land use represented by shade-
coffee. Schemes, currently non-existent in India, to certify shade-coffee plan-
tations that are good for birds (Smithsonian Migratory Bird Center 1999) need
to be explored as a means to promote conservation while directly extending
benefits to landowners (Venkatachalam 2005). Such certification should
[537]
1598
Acknowledgements
[538]
Appendix. List of birds detected in point count sampling in rainforest and plantation sites within and adjoining the Indira Gandhi Wildlife Sanctuary in the
Anamalai hills, Western Ghats. The total number of detections of each species, average detections per 25 point counts and deviations from expected detections
(significant values in bold) are presented. Category codes: Res = resident, Mig = Migrant, Pri = priority species, End = endemic to Western Ghats;
Habitat codes: OF = open-forest, RF = rainforest.
6 OF Brown-capped Pygmy Woodpecker Res 6 – 0.57 0.29 – 1.00 0.12 0.04 1.00
Dendrocopos
nanus
7 OF Chestnut-headed Bee-eater Merops Res 10 0.83 0.14 1.00 1.00 0.05 0.68 0.39 0.11
leschenaulti
8 OF Chestnut-tailed Starling Sturnus Res 2 – 0.28 – –
malabaricus
9 OF Common Hoopoe Upupa epops Res 2 – 0.14 0.14 –
10 OF Common Iora Aegithina tiphia Res 10 – 0.85 0.29 2.00 1.00 0.07 0.21 0.43
11 OF Common Tailorbird Orthotomus Res 62 – 2.13 6.71 – 1.00 0.37 0.42 1.00
sutorius
12 OF Greater Coucal Centropus sinensis Res 23 – 0.71 2.57 – 1.00 0.41 0.43 1.00
13 OF Grey-bellied Cuckoo Cacomantis Res 1 – 0.14 – –
passerinus
14 OF Grey-breasted Prinia Prinia Res 18 – 0.85 1.71 – 1.00 0.22 0.37 1.00
hodgsonii
1599
15 OF Jungle Myna Acridotheres fuscus Res 2 – – 0.29 –
Appendix. (Continued).
1600
S. No. Habitat Species Category Detections Detections/25 point counts Deviations
48 RF Eurasian Blackbird Turdus merula Res 16 0.83 1.42 0.71 – -0.19 0.09
49 RF Golden-fronted Leafbird Chloropsis Res 7 – 0.28 0.71 – 1.00 0.29 0.40 1.00
aurifrons
50 RF Greater Flameback Chrysocolaptes lucidus Res 25 1.67 2.13 0.86 2.00 0.07 0.07 0.13 0.00
51 RF Greater Racket-tailed Drongo Dicrurus Res 29 0.83 2.70 0.14 8.00 0.45 0.11 0.80 0.55
paradiseus
52 RF Grey Junglefowl Gallus sonneratii Res 7 – 0.57 0.29 1.00 1.00 0.05 0.04 0.29
53 RF Grey-headed Canary Flycatcher Culicicapa Res 55 13.33 3.41 – 15.00 0.52 0.09 1.00 0.55
ceylonensis
54 RF Hill Myna Gracula religiosa Res 62 5.83 3.98 1.86 14.00 0.11 0.07 0.19 0.48
55 RF Indian Scimitar Babbler Pomatorhinus Res 73 5.00 5.97 3.29 2.00 0.05 0.05 0.01 0.49
horsfieldii
56 RF Orange-headed Thrush Zoothera citrina Res 47 1.67 4.26 1.86 2.00 0.36 0.10 0.06 0.30
57 RF Oriental Honey-Buzzard Pemis Res 1 – 0.14 – –
ptilorhynchus
58 RF Oriental White-Eye Zosterops palpebrosus Res 119 14.17 7.67 6.57 2.00 0.22 0.07 0.11 0.65
1601
Appendix. (Continued).
1602
S. No. Habitat Species Category Detections Detections/25 point counts Deviations
59 RF Pompadour Green Pigeon Treron Res 17 0.83 0.85 0.57 6.00 0.22 0.19 0.14 0.63
pompadora
60 RF Puff-throated Babbler Pellorneum ruficeps Res 35 3.33 3.27 0.71 3.00 0.11 0.12 0.37 0.04
61 RF Red Spurfowl Galloperdix spadicea Res 9 0.83 0.99 0.14 – 0.10 0.20 0.47 1.00
62 RF Scarlet Minivet Pericrocotus flammeus Res 70 2.50 5.82 2.71 7.00 0.36 0.06 0.06 0.11
63 RF Velvet-fronted Nuthatch Sitta frontalis Res 66 5.00 5.54 2.29 5.00 0.00 0.06 0.12 0.02
64 RF Asian Fairy Bluebird Irena puella Pri 38 4.17 2.56 0.29 13.00 0.18 0.05 0.71 0.62
65 RF Black Bulbul Hypsipetes leucocephalus Pri 33 9.17 2.98 – 1.00 0.57 0.10 1.00 0.45
66 RF Black-crested Bulbul Pycnonotus Pri 12 – 1.42 0.14 1.00 1.00 0.23 0.58 0.02
melanicterus
67 RF Common Flameback Dinopium javanense Pri 6 0.83 0.57 – 1.00 0.29 0.12 1.00 0.35
[542]
68 RF Crimson-fronted Barbet Megalaima Pri 20 – 1.42 1.43 – 1.00 0.02 0.24 1.00
rubricapilla
69 RF Dark-fronted Babbler Rhopocichla Pri 21 2.50 2.27 0.14 1.00 0.22 0.19 0.73 0.25
atriceps
70 RF Dollarbird Eurystomus orientalis Pri 1 – – 0.14 –
71 RF Great Hornbill Buceros bicornis Pri 2 0.83 – 0.14 –
72 RF Heart-spotted Woodpecker Hemicircus Pri 4 0.83 0.14 – 2.00
canente
73 RF Large Woodshrike Tephrodornis gularis Pri 17 3.33 0.85 0.86 1.00 0.44 0.19 0.07 0.15
74 RF Little Spiderhunter Arachnothera Pri 69 3.33 7.10 0.14 14.00 0.22 0.16 0.91 0.44
longirostra
75 RF Malabar Trogon Harpactes fasciatus Pri 5 0.83 0.43 – 1.00 0.37 0.07 1.00 0.43
76 RF Malabar Whistling Thrush Myophonus Pri 83 8.33 7.10 2.00 9.00 0.14 0.07 0.29 0.15
horsfieldii
77 RF Mountain Imperial Pigeon Ducula Pri 22 2.50 1.42 0.14 8.00 0.20 0.07 0.74 0.64
badia
78 RF Plain Flowerpecker Dicaeum concolor Pri 290 16.67 20.74 15.71 14.00 0.14 0.02 0.10 0.24
79 RF Rufous Woodpecker Celeus brachyurus Pri 2 – 0.14 – 1.00
80 RF Vernal Hanging Parrot Loriculus Pri 70 6.67 4.83 3.00 7.00 0.11 0.03 0.01 0.11
vernalis
81 RF White-bellied Woodpecker Dryocopus Pri 1 – 0.14 – –
javensis
82 RF White-cheeked Barbet Megalaima Pri 125 11.67 8.95 5.86 7.00 0.10 0.02 0.03 0.17
viridis
83 RF Yellow-browed Bulbul Iole indica Pri 102 15.83 9.38 1.00 10.00 0.34 0.11 0.64 0.10
84 RF Ashy Drongo Dicrurus leucophaeus Mig 35 – 2.98 2.00 – 1.00 0.07 0.13 1.00
85 RF Brown-breasted Flycatcher Muscicapa Mig 3 – 0.14 0.14 1.00
muttui
86 RF Forest Wagtail Dendronanthus indicus Mig 24 – 0.57 2.71 1.00 1.00 0.51 0.44 0.31
87 RF Greenish Warbler Phylloscopus Mig 306 23.33 21.73 15.71 15.00 0.00 0.02 0.08 0.24
trochiloides
88 RF Grey Wagtail Motacilla cinerea Mig 15 – 0.28 1.86 – 1.00 0.59 0.47 1.00
89 RF Indian Blue Robin Luscinia brunnea Mig 48 0.83 5.40 0.71 4.00 0.63 0.21 0.50 0.02
[543]
1603
Appendix. (Continued).
1604
S. No. Habitat Species Category Detections Detections/25 point counts Deviations
98 RF Crimson-backed Sunbird Nectarinia End 156 22.50 12.36 2.29 26.00 0.31 0.03 0.50 0.35
minima
99 RF Grey-headed Bulbul Pycnonotus End 1 – 0.14 – –
priocephalus
100 RF Malabar Grey Hornbill Ocyceros End 33 – 2.41 2.14 1.00 1.00 0.01 0.19 0.45
griseus
101 RF Malabar Parakeet Psittacula End 30 – 3.27 0.71 2.00 1.00 0.19 0.30 0.09
columboides
102 RF Nilgiri Flycatcher Eumyias albicaudata End 29 6.67 2.84 0.14 – 0.50 0.14 0.80 1.00
103 RF Rufous Babbler Turdoides subrufus End 10 – 0.85 0.57 – 1.00 0.07 0.13 1.00
104 RF White-bellied Blue Flycatcher Cyornis End 21 3.33 1.42 – 7.00 0.35 0.04 1.00 0.62
[544]
pallipes
105 RF White-bellied Treepie Dendrocitta End 3 – 0.43 – –
leucogastra
106 RF Wynaad Laughingthrush Garrulax End 2 – 0.14 – 1.00
delesserti
1605
References
Ali S. and Ripley S.D. 1983. Handbook of the Birds of India and Pakistan. Compact edition,
Oxford University Press, Delhi.
Beehler B.M., Raju K.S.R.K. and Ali S. 1987. Avian use of man-disturbed forest habitats in the
Eastern Ghats, India. Ibis 129: 197–211.
Bibby C.J., Burgess N.D. and Hill D.A. 1992. Bird Census Techniques. Academic Press, London.
BirdLife International 2001. Threatened Birds of the World. Lynx Edicions and BirdLife Inter-
national, Barcelona and Cambridge.
Bowman D.M.J.S., Woinarski J.C.Z., Sands D.P.A., Wells A. and McShane V.J. 1990. Slash-and-
burn agriculture in the wet coastal lowlands of Papua New Guinea: response of birds, butterflies
and reptiles. J. Biogeogr. 17: 227–239.
Brown J.H. and Kodric-Brown A. 1977. Turnover rates in insular biogeography: effect of immi-
gration on extinction. Ecology 58: 445–449.
Brown S. and Lugo A.E. 1990. Tropical secondary forests. J. Trop. Ecol. 6: 1–32.
Cincotta R.P., Wisnewski J. and Engelman R. 2000. Human population in the biodiversity hot-
spots. Nature 404: 990–992.
Clarke K.R. and Gorley R.N. 2001. Primer v5: User Manual/Tutorial. PRIMER-E, Plymouth.
Clarke K.R. and Warwick R.M. 1994. Change in Marine Communities: An Approach to Statistical
Analysis and Interpretation. Plymouth Marine Laboratory, Plymouth.
Coffee Board 2001. Database on Coffee. Coffee Board, Bangalore.
Collar N.J., Crosby M.J. and Stattersfield A.J. 1994. Birds to Watch 2: The World list of Threa-
tened Birds. BirdLife International, Cambridge.
Colwell R.K. 1997. EstimateS: Statistical Estimation of Species Richness and Shared Species from
Samples. Version 6.01b. Internet URL: http://viceroy.eeb.uconn.edu/EstimateS.
Colwell R.K. and Coddington J.A. 1994. Estimating terrestrial biodiversity through extrapolation.
Philos. Trans. Roy. Soc. London Series B 345: 101–118.
Congreve H.R.T. 1942. The Anamalais. Associated Printers, Madras.
Daily G.C. 2001. Ecological forecasts. Nature 411: 245.
Daily G.C., Ehrlich P.R. and Sánchez-Azofeifa G.A. 2001. Countryside biogeography: use of
human-dominated habitats by the avifauna of southern Costa Rica. Ecol. Appl. 11: 1–13.
Daniels R.J.R., Hegde M. and Gadgil M. 1990. Birds of man-made habitats: the plantations. Proc.
Ind. Acad. Sci. (Anim. Sci.) 99: 79–89.
Daniels R.J.R., Joshi N.V. and Gadgil M. 1992. On the relationship between bird and woody plant
species diversity in the Uttara Kannada District of south India. Proc. Nat. Acad. Sci. (USA) 89:
5311–5315.
Estrada A., Coates-Estrada R. and Meritt D.Jr. 1997. Anthropogenic landscape changes and avian
diversity at Los Tuxtlas, Mexico. Biodiv. Conserv. 6: 19–43.
Gamble J.S. and Fischer C.E.C. 1935. Flora of the Presidency of Madras. Reprint, 3 volumes.
Bishen Singh Mahendra Pal Singh, Dehra Dun.
Greenberg R., Bichier P., Angon A.C. and Rietsma R. 1997a. Bird populations in shade and sun
coffee plantations in central Guatemala. Conserv. Biol. 11: 448–459.
Greenberg R., Bichier P. and Sterling J. 1997b. Bird populations in rustic and planted shade coffee
plantations of Eastern Chiapas, Mexico. Biotropica 29: 501–514.
Greenberg R., Bichier P., Angon A.C., MacVean C., Perez R. and Cano E. 2000. The impact of
avian insectivory on arthropods and leaf damage in some Guatemalan coffee plantations.
Ecology 81: 1750–1755.
Hemelrijk C.K. 1990. Models of, and tests for, reciprocity, unidirectionality and other social
interaction patterns at a groups level. Anim. Behav. 39: 1013–1029.
Johnson M.D. and Sherry T.W. 2001. Effects of food availability on the distribution of migratory
warblers among habitats in Jamaica. J. Anim. Ecol. 70: 546–560.
Kannan R. 1998. Avifauna of the Anaimalai hills (Western Ghats) of southern India. J. Bombay
Nat. Hist. Soc. 95: 193–214.
[545]
1606
Krebs C.J. 1989. Ecological Methodology. Harper and Row, New York.
Kumar M.A., Singh M., Srivastava S.K., Udhayan A., Kumara H.N. and Sharma A.K. 2002.
Distribution patterns, relative abundance and management of mammals in Indira Gandhi
Wildlife Sanctuary, Tamil Nadu, India. J. Bombay Nat. Hist. Soc. 99: 184–210.
Laurance W.F., Bierregaard R.O.Jr., Gascon C. and others 1997. Tropical forest fragmentation:
synthesis of a diverse and dynamic discipline. In: Laurance W.F. and Bierregaard R.O.Jr. (eds),
Tropical Forest Remnants: Ecology, Management, and Conservation of Fragmented Commu-
nities. University of Chicago Press, Chicago, pp. 502–514.
Leck C.F. 1979. Avian extinctions in an isolated tropical wet forest preserve. Auk 96: 343–352.
Luck G.W. and Daily G.C. 2003. Tropical countryside bird assemblages: richness, composition,
and foraging differ by landscape context. Ecol. Appl. 13: 235–247.
Mani M.S. (ed) 1974. Ecology and Biogeography of India. Dr. W. Junk Publishers, The Hague.
Manly B.F.J. 1994. Multivariate Statistical Methods: A Primer. 2nd edn. Chapman and Hall,
London, UK.
Menon S. and Bawa K.S. 1997. Applications of geographical information systems, remote sensing
and a landscape ecology approach to biodiversity conservation in the Western Ghats. Curr. Sci.
73: 134–145.
Myers N., Mittermeier R.A., Mittermeier C.G., da Fonseca G.A.B. and Kent J. 2000. Biodiversity
hotspots for conservation priorities. Nature 403: 853–858.
Nair S.C. 1991. The Southern Western Ghats: A Biodiversity Conservation Plan. INTACH, New
Delhi.
Norušis M.J. 1990. SPSS/PC+ Statistics 4.0. SPSS Inc., Chicago.
Olson D.M. and Dinerstein E. 1998. The global 200: a representation approach to conserving the
Earth’s most biologically valuable ecoregions. Conserv. Biol. 12: 502–515.
Parthasarathy N. 2001. Changes in forest composition and structure in three sites of tropical
evergreen forest around Sengaltheri, Western Ghats. Curr. Sci. 80: 389–393.
Pascal J.P. 1988. Wet Evergreen Forests of the Western Ghats of India: Ecology, Structure, Flo-
ristic Composition and Succession. Institut Français de Pondichéry, Pondicherry.
Pascal J.P. and Ramesh B.R. 1997. A Field Key to the Trees and Lianas of the Evergreen Forests of
the Western Ghats (India). Institut Français de Pondichéry, Pondicherry.
Perfecto I. and Vandermeer J. 2002. Quality of agroecological matrix in a tropical montane
landscape: ants in coffee plantations in southern Mexico. Conserv. Biol. 16: 174–182.
Perfecto I., Mas A., Dietsch T. and Vandermeer J. 2003. Conservation of biodiversity in coffee
agroecosystems: a tritaxa comparison in southern Mexico. Biodiv. Conserv. 12: 1239–1252.
Ralph C.J., Sauer J.R. and Droege S. (eds) 1995. Monitoring Bird Populations by Point Counts.
USDA, Forest Service, Pacific Southwest Research Station, Albany, California.
Raman T.R.S. 2001.Community Ecology and Conservation of Tropical Rainforest Birds in the
Southern Western Ghats, India. Ph.D. thesis, Indian Institute of Science, Bangalore.
Raman T.R.S. 2003. Assessment of census techniques for inter-specific comparisons of tropical
rainforest bird densities: a field evaluation in the Western Ghats, India. Ibis 145: 9–21.
Raman T.R.S. and Mudappa D. 2003a. Bridging the gap: sharing responsibility for ecological
restoration and wildlife conservation on private lands in the Western Ghats. Social Change 33:
129–141.
Raman T.R.S. and Mudappa D. 2003b. Correlates of hornbill distribution and abundance in
rainforest fragments in the southern Western Ghats, India. Bird Conserv. Int. 13: 199–212.
Raman T.R.S., Rawat G.S. and Johnsingh A.J.T. 1998. Recovery of tropical rainforest avifauna in
relation to vegetation succession following shifting cultivation in Mizoram, north-east India.
J. Appl. Ecol. 35: 214–231.
Raman T.R.S. and Sukumar R. 2002. Responses of tropical rainforest birds to abandoned plan-
tations, edges and logged forest in the Western Ghats, India. Anim. Conserv. 5: 201–216.
Rappole J.H., King D.I. and Rivera J.H.V. 2003. Coffee and conservation. Conserv. Biol. 17:
334–336.
[546]
1607
Renjifo L.M. 2001. Effect of natural and anthropogenic landscape matrices on the abundance of
subandean bird species. Ecol. Appl. 11: 14–31.
Ricketts T.H. 2004. Tropical forest fragments enhance pollinator activity in nearby coffee crops.
Conserv. Biol. 18: 1–10.
Ricketts T.H., Daily G.C., Ehrlich P.R. and Michener C.D. 2004. Economic value of tropical forest
to coffee production. Proc. Nat. Acad. Sci. (USA) 101: 12579–12582.
Shahabuddin G. 1997. Preliminary observations on the role of coffee plantations as avifaunal
refuges in the Palni hills of the Western Ghats. J. Bombay Nat. Hist. Soc. 94: 10–21.
Sherry T.W. 2000. Shade coffee: a good brew even in small doses. Auk 117: 563–568.
Smithsonian Migratory Bird Center 1999. Certified shade-grown coffees. <http://www.si.smbc/
coffee/cafelist.htm>.
Stattersfield A.J., Crosby M.J., Long A.J. and Wege D.C. 1998. Endemic Bird Areas of the World:
Priorities for Biodiversity Conservation. Birdlife International, Cambridge.
Stouffer P.C. and Bierregaard R.O.Jr. 1995a. Use of Amazonian forest fragments by understory
insectivorous birds. Ecology 76: 2429–2445.
Stouffer P.C. and Bierregaard R.O.Jr. 1995b. Effects of fragmentation on understory humming-
birds in Amazonian Brazil. Conserv. Biol. 9: 1085–1094.
Tea Board 2002. Tea Board home page. <http://www.teaindia.org/curr_statistics.html>.
Terborgh J. 1985. Habitat selection in Amazonian birds. In: Cody M.L. (ed.), Habitat Selection in
Birds. Academic Press, Inc., Orlando, pp. 311–340.
Tejeda-Cruz C. and Sutherland W.J. 2004. Bird responses to shade coffee production. Anim.
Conserv. 7: 169–179.
Thiollay J.M. 1995. The role of traditional agroforests in the conservation of rain forest bird
diversity in Sumatra. Conserv. Biol. 9: 335–353.
Turner I.M. and Corlett R.T. 1996. The conservation value of small isolated fragments of lowland
tropical rain forest. Trends Ecol. Evol. 11: 330–333.
Umapathy G. and Kumar A. 2000. The occurrence of arboreal mammals in the rain forest frag-
ments in the Anamalai hills, south India. Biol. Conserv. 92: 311–319.
Venkatachalam L. 2005. Coffee: need to achieve market balance. The Hindu Survey of Indian
Agriculture 2005: 113–115.
Verner J. 1985. Assessment of counting techniques. Curr. Ornithol. 2: 247–302.
Wolda H. 1981. Similarity indices, sample size and diversity. Oecologia 50: 296–302.
Wunderle J.H.Jr. 1999. Avian distribution in Dominican shade coffee plantations: area and habitat
relationships. J. Field Ornithol. 70: 58–70.
Zar J.H. 1999. Biostatistical Analysis. 4th edn. Prentice-Hall, New Jersey.
[547]