Вы находитесь на странице: 1из 11

vol. 185, no.

3 the american naturalist march 2015

E-Note

A Strong Test of the Maximum Entropy Theory of Ecology

Xiao Xiao,1,* Daniel J. McGlinn,1,2 and Ethan P. White1

1. Department of Biology, Utah State University, Logan, Utah 84322; and Ecology Center, Utah State University, Logan, Utah 84322;
2. Department of Biology, College of Charleston, Charleston, South Carolina 29424
Submitted November 25, 2013; Accepted July 18, 2014; Electronically published January 14, 2015
Online enhancements: appendixes, supplementary figures. Dryad data: http://dx.doi.org/10.5061/dryad.5fn46.

and another based on constraints. With the process-based


abstract: The maximum entropy theory of ecology (METE) is a
unified theory of biodiversity that predicts a large number of mac-
approach, characteristics of the community are captured
roecological patterns using information on only species richness, total by explicitly modeling a few key ecological processes (e.g.,
abundance, and total metabolic rate of the community. We evaluated Hanski and Gyllenberg 1997; Hubbell 2001). While this
four major predictions of METE simultaneously at an unprecedented approach has the potential to directly establish connections
scale using data from 60 globally distributed forest communities in- between patterns and processes, it has been found that the
cluding more than 300,000 individuals and nearly 2,000 species. METE
same empirical patterns can result from different processes
successfully captured 96% and 89% of the variation in the rank dis-
tribution of species abundance and individual size but performed
(Cohen 1968; Pielou 1975), and process-specific param-
poorly when characterizing the size-density relationship and intraspe- eters are often hard to obtain (Hubbell 2001; Jones and
cific distribution of individual size. Specifically, METE predicted a neg- Muller-Landau 2008). Alternatively, the constraint-based
ative correlation between size and species abundance, which is weak approach suggests that many macroecological patterns are
in natural communities. By evaluating multiple predictions with large emergent statistical properties arising from general con-
quantities of data, our study not only identifies a mismatch between
straints on the system, while processes are only indirectly
abundance and body size in METE but also demonstrates the impor-
tance of conducting strong tests of ecological theories. incorporated through their effect on the constraints (e.g.,
Harte 2011; Locey and White 2013). This approach at-
Keywords: biodiversity, body size distributions, macroecology, max- tempts to provide a general explanation of the observed
imum entropy, species abundance distribution, unified theory.
patterns that does not rely on specific processes, which
allows predictions to be made with little detailed infor-
Introduction mation about the system.
One of the newest and most parsimonious constraint-
The structure of ecological communities can be quantified based approaches is the maximum entropy theory of ecol-
using a variety of relationships, including many of the most ogy (METE; Harte et al. 2008, 2009; Harte 2011). METE
well-studied patterns in ecology, such as the distribution adopts the maximum entropy principle from information
of individuals among species (the species abundance dis- theory, which identifies the most likely (least biased) state
tribution [SAD]), the increase in species richness with area of a system given a set of constraints (Jaynes 2003). As-
(the species-area relationship [SAR]), and the distributions suming that the allocation of individuals and energy con-
of energy consumption and body size (Brown 1995; Ro- sumption within a community is constrained by three state
senzweig 1995; McGill et al. 2007; White et al. 2007). With variables (total species richness, total number of individ-
the increasing consensus that these patterns are not fully uals, and total energy consumption), METE makes pre-
independent, a growing number of unified theories have
dictions for the SAD as well as multiple patterns related
been proposed to identify links between the patterns and
to energy use. Spatial patterns such as the SAR and the
unite them under a single framework (e.g., Hanski and
endemics-area relationship can also be predicted with an
Gyllenberg 1997; Hubbell 2001; Harte 2011; for a review,
additional constraint on the area sampled (Harte et al.
see McGill 2010). Among these unified theories there are
2008, 2009; Harte 2011). METE is one of the growing
generally two different approaches, one based on processes
number of theoretical approaches that attempt to synthe-
* Corresponding author; e-mail: xiao@weecology.org.
size traditionally distinct areas of macroecology dealing
with the distributions of individuals and the distributions
Am. Nat. 2015. Vol. 185, pp. E70–E80. 䉷 2015 by The University of Chicago.
0003-0147/2015/18503-55114$15.00. All rights reserved. of energy and biomass (Dewar and Porté 2008; Morlon
DOI: 10.1086/679576 et al. 2009; O’Dwyer et al. 2009) and thus provides a very

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
Strong Test of METE E71

Figure 1: Illustration of the four patterns with data from Barro Colorado Island. A, Species abundance distribution (presented as a rank
abundance distribution); B, individual size distribution (ISD); C, size-density relationship; D, intraspecific ISD of the most abundant species,
Hybanthus prunifolius. Gray circles or bars in each panel represent empirical observations, and the magenta curve represents the maximum
entropy theory of ecology’s prediction. DBH p diameter at breast height.

general characterization of the structure of ecological sys- have shown that METE generally provides good charac-
tems. With no specific assumptions about biological pro- terizations of these patterns across geographical locations
cesses, it can potentially be applied to any community and taxonomic groups (Harte et al. 2008, 2009; Harte
where the values of the state variables can be obtained. 2011; White et al. 2012a; McGlinn et al. 2013). However,
Previous studies have evaluated the performance of these tests are relatively weak as they focus on one pattern
METE with separate data sets for different patterns and at a time (McGill 2003). As a unified theory with multiple

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
E72 The American Naturalist

predictions, METE allows stronger tests to be made by METE predicts can then be derived from R(n, ␧) (see Harte
testing the ability of the theory to characterize multiple 2011 and app. A for a detailed derivation; apps. A–F are
patterns simultaneously for the same data (McGill 2003; available online) and are given by the following four equa-
McGill et al. 2006). In this study, we conduct a strong test tions. The SAD takes the form
of the nonspatial predictions of METE using data from 60
1 ⫺(l1⫹l2 )n
globally distributed forest communities to simultaneously F(n) ≈ e , (3)
evaluate four predictions of the theory (fig. 1), including Cn
the SAD (the distribution of individuals among species) which is an upper-truncated Fisher’s log-series distribu-
and energetic analogs of the individual size distribution tion. Here, l1 and l2 are Lagrange multipliers obtained by
(ISD; the distribution of body size among individuals re- applying the maximum entropy principle with respect to
gardless of their species identity; Enquist and Niklas 2001; the constraints, and C is the proper normalization con-
Muller-Landau et al. 2006), the size-density relationship stant. The individual-level energy distribution (which is
(SDR; the correlation between species abundance and av- the energetic equivalent of the ISD) takes the form
erage individual size within species; Cotgreave 1993), and
the intraspecific ISD (iISD; the distribution of body size S0 e⫺g
W(␧) p
among individuals within a species; Gouws et al. 2011). N0 Z (1 ⫺ e⫺g)2
Our analysis shows mixed support for METE across its
four predictions, with METE successfully capturing the # [1 ⫺ (N0 ⫹ 1)e⫺gN0 ⫹ N0e⫺g(N0⫹1)], (4)
variation in some patterns while failing to do so in others.
where g p l 1 ⫹ l 2 ␧. Conditioned on abundance n, the
We discuss the ecological implications of our findings as
species-level energy distribution (which is the energetic
well as the importance of conducting strong multipattern
equivalent of the iISD) is given by
tests in the evaluation of ecological theories.
nl 2e⫺l2n␧
V(␧Fn) p , (5)
Methods e ⫺ e⫺l2nE 0
⫺l2n

Predicted Patterns of METE which is an exponential distribution with parameter l2n.


The expected value of the iISD V(␧Fn) then gives the av-
METE assumes that allocation of individuals and energy erage species energy distribution (which is the energetic
consumption within a community is constrained by three equivalent of the SDR), that is, the expected average met-
state variables: species richness (S 0), total number of indi- abolic rate (size) for individuals within a species with
viduals (N 0), and total metabolic rate summed over all in- abundance n:
dividuals in the community (E 0; Harte et al. 2008, 2009;
Harte 2011). Define R(n, ␧) as the joint probability that a 1
␧¯ (n) p
species randomly picked from the community has abun-
⫺l2n
nl 2(e ⫺ e⫺l2nE 0)
dance n and an individual randomly picked from such a
# [e⫺l2n(l 2n ⫹ 1) ⫺ e⫺l2nE 0(l 2nE 0 ⫹ 1)]. (6)
species has metabolic rate between (␧, ␧ ⫹ D␧); two con-
straints are then established on the ratio between the state It should be noted that this derivation shows that the iISD
variables: and the SDR are closely related to one another since the
SDR is the expectation of the iISD. As a result, the two

冘冕
E0
N0
N0 patterns are expected to yield similar fits to the theory and
d␧ nR(n, ␧) p , (1) provide similar insights into its performance.
np1 S0
␧p1

which represents the average abundance per species, and Data

冘冕
E0
N0 METE predicts the iISD to be an exponential distribution
E0 (eq. [5]; also see fig. 1D), where the smallest size class is
d␧ n␧R(n, ␧) p , (2)
np1 S0 the most abundant, regardless of species identity or abun-
␧p1
dance. However, most animal species exhibit interior
which represents the average total metabolic rate per spe- modes of adult body size (e.g., Koons et al. 2009; Gouws
cies. Note that the lower limit of individual metabolic rate et al. 2011; but see Harte 2011) and large variation in
is set to be 1, and all measures of metabolic rate are rescaled minimum (and maximum) body size among species as-
accordingly. sociated with these modal values (Gouws et al. 2011). In
The forms of the four macroecological patterns that other words, the body sizes of conspecifics are clustered

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
Strong Test of METE E73

around some intermediate value, while individuals that qualitative effect on our results (app. B). For individuals
are much larger or smaller are rare. Consequently, assem- with multiple stems, we adopted the pipe model to com-
bling all individuals across species in such communities bine the records, that is, D p (冘 d i2)1/2, where di’s were
often yields multimodal ISDs (Thibault et al. 2011), as diameters of individual stems (Ernest et al. 2009). Since
opposed to the monotonically decreasing form predicted metabolic rate scales as D 2, the pipe model preserves the
by METE (eq. [4]; also see fig. 1B). As such, animal com- total area as well as the total metabolic rate for all stems
munities are expected a priori to violate two of the pre- combined.
dictions of METE. Therefore, to ensure that the perfor- We obtained the Lagrange multipliers l1 and l2 in each
mance of METE was not trivially rejected because of the community with inputs S 0, N 0, and E 0 (i.e., the sum over
life-history trait of determinate growth, in our analysis we the rescaled individual metabolic rates; see app. A). Pre-
focused exclusively on trees, which are known to have dictions for the four ecological patterns were obtained
iISDs (Condit et al. 1998) and ISDs (Enquist and Niklas from equations (3)–(6) and further transformed to facil-
2001; Muller-Landau et al. 2006) that are well character- itate comparison with observations. For the SAD and the
ized by monotonically declining distributions and which ISD, we converted the predicted probability distributions
arguably have the greatest prevalence of high-quality (eqq. [3], [4]) to rank distributions of abundance (i.e.,
individual-level size data among indeterminately growing abundance at each rank from the most abundant species
taxonomic groups. to the least abundant species) and size (i.e., scaled met-
We compiled forest plot data from previous publica- abolic rate at each rank from the largest individual to the
tions, publicly available databases, and personal com- smallest individual across all species; Harte et al. 2008,
munications (table 1). All plots have been fully surveyed, 2011; White et al. 2012a), which were compared with the
with size measurements for all individuals above plot-spe- empirical rank distributions of abundance and size. For
cific minimum thresholds. For those plots where surveys the SDR, predicted average metabolic rate was obtained
have been conducted multiple times, we adopted data from from equation (6) for species with abundance n, which
the most recent survey unless otherwise specified (see table was compared with the observed average metabolic rate
1). We excluded records of ferns, palms, and herbs if they for that species. For the iISD, we converted the predicted
existed. Individuals that were dead, not identified to spe- exponential distribution (eq. [5]) into a rank distribution
cies/morphospecies, and/or missing size measurements of individual size for each species and compared the scaled
were excluded. Individuals with size measurements below metabolic rate predicted at each rank to the observed value.
or equal to the designated minimum thresholds were ex- Alternative analyses for the two continuous distributions,
cluded as well, because it is unclear whether these size the ISD and the iISD, did not change our results (app.
classes were thoroughly surveyed. Overall, our analysis en- C).
compassed 60 plots that were at least 1 ha in size and had The explanatory power of METE for each pattern was
a richness of at least 14 (table 1), with 1,943 species/mor- quantified using the coefficient of determination R 2, which
phospecies and 379,022 individuals in total. was calculated as

冘 [log (obs ) ⫺ log


i 10 i 10 (pred i)]2
Analyses R2 p 1 ⫺
冘 [log (obs ) ⫺ log
, (7)
i i 10 (obsi)]2
The scaling relationship between diameter and metabolic
rate can be described with good approximation by met- where obsi and predi were the ith observed value and
abolic theory as B ∝ D 2e⫺E/kT, where B is metabolic rate, METE’s prediction, respectively. Both observed and pre-
D is diameter, T is temperature, E is activation energy, and dicted values were log transformed for homoscedasticity.
k is Boltzmann’s constant (West et al. 1999; Gillooly et al. Note that R 2 measures the proportion of variation in the
2001). Assuming that E is constant across species and that observation explained by the prediction; it is based on the
T is constant within a community, the temperature- 1 : 1 line when the observed values are plotted against the
dependent term e⫺E/kT is constant within a community and predicted values, not the regression line. Thus, it is possible
can be dropped when the metabolic rates of individuals for R 2 to be negative, which is an indication that the pre-
are rescaled. We thus used (D/Dmin)2 as the surrogate for diction is worse than taking the average of the observation.
individual metabolic rate, where Dmin is the diameter of Simulation from null uniform distributions were con-
the smallest individual in the community, which sets the ducted to confirm that ranking both predicted and ob-
minimal individual metabolic rate to be 1 following served values did not lead to spuriously high R 2 values
METE’s assumption (see eq. [2]). Applying alternative (app. D).
models that more accurately capture nonlinearities be- While R 2 between predicted and observed values pro-
tween diameter, mass and metabolic rate did not have any vides an intuitive measure of the predictive power of the

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
Table 1: Summary of data sets
Area of
individual
Data set Description plots (ha) No. plots Survey year References
a
Serimbu Tropical rain forest 1 2 1995 Kohyama et al. 2001, 2003; Lopez-Gonzalez
2009, 2011
La Selva Tropical wet forest 2.24 5 2009 Baribault et al. 2011a, 2011b
ACA Amazon Forest Inventories Tropical moist forest 1 1 2000–2001 Pitman et al. 2005
Barro Colorado Island Tropical moist forest 50 1 2010 Condit 1998b; Hubbell et al. 1999, 2005
DeWalt Bolivia forest plots Tropical moist forest 1 2 NA DeWalt et al. 1999
Lahei Tropical moist forest 1 3 1998 Nishimura and Suzuki 2001; Nishimura et al.
2006; Lopez-Gonzalez 2009, 2011
Luquillo Tropical moist forest 16 1 1994–1996b Zimmerman et al. 1994; Thompson et al. 2002
Sherman Tropical moist forest 5.96 1 1999 Condit 1998a; Pyke et al. 2001; Condit et al.
2004
Cocoli Tropical moist forest 4 1 1998 Condit 1998a; Pyke et al. 2001; Condit et al.
2004
Western Ghats Wet evergreen/moist/dry deciduous forests 1 34 1996–1997 Ramesh et al. 2010
UCSC FERP Mediterranean mixed evergreen forest 6 1 2007 Gilbert et al. 2010
Shirakami Beech forest 1 2 2006 Nakashizuka et al. 2003; Lopez-Gonzalez
2009, 2011
Oosting Hardwood forest 6.55 1 1989 Reed et al. 1993; Palmer et al. 2007
North Carolina forest plots Mixed hardwoods/pine forest 1.3–5.65 5 1990–1993c Peet and Christensen 1987; McDonald et al.
2002; Xi et al. 2008
Note: NA p not available, UCSC FERP p University of California, Santa Cruz, Forest Ecology Research Plot.
a
One plot has a more recent survey in 1998; however, it lacks species identifiers.
b
We chose census 2 because information for multiple stems is not available in census 3, and the unit of diameter is unclear in census 4. Data from both parts a and b are included.
c
We chose a survey individually for each plot on the basis of expert opinion to minimize the effect of hurricane disturbance.

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
Strong Test of METE E75

theory, it ignores the variation that can arise from random ISD is not an artifact of ranking the data (figs. C1, D1;
sampling even when the predicted distribution is valid. To figs. B1–B5, C1–C3, D1, E1, E2 are available online). Re-
address this issue, we conducted a bootstrap analysis, sults from bootstrap analysis (app. E) are also largely con-
where we drew 500 random samples from the predicted sistent with the direct interpretation of the goodness-of-
distribution for each pattern (eq. [3] for the SAD, eq. [4] fit statistics. METE provides comparable characterization
for the ISD, and eq. [5] for the SDR and the iISD) and for the empirical and the bootstrap SADs in most com-
examined the fit of the theory to the bootstrap samples munities, while its fit to the empirical SDRs and the iISDs
using both R 2 and the Kolmogorov-Smirnov statistic (see is consistently worse than that to the bootstrap samples
app. E for details). If METE fits the empirical data as well (fig. E2). For the ISD, however, the analysis reveals that
as it fits the bootstrap samples, then the theory matches METE characterizes bootstrap samples consistently better
the data, and the residual variation is consistent with ran- than its fit to empirical data (fig. E2), which implies that
dom sampling. If instead METE fits the bootstrap samples the empirical ISD significantly deviates from METE’s pre-
better than the empirical data, it indicates that there are diction despite the theory’s ability to capture the general
meaningful deviations of empirical data from the theory’s shape of the pattern (fig. 2B; supplementary figures). This
predictions. By comparing the fits to empirical data to is consistent with model comparison in appendix F, where
those for data simulated from the theory, this analysis we show that alternative models provide a better fit to the
provides additional insights into patterns like the SAD that distribution (table F1).
are expected to be fit well by many theories (Connolly et
al. 2009; Locey and White 2013) and into patterns like the
Discussion
iISD, where large amounts of variation about the predicted
values may be expected due to sampling. Macroecological theories increasingly attempt to make
Python code to replicate our analyses together with a predictions across numerous ecological patterns (McGill
processed subset of data sets are deposited in the Dryad 2010) by either directly modeling ecological processes or
Digital Repository: http://dx.doi.org/10.5061/dryad.5fn46 imposing constraints on the system. Among the con-
(Xiao et al. 2014).1 Data included in the deposit are spe- straint-based theories, METE is unique in that it makes
cifically designed for the replication of our analyses and simultaneous predictions for two distinct sets of ecological
may lack spatial/temporal components or other useful in- patterns, synthesizing traditionally separate areas of mac-
formation in the original data. Readers interested in using roecology dealing with distributions of individuals and
the data for purposes other than replicating our analyses distributions related to body size and energy use (see also
are advised to obtain the raw data from the original Dewar and Porté 2008; Morlon et al. 2009; O’Dwyer et
sources. al. 2009). Using only information on species richness, total
abundance, and total energy use as inputs, METE attempts
to characterize various aspects of community structure
Results without additional tunable parameters or assumptions,
The results for all forest plots combined are summarized making it one of the most parsimonious of the current
in figure 2, with observations plotted against predictions unified theories.
for each macroecological pattern. METE provides excel- Our analysis shows that METE accurately captures the
lent predictions for the SAD (R 2 p 0.96) and the ISD general shape of the SAD (allocation of individuals among
(R 2 p 0.89), although the largest size classes deviate species) and the ISD (allocation of energy/biomass among
slightly but consistently in the ISD. However, the SDR individuals) within and among 60 forest communities (fig.
(R 2 p ⫺2.24) and the iISD (R 2 p 0.15) are not well char- 2A, 2B; supplementary figures). The SAD and the ISD are
acterized by the theory. among the most well-studied patterns in ecology, and nu-
Further examination of the four macroecological pat- merous models exist for both patterns. For instance, with
terns within each community (see the supplementary fig- metabolic theory and demographic equilibrium models,
ures, available online; also see insets in fig. 2) confirms Muller-Landau et al. (2006) identified four possible pre-
METE’s ability to consistently characterize the SAD (all dictions for the ISD under different assumptions of growth
R 2 1 0.60, 59 of 60 R 2 1 0.8) and the ISD (all R 2 1 0.48, and mortality rates. For the SAD, more than 20 models
49 of 60 R 2 1 0.8) as well as its inadequacy in character- have been proposed (Marquet et al. 2003; McGill et al.
izing the SDR (all R 2 ! 0) and the iISD (maximal R 2 p 2007), ranging from purely statistical to mechanistic.
0.30, 49 of 60 R 2 ! 0). The high R 2 for the SAD and the Our study demonstrates METE’s high predictive power
for these two patterns, but it does not imply that it is the
1
Code that appears in The American Naturalist is provided as a convenience best model when each pattern is considered independently.
to the readers. It has not necessarily been tested as part of the peer review. Indeed, our results reveal a consistent departure of indi-

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
E76 The American Naturalist

Figure 2: Maximum entropy theory of ecology’s predictions plotted against empirical observations across 60 communities for the species
abundance distribution (A; each data point is the abundance of a species at a single rank in one community), the individual size distribution
(ISD; B; each data point is the metabolic rate of an individual at a single rank in one community), the size-density relationship (SDR; C; each
data point is the average metabolic rate within one species in one community), and the intraspecific ISD (iISD; D; each data point is the
metabolic rate of an individual at a single rank belonging to a specific species in one community). The diagonal black line in each panel is the
1 : 1 line. The points are color coded to reflect the density of neighboring points, with warm (red) colors representing higher densities and cold
(blue) colors representing lower densities. Each inset reflects the distribution of R 2 among 60 communities from below 0 (left) to 1 (right).
DBH p diameter at breast height.

viduals in the largest size class from the ISD predicted by such deviation is more severe than expected from the effect
METE (fig. 2B, supplementary figures), which may result of random sampling alone. The discrepancy between the
from mortality unrelated to energy use (Muller-Landau et high R 2 of the ISD both within and across communities
al. 2006). Bootstrap analysis (app. E) further confirms that and the seemingly poor fit of the pattern revealed by boot-

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
Strong Test of METE E77

strapping results from the two different ways that goodness muth’s energetic equivalence rule (Damuth 1981), where
of fit is evaluated by the two analyses. While METE is able the total energy consumption within a species does not
to predict the rank size of individuals (fig. 2B; supple- depend on species identity or abundance (Harte et al. 2008;
mentary figures) as well as the relative frequency of size Harte 2011). While Damuth’s rule has been argued to apply
bins (fig. C1) with high accuracy (illustrated by the high at global scales (Damuth 1981; White et al. 2007), our results
R 2 between predicted and observed values), the empirical indicate that it does not hold locally, in concordance with
ISDs are still significantly different from the predicted dis- a number of previous studies (Brown and Maurer 1987;
tribution (illustrated by higher deviation of empirical data Blackburn and Gaston 1997; White et al. 2007).
from the predicted form when compared with bootstrap The consistency of our results across 60 forest com-
samples). Indeed, while METE has been shown to fre- munities (as well as confirmative evidence from a con-
quently outperform the most common model of the SAD current study of a single herbaceous plant community;
(the lognormal) for a variety of taxonomic groups, in- Newman et al. 2014) provides strong evidence for METE’s
cluding plants (White et al. 2012a), model comparisons mixed performance among the four macroecological pat-
for the ISD using Akaike’s information criterion suggest terns. However, several limitations of the study are worth
that the maximum likelihood Weibull distribution (one of noting. First, we analyzed only a single taxonomic group
the distributions for tree diameter in Muller-Landau et al. (trees). This was in part because individual-level size data
2006) almost always outperforms METE (although METE’s collected in standardized ways is available for a large num-
performance is comparable to that of the other two distri- ber of tree communities and in part based on a prior
butions, the exponential and the Pareto; see app. F). knowledge that the form of the ISD and the iISD (Condit
Quantitatively comparing theories that make multiple et al. 1998; Enquist and Niklas 2001; Muller-Landau et al.
predictions is challenging, and there is no general approach 2006) had a reasonable chance of being well characterized
for properly comparing models that make different num- by the theory (see “Methods”). While we know that the
bers of predictions. When comparing general theories to SAD predictions of the theory perform well in general
single prediction models with multiple tunable parameters, (White et al. 2012a), further tests are necessary to deter-
it is not surprising that theories such as METE fail to mine whether the simultaneous good fit of the ISD pre-
provide the best quantitative fit (White et al. 2012b). How- dictions is supported in other taxonomic groups. There is
ever, as a constraint-based unified theory, METE’s strength some evidence that this result holds in invertebrate com-
lies in its ability to link together ecological phenomena munities (Harte 2011). Second, we estimated the metabolic
that were previously considered distinct and to make pre- rate of individuals based on predictions of metabolic the-
dictions based on first principles with minimal inputs. The ory rather than direct measurement. While our results
general agreement between METE’s predictions and the were not sensitive to the use of other equations used for
observed SAD and ISD (as measured by the R 2 for the estimating metabolic rate (app. B), it is possible that di-
rank distributions) supports the notion that the majority rectly measured metabolic rates could result in different
of variation in these macroecological patterns can be char- fits to the theory (but see Newman et al. 2014, which
acterized by variation in the state variables S 0, N 0, and E0 adopts a different method to obtain metabolic rate yet
alone (Harte 2011; Supp et al. 2012; White et al. 2012a). reaches similar conclusions).
While METE performs well in characterizing the SAD Models and theories can be evaluated at multiple levels
and the ISD, it performs poorly when predicting the dis- that yield different strengths of inference (McGill 2003;
tribution of energy at the species level (fig. 2C, 2D; sup- McGill et al. 2006), progressing from matching theory to
plementary figures). This is not that surprising given that empirical observations on a single pattern, testing against
the iISD and the SDR (which is the expectation of the a null hypothesis, evaluating multiple a priori predictions,
iISD) provide a more detailed perspective on the com- and eventually comparing between multiple competing
munity structure by examining the intercorrelation of models. With quantitative predictions on various ecolog-
abundance and size. The deviations of the empirical pat- ical patterns, METE and other unified theories allow for
terns from the predictions reveal a mismatch between the simultaneous examination of multiple predictions, which
predicted metabolic rate of individuals and their species’ provides a much stronger test than curve fitting for a single
abundances. METE predicts a monotonically decreasing pattern and can often reveal important insight into the-
relationship between species abundance and average in- ories that are otherwise overlooked by single-pattern tests
traspecific metabolic rate, that is, species with higher abun- (e.g., Adler 2004). As a comprehensive analysis of the per-
dance are also smaller in size on average and are more formance of METE in predicting abundance and energy
likely to contain smaller individuals (eqq. [5, 6]; fig. 1C). distributions in the same data sets, our study demonstrates
Evaluating the total (instead of average) intraspecific met- the importance of moving toward stronger tests in ecology,
abolic rate, this relationship translates roughly into Da- especially when multiple intercorrelated predictions are

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
E78 The American Naturalist

available; while previous studies have shown that METE to make our conclusions more robust. N. G. Swenson
does an impressive job characterizing a single pattern provided data for wood density in the Luquillo forest plot
(White et al. 2012a; McGlinn et al. 2013), concurrently and gave insightful comments. R. K. Peet provided data
evaluating all predictions of the theory identifies a slight for the North Carolina forest plots. The Serimbu (provided
yet consistent discrepancy between the observed and the by T. Kohyama), Lahei (provided by T. B. Nishimura), and
predicted size distribution as well as a mismatch between Shirakami (provided by T. Nakashizuka) data sets were
species’ abundance and individual size. obtained from the PlotNet Forest Database. The ACA Am-
That METE fails to provide good characterization of all azon Forest Inventories (provided by N. Pitman) and
four patterns of community structure and performs more DeWalt Bolivia (provided by S. DeWalt) data sets where
poorly than alternative models in some cases can be in- obtained from SALVIAS (Synthesis and Analysis of Local
terpreted in two ways. First, the aspects of community Vegetation Inventories across Scales). The Barro Colorado
structure that are poorly characterized by the theory may Island Forest Dynamics Research Project was made pos-
be more adequately characterized by explicitly modeling sible by US National Science Foundation (NSF) grants
ecological processes. For example, O’Dwyer et al. (2009) to S. P. Hubbell (DEB-0640386, DEB-0425651, DEB-
has developed a model that incorporates individual 0346488, DEB-0129874, DEB-00753102, DEB-9909347,
demographic rates of birth, death, and growth, which like- DEB-9615226, DEB-9615226, DEB-9405933, DEB-
wise yields predictions of abundance and body size dis- 9221033, DEB-9100058, DEB-8906869, DEB-8605042,
tributions. It is worth noting, however, that the process- DEB-8206992, and DEB-7922197); by support from the
based approach and the constraint-based approach do not Center for Tropical Forest Science, the Smithsonian Trop-
have to be mutually exclusive. While O’Dwyer et al. (2009) ical Research Institute, the John D. and Catherine T. Mac-
suggested that size-related patterns may reflect ecological Arthur Foundation, the Mellon Foundation, the Small
processes, the agreement between their model and METE World Institute Fund, and numerous private individuals;
in the predicted SAD (both log series) as well as METE’s and through the hard work of more than 100 people from
performance for the ISD support the idea that information 10 countries over the last 2 decades. The University of
in the underlying processes can be summarized in con- California, Santa Cruz (UCSC), Forest Ecology Research
straints alone for some macroecological patterns. Alter- Plot was made possible by NSF grants to G. S. Gilbert
natively, the constraint-based approach may be sufficient (DEB-0515520 and DEB-084259), the Pepper-Giberson
in characterizing patterns of abundance and body size, but Chair Fund, the University of California, and the hard
the current form of METE may be incorrect. Specifically, work of dozens of UCSC students. These two projects are
the limitations revealed in our analyses may be remedied part of the Center for Tropical Forest Science, a global
either by relaxing the current constraints to remove the network of large-scale demographic tree plots. The Lu-
implicit negative correlation between species-level average quillo Experimental Forest Long-Term Ecological Research
body size and abundance (fig. 1C) from the theory or by Program was supported by grants BSR-8811902, DEB-
adding additional constraints to the system so that ener- 9411973, DEB-0080538, DEB-0218039, DEB-0620910, and
getic equivalence among species no longer holds (Harte DEB-0963447 from the NSF to the Institute for Tropical
and Newman 2014). While the success of METE in char- Ecosystem Studies, University of Puerto Rico, and to the
acterizing the general shape of the SAD and the ISD adds International Institute of Tropical Forestry, USDA Forest
to the growing support for the constraint-based approach Service, as part of the Luquillo Long-Term Ecological Re-
for studying macroecological patterns, further work is search Program. Funds were contributed for the 2000 cen-
clearly needed to develop unified theories for community sus by the Andrew Mellon Foundation and by the Center
structure whether they are based on specific biological pro- for Tropical Forest Science. The US Forest Service (De-
cesses or emergent statistical properties. partment of Agriculture) and the University of Puerto Rico
gave additional support. We also thank the many techni-
cians, volunteers, and interns who have contributed to data
collection in the field. This research was supported by a
Acknowledgments
CAREER award from the NSF to E.P.W. (DEB-0953694).
We thank J. Harte, E. Newman, and the rest of the Harte
laboratory as well as members of the Weecology laboratory
for extensive feedback on this research, for general insights Literature Cited
into MaxEnt, and for being incredibly supportive of our Adler, P. B. 2004. Neutral models fail to reproduce observed species-
efforts to evaluate the maximum entropy theory of ecology. area and species-time relationships in Kansas grasslands. Ecology
We thank S. R. Connolly and two anonymous reviewers 85:1265–1272.
for their extremely helpful comments, which have helped Baribault, T. W., R. K. Kobe, and A. O. Finley. 2011a. Tropical tree

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
Strong Test of METE E79

growth is correlated with soil phosphorus, potassium, and calcium, Harte, J. 2011. Maximum entropy and ecology: a theory of abun-
though not for legumes. Ecological Monographs 82:189–203. dance, distribution, and energetics. Oxford University Press,
———. 2011b. Data from: Tropical tree growth is correlated with Oxford.
soil phosphorus, potassium, and calcium, though not for legumes. Harte, J., and E. A. Newman. 2014. Maximum information entropy:
Ecological Monographs 82:189–203, Dryad Digital Repository, a foundation for ecological theory. Trends in Ecology and Evo-
http://dx.doi.org/10.5061/dryad.r9p70. lution 29:384–389.
Blackburn, T. M., and K. J. Gaston. 1997. A critical assessment of Harte, J., A. B. Smith, and D. Storch. 2009. Biodiversity scales from
the form of the interspecific relationship between abundance and plots to biomes with a universal species-area curve. Ecology Letters
body size in animals. Journal of Animal Ecology 66:233–249. 12:789–797.
Brown, J. H. 1995. Macroecology. University Of Chicago Press, Harte, J., T. Zillio, E. Conlisk, and A. B. Smith. 2008. Maximum
Chicago. entropy and the state-variable approach to macroecology. Ecology
Brown, J. H., and B. A. Maurer. 1987. Evolution of species assem- 89:2700–2711.
blages: effects of energetic constraints and species dynamics on the Hubbell, S. P. 2001. The unified neutral theory of biodiversity and
diversification of the North American avifauna. American Natu- biogeography. Princeton University Press, Princeton, NJ.
ralist 130:1–17. Hubbell, S. P., R. Condit, and R. B. Foster. 2005. Barro Colorado
Cohen, J. E. 1968. Alternate derivations of a species-abundance re- Forest Census plot data. http://ctfs.arnarb.harvard.edu/webatlas
lation. American Naturalist 102:165–172. /datasets/bci/. Accessed April 9, 2012.
Condit, R. 1998a. Ecological implications of changes in drought pat- Hubbell, S. P., R. B. Foster, S. T. O’Brien, K. E. Harms, R. Condit,
terns: shifts in forest composition in Panama. Climatic Change B. Wechsler, S. J. Wright, et al. 1999. Light-gap disturbances, re-
39:413–427. cruitment limitation, and tree diversity in a Neotropical forest.
———. 1998b. Tropical forest census plots. Springer, Berlin, and Science 283:554–557.
R. G. Landes, Georgetown, TX. Jaynes, E. T. 2003. Probability theory: the logic of science. G. L.
Condit, R., S. Aguilar, A. Hernández, R. Pérez, S. Lao, G. Angehr, Bretthorst, ed. Cambridge University Press, Cambridge.
S. P. Hubbell, et al. 2004. Tropical forest dynamics across a rainfall Jones, F. A., and H. C. Muller-Landau. 2008. Measuring long-distance
gradient and the impact of an El Niño dry season. Journal of seed dispersal in complex natural environments: an evaluation and
Tropical Ecology 20:51–72. integration of classical and genetic methods. Journal of Ecology
Condit, R., R. Sukumar, S. P. Hubbell, and R. B. Foster. 1998. Pre- 96:642–652.
dicting population trends from size distributions: a direct test in Kohyama, T., E. Suzuki, T. Partomihardjo, and T. Yamada. 2001.
a tropical tree community. American Naturalist 152:495–509. Dynamic steady state of patch-mosaic tree size structure of a mixed
Connolly, S. R., M. Dornelas, D. R. Bellwood, and T. P. Hughes. dipterocarp forest regulated by local crowding. Ecological Research
2009. Testing species abundance models: a new bootstrap approach 16:85–98.
applied to Indo-Pacific coral reefs. Ecology 90:3138–3149. Kohyama, T., E. Suzuki, T. Partomihardjo, T. Yamada, and T. Kubo.
Cotgreave, P. 1993. The relationship between body size and popu- 2003. Tree species differentiation in growth, recruitment and al-
lation abundance in animals. Trends in Ecology and Evolution 8: lometry in relation to maximum height in a Bornean mixed dip-
244–248. terocarp forest. Journal of Ecology 91:797–806.
Damuth, J. 1981. Population density and body size in mammals. Koons, D. N., R. D. Birkhead, S. M. Boback, M. I. Williams, and
Nature 290:699–700. M. P. Greene. 2009. The effect of body size on cottonmouth (Ag-
DeWalt, S. J., G. Bourdy, L. R. Chávez de Michel, and C. Quenevo. kistrodon piscivorus) survival, recapture probability, and behavior
1999. Ethnobotany of the Tacana: quantitative inventories of two in an Alabama swamp. Herpetological Conservation and Biology
permanent plots of northwestern Bolivia. Economic Botany 53: 4:221–235.
237–260. Locey, K. J., and E. P. White. 2013. How species richness and total
Dewar, R. C., and A. Porté. 2008. Statistical mechanics unifies dif- abundance constrain the distribution of abundance. Ecology Let-
ferent ecological patterns. Journal of Theoretical Biology 251:389– ters 16:1177–1185.
403. Lopez-Gonzalez, G., S. L. Lewis, M. Burkitt, T. R. Baker, and O. L.
Enquist, B. J., and K. J. Niklas. 2001. Invariant scaling relations across Phillips. 2009. ForestPlots.net database. http://www.forestplots.net.
tree-dominated communities. Nature 410:655–660. Date of extraction: July 6, 2012.
Ernest, S. K. M., E. P. White, and J. H. Brown. 2009. Changes in a Lopez-Gonzalez, G., S. L. Lewis, M. Burkitt, and O. L. Phillips. 2011.
tropical forest support metabolic zero-sum dynamics. Ecology Let- ForestPlots.net: a web application and research tool to manage and
ters 12:507–515. analyse tropical forest plot data. Journal of Vegetation Science 22:
Gilbert, G. S., E. Howard, B. Ayala-Orozco, M. Bonilla-Moheno, J. 610–613.
Cummings, S. Langridge, I. M. Parker, et al. 2010. Beyond the Marquet, P. A., J. A. Keymer, and H. Cofre. 2003. Breaking the stick
tropics: forest structure in a temperate forest mapped plot. Journal in space: of niche models, metacommunities and patterns in the
of Vegetation Science 21:388–405. relative abundance of species. Pages 64–86 in T. M. Blackburn and
Gillooly, J. F., J. H. Brown, G. B. West, V. M. Savage, and E. L. K. J. Gaston, eds. Macroecology: concepts and consequences.
Charnov. 2001. Effects of size and temperature on metabolic rate. Blackwell Science, Oxford.
Science 293:2248–2251. McDonald, R. I., R. K. Peet, and D. L. Urban. 2002. Environmental
Gouws, E. J., K. J. Gaston, and S. L. Chown. 2011. Intraspecific body correlates of aak decline and red maple increase in the North
size frequency distributions of insects. PLoS ONE 6:e16606. Carolina piedmont. Castanea 67:84–95.
Hanski, I., and M. Gyllenberg. 1997. Uniting two general patterns McGill, B. 2003. Strong and weak tests of macroecological theory.
in the distribution of species. Science 275:397–400. Oikos 102:679–685.

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).
E80 The American Naturalist

———. 2010. Towards a unification of unified theories of bio- Ramesh, B. R., M. H. Swaminath, S. V. Patil, R. Pélissier, P. D.
diversity. Ecology Letters 13:627–642. Venugopal, S. Aravajy, C. Elouard, et al. 2010. Forest stand struc-
McGill, B. J., R. S. Etienne, J. S. Gray, D. Alonso, M. J. Anderson, ture and composition in 96 sites along environmental gradients
H. K. Benecha, M. Dornelas, et al. 2007. Species abundance dis- in the central Western Ghats of India. Ecology 91:3118.
tributions: moving beyond single prediction theories to integration Reed, R. A., R. K. Peet, M. W. Palmer, and P. S. White. 1993. Scale
within an ecological framework. Ecology Letters 10:995–1015. dependence of vegetation-environment correlations: a case study
McGill, B. J., B. A. Maurer, and M. D. Weiser. 2006. Empirical eval- of a North Carolina piedmont woodland. Journal of Vegetation
uation of neutral theory. Ecology 87:1411–1423. Science 4:329–340.
McGlinn, D. J., X. Xiao, and E. P. White. 2013. An empirical eval- Rosenzweig, M. L. 1995. Species diversity in space and time. Cam-
uation of four variants of a universal species-area relationship. bridge University Press, Cambridge.
PeerJ 1:e212. Supp, S. R., X. Xiao, S. K. M. Ernest, and E. P. White. 2012. An
Morlon, H., E. P. White, R. S. Etienne, J. L. Green, A. Ostling, D. experimental test of the response of macroecological patterns to
Alonso, B. J. Enquist, et al. 2009. Taking species abundance dis- altered species interactions. Ecology 93:2505–2511.
tributions beyond individuals. Ecology Letters 12:488–501. Thibault, K. M., E. P. White, A. H. Hurlbert, and S. K. M. Ernest.
Muller-Landau, H. C., R. S. Condit, K. E. Harms, C. O. Marks, 2011. Multimodality in the individual size distributions of bird
S. C. Thomas, S. Bunyavejchewin, G. Chuyong, et al. 2006. Com- communities. Global Ecology and Biogeography 20:145–153.
paring tropical forest tree size distributions with the predictions Thompson, J., N. Brokaw, J. K. Zimmerman, R. B. Waide, E. M.
of metabolic ecology and equilibrium models. Ecology Letters 9: Everham, D. J. Lodge, C. M. Taylor, et al. 2002. Land use history,
589–602. environment, and tree composition in a tropical forest. Ecological
Nakashizuka, T., M. Saito, K. Matsui, A. Makita, T. Kambayashi, T. Applications 12:1344–1363.
Masaki, T. Nagaike, et al. 2003. Monitoring beech (Fagus crenata) West, G. B., J. H. Brown, and B. J. Enquist. 1999. A general model
forests of different structure in Shirakami Mountains. Tohoku for the structure and allometry of plant vascular systems. Nature
Journal of Forest Science 8:67–74. 400:664–667.
Newman, E. A., M. E. Harte, N. Lowell, M. Wilber, and J. Harte. White, E. P., S. K. M. Ernest, A. J. Kerkhoff, and B. J. Enquist. 2007.
2014. Empirical tests of within- and across-species energetics in a Relationships between body size and abundance in ecology. Trends
diverse plant community. Ecology 95:2815–2825. in Ecology and Evolution 22:323–330.
Nishimura, T. B., and E. Suzuki. 2001. Allometric differentiation White, E. P., K. M. Thibault, and X. Xiao. 2012a. Characterizing
among tropical tree seedlings in heath and peat-swamp forests. species abundance distributions across taxa and ecosystems using
Journal of Tropical Ecology 17:667–681. a simple maximum entropy model. Ecology 93:1772–1778.
Nishimura, T. B., E. Suzuki, T. Kohyama, and S. Tsuyuzaki. 2006. White, E. P., X. Xiao, N. J. B. Issac, and R. M. Sibly. 2012b. Meth-
Mortality and growth of trees in peat-swamp and heath forests in odological tools. Pages 9–20 in R. M. Sibly, J. H. Brown, and A.
central Kalimantan after severe drought. Plant Ecology 188:165– Kodric-Brown, eds. Metabolic ecology: a scaling approach. Wiley,
177. Chichester.
O’Dwyer, J. P., J. K. Lake, A. Ostling, V. M. Savage, and J. L. Green. Xi, W., R. K. Peet, J. K. Decoster, and D. L. Urban. 2008. Tree damage
2009. An integrative framework for stochastic, size-structured risk factors associated with large, infrequent wind disturbances of
community assembly. Proceedings of the National Academy of Carolina forests. Forestry 81:317–334.
Sciences of the USA 106:6170–6175. Xiao, X., S. Aravajy, T. W. Baribault, N. Brokaw, N. L. Christensen,
Palmer, M. W., R. K. Peet, R. A. Reed, W. Xi, and P. S. White. 2007. Dasappa, S. J. DeWalt, et al. 2014. Data from: A strong test of the
A multiscale study of vascular plants in a North Carolia Piedmont maximum entropy theory of ecology. American Naturalist, Dryad
forest. Ecology 88:2674. Digital Repository, http://dx.doi.org/10.5061/dryad.5fn46.
Peet, R. K., and N. L. Christensen. 1987. Competition and tree death. Zimmerman, J. K., E. M. Everham III, R. B. Waide, D. J. Lodge,
BioScience 37:586–595. C. M. Taylor, and N. V. L. Brokaw. 1994. Responses of tree species
Pielou, E. C. 1975. Ecological diversity. Wiley, New York. to hurricane winds in subtropical wet forest in Puerto Rico: im-
Pitman, N. C. A., C. E. Cerón, C. I. Reyes, M. Thurber, and J. plications for tropical tree life histories. Journal of Ecology 82:
911–922.
Arellano. 2005. Catastrophic natural origin of a species-poor tree
community in the world’s richest forest. Journal of Tropical Ecol-
ogy 21:559–568.
Pyke, C. R., R. Condit, S. Aguilar, and S. Lao. 2001. Floristic com-
position across a climatic gradient in a Neotropical lowland forest. Associate Editor: Sean R. Connolly
Journal of Vegetation Science 12:553–566. Editor: Judith L. Bronstein

This content downloaded from 200.008.076.068 on July 06, 2016 00:00:38 AM


All use subject to University of Chicago Press Terms and Conditions (http://www.journals.uchicago.edu/t-and-c).

Вам также может понравиться