ArticlePDF Available

Abstract and Figures

Species immigration history can structure ecological communities through priority effects, which are often mediated by competition. As competition tends to be stronger between species with more similar niches, we hypothesize that species phylogenetic relatedness, under niche conservatism, may be a reasonable surrogate of niche similarity between species, and thus influence the strength of priority effects. We tested this hypothesis using a laboratory microcosm experiment in which we established bacterial species pools with different levels of phylogenetic relatedness and manipulated the immigration history of species from each pool into microcosms. Our results showed that strong priority effects, and hence multiple community states, only emerged for the species pool with the greatest phylogenetic relatedness. Community assembly also resulted in a significant positive relationship between bacterial phylogenetic diversity and ecosystem functions. Interestingly, these results emerged despite a lack of phylogenetic conservatism for most of the bacterial functional traits considered. Our results highlight the utility of phylogenetic information for understanding the structure and functioning of ecological communities, even when phylogenetically conserved functional traits are not identified or measured.
Content may be subject to copyright.
Ecology, 93(5), 2012, pp. 1164– 1172
Ó2012 by the Ecological Society of America
Species phylogenetic relatedness, priority effects,
and ecosystem functioning
JIAQI TAN,
1,3
ZHICHAO PU,
1
WADE A. RYBERG,
1,2
AND LIN JIANG
1
1
School of Biology, Georgia Institute of Technology, 310 Ferst Drive, Atlanta, Georgia 30332 USA
2
Department of Wildlife and Fisheries Sciences, Texas A&M University, 210 Nagle Hall, College Station, Texas 77843 USA
Abstract. Species immigration history can structure ecological communities through
priority effects, which are often mediated by competition. As competition tends to be stronger
between species with more similar niches, we hypothesize that species phylogenetic relatedness,
under niche conservatism, may be a reasonable surrogate of niche similarity between species,
and thus influence the strength of priority effects. We tested this hypothesis using a laboratory
microcosm experiment in which we established bacterial species pools with different levels of
phylogenetic relatedness and manipulated the immigration history of species from each pool
into microcosms. Our results showed that strong priority effects, and hence multiple
community states, only emerged for the species pool with the greatest phylogenetic
relatedness. Community assembly also resulted in a significant positive relationship between
bacterial phylogenetic diversity and ecosystem functions. Interestingly, these results emerged
despite a lack of phylogenetic conservatism for most of the bacterial functional traits
considered. Our results highlight the utility of phylogenetic information for understanding the
structure and functioning of ecological communities, even when phylogenetically conserved
functional traits are not identified or measured.
Key words: bacteria; community assembly; ecosystem function; multiple community states; phyloge-
netic relatedness; priority effects.
INTRODUCTION
Understanding mechanisms underlying the assembly
of ecological communities is one of the central goals of
community ecology (Gleason 1927, Diamond 1975).
Ecologists now recognize that both niche-based deter-
ministic processes (Chase and Leibold 2003) and neutral
stochastic processes (Bell 2001, Hubbell 2001) can
operate during the process of community assembly.
Niche-based processes involve the interaction between
species’ niches and the conditions of the environment
where they live, which can jointly regulate the structure
of the assembling communities. In habitats with similar
environmental conditions and under the same regional
species pool, such processes often result in convergent
communities with similar species composition and
abundance. Stochastic processes, highlighted by the
neutral theory (Bell 2001, Hubbell 2001), can also
strongly impact ecological communities. In particular,
stochasticity in the order and timing of species
colonization events, as demonstrated by both theoretical
and empirical studies (e.g., Drake 1991, Law and
Morton 1993, Jiang and Patel 2008, Fukami et al.
2010; reviewed by Chase 2003), can result in divergent
communities dominated by different species. These
multiple community states associated with different
species colonization histories frequently arise from
priority effects, in which early colonizing species affect
the establishment and abundance of later colonizers.
One factor that can potentially influence the relative
importance of deterministic and stochastic processes, and
hence the strength of priority effects, is ecological
similarity of species in the regional species pool. Both
theory (e.g., MacArthur and Levins 1967) and experi-
ments (e.g., Gause 1934) have demonstrated the difficulty
for species with similar niches to coexist, which prompted
Hardin (1960) to coin the competitive exclusion principle.
A corollary of this principle, applying to community
assembly, is that increasing ecological similarity of species
in the regional pool may make it more likely for species
already established at a locality to have strong negative
impacts on newly colonizing species, promoting inhibitive
priority effects. As species niches are often difficult to
quantify and phylogenetically closely related species tend
to possess similar niches (i.e., phylogenetic niche conser-
vatism; Harvey and Pagel 1991, Prinzing et al. 2001,
Webb et al. 2002, Donoghue 2008), we suggest that
species phylogenetic relatedness may be used as a
surrogate of niche similarity to predict the strength of
competition and priority effects. The positive relationship
between species phylogenetic relatedness and competition
was in fact first hypothesized by Darwin (1859), and
supported by a recent experiment (Violle et al. 2011).
However, whether phylogenetic relatedness of the region-
Manuscript received 24 August 2011; revised 28 November
2011; accepted 30 November 2011. Corresponding Editor: J. B.
Yavitt.
3
E-mail: jtan@gatech.edu
1164
al species pool influences the strength of priority effects
during community assembly remains an open question.
Phylogenetic relatedness of the regional species pool
may also have consequences for the functioning of the
assembled communities. For example, if phylogenetic
relatedness serves as a reasonable surrogate for species
ecological similarity, then low phylogenetic relatedness
(i.e., high phylogenetic diversity) may translate into
increased niche complementarity among species in the
assembled communities, potentially resulting in high
levels of ecosystem functioning (Cavender-Bares et al.
2009). On the other hand, high phylogenetic relatedness
among species within the regional species pool would
indicate possible redundancy in species’ niches, likely
leading to reduced ecosystem functioning. So far only a
handful of studies have investigated the relevance of
species phylogenetic relatedness for ecosystem function-
ing (Maherali and Klironomos 2007, Cadotte et al. 2008,
2009, Jiang et al. 2010), but the potential interactive
effects of phylogenetic relatedness and assembly history
on ecosystem functions have not been explored.
Here, we describe an experimental study examining
how species phylogenetic relatedness affects priority
effects and ecosystem functioning by using a laboratory
model of bacterial communities. We established bacte-
rial species pools with different levels of species
phylogenetic relatedness and manipulated the immigra-
tion history of bacteria from each species pool into the
assembled communities. We showed that significant
dissimilarity among communities subjected to different
assembly histories emerged only when bacteria in the
species pool were phylogenetically closely related. We
also found significant effects of phylogenetic relatedness
and assembly history on bacterial ecosystem functions
(i.e., bacterial production and decomposition).
MATERIAL AND METHODS
Our experiment used eight strains of common
environmental bacteria from freshwater ecosystems
(Fig. 1), all of which can form colonies with unique
morphological characteristics on agar plates. To esti-
mate phylogenetic relatedness between these bacteria,
we constructed phylogeny based on bacterial 16S rRNA
sequences (Fig. 1a). We sequenced the 16S rRNA gene
of each bacterial strain, aligned the sequences with
Clustal X (version 2.0; Larkin et al. 2007), selected the
best sequence evolution model, GTRþG with MrModel-
test (version 2.3; Nylander 2004), by using the Akaike
information criterion, and built the phylogenetic tree
with Bayesian method in MrBAYES (version 3.1.2;
Huelsenbeck and Ronquist 2001). Three archaea were
used as the out-group. The phylogenetic distance
between bacteria was obtained by summing lengths of
the intervening branches between the two species on the
phylogeny; smaller phylogenetic distance between bac-
teria indicates greater phylogenetic relatedness. Using
these eight strains of bacteria, we established four
species pools: Serratia,Staphylococcus,Bacillus, and a
mixed-genus pool with one bacterium randomly selected
from each of the single-genus pools (Fig. 1a; see Plate 1).
The phylogenetic diversity (hereafter PD) of each species
pool was calculated by summing the lengths of all the
intervening branches of all the species in each pool
(Faith 1992). PD is thus an aggregate measure of the
phylogenetic relatedness of each species pool; higher PD
values indicate larger phylogenetic distances and thus
weaker phylogenetic relatedness among species.
We estimated functional trait diversity of each species
pool based on the bacteria’s ability to utilize a variety of
carbon substrates that may appear in the bacterial
growth medium used in our experiment. We measured
55 bacterial traits with Biolog MicroPlates (Biolog,
Hayward, California, USA). Following the manufactur-
er’s instructions, we prepared and inoculated Gram-
positive and Gram-negative bacterial cultures into their
corresponding type of Biolog MicroPlates. Gram-
positive and Gram-negative microplates, each contain-
ing 96 wells, share 55 carbon substrates in common, so
we only recorded the results of these 55 traits. We scored
positive results, indicating that the species was able to
use carbon sources in the wells, as 1 and negative results
as 0. In addition, we tested the ability of these bacteria to
utilize two common carbon substrates: cellulose and
starch. We spread diluted cultures of each bacterial
strain on carboxymethylcellulose (Wohl et al. 2004) and
starch agar plates, incubated them at room temperature
(;228C) for 5 days, and flooded plates with 1%Gango
Red and Lugol’s iodine solutions, respectively. Colorless
zones around bacterial colonies on agar plates were
observed if bacteria utilize cellulose or starch. Based on
the total 57 traits, we calculated functional trait diversity
of each species pool in two ways. First, we calculated
functional richness (hereafter FR) by counting the total
number of carbon substrates that bacteria from a species
pool could utilize. Second, we calculated functional
diversity (hereafter FD) of each species pool. We
performed a UPGMA-based cluster analysis (unweight-
ed pair group method with arithmetic mean) with the
Euclidean distance between bacteria in the 57-dimen-
sional trait space to produce the functional dendrogram
(Fig. 1b), and calculated FD of each species pool as the
total intervening branch lengths of all the species in the
pool on the dendrogram (Petchey and Gaston 2002). To
test for phylogenetic conservatism of the measured
traits, we conducted a Mantel test based on 10 000
permutations that evaluated the correlation between
bacterial phylogenetic distance and trait Euclidean
distance. We also tested the phylogenetic signal of each
trait with Blomberg’s K (Blomberg et al. 2003), using
the multiPhylosignal function in the Picante package
(Kimbel et al. 2010).
Our experiment used 25-mL capped test tubes as
microcosms, each of which contained 10 mL of medium.
The medium contained 0.55 g of crushed protozoan
pellets (Carolina Biological Supply, Burlington, North
Carolina, USA) per liter of deionized water. Protozoan
May 2012 1165PHYLOGENY AND COMMUNITY ASSEMBLY
pellets are made from grass and include a variety of
common carbon resources for bacterial growth. Medium
was autoclaved in large flasks and filtered to remove
insoluble particles, then transferred into experimental
microcosms and autoclaved again before the experiment
started. The microcosms were incubated on a shaker at
200 rpm under room temperature (;228C).
The experiment included all the possible assembly
sequences for each bacterial pool. Thus, we had two
sequential assembly history treatments for the Serratia
pool that contained two species, and six for the
Staphylococcus, Bacillus, and mixed pools that each
contained three species (Fig. 1). Each treatment was
replicated three times. Prior to the experiment, we
prepared stock cultures of each bacterial strain in 8%
nutrient broth. At the beginning of the experiment (day
0), we introduced the first species into microcosms by
transferring a small volume (,5lL) of stock culture
with an aseptic loop. In the same way, on days 7 and 14,
we introduced the second and third immigrants (no third
immigrant for the Serratia communities), respectively.
The weekly interval between species introduction
allowed the assembled communities to equilibrate before
the next introduction event. Our pilot experiment, albeit
using only half of the eight bacterial strains used in this
study, indicated that bacterial populations of individual
species, initiated at small size in isolation from other
species, require 2–3 days to reach carrying capacity and
persist at the stationary phase for at least our
experimental duration; bacterial communities contain-
ing multiple species generally reach equilibrium in one
week and can persist for similarly long periods of time
(J. Tan, unpublished data). On day 21, we added a dried,
weighed, and autoclaved wheat seed to each microcosm.
On day 49, we terminated the experiment and destruc-
tively sampled the microcosms. The samples from
microcosms were serially diluted and spread on nutrient
agar plates. After 7-day incubation, we counted the
number of bacterial colonies on plates to determine
population density (colony formation units per milliliter
[CFU/mL]) of each bacterial strain. Wheat seeds were
retrieved from microcosms, oven dried, and weighed.
Two ecosystem functions were recorded. Total bacterial
production in each microcosm was obtained by sum-
ming the density of each bacterial strain. Decomposition
was measured as the fraction of wheat seed mass lost
during the experiment.
FIG. 1. (a) Phylogeny based on Bayesian methods and (b) functional dendrogram based on cluster analysis (via unweighted pair
group method with arithmetic mean [UPGMA]) of 57 traits for the study bacteria. Four species pools, Serratia (initial phylogenetic
diversity [PD], 0.0065; initial functional richness [FR], 36; initial functional diversity [FD], 2.645), Staphylococcus (initial PD,
0.0274; initial FR, 42; initial FD, 7.224), Bacillus (initial PD, 0.0959; initial FR, 35; initial FD, 6.959), and the mixed-species pool
(initial PD, 0.4854; initial FR, 50; initial FD, 7.550) were formed by these bacteria. Daggers indicate the bacteria constituting the
mixed species pool. The scales for branch lengths are shown beneath the phylogenetic tree (panel a) and the functional dendrogram
(panel b).
JIAQI TAN ET AL.1166 Ecology, Vol. 93, No. 5
We calculated realized community PD, FR, and FD,
based on the realized species composition measured at the
end of the experiment. We calculated bdiversity between
communities sharing the same species pool but subjected
to different assembly histories, by first calculating the
modified Morisita similarity index (Horn 1966), then
subtracting it from 1. Calculation of Morisita indices was
based on untransformed bacterial density data. For
subsequent statistical analyses, all the bacteria density
data were log
10
-transformed (log
10
[xþ1]) to improve
normality. We used one-way ANOVA with bdiversity as
the dependent variable and species pool as the class
variable to assess the effect of varying phylogenetic
relatedness among species pools on history-induced
differences in community structure, as represented by b
diversity. Tukey’s HSD was further conducted as the post
hoc test. To test the effect of assembly history on the
density of bacteria in communities sharing the same
species pool, we used MANOVA with bacteria densities
for each species pool as the dependent variable and
history sequence as the class variable. To test the effect of
assembly history on bacterial production and decompo-
sition in different species pools, we used nested ANOVA
with production and decomposition as the dependent
variables and history sequences as a factor nested within
species pools. To further test the effect of assembly
history, we used one-way ANOVA within each species
pool, with production and decomposition as the depen-
dent variables and assembly history sequence as the
independent variable. To test the effect of phylogenetic
and functional diversity on bacterial production and
decomposition, we used simple and backward-selection
multiple linear regressions to model the ecosystem
functions (i.e., bacterial production and decomposition)
as functions of realized PD, FR, and FD. In all the
regressions, explanatory variables were deemed signifi-
cant if P0.05.
RESULTS
Our study bacteria did not exhibit significant phyloge-
netic conservatism when all the 57 traits were considered
together (Mantel test, P¼0.152). When examined
individually, 9 of 57 traits (15%), including D-fructose,
L-fucose, a-D-glucose, a-D-lactose, D-melibiose, D-
alanine, D, L, a-glycerol phosphate, a-D-glucose-1-phos-
phate, and D-glucose-6-phosphate, showed significant
phylogenetic signals (multiPhylosignal function, P,
0.05).
The bdiversity among communities subjected to
different histories varied significantly among the four
species pools (ANOVA, F
3, 411
¼443.081, P,0.001).
This significant variation mainly resulted from the larger
values of bdiversity observed in the Serratia pool (see
Fig. 2; Tukey’s HSD). The dominant species in
communities of the Serratia pool differed depending
on history treatments (Fig. 3a). In contrast, in the
Staphylococcus,Bacillus, and mixed-species pools, the
dominant species remained the same in different history
treatments (Fig. 3b–d). Nevertheless, MANOVA still
revealed a significant effect of assembly history on
species densities in those species pools (Staphylococcus,
Wilks’ lambda ¼0.010, F
15,28
¼7.882, P,0.001;
Bacillus, Wilks’ lambda ¼0.029, F
15,28
¼4.867, P,
0.001; mixed, Wilks’ lambda ¼0.018, F
15,28
¼6.097, P,
0.001), in addition to the significant effect of history for
the Serratia pool (Wilks’ lambda ¼0.017, F
1,4
¼234.1, P
,0.001).
Nested ANOVA revealed a significant effect of
assembly history on bacterial production (F
5,40
¼
14.449, P,0.001), but no effect of assembly history
on decomposition (F
5,40
¼0.886, P¼0.499). One-way
ANOVA indicated that assembly history had a signif-
icant effect on bacterial production in communities of
the Staphylococcus (F
5,12
¼30.086, P,0.001), Bacillus
(F
5,12
¼16.888, P,0.001), and mixed (F
5,12
¼3.601, P¼
0.032) pools, but had no effects in communities of the
Serratia pool (F
1,4
¼1.136, P¼0.346). In contrast,
assembly history significantly affected decomposition in
the Staphylococcus communities only (Staphylococcus,
F
5,12
¼37.615, P,0.001; Serratia,F
1,4
¼1.136, P¼
0.346; Bacillus,F
5,12
¼0.670, P¼0.654; mixed, F
5,12
¼
2.348, P¼0.105). Nested ANOVA also revealed that
ecosystem function level differed significantly in com-
munities of different species pools (production, F
14,40
¼
41.161, P,0.001; decomposition, F
14,40
¼6.288, P,
0.001).
Simple linear regressions showed that both bacterial
production and decomposition increased with realized
FIG. 2. The bdiversity among communities
assembled from the four species pools with
varying phylogenetic relatedness; bdiversity is
calculated as 1 (Morisita similarity index).
Values are means þSE. Treatments sharing the
same letters do not differ according to Tukey’s
HSD test.
May 2012 1167PHYLOGENY AND COMMUNITY ASSEMBLY
PD (Fig. 4a; R
2
¼0.461, P,0.001; Fig. 4b; R
2
¼0.212,
P,0.001), FR (Fig. 4c; R
2
¼0.586, P,0.001; Fig. 4d;
R
2
¼0.410, P,0.001), and FD (Fig. 4e; R
2
¼0.268, P
,0.001; Fig. 4f; R
2
¼0.415, P,0.001), respectively.
Multiple regression models retained realized FR and FD
as best predictors of both bacterial production and
decomposition.
DISCUSSION
The results of our experiment demonstrated the
importance of understanding species phylogenetic relat-
edness when predicting the strength of priority effects.
We observed the highest bdiversity among communities
in the Serratia pool (Fig. 2), which contained phyloge-
netically most closely related bacterial strains (Fig. 1a).
FIG. 3. Population density of each bacterium from the four species pools: (a) Serratia,(b)Staphylococcus, (c) Bacillus, and (d)
mixed-species pool, at the end of the experiment. Values are means þSE with density measured as colony formation units (CFU )
per mL and was log
10
(xþ1)-transformed prior to analysis.
JIAQI TAN ET AL.1168 Ecology, Vol. 93, No. 5
Different Serratia marcescens strains were dominant in
these communities when subjected to different assembly
histories (Fig. 3a). In contrast, communities from each
of the other pools with lower phylogenetic relatedness
were structurally similar (Fig. 2), containing the same
dominant species regardless of history (Fig. 3b–d). This
difference emerged despite the fact that history had a
significant effect on the structure of the assembled
communities for all species pools, as revealed by
MANOVA. These results appear consistent with our
hypothesis that stronger competition may occur between
species that are more closely related phylogenetically
FIG. 4. Relationships between (a, b) realized phylogenetic diversity (PD), (c, d) functional richness (FR), (e, f ) functional
diversity (FD) and (a, c, e) bacterial production and (b, d, f) decomposition. PD and FD attained zero values in communities with
one species. Data are plotted with linear regression lines. Bacterial production was measured as colony formation units (CFU) per
mL and was log
10
(xþ1)-transformed prior to analysis.
May 2012 1169PHYLOGENY AND COMMUNITY ASSEMBLY
(Maherali and Klironomos 2007, Violle et al. 2011),
leading to stronger priority effects that generate multiple
community states (Chase and Leibold 2003, Fukami and
Lee 2006). However, phylogenetic conservatism was not
detected when all bacterial traits were considered
together, and nonsignificant phylogenetic signals were
detected for the majority of measured traits. At least
three mutually nonexclusive explanations can account
for these results. One possibility is that at least some of
the phylogenetically conserved traits that we measured
are important in defining the ecological niches of our
study bacteria in our experiment. This is supported by
the fact that phylogenetic diversity and functional
diversity based on measured traits (including FR and
FD) were both positively related to bacterial production
and decomposition in our experiment. Another possi-
bility is that some unmeasured traits that are important
in defining species niches may be phylogenetically
conserved, making phylogenetic relatedness a reason-
able proxy of functional similarity with regard to these
traits. A third explanation is that phylogenetic relation-
ships based on the 16S rRNA gene, which is known to
be highly conserved between different species of bacteria
(Coenye and Vandamme 2003), may not adequately
capture the potentially large variation in traits coded by
less conserved genes (see Dahle et al. 2011 for a
counterexample). Note that this issue can be circum-
vented in the future by constructing phylogeny based on
whole genomes, which are currently unavailable for
most organisms. Regardless, our results highlight the
utility of phylogenetic information for understanding
the structure and functioning of ecological communities,
even when phylogenetically conserved functional traits
are not identified or measured.
Our results indicated that phylogenetic diversity
positively affected ecosystem functions (i.e., bacterial
production and decomposition), but that ecosystem
functioning was better predicted by functional diversity.
Using data from plant experiments, Cadotte et al. (2008,
2009) also showed that primary productivity was
positively correlated with both plant phylogenetic and
functional diversity. However, their results indicated
that phylogenetic diversity explained more variation in
plant productivity than several measurements of func-
tional diversity. This discrepancy between the results of
the two studies may be due to the fact that horizontal
PLATE 1. Colonies of the eight bacteria studied, on agar plates. Photo credits: J. Tan and L. Jiang.
JIAQI TAN ET AL.1170 Ecology, Vol. 93, No. 5
gene transfer, which may increase trait similarity among
distantly related species and weaken the correlation
between phylogenetic relatedness and trait similarity, is
much more common for bacteria than for plants
(Andersson 2005, Richardson and Palmer 2007). Note
that phylogenetic diversity nevertheless remained signif-
icant in explaining the functioning of bacterial commu-
nities in our experiment.
Our results also showed that community assembly
history had significant effects on bacterial production in
the Staphylococcus,Bacillus, and mixed communities,
and on decomposition in the Staphylococcus communi-
ties. Likewise, Fukami et al. (2010) manipulated the
assembly history of wood-decay fungal communities and
found a significant effect of assembly history on fungal
decomposition. They showed that community diver-
gence in species richness and composition, resulting
from different assembly histories, led to the differenti-
ation of ecosystem functioning. However, this mecha-
nism cannot explain the divergence/convergence of
ecosystem functioning in communities subjected to
different assembly histories in our study. Two distinct
alternative states were formed in communities of the
Serratia pool (Fig. 3a), but ecosystem functions of these
two community states were similar. In contrast, a single
community state was observed in the Staphylococcus
pool, but ecosystem functions differed among the
assembled communities (Fig. 3b). One explanation for
the lack of historical effects on ecosystem functioning in
the Serratia communities is that the two strains of
Serratia marcescens may play similar ecological roles
since they are phylogenetically closely related (99%
similarity based on phylogeny) and functionally similar
(sharing 50 of 57 traits). The two Serratia strains may
thus be largely functionally substitutable, resulting in the
same levels of ecosystem functions in communities
dominated by different Serratia strains. In other species
pools, although the historical effect was not strong
enough to generate multiple community states, the
abundance of subdominant species differed under
different assembly histories (hence the significant effect
of assembly history on species densities in MANOVA),
especially in the Staphylococcus pool (Fig. 3b), which
may have caused the differentiation of ecosystem
functioning in those species pools. All together, our
results showed that assembly history affected ecosystem
functioning in some communities, but not in others.
Understanding the conditions that promote the rela-
tionship between assembly history and ecosystem
functioning remains an important topic of future
research.
One concern is that each phylogenetic relatedness
level in our experiment included only one species
combination, so one could argue that the effect of
phylogenetic relatedness may have been confounded
with the effect of species identity. An ideal solution to
this problem would be to use as many species
combinations at each phylogenetic level as possible,
but this may not be logistically possible. In particular,
finding many combinations of phylogenetically closely
related bacteria with different colony morphologies
(e.g., the red and white Serratia marcescens) is difficult.
In this experiment, although we cannot exclude the
possibility that the effects of species phylogenetic
relatedness and identity are confounded, results from a
related experiment suggests that this is not the case. That
experiment produced results similar to the current
experiment. In particular, strong priority effects were
also observed in bacterial communities containing
closely related species, specifically those with three
strains of Bacillus pumilus; weaker priority effects were
detected in communities with less related species (J. Tan,
unpublished data). In the present experiment, weak
priority effects also emerged in all communities of the
three species pools with relatively low levels of
phylogenetic relatedness, resulting in single community
states. Together, these results strongly suggest a linkage
between species phylogenetic relatedness and the
strength of priority effects. Nevertheless, future studies
that manipulate phylogenetic relatedness or diversity
should aim to establish multiple species combinations
within each treatment, in order to eliminate the potential
confounding effects from species identity.
In this study, different bacterial species pools exhib-
ited different levels of phylogenetic relatedness, permit-
ting an evaluation of how phylogenetic relatedness
might govern the relative contributions of niche-based
deterministic processes (Chase and Leibold 2003) and
neutral stochastic processes (Bell 2001, Hubbell 2001) to
community assembly. In the experiment we conducted
to accomplish this evaluation, multiple community
states resulting from strong stochastic assembly pro-
cesses (i.e., priority effects) were only observed in the
species pool with the highest phylogenetic relatedness
and highest functional similarity. Alternatively, single-
community states resulting from strong deterministic
assembly processes were observed in communities
assembled from less phylogenetically related species
pools. As such, these observations support our hypoth-
esis that priority effects are stronger between species that
are more closely related phylogenetically, although some
caution must be exercised when generalizing these
results given the limitation of our experimental design
(see last paragraph). Further, our study demonstrates a
positive relationship between phylogenetic diversity and
ecosystem functions in an experiment that directly
manipulated phylogenetic diversity. Importantly, we
obtained these results despite the fact that many
functional traits measured in our experiment exhibited
nonsignificant phylogenetic signals. Our results thus
highlight the difficulty of identifying species functional
traits relevant for community assembly and ecosystem
functioning, and at the same time, the utility of basic
phylogenetic information in predicting the structure and
functioning of ecological communities.
May 2012 1171PHYLOGENY AND COMMUNITY ASSEMBLY
ACKNOWLEDGMENTS
We thank Michael Cortez, Cyrille Violle, Edward Burdette,
and two anonymous reviewers for their insightful comments,
which improved the manuscript. This project was supported by
a British Ecological Society early career project grant and U.S.
NSF grant (DEB-1120281) awarded to L. Jiang.
LITERATURE CITED
Andersson, J. O. 2005. Lateral gene transfer in eukaryotes.
Cellular and Molecular Life Sciences 62:1182–1197.
Bell, G. 2001. Ecology—neutral macroecology. Science
293:2413–2418.
Blomberg, S. P., T. Garland, Jr., and A. R. Ives. 2003. Testing
for phylogenetic signal in comparative data: behavioral traits
are more labile. Evolution 57:717–745.
Cadotte, M. W., B. J. Cardinale, and T. H. Oakley. 2008.
Evolutionary history and the effect of biodiversity on plant
productivity. Proceedings of the National Academy of
Sciences USA 105:17012–17017.
Cadotte, M. W., J. Cavender-Bares, D. Tilman, and T. H.
Oakley. 2009. Using phylogenetic, functional, and trait
diversity to understand patterns of plant community
productivity. PLoS ONE 4(5):e5695.
Cavender-Bares, J., K. H. Kozak, P. V. A. Fine, and S. W.
Kembel. 2009. The merging of community ecology and
phylogenetic biology. Ecology Letters 12:693–715.
Chase, J. M. 2003. Community assembly: when should history
matter? Oecologia 136:489–498.
Chase, J. M., and M. A. Leibold. 2003. Ecological niches:
linking classical and contemporary approaches. University of
Chicago Press, Chicago, Illinois, USA.
Coenye, T., and P. Vandamme. 2003. Diversity and significance
of Burkholderia species occupying diverse ecological niches.
Environmental Microbiology 5:719–729.
Dahle, H., B. Hannisdal, B. O. Steinsbu, H. Ommedal, J. Einen,
S. Jensen, O. Larsen, L. Ovreas, and S. Norland. 2011.
Evolution of temperature optimum in Thermotogaceae and
the prediction of trait values of uncultured organisms.
Extremophiles 15:509–516.
Darwin, C. 1859. On the origin of Species. Murray, London,
UK.
Diamond, J. M. 1975. Assembly of species communities. Pages
342–344 in M. L. Cody and J. M. Diamond, editors. Ecology
and evolution of communities. Harvard University Press,
Cambridge, Massachusetts, USA.
Donoghue, M. J. 2008. A phylogenetic perspective on the
distribution of plant diversity. Proceedings of the National
Academy of Sciences USA 105:11549–11555.
Drake, J. A. 1991. Community-assembly mechanics and the
structure of an experimental species ensemble. American
Naturalist 137:1–26.
Faith, D. P. 1992. Conservation evaluation and phylogenetic
diversity. Biological Conservation 61:1–10.
Fukami, T., I. A. Dickie, J. P. Wilkie, B. C. Paulus, D. Park, A.
Roberts, P. K. Buchanan, and R. B. Allen. 2010. Assembly
history dictates ecosystem functioning: evidence from wood
decomposer communities. Ecology Letters 13:675–684.
Fukami, T., and W. G. Lee. 2006. Alternative stable states, trait
dispersion, and ecological restoration. Oikos 113:353–356.
Gause, G. F. 1934. The struggle for existence. Williams and
Wilkins, Baltimore, Maryland, USA.
Gleason, H. A. 1927. Further views on the succession concept.
Ecology 8:229–326.
Hardin, G. 1960. The competitive exclusion principle. Science
131:1292–1297.
Harvey, P.H., and M. D. Pagel. 1991. The comparative method in
evolutionary biology. Oxford University Press, Oxford, UK.
Horn, H. S. 1966. Measurement of overlap in comparative
ecological studies. American Naturalist 100:419–424.
Hubbell, S. P. 2001. The unified neutral theory of biodiversity
and biogeography. Princeton University Press, Princeton,
New Jersey, USA.
Huelsenbeck, J. P., and F. Ronquist. 2001. MRBAYES:
Bayesian inference of phylogenetic trees. Bioinformatics
17:754–755.
Jiang, L., and S. N. Patel. 2008. Community assembly in the
presence of disturbance: a microcosm experiment. Ecology
89:1931–1940.
Jiang, L., J. Tan, and Z. Pu. 2010. An experiment test of
Darwin’s naturalization hypothesis. American Naturalist
175:415–423.
Kimbel, S. W., P. D. Cowan, M. R. Helmus, W. K. Cornwell,
H. Morlon, D. D. Ackerly, S. P. Blomberg, and C. O. Webb.
2010. Picante: R tools for integrating phylogenies and
ecology. Bioinformatics 26:1463–1464.
Larkin, M. A., et al. 2007. Clustal W and clustal X version 2.0.
Bioinformatics 23:2947–2948.
Law, R., and R. D. Morton. 1993. Alternative permanent states
of ecological communities. Ecology 74:1347–1361.
MacArthur, R. H., and R. Levins. 1967. The limiting similarity,
convergence, and divergence of coexisting species. American
Naturalist 101:377–385.
Maherali, H., and J. N. Klironomos. 2007. Influence of
phylogeny on fungal community assembly and ecosystem
functioning. Science 316:1746–1748.
Nylander, J. A. A. 2004. MrModeltest. Version 2. Evolutionary
Biology Centre, Uppsala University, Uppsala, Sweden.
Petchey, O. L., and K. J. Gaston. 2002. Functional diversity
(FD), species richness and community composition. Ecology
Letters 5:402–411.
Prinzing, A., W. Durka, S. Klotz, and R. Brandl. 2001. The
niche of higher plants: evidence for phylogenetic conserva-
tism. Proceedings of the Royal Society B 268:2383–2389.
Richardson, A. O., and J. D. Palmer. 2007. Horizontal gene
transfer in plants. Journal of Environmental Botany 58:1–9.
Violle, C., D. R. Nemergut, Z. Pu, and L. Jiang. 2011.
Phylogenetic limiting similarity and competitive exclusion.
Ecology Letters 14:782–787.
Webb, C. O., D. D. Ackerly, M. A. McPeek, and M. J.
Donoghue. 2002. Phylogenies and community ecology.
Annual Review of Ecology, Evolution and Systematics
33:475–505.
Wohl, D. L., S. Arora, and J. R. Gladstone. 2004. Functional
redundancy supports biodiversity and ecosystem function in
a closed and constant environment. Ecology 85:1534–1540.
JIAQI TAN ET AL.1172 Ecology, Vol. 93, No. 5
... For instance, in many microbial systems, priority effects are only measured by changes in community composition to differences in arrival times. While some studies point to more resolved mechanisms such as sugar concentration or pH [3,24], others are often attributed to general factors like resource pre-emption [25] or phylogenetic relatedness [26,27], with the specific mechanisms often unknown. In coral reef fish, priority effects often arise because early arrivers become more aggressive toward late arrivers [28][29][30]. ...
... In addition to system selection and experimental design, priority effects are also measured in diverse ways. The most common response variables include community composition [49], functional traits [26,27], demographic rates such as growth, survival, and fecundity [2,42,61], and ecosystem function [96,97], but studies can also measure highly specific responses such as aggression, site occupancy [28,30,98], and in coinfection experiments, pathogen load, and host response [38,99]. These measurements in part highlight the importance of priority effects in nature but also increase the difficulty of generalizing, quantifying, and comparing the consequences of priority effects across systems. ...
Article
Priority effects play a key role in structuring natural communities, but considerable confusion remains about how they affect different ecological systems. Synthesizing previous studies, we show that this confusion arises because the mechanisms driving priority and the temporal scale at which they operate differ among studies, leading to divergent outcomes in species interactions and biodiversity patterns. We suggest grouping priority effects into two functional categories based on their mechanisms: frequency-dependent priority effects that arise from positive frequency dependence, and trait-dependent priority effects that arise from time-dependent changes in interacting traits. Through easy quantification of these categories from experiments, we can construct community models representing diverse biological mechanisms and interactions with priority effects, therefore better predicting their consequences across ecosystems.
... The direction and magnitude of interactions among parasites and host-associated microbiota may depend on the sequence in which they infect host individuals (i.e. priority effects; [37][38][39][40][41][42]). It has become increasingly clear across ecological communities that assembly history can have lasting consequences for community dynamics [39] which can impact community structure, as well as ecosystem function [40,41]. ...
... The direction and magnitude of interactions among parasites and host-associated microbiota may depend on the sequence in which they infect host individuals (i.e. priority effects; [37][38][39][40][41][42]). It has become increasingly clear across ecological communities that assembly history can have lasting consequences for community dynamics [39] which can impact community structure, as well as ecosystem function [40,41]. In the context of within-host environments, simultaneous co-infections can occur in nature [42], but more often, microbes infect a host at different time points [43]. ...
Article
Full-text available
Interactions among parasites and other microbes within hosts can impact disease progression, yet study of such interactions has been mostly limited to pairwise combinations of microbes. Given the diversity of microbes within hosts, indirect interactions among more than two microbial species may also impact disease. To test this hypothesis, we performed inoculation experiments that investigated interactions among two fungal parasites, Rhizoctonia solani and Colletotrichum cereale, and a systemic fungal endophyte, Epichloë coenophiala, within the grass, tall fescue (Lolium arundinaceum). Both direct and indirect interactions impacted disease progression. While the endophyte did not directly influence R. solani disease progression or C. cereale symptom development, the endophyte modified the interaction between the two parasites. The magnitude of the facilitative effect of C. cereale on the growth of R. solani tended to be greater when the endophyte was present. Moreover, this interaction modification strongly affected leaf mortality. For plants lacking the endophyte, parasite co-inoculation did not increase leaf mortality compared to single-parasite inoculations. By contrast, for endophyte-infected plants, parasite co-inoculation increased leaf mortality compared to inoculation with R. solani or C. cereale alone by 1.9 or 4.9 times, respectively. Together, these results show that disease progression can be strongly impacted by indirect interactions among microbial symbionts.
... A recent study found the opposite result by showing that closely related green algal species had weaker competition and more facilitation than distantly related species (Narwani et al., 2017). While relationships between phylogenetic relatedness among community members and ecosystem function have been assessed in bacterial systems (Tan et al., 2012;Galand et al., 2015;Roger et al., 2016), most work has focused on low-diversity, experimentally-assembled communities with bacteria that can be grown in culture. We need to expand these findings to communities with richness levels typically found in natural communities. ...
... and Russel et al. (2017) inherently break up potential interdependent relationships between bacteria either by creating artificial communities or evaluating pairwise interactions and remove the natural effect of spatial heterogeneity, environmental fluctuations, and the rest of the bacterial community. As a result, future studies on bacterial interactions and the role of phylogenetic diversity will need to maintain natural structure and complexity in bacterial communities.Previous studies on bacterial BEF relationships have used three approaches to manipulate bacterial diversity(Krause et al., 2014): (1) dilution to extinction in which complex communities are diluted to more simple communities(Wertz et al., 2006;Peter et al., 2011;Philippot et al., 2013; see Roger et al., 2016 for a review of this approach), (2) manually assembled communities in culture(Tan et al., 2012;Salles et al., 2009), or (3) natural or manipulated environmental communities ...
Thesis
This dissertation helps to integrate bacteria into the broader field of ecology by investigating bacterial community composition and diversity as it relates to ecosystem function in microhabitats within freshwater systems of the Great Lakes Region. Here, I combine field- and laboratory-based measurements of observational data collected from three major types of lake ecosystems: inland lakes, a freshwater estuary (Muskegon Lake), and a Great Lake (Lake Michigan). First, to determine the primary controls on lake bacterial community composition, I assessed the influence of lake layer (i.e. stratification), lake productivity, and particle-association on the bacterial community across 11 inland lakes with varying productivity in Southwestern Michigan. I found that particle-association very strongly structures freshwater bacterial community composition. Second, I studied a freshwater estuarine lake, Muskegon Lake, which has a large spatio-temporal variation in bacterial heterotrophic productivity, to test whether there was an association between heterotrophic production and bacterial biodiversity (defined as the number of taxa and taxon abundance). I specifically focused on two co-occurring freshwater habitats that my first chapter showed to be populated by very distinct communities: particle-associated and free-living. Positive biodiversity-heterotrophic productivity relationships were found only in particles. Third, I performed a genome-based analysis of free-living specialists, particle-associated bacterial specialists, and generalists to characterize the genomic architecture and genetic traits that are associated with adaptations to these specific habitats. The genomes of particle-associated specialist bacteria were about twice the size of the genomes of free-living specialists and generalists, which had streamlined genomes. Fourth, to identify the bacterial taxa driving heterotrophic productivity across the large set of lake samples, I found that high nucleic acid (i.e., HNA) functional groups identified by flow cytometry can serve as a proxy for freshwater bacterial heterotrophic productivity, whereas low nucleic acid (i.e., LNA) functional groups cannot. Then, I used a machine learning approach to identify bacterial taxa associated with HNA and LNA. This allowed me to identify the bacterial taxa, which were often members of the Phylum Bacteroidetes, that are associated heterotrophic productivity. Finally, I investigated patterns of lake specificity and phylogenetic conservation of taxonomic groups. Throughout my dissertation, I found that there was very deep (Class to Phylum-level) phylogenetic conservation of which bacteria lived in which habitats, but not of what bacterial taxa contributed to HNA and LNA functional groups, and thus heterotrophic productivity. Positive biodiversity-heterotrophic productivity relationships only existed in particle-associated, and not free-living communities, and communities composed of more phylogenetically related organisms were more productive per-capita. These differences in biodiversity-ecosystem function relationships may in part be explained by particle-associated bacteria having larger genomes, higher nitrogen content, and more unique genes that provide the potential for niche complementarity. The taxa that drove HNA and LNA cell numbers, and by proxy heterotrophic productivity, were lake and time-specific and indicated that taxa could switch between the two functional groups. Overall, my dissertation elucidates the ecological and evolutionary effects of microhabitat structure on bacterial communities and genomes in natural systems.
... Additionally, phylogenetic diversity enhances the quality of litter and plays a critical role in nutrient cycling within ecosystems (Xiao et al., 2020). Therefore, phylogenetic diversity is a more significant predictor of soil multifunctionality in studies of forest-grassland transition zones than taxonomic and functional diversity (Tan et al., 2012;Venail et al., 2015). Henceforth, emphasis should be placed on phylogenetic diversity in the examination of arid forest-grassland ecosystems, rather than solely on taxonomic and functional diversity. ...
Article
Full-text available
Exploring plant diversity and ecosystem functioning in different dimensions is crucial to preserve ecological balance and advance ecosystem conservation efforts. Ecosystem transition zones serve as vital connectors linking two distinct ecosystems, yet the impact of various aspects of plant diversity (including taxonomic, functional, and phylogenetic diversity) on soil multifunctionality in these zones remains to be clarified. This study focuses on the forest-grassland transition zone in the mountains on the northern slopes of the Tianshan Mountains, and investigates vegetation and soil characteristics from forest ecosystems to grassland ecosystems to characterize plant diversity and soil functioning, as well as the driving role of plant diversity in different dimensions. In the montane forest-grassland transition zone, urease (URE) and total nitrogen (TN) play a major role in regulating plant diversity by affecting the soil nutrient cycle. Phylogenetic diversity was found to be the strongest driver of soil multifunctionality, followed by functional diversity, while taxonomic diversity was the least important driver. Diverse species were shown to play an important role in maintaining soil multifunctionality in the transition zone, especially distantly related species with high phylogeny. The study of multidimensional plant diversity and soil multifunctionality in the montane forest-grassland transition zone can help to balance the relationship between these two elements, which is crucial in areas where the ecosystem overlaps, and the application of the findings can support sustainable development in these regions.
... During the process of microbiome assembly, organisms arriving early often undergo niche preemption, particularly through competitive utilization of nutrients, a phenomenon known as exploitative competition. This process plays a crucial role in shaping the composition and dynamics of the microbiota [13,14]. To summarize, microorganisms that establish themselves in the host's early-life are more likely to secure suitable niche opportunities compared to external colonizers when the gut microbiota has reached a state of dynamic equilibrium. ...
Article
Full-text available
Background The establishment of a robust gut microbiota in piglets during their early developmental stage holds the potential for long-term advantageous effects. However, the optimal timeframe for introducing probiotics to achieve this outcome remains uncertain. Results In the context of this investigation, we conducted a longitudinal assessment of the fecal microbiota of 63 piglets at three distinct pre-weaning time points. Simultaneously, we gathered vaginal and fecal samples from 23 sows. Employing 16S rRNA gene and metagenomic sequencing methodologies, we conducted a comprehensive analysis of the fluctuation patterns in microbial composition, functional capacity, interaction networks, and colonization resistance within the gut microbiota of piglets. As the piglets progressed in age, discernible modifications in intestinal microbial diversity, composition, and function were observed. A source-tracking analysis unveiled the pivotal role of fecal and vaginal microbiota derived from sows in populating the gut microbiota of neonatal piglets. By D21, the microbial interaction network displayed a more concise and efficient configuration, accompanied by enhanced colonization resistance relative to the other two time points. Moreover, we identified three strains of Ruminococcus sp. at D10 as potential candidates for improving piglets' weight gain during the weaning phase. Conclusions The findings of this study propose that D10 represents the most opportune juncture for the introduction of external probiotic interventions during the early stages of piglet development. This investigation augments our comprehension of the microbiota dynamics in early-life of piglets and offers valuable insights for guiding forthcoming probiotic interventions.
... Luc kil y, k e y predicti ve traits are often phylogeneticall y conserv ed (Martin y et al. 2015 ) and thus v ary pr edictabl y with phylogen y. As a consequence, phylogenetic similarity can increase competition by intensifying niche overlap, as has been shown in bacterial (Tan et al. 2012 ), fungal (Taylor et al. 2014 ), y east (P eay et al. 2012 ), and arbuscular mycorrhizal (Maher ali and Klir onomos 2012 ) comm unities . T her efor e, phylogenetic relatedness has been suggested as a strong predictor of competitive outcomes . ...
Article
Full-text available
Ectomycorrhizal fungi are among the most prevalent fungal partners of plants and can constitute up to one-third of forest microbial biomass. As mutualistic partners that supply nutrients, water, and pathogen defense, these fungi impact host plant health and biogeochemical cycling. Ectomycorrhizal fungi are also extremely diverse, and the community of fungal partners on a single plant host can consist of dozens of individuals. However, the factors that govern competition and coexistence within these communities are still poorly understood. In this study, we used in vitro competitive assays between five ectomycorrhizal fungal strains to examine how competition and pH affect fungal growth. We also tested the ability of evolutionary history to predict outcomes of fungal competition. We found that the effects of pH and competition on fungal performance varied extensively, with changes in growth media pH sometimes reversing competitive outcomes. Furthermore, when comparing the use of phylogenetic distance and growth rate in predicting competitive outcomes, we found that both methods worked equally well. Our study further highlights the complexity of ectomycorrhizal fungal competition and the importance of considering phylogenetic distance, ecologically relevant traits, and environmental conditions in predicting the outcomes of these interactions.
... For this reason, it is media widely used in microbial experiments. Protozoan pellets were used as they are created from plant material that encompasses a large diversity of carbon sources that facilitate bacterial growth (Tan et al., 2012). Samples were then diluted to a common density and pipetted into a 96-well plate with 200 μl of filtered and autoclaved protozoan medium. ...
Article
Full-text available
Understanding the ecological processes that underpin the dynamics of community turnover in response to environmental change is critical to predicting how warming will influence ecosystem functioning. Here, we quantify the effect of changing temperature on community composition and ecosystem functioning via the action of ecological selection on population-level thermal traits. To achieve this, we use microbes isolated from a network of geothermal streams in Iceland where in situ temperatures span 8–38°C within a single catchment. We first quantified variability in thermal tolerance between taxa, and then assembled synthetic communities along a broad thermal gradient to explore how temperature-driven selection on thermal tolerance traits shaped the emergent community structures and functions. We found marked changes in community structure and composition with temperature, such that communities exposed to extreme temperatures (10, 35°C) had highly asymmetric biomass distributions and low taxonomic richness. Thermal optima were a good predictor of the presence and relative abundance of taxa in the high-temperature treatments. We also found that the evenness of the abundance distribution was related to ecosystem production, such that communities with more equitable abundance distribution were also the most productive. Our results highlight the utility of using a multi-level approach that links population-level traits with community structure and ecosystem functioning to better understand how ecological communities will respond to global warming.
... A major limitation of our approach is that we have focused on species richness or species composition. The general concept of priority effects also covers the functional properties of ecological communities, such as energy flow and productivity (Dickie et al., 2012;Fukami & Morin, 2003;Tan et al., 2012). A possible solution is to establish a functional map from the nodes or the links in the assembly graph onto the functional property (e.g. ...
Article
Full-text available
The history of species immigration can dictate how species interact in local communities, thereby causing historical contingency in community assembly. Since immigration history is rarely known, these historical influences, or priority effects, pose a major challenge in predicting community assembly. Here, we provide a graph‐based, non‐parametric, theoretical framework for understanding the predictability of community assembly as affected by priority effects. To develop this framework, we first show that the diversity of possible priority effects increases super‐exponentially with the number of species. We then point out that, despite this diversity, the consequences of priority effects for multispecies communities can be classified into four basic types, each of which reduces community predictability: alternative stable states, alternative transient paths, compositional cycles and the lack of escapes from compositional cycles to stable states. Using a neural network, we show that this classification of priority effects enables accurate explanation of community predictability, particularly when each species immigrates repeatedly. We also demonstrate the empirical utility of our theoretical framework by applying it to two experimentally derived assembly graphs of algal and ciliate communities. Based on these analyses, we discuss how the framework proposed here can help guide experimental investigation of the predictability of history‐dependent community assembly. How predictable is the assembly of species‐rich communities? We provide a graph‐based, non‐parametric, theoretical framework for understanding the predictability of community assembly as affected by priority effects.
... Peay et al. (2012) observed priority effects in nectar yeast communities, and used pairwise phylogenetic distances to show that effects were stronger between close relatives, supporting the niche-pre-emption hypothesis. Later arrivals were less likely to establish and attain high densities when initial colonisers were close relatives, but early arriving species experienced little negative impacts from later arrivals, irrespective of relatedness (see Tan et al. [2012] for similar findings). Migrant communities offer another useful study system, effectively allowing observations of priority effects following each seasonal migration. ...
Article
Full-text available
Species' evolutionary histories shape their present-day ecologies, but the integration of phylogenetic approaches in ecology has had a contentious history. The field of ecophylogenetics promised to reveal the process of community assembly from simple indices of phylogenetic pairwise distances - communities shaped by environmental filtering were composed of closely related species, whereas communities shaped by competition were composed of less closely related species. However, the mapping of ecology onto phylogeny proved to be not so straightforward, and the field remains mired in controversy. Nonetheless, ecophylogenetic methods provided important advances across ecology. For example the phylogenetic distances between species is a strong predictor of pest and pathogen sharing, and can thus inform models of species invasion, coexistence and the disease dilution/amplification effect of biodiversity. The phylogenetic structure of communities may also provide information on niche space occupancy, helping interpret patterns of facilitation, succession and ecosystem functioning - with relevance for conservation and restoration - and the dynamics among species within foodwebs and metacommunities. I suggest leveraging advances in our understanding of the process of evolution on phylogenetic trees would allow the field to progress further, while maintaining the essence of the original vision that proved so seductive.
Article
Advances in next-generation sequencing have enabled the widespread measurement of microbiome composition across systems and over the course of microbiome assembly. Despite substantial progress in understanding the deterministic drivers of community composition, the role of historical contingency remains poorly understood. The establishment of new species in a community can depend on the order and/or timing of their arrival, a phenomenon known as a priority effect. Here, we review the mechanisms of priority effects and evidence for their importance in microbial communities inhabiting a range of environments, including the mammalian gut, the plant phyllosphere and rhizosphere, soil, freshwaters and oceans. We describe approaches for the direct testing and prediction of priority effects in complex microbial communities and illustrate these with re-analysis of publicly available plant and animal microbiome datasets. Finally, we discuss the shared principles that emerge across study systems, focusing on eco-evolutionary dynamics and the importance of scale. Overall, we argue that predicting when and how current community state impacts the success of newly arriving microbial taxa is crucial for the management of microbiomes to sustain ecological function and host health. We conclude by discussing outstanding conceptual and practical challenges that are faced when measuring priority effects in microbiomes. The order and timing of the arrival (priority effects) of members of a microbiome can influence microbiome composition and function. In this Review, Debray and colleagues provide an overview of the mechanisms of priority effects, highlight examples in host-associated and environmental communities, and discuss methods to detect priority effects in microbial communities.
Article
Full-text available
Functional diversity is an important component of biodiversity, yet in comparison to taxonomic diversity, methods of quantifying functional diversity are less well developed. Here, we propose a means for quantifying functional diversity that may be particularly useful for determining how functional diversity is related to ecosystem functioning. This measure of functional diversity ''FD'' is defined as the total branch length of a functional dendrogram. Various characteristics of FD make it preferable to other measures of functional diversity, such as the number of functional groups in a community. Simulating species' trait values illustrates how the relative importance of richness and composition for FD depends on the effective dimensionality of the trait space in which species separate. Fewer dimensions increase the importance of community composition and functional redundancy. More dimensions increase the importance of species richness and decreases functional redundancy. Clumping of species in trait space increases the relative importance of community composition. Five natural communities show remarkably similar relationships between FD and species richness.
Article
The role functionally redundant species play in ecosystem function has not been adequately investigated. To study this, we examined species richness and an ecosystem function, cellulose decomposition, while environmental conditions were held constant. Our hypotheses were (1) increasing species richness will have no effect on rates of cellulose decomposition and (2) species richness will decline over time in functionally redundant communities. A relatively simple microcosm-based system to manipulate complex microbial interactions was employed. Microcosms containing cellulose as the sole carbon source were inoculated at equal densities with none, one, two, four, or eight species of cellulolytic bacteria. At 5-d intervals for 25 d, community composition and cellulose decomposition were determined. We rejected both of our hypotheses. In a constant environment, greater species richness supported a greater number of individuals and subsequently greater rates of total cellulose decomposition. Furthermore, greater initial species richness maintained greater richness over time. These results provide experimental evidence that functionally redundant species may play an integral role in ecosystem function.
Article
Objective, empirical measures of overlap between samples of items distributed proportionally into various qualitative categories are presented and reviewed. These indices of overlap, derived from either probability or information theory, should prove useful to the ecologist in comparative studies of diet, habitat preference, seasonal patterns of abundance, faunal lists, or similar data.
Article
-It was,relatively easy to produce,alternative community,states as a function,of variability in a sequence,of species invasions,employed,to assemble,a community.,Numerous mechanisms,and processes,are capable,of producing,both temporary,and nonrecoverable,differ- ences in community structure. They include priority effects, intransitivities, emergent properties (e.g., vulnerability to invasion, specific topologies), and effects specific to differences in assem- bly sequences themselves. In ecological systems, the existence of alternative states presents a difficult comparative,problem,in the,search,for unifying,principles,that,might,underlie community-level,organization. This problem,can be solved only if we understand,the cause of alternative states and determine whether such states are persistent or transient. Otherwise,
Article
Didham et al. suggested that abiotically structured communities showing trait under-dispersion are more likely to exhibit alternative stable states than are competitively structured communities showing trait over-dispersion. We argue that the opposite is the logical expectation, and discuss implications for ecological restoration at local and regional scales.
Article
Key Words community assembly and organization, phylogenetic conservatism, biogeography, species diversity, niche differentiation s Abstract As better phylogenetic hypotheses become available for many groups of organisms, studies in community ecology can be informed by knowledge of the evo-lutionary relationships among coexisting species. We note three primary approaches to integrating phylogenetic information into studies of community organization: 1. examining the phylogenetic structure of community assemblages, 2. exploring the phylogenetic basis of community niche structure, and 3. adding a community context to studies of trait evolution and biogeography. We recognize a common pattern of phy-logenetic conservatism in ecological character and highlight the challenges of using phylogenies of partial lineages. We also review phylogenetic approaches to three emer-gent properties of communities: species diversity, relative abundance distributions, and range sizes. Methodological advances in phylogenetic supertree construction, charac-ter reconstruction, null models for community assembly and character evolution, and metrics of community phylogenetic structure underlie the recent progress in these ar-eas. We highlight the potential for community ecologists to benefit from phylogenetic knowledge and suggest several avenues for future research.
Article
Protecting biological diversity with limited resources may require placing conservation priorities on different taxa. A system of priorities that reflects the value of taxonomic diversity can be achieved by setting priorities such that the subset of taxa that is protected has maximum underlying feature diversity. Such feature diversity of taxon subsets is difficult to estimate directly, but can be predicted by the cladistic/phylogenetic relationships among the taxa. In this study, a simple measure of phylogenetic diversity is defined based on cladistic information. The measure of phylogenetic diversity, PD, is contrasted with a measure of taxic diversity recently developed by Vane-Wright et al. (Biol. Conserv., 55, 1991). In re-examining reserve-selection scenarios based on a phylogeny of bumble bees (Apidae), PD produces quite different priorities for species conservation, relative to taxic diversity. The potential application of PD at levels below that of the species is then illustrated using a mtDNA phylogeny for populations of crested newts Triturus cristatus. Calculation of PD for different population subsets shows that protection of populations at either of two extremes of the geographic range of the group can significantly increase the phylogenetic diversity that is protected.