You are on page 1of 8

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/265857058

Mahogany Growth and Mortality and the Relation of Growth to Site


Characteristics in a Natural Forest in Quintana Roo, Mexico

Article  in  European Journal of Marketing · October 2014


DOI: 10.5849/forsci.14-031

CITATIONS READS

5 1,401

2 authors:

Patricia Negreros-Castillo Carl W. Mize


Academia Nacional de Ciencias Forestales, Mexico. Iowa State University
60 PUBLICATIONS   1,167 CITATIONS    61 PUBLICATIONS   1,595 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Magnetic soil properties and their applications. Las propiedades magnéticas de los suelos y sus aplicaciones: contaminación, clasificación, cartografía y uso del fuego en la
agricultura View project

Proyecto integral de producción sustentable de Papel Amate en San Pablito Puebla View project

All content following this page was uploaded by Carl W. Mize on 19 November 2014.

The user has requested enhancement of the downloaded file.


For. Sci. 60(5):907–913
FUNDAMENTAL RESEARCH http://dx.doi.org/10.5849/forsci.14-031
Copyright © 2014 Society of American Foresters

silviculture

Mahogany Growth and Mortality and the Relation


of Growth to Site Characteristics in a Natural
Forest in Quintana Roo, Mexico
Patricia Negreros-Castillo and Carl W. Mize

The management of mahogany (Swietenia macrophylla King) in the Yucatan peninsula of Mexico is based on the assumption that diameter at breast height (dbh) growth
averages 0.73 cm yrⴚ1 so mahogany will reach a harvestable dbh of 55 cm in 75 years. In 2002, transects were established in the forested area of Laguna Kaná
in Quintana Roo, Mexico. All mahogany with a dbh > 10.0 cm were measured and marked. In 2008/2009, the trees were remeasured, and tree, soil, and site
characteristics and competition around each tree were recorded. Of the 499 trees studied, average dbh growth of surviving trees was 0.22 cm yrⴚ1, far below the rate
assumed by managers. Only 21 trees grew 0.73 cm yrⴚ1 or more. Dbh growth was significantly correlated to initial dbh and the percentage of the soil that was Box
lu⬘um, varied significantly among soil types and colors and was negatively correlated to stoniness. Growth was not significantly influenced by crown radius, basal area
of competition, crown growing space, topographic position, microrelief, and soil depth (for soils 30 cm in depth or less). Mortality averaged 1% per year and was not
related to initial dbh or initial commercial height. Forest management in Quintana Roo requires new science-based practices that increase the growth of standing trees
and that ensure the establishment and continued recruitment of mahogany reproduction.

Keywords: site quality, Maya Forest, big leaf mahogany

C
ommercially, big-leaf or Honduras mahogany (Swietenia duction in selectively logged tropical forests include fostering ade-
macrophylla King) has been the most important tree species quate regeneration and managing high species diversity and slow
in tropical forests from Mexico to Brazil for centuries growth of commercially important species, such as mahogany
(Lamb 1966) and is one of the most valuable timber species in the (Negreros-Castillo and Mize 2006, Keller et al. 2007). Considerable
world (Mayhew and Newton 1998). It is so important that it could attention has been given to reduced impact logging (RIL) as a way to
be considered a financial keystone species in those countries. When do selective logging in tropical forests (Rice et al. 2001). RIL, al-
forests in some countries do not contain commercial species, they though much preferable to conventional logging, only incidentally
are more likely to be converted to pasture or other more profitable benefits residual forest increment (Wadsworth and Zweede 2006)
land uses (Fredericken and Putz 2003). In Quintana Roo, some- and regeneration.
thing similar happens when forests do not contain mahogany. The Some tree species grow on a wide range of microsites; however,
continuous presence of mahogany in a forest increases the likelihood growth varies with soil conditions, land form, hydrology, and other site
of its existence. factors. In temperate regions, characteristics of sites where tree species,
Mahogany is selectively logged in Mexico and in the rest of Latin especially economically important ones, tend to grow faster are often
America (Whitman et al. 1997, Snook 1998). From a biological well known and are used to profitably manage forests (Lu et al. 2002).
standpoint, primary challenges to achieving sustainable timber pro- In tropical regions, however, such information is less available.

Manuscript received February 23, 2014; accepted July 1, 2014; published online July 24, 2014.
Affiliations: Patricia Negreros-Castillo (patri_nc@yahoo.com), University of Veracruz, Veracruz, Mexico. Carl W. Mize (carlmize@gmail.com), Iowa State Uni-
versity, retired.
Acknowledgments: Special thanks to the community of Laguna Kana⬘ and many of its members who directly participated in the study. Thanks to the Ford and
Hewlett Foundations for research support in 2002 through the Department of Environmental Studies, Florida International University, David Bray,
principal investigator. Thanks to CONACYT and Universidad Veracruzana for research support in 2008 –2009. Thanks to the Organización de Ejidos
Productores Forestales de la Zona Maya (OEPFZM), particularly to Victoria Santos Jiménez, technical director of OEPFZM. Thank you to Imelda
Martínez Salazar, Gabriela Uc Cua, and Juan Damas Damas for their help with the fieldwork in 2008. Thank you with special gratitude to Dr. Frank
Wadsworth, whose advice and guidance were the idea behind this study and contributed to the manuscript. Finally, thanks to Steven Shifley and anonymous
reviewers for helping us markedly improve the manuscript.

Forest Science • October 2014 907


Figure 1. Location of study area, Laguna Kaná, a forestry community.

Growth and/or mortality rates of some tropical tree species de- tropical forest type in Central America. The forest in Laguna Kaná is
pend on edaphic and/or topographic factors (Baraloto et al. 2006, a mosaic forest resulting from past and current human utilization of
2007, Sorensen 2006, Yamada et al. 2007). The growth response of the forest, including centuries of slash and burn agriculture (Gliess-
tropical tree species to light availability has been shown to depend man et al. 1981, Edwards 1986, Gómez-Pompa et al. 1987) and
on soil moisture availability (Poorter and Hayashida-Oliver 2000, periodic hurricane events (Dickinson et al. 2001, Goode and Allen
Fotelli et al. 2001, Wichmann 2001) and nutrients (Palmiotto et al. 2008).
2004, Russo et al. 2005, 2008). Grogan et al. (2003) studied the Annual precipitation averages 1,300 mm, with 75% falling be-
growth and survival of planted nursery-grown mahogany seedlings tween May and November, and temperature averages 26° C with a
on upland and bottomland topographic positions in Parà, Brazil, minimum of 8° C (García 1973, Tamayo 1981). Soils have devel-
and concluded that soil nutrient changes across topographic relief oped from a limestone parent material characteristic of the Yucatan
complements increased light levels that shaped adult distribution Peninsula (Tellez et al. 1980, Wilson 1980). Topographic change
patterns and population structure. across the state is minimal; the terrain often is undulated with gentle
Mahogany is found in tropical forests from medium evergreen grades (1%) over kilometers and many small mounds with heights
forests to medium semi-evergreen forests (Mayhew and Newton rarely exceeding 15 m (Vester and Navarro-Martínez 2005). Soil
1998). Mahogany grows in diverse soil types from soils rich to poor heterogeneity associated with mounds and depressions on generally
in nutrients with humidity regimens from dry to very humid with an flat limestone soils is high (Hernàndez-Xolocotzi 1958, Furley and
excess of moisture and in karst soils rich in calcium and magnesium. Newey 1979, in Belize Bautista and Zinck 2010). Soils on the tops
In general, mahogany develops better on soils low in calcium and of very low rises (a few m) tend to be shallower than those at the
magnesium, which is reflected in the adult distribution pattern bottom (Hernández-Xolocotzi 1958, Furley and Newey 1979,
(Grogan et al. 2003). In Quintana Roo, mahogany occurs preferen- Sánchez-Sánchez and Islebe 2002, Bautista et al. 2003). Black soils
tially on level sites compared to sloped and shows a preference for are generally shallow, occurring on mounds, while red soils are
black soils (Negreros-Castillo and Mize 2012). found at bottom relief positions and are generally deeper, with some
The objectives of this study were to estimate the dbh growth and variation (Bautista and Zinck 2010). Among the main soil types in
mortality rates of mahogany in a natural forest and to evaluate the the region are histosols, lithisols, rendzisols, mollisols, and gleysols
relationship of growth rate to soil-site characteristics and competi- (Bautista and Zinck 2010).
tion with the intention of developing inexpensive and easy-to-apply The study was established in the 11,000 ha forested area owned
recommendations for forest managers in Quintana Roo. by Laguna Kaná. The forest was managed for at least 25 years by a
concessionaire until 1983. The concessionaire focused on harvest-
Methodology ing high-quality mahogany with dbh ⱖ 65 cm. Since 1983, the
This study was conducted in the community forests of Laguna community, with the help of forest consultants, has managed the
Kaná, near the center of the state of Quintana Roo, Mexico (Figure forest with a primary objective of harvesting mahogany with dbh ⱖ
1). The forests of central and southern Quintana Roo are known as 55 cm. No thinning has ever been attempted, and limited, mostly
seasonal tropical forests or medium-height semi-evergreen forests unsuccessful, planting of mahogany seedlings has been done
(Pennington and Sarukhan 1998), examples of the most important (Negreros-Castillo and Mize 2003).

908 Forest Science • October 2014


Table 1. Characteristics of Box luⴕum and Chac luⴕum soils of Quintana Roo, Mexico.a

Soil type (three systems) Maya soil General location within % of Quintana Roo
classification FAO/UNESCO WRB Soil color Characteristics the state with soil
Box lu⬘um (NA) (mollic leptosols) Black Black soils, with little fine earth, shallow, 20–60% NAb NA
gravel and stones, ⬎10% organic matter, well
drained, with or without calcium carbonate.
Chac lu⬘um (cambisols) (NA) Red Less than 40 cm thick. Found on calcareous rocks. Mainly in the northwest 20.2
Efficient internal drainage, superficial drainage and center of the
moderately efficient. Soils do not retain state, on flat areas
humidity. and slopes between
1.5 and 10%
a
Adapted from Sánchez-Sánchez and Islebe (2002) and Bautista and Zinck (2010).
b
WRB ⫽ World Reference Base for Soil Resources. NA, not available.

In 2002, 20 transects of varying lengths (total length ⫽ 119 km, Table 2. Number of trees by dbh class and number of years for
total area ⫽ 448.4 ha) were established at intervals of 500 m across trees in each class to reach 55 cm dbh.
the forest. The transects ran north-south, with the first transect Number of Years to 55
starting 500 m from the northwest corner of the forest and the last Dbh class trees cm dbha
transect starting 100 m from the northeast corner of the forest. Low 10.1–20.0 37 182
areas, called bajos, that frequently flood and can remain flooded for 20.1–30.0 97 136
months were not included in the transects as mahogany seldom 30.1–40.0 131 91
40.1–50.0 143 45
grows there and moving across them is problematic. Each transect 50.1–60.0 61 0
was divided into 50 m long contiguous plots, which were initially ⬎60.1–70.0 26 0
30 m wide. After finishing two transects, width was increased to a
Based on a dbh growth rate ⫽ 0.22 cm yr⫺1 and projecting the midpoint of each
40 m to increase the number of mahogany that would be measured. class.
All mahogany with dbh ⬎ 10.0 cm found along the transects were
tagged using an aluminum nail placed 5 cm below 1.3 m to guide
Table 3. Characteristics of mahogany remeasured in 2008.
remeasurement, then dbh and commercial height were measured.
No trees had buttresses above 1.0 m, so dbh was measured at 1.3 m Number of
on all trees. Characteristic trees Average Minimum Maximum
In 2008/2009, all mahogany was searched for, and for those Dbh in 2002 (cm) 499 39.1 11.1 110.8
found alive, dbh and commercial height were remeasured. Mahog- Dbh growth (cm yr⫺1) 499 0.22 ⫺0.28 1.97
Basal area growth (cm2 yr⫺1) 499 13.3 ⫺26.0 124.1
any that was too small to measure in 2002 but had reached the Commercial height in 2002 (m) 297 9.7 3 42.5
minimum dbh measured, 10.1 cm, was not added to the study Crown radius in 2008 (m) 207 3.4 1.0 8.1
because of limited resources. To evaluate the soil-site characteristics Basal area of competition in 498 9.9 2.3 25.3
around each live mahogany, the tree crown was projected onto the 2008 (m2 ha⫺1)
No. of free to grow quadrants in 496 3.0 0 4
ground and divided in four quarters. In each quarter, percentage 2008 (no. open)
coverage by soil type and soil color, stoniness as a percentage of the
soil surface that was exposed rock, and depth to a solid surface to a
maximum depth of 30 cm were recorded. For the projected crown
tion coefficient were used to evaluate the relationships of dbh
area as a whole, microrelief (flat, concave, convex, and sloped), and
growth and mortality to individual tree and site characteristics.
topography (top of a mound, shoulder of mound, bottom of
Given the possible interrelationships among variables, graphs
mound, flat terrain, and undulating flat terrain) were recorded. Be-
and multiple linear regression were used to evaluate the relation-
cause one of the objectives of this research was to develop low cost
ship of dbh growth to various combinations of tree and site
and relatively easy-to-apply recommendations for forest managers,
variables.
we used color and soil type from the Mayan soil classification sys-
tem (Table 1) (Duch 1988, Sánchez-Sánchez and Islebe 2002,
Negreros-Castillo and Mize 2012), which is well known and cur- Results
rently used by local people. No soil chemical analysis was done. In 2002, 573 mahogany were identified. In 2008/2009, 504
Competition was evaluated using two methods: tree crown free-to- trees were remeasured. Of the missing 69 trees, 15 were eliminated
grow quadrants (FTG) and basal area. For FTG quadrants, trees because they were incorrectly tagged or the original dbh was miss-
were scored from 0 to indicate no growing space (meaning all four ing, 6 were not found and appeared to have been harvested, 8 were
sides of the crown were touching crowns of other trees) to 4 which the wrong species, 4 were not found, and 36 were found but dead.
meant that all four sides were not touching crowns of other trees The density of mahogany was 1.3 trees per ha, and the dbh distri-
(Lamson et al. 1990). Basal area of competition surrounding the bution of the trees is shown in Table 2.
mahogany tree was measured using a 2.3 m2 ha⫺1 factor prism with Of the 504 trees, five were excluded from growth analyses be-
the tree as the plot center. Crown diameter was estimated, when the cause of what we believe were significant errors in measurement or
crown outline was visible, as the average of two perpendicular recording. Of the remaining 499 trees, average dbh growth was 0.22
measurements. cm yr⫺1 (SD ⫽ 0.27) (Table 3). Management plans for the region
T-tests, analysis of variance (ANOVA), and Pearson correla- assume mahogany will reach 55 cm in dbh in 75 years (Carreón et al.

Forest Science • October 2014 909


Table 5. Average dbh growth among five topographic positions.

Number of Average dbh growth


Position trees (cm yr⫺1)
Level, not undulating 49 0.30 aa
Shoulder of mound 66 0.26 a
Level, undulating 344 0.21 a
Bottom of mound 14 0.16 a
Top of mound 25 0.15 a
a
Means with same letters are not significantly different, Tukey’s Honestly signifi-
cant difference (HSD) 5%, SE ⫽ 0.070.

bottom of hills being lower than the other three positions (Table 5),
but microrelief clearly did not influence dbh growth (F ⫽ 1.0, df ⫽
Figure 2. Dbh growth versus initial dbh. Horizontal line is at 0.73 3,493, P ⫽ 0.38).
cm/yr, dbh growth rate assumed by forest managers. For soils less than 30 cm deep, the correlation of dbh growth and
soil depth was r ⫽ 0.01 (P ⫽ 0.88, n ⫽ 404). For another perspec-
Table 4. Average dbh growth among sites that were at least 75% tive, the correlation was calculated separately for 100% Box lu⬘um
one soil color and the soil color was found for at least five and 100% Chak lu⬘um soils and the results were similar (r ⫽ 0.02,
mahogany. P ⫽ 0.72, n ⫽ 318 and r ⫽ ⫺0.01, P ⫽ 0.95, n ⫽ 88, respectively).
Number of Average dbh growth
To include soils more than 30 cm deep, a t-test was done to compare
Soil color trees (cm yr⫺1) the average dbh growth on all soils less than 30 cm deep to all soils
more than 30 cm deep (shallow soils—mean ⫽ 0.20 cm, n ⫽ 406,
Café oscuro 45 0.29 ba
Negro 282 0.24 b SE ⫽ 0.01; deep soils—mean ⫽ 0.28 cm, n ⫽ 93, SE ⫽ 0.03; t ⫽
Café rojizo 28 0.16 a b 2.6, P ⫽ 0.009).
Rojo 11 0.08 a b For a more in-depth evaluation, data were separated into two
Cafo 63 0.05 a
groups based on soil type, 75% or more Box lu⬘um and 75% or
a
Means with same letters are not significantly different, Tukey 5%, SE ⫽ 0.056. more Chak lu⬘um because soil type had the largest impact on dbh
growth. For both groups, graphs (not included) of dbh growth
against the previously mentioned characteristics showed that the
1990), which requires a growth rate of 0.73 cm yr⫺1. Only 21 (4%)
fastest-growing trees grew more commonly on soils with a low per-
trees grew that fast or faster (Figure 2).
centage of stoniness and for the trees with 3 or 4 FTG quadrants and
The relationship of initial dbh, stoniness, basal area of competi-
that the fastest-growing trees showed no relationship to levels of the
tors, and crown radius to dbh growth was evaluated with the Pear-
other variables.
son correlation coefficient. For initial dbh, r ⫽ ⫺0.12 (P ⫽ 0.008,
Multiple regression was used to evaluate the significance of the
n ⫽ 499) (Figure 2), for stoniness r ⫽ ⫺0.11 (P ⫽ 0.018, n ⫽ 498),
relationship of stoniness and number of FTG quadrants to dbh
for basal area of competition r ⫽ 0.00 (P ⫽ 0.98, n ⫽ 498), and for
growth for both groups. Percentage cover was converted to a
crown radius r ⫽ ⫺0.086 (P ⫽ 0.22, N ⫽ 207). ANOVA, consid-
dummy variable with a value of 0 for percentage rock cover of 5% or
ering the five classifications of FTG as treatments, and correlation,
less and a value of 1 for more than 5% cover. FTG was also con-
considering the five classifications as a continuous variable, were
verted to a dummy variable with 0 to 2 quadrants given a value of 0
used to evaluate the relationship of FTG (0 – 4) and dbh growth,
and 3 and 4 a value of 1. For both groups the coefficient for stoniness
and both analyses showed no effect (P ⫽ 0.38 and 0.28,
showed an increase in dbh growth for plots with 5% and less stoni-
respectively).
ness (for Chak lu⬘um P ⬍ 0.001 and for Box lu⬘um P ⫽ 0.052). For
Of the 499 tree plots, the soils for 493 were a mixture of Box
both groups the coefficient for FTG was not significant (for Chak
lu⬘um and Chak lu⬘um soil types only. Using those plots, correla-
lu⬘um P ⫽ 0.76 and for Box lu⬘um P ⫽ 0.34). Although some of the
tion between percentage of the soil that was Box lu⬘um and dbh
trees with 3 or 4 FTG quadrants had the fastest dbh growth, many
growth was 0.22 (P ⬍ 0.001, n ⫽ 493). Of the 376 plots that were
other trees in that group grew very slowly.
at least 75% Box lu⬘um 18 mahogany grew more than 0.73 cm yr⫺1
Mortality was 1% per year with 36 of 555 trees (573 total, de-
while only one mahogany on the 108 plots that were at least 75%
leting six probably harvested, eight wrong species, and four not
Chak lu⬘um grew more than 0.73 cm yr⫺1. Of the eight colors used
found) dying over 6 years. A comparison of the trees that died before
to classify the soils, a few color categories were found infrequently,
the 2008/2009 measurement with those that survived was limited
and many plots had a mixture of colors that would make analysis
because most characteristics evaluated in 2008/2009 were done for
quite complicated. But the majority of plots were mostly one color,
live trees. Based on a t-test, average initial dbhs were not different for
so ANOVA was used to compare the average dbh growth among
the two groups (P ⫽ 0.68) nor were average initial commercial
those plots that were at least 75% one color (F ⫽ 10.3, df ⫽ 4,424,
heights (P ⫽ 0.14).
P ⬍ 0.001) (Table 4). For soils that were 100% Box lu⬘um, café
oscuro and negro were the only soil colors found and for soils that
were 100% Chak lu⬘um, café rojizo, rojo, and cafo were the only soil Discussion
colors found. ANOVA of dbh growth for the five topographic po- Big-leaf mahogany dbh growth rate is quite variable across its
sitions showed an indication of a difference among positions (F ⫽ distribution range. In some locations it grows rapidly (0.6 –2.5 cm
2.1, df ⫽ 4,493, P ⫽ 0.080) with the averages for the top and yr⫺1) (Lamb 1966, Shono and Snook 2006, Grogan et al. 2010),

910 Forest Science • October 2014


while in other locations, like the Yucatan Peninsula, it grows rela- etation is also a barrier for regeneration and recruitment of new
tively slowly (0.16 – 0.38 cm yr⫺1) (Snook 2000, Negreros-Castillo mahogany trees. Therefore, our second recommendation is (i) to
and Mize 2006), which is consistent with 0.22 cm yr⫺1 observed in create slash and burn clearings in the forest and seed or plant
this study. Variability in-growth rate has been attributed to various mahogany into the clearings, preferably planting on soils that are
factors, such as age, size class, rainfall, soil nutrient content, genetic at least 75% Box lu⬘um, (ii) clear seedlings that develop in these
variation, and disturbances (natural and anthropogenic) (Lamb clearings of vines and liberate the mahogany every 2 years for at
1966, Guillison et al. 1996, Whigham et al. 1998, Grogan et al. least 5 years (Negreros-Castillo et al. 2014), (iii) release larger
2003, Shono and Snook 2006, Grogan et al. 2010). The fastest mahogany established in these clearings every 5–10 years as lib-
growth rates have been reported for trees growing on fertile, well- eration has rendered positive results with other species in eastern
drained soils in Belize and Brazil (Lamb 1966, Grogan et al. 2003), Amazonia (Wadsworth and Zweede 2006), (iv) liberate mahog-
as well as in sites with a combination of good soil characteristics and anies already established with the hope of increasing their growth
logging or competition reduction (Shono and Snook 2006, Grogan rates, and (v) establish studies to evaluate the impact of ii, iii,
et al. 2010). The Yucatan peninsula, where Laguna Kana is located, and iv.
has calcareous soils that are often associated with slow tree growth. The observed 1% annual mortality is consistent with observa-
Rainfall decreases from south to north in the peninsula, and Laguna tions in previous studies (Negreros-Castillo and Mize 2006 in Quin-
Kana is near the center of the peninsula, which is near the northern tana Roo and Grogan and Landis 2009 in Pará, Brazil). If mahogany
limit of mahogany in the area. was established in slash and burn agricultural sites and given some
It was surprising that FTG quadrants and basal area of competi- liberation from competing trees and vines, mortality would likely
tion did not significantly impact dbh growth. It is possible that decrease.
rainfall is low enough that moisture competition is a more impor- Only 4% of mahogany grew fast enough to reach the desired 55
tant factor in Quintana Roo than it is in other areas where mahog- cm dbh in 75 years, but undoubtedly, others can grow that fast with
any grows. tending (clearing vines, liberation, and thinning of competition).
Many of the characteristics evaluated had no apparent influence Retaining tropical forests and sustaining their ecosystem services is a
on dbh growth, but mahogany growth was influenced by soil type goal for many countries and organizations. In southern Quintana
and color as other authors have found (Lamb 1966, Grogan et al. Roo, the continuing existence of tropical forests depends on the
2010) and trees on plots with at least 75% Box lu⬘um grew substan- economically viable production of mahogany. Based on this study,
tially faster than trees on plots with at least 75% Chac lu⬘um. Ma- sustaining a productive mahogany resource will require new science-
hogany on soils with 5% or less stoniness grew faster than stonier based practices that increase the growth of standing trees and that
soils. ensure the establishment and continued recruitment of mahogany
With a goal of harvesting trees with a dbh of 55 cm or more, the reproduction.
average growth rate for all trees of 0.22 cm yr⫺1 would require a
commercially impractical rotation of 250 years for this region, and
as shown in Table 2, for trees in the 40 –50 cm dbh class, they will Literature Cited
not be harvestable for 45 years. For trees growing on plots that were
BARALOTO, C., D. BONAL, AND D.E. GOLDBERG. 2006. Differential seed-
at least 75% Box lu⬘um soil with 5% or less stoniness, dbh growth of
ling growth response to soil resource availability among nine neotropical
the 153 trees that met that criteria averaged 0.28 cm yr⫺1 (SE ⫽ tree species. J. Trop. Ecol. 22(5):487– 497.
0.05), which would require a slightly less impractical rotation of 196 BARALOTO, C., F. MORNEAU, D. BONAL, L. BLANC, AND B. FERRY. 2007.
years. Thus, our first recommendation is that the forest owners and Seasonal water stress tolerance and habitat associations within four neo-
their managers reconsider their management plans because their tropical tree genera. Ecol. 88(2):478 – 489.
assumed 75 year rotation is about a third of what it should be based BAUTISTA-ZUÑIGA, F., J. JIMÉNEZ-OSORNIO, J. NAVARRO-ALBERTO,
on the growth rates observed. A. MANU, AND R. LOZANO. 2003. Microrelieve y color del suelo
However, a small percentage of the trees in the study did grow como propiedades de diagnóstico en leptosoles cársticos [Microrelief
0.73 cm yr⫺1 or faster, which shows that mahogany is capable of and soil color as diagnostic properties in carstic leptosols]. Terra.
growing fast enough to achieve 55 cm in 75 years. Mahogany 21(1):1–11.
regeneration that establishes under a natural forest or in a lightly BAUTISTA, F., AND J.A. ZINCK. 2010. Construction of a Yucatec Maya soil
harvested forest grows slowly and often dies within a few years classification and comparison with the WRB framework. J. Ethnobiol.
(Negreros-Castillo et al. 2003, Snook and Negreros-Castillo Ethnomed. 6:7. (11 p.).
2004), but mahogany regeneration obtained by planting seed or CARREÓN, M., H. GALLETTI, AND M.V. SANTOS. 1990. Plan de manejo
seedlings into slash and burn agriculture sites has adequate sur- forestal integral de los bosques del ejido Naranjal Poniente [Integrated forest
management plan of the Naranjal Poniente community]. Organization of
vival and growth (Negreros-Castillo et al. 2003, Snook and
Maya Forestry Ejidos of Quintana Roo (OEPFZM). 150 p.
Negreros-Castillo 2004). The only forest management con-
DICKINSON, M.B., S.M. HERMANN, AND D.F. WHIGHAM. 2001. Low
ducted in Laguna Kana⬘s forest for at least the last 60 years has rates of background canopy-gap disturbance in a seasonally dry forest in
been harvesting of a limited number of trees per hectare about the Yucatan Peninsula with a history of fires and hurricanes. J. Trop.
every 20 years. Thus, many of the mahoganies that were evalu- Ecol. 17(6):895–902.
ated have passed much, if not all, of their lives growing with DUCH, G.J. 1988. La conformación territorial del estado de Yucatán. Los
substantial competition that likely prevented them from devel- componentes del medio físico [Territorial conformation of the Yucatan. The
oping full crowns that would allow them to grow faster than they components of the physical environment]. Centro Regional de la Península
have grown, and silvicultural treatments are only effective when de Yucatán, Universidad Autónoma de Chapingo, Edo. de México,
mahogany trees are small (Verwer et al. 2008). Competing veg- México. 473 p.

Forest Science • October 2014 911


EDWARDS, C.R. 1986. The human impact on the forest in Quintana Roo, NEGREROS-CASTILLO, P., AND C. MIZE. 2006. Stand and species growth of
Mexico. J. For. Hist. 30(2):120 –127. a tropical forest in Quintana Roo, Mexico. J. Sust. For. 23(6):83–96.
FOTELLI, M.N., A. GEßLER, A.D. PEUKE, AND H. RENNENBERG. 2001. NEGREROS-CASTILLO, P., AND C. MIZE. 2012. Soil-site preferences for
Drought affects the competitive interactions between Fagus sylvatica mahogany (Swietenia macrophylla King) in the Yucatan Peninsula. New
seedlings and an early successional species, Rubus fruticosus: Re- For. 44(1):85–99.
sponses of growth, water status and 13C composition. New Phytol. NEGREROS-CASTILLO, P., L. SNOOK, AND C.W. MIZE. 2003. Regenerating
151(2):427– 435. mahogany (Swietenia macrophylla) from seed in Quintana Roo, Mexico:
FREDERICKSEN, T.S., AND F.E. PUTZ. 2003. Silvicultural intensification for The effects of sowing method and clearing treatment. For. Ecol. Manage.
tropical forest conservation. Biodiv. Cons. 12(7):1445–1453. 183(1):351–362.
FURLEY, P.A., AND W.W. NEWEY. 1979. Variations in plant communities PALMIOTTO, P.A., S.J. DAVIES, K.A. VOGT, M.S. ASHTON, D.J. VOGT,
with topography over tropical limestone soils. J. Biogeog. 6(1):1–15. AND P.S. ASHTON. 2004. Soil-related habitat specialization in diptero-
GARCÍA, E. 1973. Modificaciones al sistema de classificación climática de carp rain forest tree species in Borneo. J. Ecol. 92(4):609 – 623.
Köppen [Modification to the Köppen climate classification system]. Instituto PENNINGTON, T.D., AND J. SARUKHAN. 1998. Árboles tropicales de México
de Geografía, Universidad Nacional Autónoma de México, Mexico [Tropical trees of Mexico]. UNAM y Fondo de Cultura Economica,
D.F., Mexico. 71 p. Mexico City, Mexico. 521 p.
GLIESSMAN, S.R., R.E. GARCIA, AND M.A. AMADOR. 1981. The eco- POORTER, L., AND Y. HAYASHIDA-OLIVER. 2000. Effects of seasonal
logical basis for the application of traditional agricultural technology drought on gap and understorey seedlings in a Bolivian moist forest. J.
in the management of tropical agro-ecosystems. Agro-Ecosys. 7(3): Trop. Ecol. 16(4):481– 498.
173–185. RICE, R.E., C.A. SUGAL, S.M. RATAY, AND G.A. FONSECA. 2001. Sustain-
GOODE, L.K., AND M.F. ALLEN. 2008. The impacts of Hurricane Wilma able forest management: A review of conventional wisdom. Center for
on the epiphytes of El Edén Ecological Reserve, Quintana Roo, Mexico Applied Biodiversity Science and Conservation International, No. 3,
1. J. Torrey Bot. Soc. 135(3):377–387. Washington, DC. 29 p.
GÓMEZ-POMPA, A., J.S. FLORES, AND V. SOUSA. 1987. The “Pet Kot:” A RUSSO, S.E., P. BROWN, S. TAN, AND S.J. DAVIES. 2008. Interspecific
man made tropical forest of the Maya. Intercien. 12(1):10 –15. demographic trade-offs and soil-related habitat associations of tree spe-
GROGAN, J., M.S. ASHTON, AND J. GALVÃO. 2003. Big-leaf mahogany cies along resource gradients. J. Ecol. 96(1):192–203.
(Swietenia macrophylla) seedling survival and growth across a topo- RUSSO, S.E., S.J. DAVIES, D.A. KING, AND S. TAN. 2005. Soil-related
graphic gradient in southeast Pará, Brazil. For. Ecol. Manage. 186(1–3): performance variation and distributions of tree species in a Bornean rain
311–326. forest. J. Ecol. 93(5):879 – 889.
GROGAN, J., AND R.M. LANDIS. 2009. Growth history and crown vine SÁNCHEZ-SÁNCHEZ, O., AND G.A. ISLEBE. 2002. Tropical forest commu-
coverage are principal factors influencing growth and mortality rates of nities in southeastern Mexico. Plant Ecol. 158(2):183–200.
big-leaf mahogany Swietenia macrophylla in Brazil. J. Appl. Ecol. 46(6): SHONO, K., AND L. SNOOK. 2006. Growth of big-leaf mahogany (Swi-
1283–1291. etenia macrophylla) in natural forests in Belize. J. Trop. For. Sci.
GROGAN, J., M. SCHULZE, AND J. GALVAO. 2010. Survival, growth and 18(1):66 –73.
reproduction by big-leaf mahogany (Swietenia macrophylla) in open SNOOK, L.K. 1998. Sustaining harvests of mahogany (Swietenia macro-
clearing vs. forested conditions in Brazil. New For. 40(3):335–347. phylla King) from Mexico’s Yucatán forests: Past, present and future.
GULLISON, R.E., S.N. PANFIL, J.J. STROUSE, AND S.P. HUBBELL. 1996. P. 61– 80 in Timber, tourists, and temples: Conservation and develop-
Ecology and management of mahogany in the Chimanes Forest, Beni, ment in the Maya forest of Belize, Guatemala, and Mexico, Primack,
Bolivia. Bot. J. Linn. Soc. 122(1):9 –34. R., D. Bray, H. Galletti, and I. Ponciano (eds.). Island Press, Wash-
HERNÁNDEZ-XOLOCOTZI, E. 1958. La agricultura [Agriculture]. P. 1–57 ington, DC.
in Los recursos naturales del sureste y su aprovechamiento, Beltrán, E. (ed.). SNOOK, L.K. 2000. Regeneration and growth of mahogany (Swietenia mac-
IMRNR, México, D.F., Mexico. rophylla King) in the forests of Quintana Roo, Mexico. Cien. For. Mex.
KELLER, M., G.P. ASNER, G. BLATE, J. MCGLOCKLIN, F. MERRY, M. 25(87):59 –76.
PEÑA-CLAROS, AND J. ZWEEDE. 2007. Timber production in selec- SNOOK, L., AND P. NEGREROS-CASTILLO. 2004. Regenerating mahogany
tively logged tropical forests in South America. Front. Ecol. Env. (Swietenia macrophylla King) in clearings in Mexico’s Maya forest: The
5(1):213–216. effects of clearing method and cleaning of seedlings survival and growth.
LAMB, F.B. 1966. Mahogany of tropical America: Its ecology and management. For. Ecol. Manage. 189(1):143–160.
University of Michigan Press, Ann Arbor, MI. 220 p. SORENSEN, N. 2006. Regeneration and growth of several canopy tree species in
LAMSON, N.I., H.C. SMITH, A.W. PERKEY, AND S.M. BROCK. 1990. the Maya forest of Quintana Roo, Mexico: The role of competition and
Crown release increases growth of crop trees. USDA For. Serv., Res. Pap. microhabitat conditions. Oregon State University, PhD thesis, Corvallis,
NE-635. 8 p. OR. 307 p.
LU, D., E. MORAN, AND P. MAUSEL. 2002. Linking Amazonian secondary TAMAYO, J.L. 1981. Geografía Moderna de México [Modern Geography of
succession forest growth to soil properties. Land Degrad. Develop. Mexico]. Editorial Trillas, Mexico, D.F. Mexico. 400 p.
13(4):331–343. TELLEZ-V., O.E. CABRERI, AND M. SOUSA-SANCHEZ. 1980. Inventario de
MAYHEW, J.E., AND A.C. NEWTON. 1998. The silviculture of mahogany. los recursos vegetales de Quintana Roo [Inventory of the vegetation
CABI Publishing, Wallingford, England. 240 p. resources of Quintana Roo]. P. 97–104 in Memorias del simposio Quin-
NEGREROS-CASTILLO, P., L. CÁMARA-CABRALES, M.S. DEVALL, M.A. tana Roo: Problematica y perspectiva. Instituto de Geografía, UNAM y
MENDOZA-BRISEÑO, C.W. MIZE, AND A. NAVARRO MARTÍNEZ. 2014. Centro de Investigaciones de Quintana Roo (CIQRO) (eds.). Cancun,
Silviculture guide for the mahogany forests of Quintana Roo, Mexico. North Quintana Roo, Mexico.
American Forestry Commission–FAO, Guadalajara, Mexico. 75 p. VERWER, C., M. PEÑA-CLAROS, D. VAN DER STAAK, K. PHLSON-
NEGREROS-CASTILLO, P., AND C. MIZE. 2003. Enrichment planting of Swi- KIEHN, AND F.J. STERK. 2008. Silviculture enhances the recovery of
etenia macrophylla (big leaf mahogany) and Cedrela odorata (Spanish cedar) overexploited mahogany Swietenia macrophylla. J. Appl. Ecol. 45(6):
in Quintana Roo, Mexico. P. 278 –287 in Big-leaf mahogany: Genetics, 1770 –1779.
ecology, and management, Lugo, A.E., J.C. Figueroa Colon, and M. Alayon VESTER, H.M., AND M.A. NAVARRO-MARTÍNEZ. 2005. Ecological issues
(eds.) Ecological Studies, Vol. 159, Springer-Verlag, New York. in community tropical forest management in Quintana Roo, Mexico. P.

912 Forest Science • October 2014


183–213 in The community forests of Mexico: Managing for sustainable age caused by selection logging of mahogany (Swietenia macrophylla) in
landscapes, Bray, D.B., L. Merino-Pérez, and D. Barry (eds.). University northern Belize. For. Ecol. Manage. 92(1):87–96.
of Texas Press, Austin, TX. WICHMANN, L. 2001. Annual variations in competition symmetry in even-
WADSWORTH, F.H., AND J.C. ZWEEDE. 2006. Liberation: Acceptable pro- aged Sitka spruce. Ann. Bot. 88(1):145–151.
duction of tropical forest timber. For. Ecol. Manage. 233(1):45–51. WILSON, E.M. 1980. Physical geography of the Yucatan Peninsula. P. 5– 40
WHIGAN, D.F., J.F. LYNCH, AND M.B. DICKINSON. 1998. Dynamics and in Yucatan: A world apart, Mosley, E.H., and E.D. Terry (eds.). Uni-
ecology of natural and managed forests in Quintana Roo, Mexico. P. versity of Alabama Press, Tuscaloosa, AL.
267–281 in Timber, tourists and temples: Conservation and development in the YAMADA, T., P.A. ZUIDEMA, A. ITOH, T. YAMAKURA, T. OHKUBO, M.
Maya forests of Belize, Guatemala and Mexico, Primack, R.B., D. Bray, H.A. KANZAKI, S. TAN, AND P.S. ASHTON. 2007. Strong habitat preference
Galletti, and I. Ponciano. (eds.). Timber Island Press, Washington, DC. of a tropical rain forest tree does not imply large differences in popula-
WHITMAN, A.A., N.V.L. BROKAW, AND J.M. HAGAN. 1997. Forest dam- tion dynamics across habitats. J. Ecol. 95(2):332–342.

Forest Science • October 2014 913

View publication stats

You might also like