ArticlePDF Available

Wasp Gene Expression Supports an Evolutionary Link Between Maternal Behavior and Eusociality

Authors:

Abstract and Figures

The presence of workers that forgo reproduction and care for their siblings is a defining feature of eusociality and a major challenge for evolutionary theory. It has been proposed that worker behavior evolved from maternal care behavior. We explored this idea by studying gene expression in the primitively eusocial wasp Polistes metricus. Because little genomic information existed for this species, we used 454 sequencing to generate 391,157 brain complementary DNA reads, resulting in robust hits to 3017 genes from the honey bee genome, from which we identified and assayed orthologs of 32 honey bee behaviorally related genes. Wasp brain gene expression in workers was more similar to that in foundresses, which show maternal care, than to that in queens and gynes, which do not. Insulin-related genes were among the differentially regulated genes, suggesting that the evolution of eusociality involved major nutritional and reproductive pathways.
P. metricus wasp brain gene expression analysis tests the prediction that maternal and worker (eusocial) behavior share a common molecular basis. (A) Similarities and differences in reproductive and brood provisioning status for the four behavioral groups analyzed in this study: foundresses (n = 22), gynes (n = 20), queens (n = 23), and workers (n = 22). Each individual wasp (total of 87) was assigned to a behavioral group on the basis of physiological measurements (14). (B to D) Results for 28 genes selected for their known involvement in worker (honey bee) behavior. (B) Heatmap of mean expression values by group and a summary of analysis of variance (ANOVA) results for each gene. Genes were clustered by K-means clustering (37); those in red showed significant differences (ANOVA, P < 0.05; table S1) between the behavioral groups. P. metricus gene names were assigned on the basis of orthology to honey bee genes (reference in parentheses); putative functions were assigned on the basis of similarity to Drosophila melanogaster genes. (C) Results of linear discriminant analysis show that foundress and worker brain profiles are more similar to each other than to the other groups. (D) Results of hierarchical clustering show the same result (based on group mean expression value for each gene). Four genes (PmVg, Pmg5sd, PmGlyP, and PmRfaBp) were excluded from these analyses because they showed high levels of expression in tissue adjacent to the brain (fig. S2); results for all three analyses were similar with and without these four genes (fig. S3).
… 
Content may be subject to copyright.
DOI: 10.1126/science.1146647
, 441 (2007); 318Science
et al.Amy L. Toth,
Link Between Maternal Behavior and Eusociality
Wasp Gene Expression Supports an Evolutionary
www.sciencemag.org (this information is current as of March 22, 2009 ):
The following resources related to this article are available online at
http://www.sciencemag.org/cgi/content/full/318/5849/441
version of this article at:
including high-resolution figures, can be found in the onlineUpdated information and services,
http://www.sciencemag.org/cgi/content/full/1146647/DC1
can be found at: Supporting Online Material
http://www.sciencemag.org/cgi/content/full/318/5849/441#otherarticles
, 10 of which can be accessed for free: cites 25 articlesThis article
1 article(s) on the ISI Web of Science. cited byThis article has been
http://www.sciencemag.org/cgi/content/full/318/5849/441#otherarticles
1 articles hosted by HighWire Press; see: cited byThis article has been
http://www.sciencemag.org/cgi/collection/genetics
Genetics
: subject collectionsThis article appears in the following
http://www.sciencemag.org/about/permissions.dtl
in whole or in part can be found at: this article
permission to reproduce of this article or about obtaining reprintsInformation about obtaining
registered trademark of AAAS.
is aScience2007 by the American Association for the Advancement of Science; all rights reserved. The title
CopyrightAmerican Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005.
(print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by theScience
on March 22, 2009 www.sciencemag.orgDownloaded from
Atlantic are available, except a recent record
from the Irish margin by Peck et al.(23). The
authors recorded identical d
18
O N. pachyderma (s)
and G. bulloides values during the LGM that were
attributed to a continuous discharge of meltwater
from the British Ice Sheet and year round mixing
that homogenized the upper waters.
Several processes can create variability in the
d
18
O of foraminifera. Seasonal variability was
interpreted by Ganssen and Kroon (19) to explain
why G. bulloides d
18
O was more positive than
G. inflata d
18
O in the modern North Atlantic at
57°N, which was attributed to a later seasonal
period of G. bulloides production further south.
The uniform d
18
O of the foraminifera during HEs
would require improbable ecological changes in
preferred depth-habitat zones or in seasonal
behavior if these values were not the result of
uniform upperwater-mass conditions. Upwell-
ing of
18
O-depleted water produced by brine-
rejection at higher latitudes might affect the d
18
O
of deeper-dwelling foraminifera; however , it is
difficult to imagine it influencing d
18
Ovaluesin
all three taxa. W ithout a decrease in temperature
by ~2.5°C, it is impossible to lower the salinity
by 0.8 per mil ()(24)(andbyextensionto
lower d
18
Oby0.5) while remaining along the
same isopycnal surface, which by itself would
correspond to a 0.5 increase in the d
18
Oof
calcite. Because these effects cancel each other
out, such a mechanism is inadequate to explain
the anomalously low d
18
O values in all three
planktonic foraminifera (Fig. 3B).
Intensified vertical mixing and deepening of
the mixed layer during HEs is the mechanism
remaining to explain the data. Atmospheric con-
ditions directly influence the mixed layer through
turbulence, and wind driven Langmuir circula-
tion could be the prime driver of the turbulence
(25, 26). As a result, the upper ocean often be-
comes well mixed to depths as great as 600 m
(27). Our d
18
O data from planktonic foramini-
fera that live at different depth ranges illustrate
the extent of this process, suggesting that during
the times of HEs the near-surface waters were
homogenized by stronger mixing.
During the last glacial cycle, large ice sheets
in the Northern Hemisphere and steeper merid-
ional temperature gradients in the atmosphere
must have reorganized atmospheric circulation.
As a result, winter sea-ice cover extended fur-
ther south, and glacial winds were stronger and
more zonal (28, 29). These winds would have
intensified the vertical mixing and turbulence in
the upper water masses. It is counterintuitive to
visualize such a mechanism during HEs when
the glacial North Atlantic was flooded with melt-
water resulting in stronger stratification. However,
our d
18
O data demonstrate that homogeniza-
tion of upper water masses did occur, suggesting
that this mechanism functioned at the core site.
Additional evidence of this mechanism comes
from the measurements of Ca
+2
and Na
+
ions
derived from sea salt and continental dust in the
Greenland ice core (30). Both Ca
+2
and Na
+
ions
in the ice core show rapid increases from their am-
bient concentration during stadials. Abrupt, many-
fold increases of these chemical species suggest
that storminess during the glacial period caused
stronger vertical mixing at the atmosphere/ocean
boundary at the subpolar and subtropical fronts.
The question of why weaker homogenization
of near-surface waters occurred during other D/O
cycles not associated with HEs could be raised,
because Greenland ice core data show similar
patterns of glaciochemical species. One possibil-
ity is that unfavorable composition or insufficient
volumes of meltwater were available to perturb
the near-surface waters during these D/O ice-
rafting cycles as the icebergs originated from
the smaller ice sheets. Hence, even though the
glacial climate was windier and stormier, the
near-surface waters continued to be stratified.
References and Notes
1. H. Heinrich, Quat. Res. 29, 143 (1988).
2. G. C. Bond et al., Nature 365, 143 (1993).
3. H. Rashid, R. Hesse, D. J. W. Piper, Paleoceanography
18, 1077 (2003).
4. R. B. Alley, D. R. MacAyeal, Paleoceanography 9, 503
(1994).
5. C. Huber et al., Earth Planet. Sci. Lett. 243, 504 (2006).
6. S. Manabe, R. J. Stouffer, Nature 378, 165 (1995).
7. D. Rind et al., J. Geophys. Res. 106, 27335 (2001).
8. J. Flückiger, R. Knutti, J. W. C. White, Paleoceanography
21, 1204 (2006).
9. E. A. Boyle, L. Keigwin, Earth Planet. Sci. Lett. 76, 135
(1985/86).
10. W. F. Ruddiman, Geol. Soc. Am. Bull. 88, 1813
(1977).
11. L. Keigwin, S. Lehman, Paleoceanography 9, 185 (1994).
12. M. Stuiver, P. J. Reimer, Radiocarbon 35, 215 (1993).
13. R. G. Fairbanks et al., Quat. Sci. Rev. 24, 1781 (2005).
14. H. Rashid, R. Hesse, D. J. W. Piper, Earth Planet. Sci. Lett.
208, 319 (2003).
15. H. Rashid, E. Grosjean, Paleoceanography 21, 1240
(2006).
16. J. J. Ottens, Oceanol. Acta 14, 123 (1991).
17. W. G. Deuser, J. Foraminifer. Res. 17, 14 (1987).
18. R. G. Fairbanks, P. H. Wiebe, A. W. , Science 207,61
(1980).
19. G. Ganssen, D. Kroon, J. Geol. Soc. 157, 693 (2000).
20. S. Mulitza, A. rkoop, W. Hale, G. Wefer, H. S. Niebler,
Geology 25, 335 (1997).
21. J.R.Luyten,A.J.Pedlosky,H.Stommel,J. Phys. Oceanogr.
13, 192 (1983).
22. L. Labeyrie et al., AGU Monogr. 112, 77 (1999).
23. V. Peck et al., Earth Planet. Sci. Lett. 243, 476 (2006).
24. G. L. Pickard, W. J. Emery, Descriptive Physical
Oceanography (Pergamon, Oxford, ed. 5, 1990).
25. S. K. Gulev, B. Barnier, H. Knochel, J.-M. Molines, M. Cottet,
J. Clim. 16, 3085 (2003).
26. K. Hanawa, T. Suga, in Ocean-Atmosphere Interactions,
Y. Toba, Ed. (Kluwer Academic, Tokyo, 2003), pp. 63109.
27. M. K. Robinson et al., Atlas of the North Atllantic-Indian
Ocean Monthly Mean Temperatures and Mean Salinities
of the Surface Layer (U.S. Naval Oceanogr. Office
Reference Publication 18, Washington, DC, 1978).
28. M. Sarnthein, U. Pflaumann, M. Weinelt,
Paleoceanography 18, 771 (2003).
29. H. Gildor, E. Tzipperman, Philos. Trans. R. Soc. London
Ser. A 361, 1935 (2003).
30. P. A. Mayewski et al., Science 263, 1747 (1994).
31. We thank E. Goddard for helping to acquire part of the
isotope data and D. J. W. Piper and B. P. Flower for
discussions to improve an initial version of the manu-
script. H.R. thanks Fonds pour la Formation de
Chercheurs et l'Aide à la Recherche, Québec, for its
support through a postdoctoral fellowship. E.A.B. was
supported by grants from NSF and the Cambridge
Massachusetts Institute of Technology.
Supporting Online Material
www.sciencemag.org/cgi/content/full/1146138/DC1
SOM Text
Fig. S1
Table S1
References
6 June 2007; accepted 11 September 2007
Published online 20 September 2007;
10.1126/science.1146138
Include this information when citing this paper.
Wasp Gene Expression Supports an
Evolutionary Link Between Maternal
Behavior and Eusociality
Amy L. Toth,
1
*
Kranthi Varala,
2
Thomas C. Newman,
1
Fernando E. Miguez,
2
Stephen K. Hutchison,
3
David A. Willoughby,
3
Jan Fredrik Simons,
3
Michael Egholm,
3
James H. Hunt,
4
Matthew E. Hudson,
2
Gene E. Robinson
1,5
The presence of workers that forgo reproduction and care for their siblings is a defining feature of
eusociality and a major challenge for evolutionary theory. It has been proposed that worker behavior
evolved from maternal care behav ior. We explor ed this idea by studying gene expression in the
primitively eusocial wasp Polistes metricus . Because little genomic information existed for this species,
we used 454 sequencing to generate 391,157 brain complementary DNA reads, resulting in robust hits
to 3017 genes from the honey bee genome, from which we identified and assayed orthologs of 32
honey bee behaviorally related genes. Wasp brain gene expression in workers was more similar to
that in foundresses, which show maternal care, than to that in queens and gynes, which do not.
Insulin-related genes were among the differentially regulated genes, suggesting that the evolution of
eusociality involved major nutritional and reproductive pathways.
A
major challenge in biology is to under-
stand the evolution of animal society in
molecular terms. Eusociality is the most
extreme form of cooperation, typified by indi-
viduals that care for siblings rather than repro-
duce themselves, i.e., workers. The evolution
www.sciencemag.org SCIENCE VOL 318 19 OCTOBER 2007
441
REPORTS
on March 22, 2009 www.sciencemag.orgDownloaded from
of eusociality has been ascribed to kin or colony-
level selection (1, 2), but these explanations do
not specify mechanistic routes.
It has long been suggested (35) that sibling
care by hymenopteran (ant, bee, wasp) workers
evolved from maternal care, which involves pro-
visioning brood by foraging for food and then
feeding them. According to this idea, two prin-
cipal behaviors exhibited by solitary Hymenop-
tera, reproduction (egg-laying) and maternal care
(brood provisioning), became uncoupled during
the early stages of social evolution (6), and these
behaviors eventua lly occurred in separate castes,
queens and workers, respectively (7). Linksvayer
and Wade (8) added a molecular dimension to
this idea by predicting that sibling care and
maternal care behaviors should be regulated by
similar patterns of gene expression.
We used Polistes paper wasps to test Linksvayer
and Wades idea. Polistes are primitively eusocial,
which means that although individuals special-
ize as either workers or reproductive individu-
als, these two castes are less distinct than in
advanced eusocial species. In Polistes, both
workers and reproductives display provision-
ing behavior, but at different points in the life
of a colony. Advanced eusocial insects, by con-
trast, have morphologically distinct queen and
worker castes, and in some species, such as the
honey bee, queens no longer exhibit any ma-
ternal care, which precludes comparing the
molecular basis of sibling and maternal care.
Primitively eusocial insects like Polistes afford
the opportunity to explore the molecular basis
of maternal and worker behavior within a sin-
gle species.
We measured brain gene expression in 87
individuals from four distinct behavioral groups
of females from naturally occurring colonies of
the temperate species Polistes metricus (Fig. 1A).
Foundresses are females that establish new col-
onies in the spring, often as solitary individuals.
Foundresses exhibit both reproductive (egg-
laying) and maternal (foraging and brood-
feeding) behavior . After rearing a first generation
of female brood that develop into workers,
successful foundresses become queens and cease
caring for brood. W orkers take over provisioning
the broodtheir siblingsby foraging for food
and then feeding them; workers show little, if
any, reproductive behavior . By contrast, queens
focus exclusively on reproductive behavior.
Gynes are reared late in the season; they engage
in no reproductive or maternal care behavior (9).
After successfully mating, gynes overwinter and
then become foundresses (10). We hypothesized
that brain gene expression patterns in P. m et r ic u s
workers and foundresses should be most similar
to each other from among these four groups, be-
cause they both show brood provisioning behav-
ior despite their different reproductive status.
Alternatively, if brain gene expression more
closely reflects reproductive behavior, expression
in foundresses and queens should be most similar
to each other .
Social behavior is a complex and polygenic
trait, so an appropriate test of the idea that ma-
ternal and worker behavior share a common mo-
lecular basis requires analysis of multiple genes
in different pathways. But Polistes wasps, though
venerable models for studies of social evolution
(11, 12), have until recently lacked genomic
sequence information (13). T o provide a ready
source of test genes for quantitative reverse
transcriptionpolymerase chain reaction analysis,
we used 454 sequencing to obtain 45 megabases
(Mb) in 391,157 cDNA sequence fragments
from the P. m et r i cu s brain transcriptome (14).
We were interested to see whether this low-cost,
high-throughput sequencing method would be
successful for this purpose, despite short se-
1
Department of Entomology and Institute for Genomic
Biology, University of Illinois at Urbana-Champaign, Urbana,
IL 61801, USA.
2
Department of Crop Sciences, University of
Illinois at Urbana-Champaign, Urbana, IL 61801, USA.
3
454
Life Sciences, Branford, CT 06405, USA.
4
Department of Bi-
ology, University of Missouri at St. Louis, St. Louis, MO 63121,
USA.
5
Neuroscience Program, University of Illinois at Urbana-
Champaign, Urbana, IL 61801, USA.
*To whom correspondence should be addressed. E-mail:
amytoth@uiuc.edu
Fig. 1. P. metricus wasp
brain gene expression
analysis tests the predic-
tion that maternal and
worker (eusocial) behavior
share a common molecu-
lar basis. (A) Similarities
and differences in repro-
ductive and brood provi-
sioning status for the four
behavioral groups ana-
lyzed in this study: found-
resses (n = 22), gynes (n =
20), queens (n =23),and
workers (n = 22). Each
individual wasp (total of
87) was assigned to a
behavioral group on the
basis of physiological
measurements (14). (B
to D) Results for 28 genes
selected for their known
involvement in worker
(honey bee) behavior. (B)
Heatmap of mean expres-
sion values by group and
a summary of analysis of
variance (ANOVA) results
for each gene. Genes were
clustered by K-means
clustering (37); those in
red showed significant
differences (ANOVA, P <
0.05; table S1) between
the behavioral groups.
P. metricus gene names
were assigned on the ba-
sis of orthology to honey
bee genes (reference in
parentheses); putative
functions were assigned
on the basis of similarity
to Drosophila melanogas-
ter genes. (C) Results of
linear discriminant anal-
ysis show that foundress
and worker brain profiles
are more similar to each
other than to the other
groups. (D) Results of hierarchical clustering show the same result (based on group mean expression
value for each gene). Four genes (PmVg, Pmg5sd, PmGlyP,andPmRfaBp) were excluded from these
analyses because they showed high levels of expression in tissue adjacent to the brain (fig. S2); results
for all three analyses were similar with and without these four genes (fig. S3).
19 OCTOBER 2007 VOL 318 SCIENCE www.sciencemag.org
442
REPORTS
on March 22, 2009 www.sciencemag.orgDownloaded from
quence read lengths (average of 120 bp) and an
estimated 100- to 150-million-year divergence
time between P. metricus and the honey bee,
Apis mellifera (15), the most closely related spe-
cies with a sequenced genome to use as refer-
ence (16).
We generated a map of the honey bee genome
combined with known transcripts and their
relative abundance in the combined bee ex-
pressed sequence tag (EST) data sets (1618).
P. me t r icu s transcript fragments predicted to
encode proteins orthologous to those encoded
by A. mellifera genes were plotted on the map
according to the number of fragments identified
for a particular locus; matches were found for
39% of all honey bee mRNAs. The relative abun-
dance of P. metricus sequence fragments corre-
sponded well with the abundance of A. mellifera
ESTs for the respective loci (Fig. 2). The com-
bined P. metricusA. mellifera transcriptome data
set was then used to select the genes for this study .
Prior information allowed us to focus on
genes implicated in honey bee foraging and pro-
visioning behavior, rather than a set of random-
ly chosen genes that might be less informative.
We selected 32 genes (Fig. 1B and table S1)
from the P. metricus EST set that are orthologs
of A. mellifera genes known to be associated in
some way with worker bee behavior, based on
results from studies with microarrays (22 genes)
(19, 20) and candidate genes (10 genes) (2129).
Twenty-two of the genes have been shown by
microarray analysis to be both differentially
expressed in the brains of honey bees engaged
in foraging or feeding brood [on the list of the
top 100 genes most consistently associated
with bee foraging behavior (19)] and regulated
by juvenile hormone (20), which also causes
worker bee foraging behavior (30). Five can-
didate genes are differentially expressed in hon-
ey bees engaged in foraging or feeding brood
(2124, 29), three of which also have been shown
to play causal roles in the regulation of worker
bee foraging behavior (21, 22, 31). Five addi-
tional candidate genes involved in insulin sig-
naling were selected because this pathway is
implicated in honey bee queen-worker caste
determination (25, 26, 32) and worker foraging
behavior (27, 28). Patterns of gene expression in
P. metricus were not used as criteria for gene
selection.
There was a robust association between indi-
vidual wasp brain gene expression and naturally
occurring behavioral differences among the wasp
groups. Leave-one-out cross-validation analysis
(19) resulted in 68, 69, 70, and 47% correct
assignments to the foundress, gyne, queen, and
worker groups, respectively . For the less con-
servative resubstitution method (33), the results
were 89, 100, 100, and 95%. The predictions
from both classification methods were signifi-
cantly better than random (Chi-square tests, P <
0.0001, 25% expected). This honey beederived
gene set thus demonstrates extensive brain regu-
lation across the four wasp groups, making it an
informative set to explore the molecular relation-
ship between maternal and worker behavior in
P. metricus.
Sixty-two percent of the genes in the gene set
were differentially regulated in P. m e tr i c u s as a
function of reproductive or provisioning behavior
(Fig. 1B and table S1). Multivariate analysis of
variance showed that brain gene expression
varied significantly with reproduction (F = 3.28,
P = 0.0002) and provisioning (F = 4.76, P <
0.0001), with a signi ficant provi sioning ×
reproduction interaction (F = 2.48, P = 0.002).
Three out of the five insulin-related genes
showed significant associations with provision-
ing and/or reproductive behavior, consistent with
known nutritional effects on behavior and phys-
iology in honey bees and other social insects (34).
Three statistical analyses demonstrated that
brain gene expression for worker wasps was
more similar to that of maternal females (found-
resses) than to that of females not showing
maternal care (queens and gynes). First, K-means
clustering (Fig. 1B and fig. S1) revealed five
clusters of coexpressed genes. The first cluster
contained genes (n = 8) that showed coex-
pression in foundresses and workers compared
to queens and gynes. The second cluster of
genes (n = 7) was mainly characterized by up-
regulation in gynes, and the third (n =6)by
down-regulation in queens, but in both of these
clusters, foundresses and workers also showed
patterns of expression that were similar to each
other (Fig. 1B and fig. S1).
The second statistical analysis, linear discrim-
inant (LD) analysis, also showed similarities be-
tween foundress and worker brain gene expression
(Fig. 1C). A plot of LD1 versus LD2 (which
accounted for 90% of the variation in brain gene
expression across all four groups) revealed group-
specific expression patterns, but foundresses and
workers showed the greatest overlap. This is
consistent with the poorer performance of clas-
sification methods (described above) for those
two groups; overlap in gene expression patterns
made them difficult to distinguish from each
other. Gynes, which engage in neither reproduc-
tive nor provisioning behavior , were the most
distinct group. The third statistical analysis,
hierarchical clustering by group, supported the
patterns found in the other two analysesbrain
gene expression of workers and foundresses was
most similar, and that of gynes was most
divergent (Fig. 1D).
There are marked temporal changes in brain
gene expression as females shift from found-
ress to queen status, i.e., from maternal to repro-
ductive behavior. These findings demonstrate
heterochronic expression of genes associated
with maternal behavior, a form of plasticity that
is considered to be necessary for the evolution of
worker behavior (8). They also reflect the ap-
parent modularity of egg-laying and brood pro-
visioning behavior and their underlying regulatory
networks; this type of modularity also is thought to
be important in the evolution of novel traits (35).
We used the honey bee genome, together with
next-generation sequencing technology, to
rapidly bring genomics to the relatively closely
Fig. 2. A representation of
P. metricus brain transcripts
overlaid on a honey bee
genome template (16)
shows wide coverage and
similar transcript abundance
for P. metricus relative to
known honey bee tran-
scripts. P. metricus brain
cDNA sequence fragments
were matched as predicted
proteins to A. mellifera tran-
scripts with experimental
support (known cDNA or
EST sequences). A. mellifera
transcripts (red points, right
of axis) and their closest
P. metricus orthologs from
our survey (blue points, left
of axis) were then mapped
to the corresponding ge-
nomic locus in the A.
mellifera genome. The verti-
cal lines represent A. melli-
fera chromosomes 1 to 16.
The distance of each point from the midline is proportional to the logarithm of the abundance of the
mRNA (the number of sequences for each P. metricus or A. mellifera transcript corresponding to the
A. mellifera gene at that locus) (1618). P. metricus orthologs were obtained for a total of 3017 A.
mellifera transc ripts. The P. metricus transcriptome data contained putative orthologs for 39% of
known A. mellifera mRNAs. An additional 252,556 transcript sequence fragments obtained from
P. metricus did not have a clearly orthologous transcript in A. mellifer a.
10 11 12 13 14 15 16
Apis mellifera chromosome
23 4 56 7819
number of
transcripts
10000
1000
100
10
1
Apis mellifera transcript
Polistes metricus transcript
10
6
base pairs
www.sciencemag.org SCIENCE VOL 318 19 OCTOBER 2007 443
REPORTS
on March 22, 2009 www.sciencemag.orgDownloaded from
related wasp P. m et r ic u s; this is an early example
of the utility of 454 sequencing for transcriptomics
(36). Our results demonstrate that it is possible
to use species that have had their genomes se-
quenced as hubs to efficiently generate genomic
resources for clusters of related species that might
eachbeespeciallywellsuitedtoaddressparticular
evolutionary problems. This h ub and spokes
approach should enable genomics to be deployed
for a broader range of species than is currently
being done, until whole-genome sequencing of
eukaryote genomes becomes routine.
References and Notes
1. E. O. Wilson, B. lldobler, Proc. Natl. Acad. Sci. U.S.A.
102, 13367 (2005).
2. L. Lehmann, L. Keller, J. Evol. Biol. 19, 1365 (2006).
3. W. M. Wheeler, The Social Insects: Their Origin and
Evolution (Harcourt, Brace, New York, 1928).
4. H. E. Evans, M. J. West-Eberhard, The Wasps (Univ. of
Michigan Press, Ann Arbor, MI, 1970).
5. J. H. Hunt, Evolution 53, 225 (1999).
6. M. J. West-Eberhard, in Natural History and Evolution of
Paper-Wasps, S. Turillazzi, M. J. West-Eberhard, Eds.
(Oxford Univ. Press, New York, 1996), pp. 290317.
7. E. O. Wilson, The Insect Societies (Belknap, Cambridge,
MA, 1971).
8. T. A. Linksvayer, M. J. Wade, Q. Rev. Biol. 80, 317
(2005).
9. M. J. West-Eberhard, Misc. Publ. Mus. Zool. Univ. Mich.
140, 1 (1969).
10. J. H. Hunt, The Evolution of Social Wasps (Oxford Univ.
Press, New York, 2007).
11. S. Turillaz zi, M. J. West-Eberhard, Eds., Natural History
and Evolution of Paper-Wasps (Oxford Univ. Press,
New York, 1996).
12. H. K. Reeve, in The Social Biology of Wasps,K.G.Ross,
R. W. Matthews, Eds. (Cornell Univ. Press, Ithaca, NY, 1991).
13. S. Sumner, J. J. M. Pereboom, W. C. Jordan, Proc. R. Soc.
London B Biol. Sci. 273, 19 (2006).
14. Materials and methods are available as supporting
material on Science Online.
15. B. N. Danforth, S. G. Brady, S. D. Sipes, A. Pearson,
Syst. Biol. 53, 309 (2004).
16. Honey Bee Genome Sequencing Consortium, Nature 443,
931 (2006).
17. F. M. F. Nunes et al., BMC Genomics 5, 84 (2004).
18. C. W. Whit field et al., Genome Res. 12, 555 (2002).
19. C. W. Whitfield, A. M. Cziko, G. E. Robins on, Science 302,
296 (2003).
20. C. W. Whit field et al., Proc. Natl. Acad. Sci. U.S.A. 103,
16068 (2006).
21. Y. Ben-Shahar, A. Robichon, M. B. Sokolowski,
G. E. Robinson, Science 296, 741 (2002).
22. Y. Ben-Shahar, N. L. Dudek, G. E. Robinson, J. Exp. Biol.
207, 3281 (2004).
23. G. V. Amdam, K. Norberg, M. K. Fondrk, R. E. Page,
Proc. Natl. Acad. Sci. U.S.A. 101, 11350 (2004).
24. A. R. Barchuk, R. Maleszka, Z. L. P. Simoes, Insect Mol.
Biol. 13, 459 (2004).
25. D. E. Wheeler, N. Buck, J. D. Evans, Insect Mol. Biol. 15,
597 (2006).
26. M. Corona et al., Proc. Natl. Acad. Sci. U.S.A. 104, 7128
(2007).
27. S. A. Ament, R. A. Verlarde, G. E. Robinson, Society for
Neuroscience Itinerary Planner and Abstract Viewer,
Neuropeptide Y signaling and nutritionally mediated
social behavior in the honey bee (2006).
28. G. J. Hunt et al., Naturwissenschaften 94, 247 (2007).
29. G. Bloch, D. P. Toma, G. E. Robinson, J. Biol. Rhythms 16,
444 (2001).
30. G. Bloch, D. E. Wheeler, G. E. Robinson, in Hormones,
Brain and Behavior, D. W. Pfaff, A. Arnold, A. Etgen,
S. Fahrbach, R. Rubin, Eds. (Elsevier Science, St. Louis,
MO, 2002), vol. 3, pp. 195235.
31. M. Nelson, K. Ihle, M. K. Fondrk, R. E. Page,
G. V. Amdam, PLoS Biol. 5, e62 (2007).
32. A. Patel et al., PLoS One 2, e509 (2007).
33. W. N. Ven ables, B. D. Ripley, Modern Applied Statistics
with S (Springer, New York, ed. 4, 2002).
34. A. L. Toth, S. Kantarovich, A. F. Meisel, G. E. Robinson,
J. Exp. Biol. 208, 4641 (2005).
35. M. J. West-Eberhard, Developmental Plasticity and
Evolution (Oxford Univ. Press, New York, 2003).
36. M. E. Hudson, Mol. Ecol. Notes, 10.1111/j.1471-
8286.2007.02019.x (2007).
37. A. Sturn, J. Quackenbush, Z. Trajanoski, Bioinformatics
18, 207 (2002).
38. We thank A. S. Escalante, A. Bowling, S. Kantarovich,
K. J. Bilof, and S. Buck for assistance in the field;
D. Schejbal and S. Buck for permission to collect wasps at
field sites owned by the University of Illinois;
A. S. Escalante and K. J. Bilof for physiological
measurements; M. T. Henshaw for microsatellite analyses;
R. A. Gibbs for strategic assistance; C. W. Whitfield for
assistance with gene identific ation; R. Rego for brain
dissections and RNA extractions; Y. Li, Y. Lu, and
S. Zhong for assistance with statistical analysis;
E. L. Hadley for assisting with figure preparation; and
M. B. Sokolowski, H. M. Robertson, C. M. Grozinger,
C. W. Whitfield, M. R. Berenbaum, S. A. Cameron,
J. L. Beverly, members of the Robinson laboratory, and
members of the University of Illinois Social Insect
Training Initiative for constructive comments on the
manuscript. Sup ported by the Illinois Sociogenomic
Initiative and NSF grant IOS-0641431 (G.E.R.). The
individual P. metricus sequences and flowgram data have
been uploaded to NCBI Trace Archive, TI range
1888756160 to 1889135944.
Supporting Online Material
www.sciencemag.org/cgi/content/full/1146647/DC1
Materials and Methods
Figs. S1 to S3
Tables S1 and S2
References
18 June 2007; accepted 19 September 2007
Published online 27 September 2007;
10.1126/science.1146647
Include this information when citing this paper.
JMJD6 Is a Histone
Arginine Demethylase
Bingsheng Chang, Yue Chen, Yingming Zhao, Richard K. Bruick*
Arginine methylation occurs on a number of proteins involved in a variety of cellular functions.
Histone tails are known to be mono- and dimethylated on multiple arginine residues where they
influence chromatin remodeling and gene expression. To date, no enzyme has been shown to
reverse these regulatory modifications. We demonstrate that the Jumonji domaincontaining 6
protein (JMJD6) is a JmjC-containing iron- and 2-oxoglutaratedependent dioxygenase that
demethylates histone H3 at arginine 2 (H3R2) and histone H4 at arginine 3 (H4R3) in both
biochemical and cell-based assays. These findings may help explain the many developmental
defects observed in the JMJD6
/
knockout mice.
I
ron- and 2-oxoglutaratedependent dioxy-
genases have been shown to oxidize a variety
of substrates including metabolites, nucleic
acids, and proteins (1). A candidate dioxygenase,
JMJD6, shares extensive sequence and predicted
structural homology with an asparaginyl hydrox-
ylase (2, 3) as well as the JmjC domains found in
several histone lysine demethylases (fig. S1A)
(48). Given the predicted conservation of struc-
tural elements and key residues (911), it is likely
that JMJD6 retains an analogous catalytic ac-
tivity. Here we report in vitro and in vivo data that
clearly indicate that JMJD6 functions as an ar-
ginine demethylase.
To test whether JMJD6 demethylates the
N-terminal tails of histone H3 or H4, we incu-
bated bulk histones with JMJD6 in the presence
of Fe(II), 2-oxoglutarate, and ascorbate (12). An-
tibodies specific for various methylated sites on
histones H3 and H4 were used to assess demeth-
ylation. Although no lysine demethylation was
observed, a substantial reduction in H3R2me2
and H4R3me2 was observed in the presence of
JMJD6 compared with buffer alone (Fig. 1A).
These effects were site-specific as no changes in
dimethylarginine were seen at positions H3R17
or H3R26. Previously, no enzyme had been
shown to reverse regulatory arginine methylation,
although deiminases can convert methylarginin e
to citrulline via demethylimination (13, 14). How-
ever , the requisite chemistry is analogous to that
demonstrated for demethylation of alkylated ni-
trogens by other dioxygenases (fig. S1C).
To investigate the preference for the substrate
methylation state, we used antibodies specific
for either mono- or dimethylated (symmetric)
H4R3 (Fig. 1B). The recombinant JMJD6 was
able to demethylate H4R3me2 when either het-
erogeneous bulk histones or synthetic peptides
encompassing the N-terminal 30 residues of his-
tone H4 were used as substrates (Fig. 1C). To a
lesser extent, JMJD6 could also demethylate
H4R3me1-containing substrates (Fig. 1C). Mu-
tation of the residues predicted to mediate Fe(II)
binding (mut JMJD6) prevented demethylation
(Fig. 1C).
Department of Biochemistry, University of Texas South-
western Medical Center, 5323 Harry Hines Boulevard,
Dallas, TX 753909038, USA.
*To whom correspondence should be addressed. E-mail:
richard.bruick@utsouthwestern.edu
19 OCTOBER 2007 VOL 318 SCIENCE www.sciencemag.org444
REPORTS
on March 22, 2009 www.sciencemag.orgDownloaded from
... Only two studies have addressed the issue of caste-related differences postulated by the OGPH (Table 1). In the metric paper wasp (Polistes metricus Say, 1831), which has behavioural castes, for expression levels were higher in foragers (foundress queens and workers) than non-foragers (queens and gynes) but were not associated with caste per se (Toth et al. 2007(Toth et al. , 2010. In the buff-tailed bumblebee (Bombus terrestris (Linnaeus, 1758)), which has morphological castes (Holland and Bloch 2020), for expression levels were similar among foragers (nest foundresses and workers) and nonforagers (queens and gynes), implying that neither foraging behaviour nor caste were associated with differences in for expression (Woodard et al. 2014). ...
... Foraging expression levels have been compared between queens and workers in just two other hymenopteran species, P. metricus (Toth et al. 2007(Toth et al. , 2010 and B. terrestris (Woodard et al. 2014). Since these studies pre-dated the discovery that the foraging gene is alternatively spliced, some caution is required when making comparisons. ...
... Since these studies pre-dated the discovery that the foraging gene is alternatively spliced, some caution is required when making comparisons. In P. metricus, for expression levels are higher in workers than non-foraging queens (Toth et al. 2007(Toth et al. , 2010. In contrast, B. terrestris workers and non-foraging queens had similar for expression levels (Woodard et al. 2014). ...
Article
Full-text available
Reproductive division of labour is based on biased expression of complementary parental behaviours, brood production (egg-laying) by queens and brood care (in particular, brood-provisioning) by workers. In many social insect species, queens provision brood when establishing colonies at the beginning of a breeding season and reproductive division of labour begins with the emergence of workers. In many social insect species, the expression of foraging (for) mRNA is associated with the intensity of foraging behaviour and therefore brood-provisioning. However, only two studies have compared queen and worker for expression levels and neither accounted for transcript splice variation. In this study, we compare the expression level of the for-α transcript variant across four life stages of the queen caste, two behavioural groups of workers, and males of a eusocial sweat bee Lasioglossum laevissimum (Smith, 1853). Foundresses collected prior to the onset of the foraging season and males had the highest for-α expression levels. All active (post-hibernatory) queens and workers had similar for-α expression levels independent of behaviour. These results suggest that the for gene in L. laevissimum acts as a primer before foraging activity and that caste-specific expression patterns correlate with the timing of foraging activity in queens and workers.
... In recent years, the proposal that nutrition is an important factor in regulating the division of labor among social insects has received increased support (Schulz et al., 1998;Blanchard et al., 2000;Toth and Robinson, 2005;Toth et al., 2007;Ament et al., 2008;Ament et al., 2010;Daugherty et al., 2011). Pollen consumption is low in young bees, largest in about 9-day-old nurses, and then declines to minimal amounts in foragers (Crailsheim et al., 1992), which rely primarily on honey intake. ...
Article
Full-text available
Nutritional stress, especially a dearth of pollen, has been linked to honey bee colony losses. Colony-level experiments are critical for understanding the mechanisms by which nutritional stress affects individual honey bee physiology and pushes honey bee colonies to collapse. In this study, we investigated the impact of pollen restriction on key markers of honey bee physiology, main elements of the immune system, and predominant honey bee viruses. To achieve this objective, we uncoupled the effects of behavior, age, and nutritional conditions using a new colony establishment technique designed to control size, demography, and genetic background. Our results showed that the expression of storage proteins, including vitellogenin (vg) and royal jelly major protein 1 (mrjp1), were significantly associated with nursing, pollen ingestion, and older age. On the other hand, genes involved in hormonal regulation including insulin-like peptides (ilp1 and ilp2) and methyl farnesoate epoxidase (mfe), exhibited higher expression levels in young foragers from colonies not experiencing pollen restriction. In contrast, pollen restriction induced higher levels of insulin-like peptides in old nurses. On the other hand, we found a strong effect of behavior on the expression of all immune genes, with higher expression levels in foragers. In contrast, the effects of nutrition and age were significant only the expression of the regulatory gene dorsal. We also found multiple interactions of the experimental variables on viral titers, including higher Deformed wing virus (DWV) titers associated with foraging and age-related decline. In addition, nutrition significantly affected DWV titers in young nurses, with higher titers induced by pollen ingestion. In contrast, higher levels of Black queen cell virus (BQCV) were associated with pollen restriction. Finally, correlation, PCA, and NMDS analyses proved that behavior had had the strongest effect on gene expression and viral titers, followed by age and nutrition. These analyses also support multiple interactions among genes and virus analyzed, including negative correlations between the expression of genes encoding storage proteins associated with pollen ingestion and nursing (vg and mrjp1) with the expression of immune genes and DWV titers. Our results provide new insights into the proximal mechanisms by which nutritional stress is associated with changes in honey bee physiology, immunity, and viral titers.
... Another investigation utilized trans-alignment method and 391,157 ESTs were produced from the brain transcriptome of the wasp P. metricus using NGS method, then P. metricus transcripts were annotated by the trans-alignment of sequencing reads to the genomic segments and expressed sequence tags generated from the honeybee, Apis mellifera. This investigation found that ESTs produced from P. metricus corresponds to 39% of Apis mellifera transcripts and noticed a potent relationship between the expression degree of mRNAs from the two species [28] . ...
Chapter
Next generation sequencing has been reforming biological and medical sciences over the years, offering exactness at single base level to the researchers’ comprehension of sequences of the DNA/RNA in highly effective manner. Next generation-based transcriptome sequencing is presently a typical technique for analyzing the expression of genes and revealing potential new RNAs. In contrast to microarray dependent methods, NGS technologies provide a more prominent dynamic range for detection and less background noise. The utilization of this innovation for transcriptome analysis has offered new perceptions in advancement as well as in the investigation of various disease forms and has currently relocated to clinical area including symptomatic examinations in an effort to recognize fusion transcripts associated with oncogenes and play a job in analysis and treatment of patients suffering from a cancer.
... There is also evidence that eusocial insects can decouple the fertility/longevity trade-off due to a different 'wiring', i.e. changes in the regulatory architecture of the IIS-JH-Vg/YP network (Corona et al. 2007;von Wyschetzki et al. 2015;Pamminger et al. 2016;Rodrigues & Flatt 2016), possibly by means of epigenetic factors (Yan et al. 2015;Kozeretska et al. 2017;Cardoso-Júnior et al. 2018. Moreover, in a variety of caregiving in insect systems such as burying beetles, European earwig, ants, and bees (Corona et al. 2007;Toth et al. 2007;Libbrecht et al. 2013;Roy-Zokan et al. 2015;Engel et al. 2016;Kohlmeier et al. 2018;Wu et al. 2020) evidence for an evolutionary link between parenting genes IIS, JH, Vg and genes shaping eusociality has been found (Rehan & Toth 2015). In the ponerine ant Harpegnathos saltator, adult workers can shift to reproduction as a gamergate, changing the regulation of stress resistance, which may be causally linked to increased lifespan (Schneider et al. 2011). ...
... Previous findings of other social insects have suggested that plasticity in an ancestral life cycle was recruited to the novel phenotype of a sterile caste (16,17,19,20,66), but evidence was based on comparisons of morphs within a single species or distant clades, and thus its detailed evolutionary process has been largely unknown. In Colophina aphids, by contrast, the putative ancestral phenotype, Gall defender, persists within their life cycle, thus correlations in the two evolutionary trajectories between the ancestral and derived phenotypes can be investigated quantitatively using multiple species in the same genus (Fig. 3C, 4C and D). ...
Preprint
Full-text available
The origin of a sterile caste among eusocial animals has been a fundamental but still unresolved problem in understanding the evolution of biological complexity. At the origin of a sterile caste, recruitment of pre-existing plasticity may lead to produce physiologically, morphologically and behaviorally distinct caste phenotypes. Here, we provide convincing evidence that preexisting seasonal polyphenism has been recruited to generate a sterile soldier caste in host-alternating social aphids. We demonstrate that sterile soldier nymphs of Colophina aphids resemble those of monomorphic defensive nymphs produced in a different host-plant generation. Notably, the two morphs in the basal species show the closest similarity in morphology and gene expression among all morph pairs. Moreover, their evolutionary phenotypic changes along the phylogeny of four Colophina species are significantly correlated positively. These results suggest that they may share the common regulatory mechanisms of development, which underpin the heterochronic expression of monomorphic defenders on the different host plant leading to the evolution of a novel soldier phenotype. We further demonstrate that the monomorphic defenders can increase their inclusive fitness by killing predator's eggs on a seasonally different host plant. Taken together, our findings suggest that preexisting plasticity that can gain indirect fitness benefits facilitates the early evolution of a sterile caste.
... We further hypothesised that in contrast to workers, where queen mandibular pheromone and brood pheromones are the cue for workers to stop oogenesis, in queens it is nutritional cues that regulate oogenesis. More broadly, we note that in solitary Hymenopterans, nutritional cues are intimately related to oogenesis (Toth et al., 2007). Thus, nutritional cues may also have a role in regulating oogenesis in the queens of advanced eusocial insects Amdam, 2005, Dolezal et al., 2013). ...
Article
In the honey bee (Apis mellifera), queen and worker castes originate from identical genetic templates but develop into different phenotypes. Queens lay up to 2,000 eggs daily whereas workers are sterile in the queen’s presence. Periodically queens stop laying: during swarming, when resources are scarce in winter, and when they are confined to a cage by beekeepers. We used confocal microscopy and gene expression assays to investigate the control of oogenesis in the ovaries of honey bee queens that were caged inside and outside the colony. We find evidence that queens use a different combination of ‘checkpoints’ to regulate oogenesis compared to honey bee workers and other insect species. However, both queen and worker castes likely use the same programmed cell death pathways to terminate oocyte development at their caste-specific checkpoints. Our results also suggest that a key factor driving the termination of oogenesis in queens is nutritional stress. Thus, queens may regulate oogenesis via the same regulatory pathways that were utilised by ancestral solitary species but likely have adjusted physiological checkpoints to suit their highly-derived life history.
... Losers in our experiment would potentially become subordinate foundresses in a natural nesting context and not workers, though subordinates do more foraging than dominants [31]. Despite reduced reproduction and greater foraging relative to dominant foundresses, subordinate foundresses are not the same as workers and have been shown to have distinct neurogenomic profiles compared to dominant foundresses and workers in microarray and candidate-gene studies [38,86,89,90]. Nevertheless, the expression patterns of these genes suggest that within a few hours of emerging from a social encounter as a subordinate, multiple genes are dynamically regulated in a manner suggesting changes to aggression, reproduction, and metabolism (Fig 6). ...
Article
Full-text available
Social interactions have large effects on individual physiology and fitness. In the immediate sense, social stimuli are often highly salient and engaging. Over longer time scales, competitive interactions often lead to distinct social ranks and differences in physiology and behavior. Understanding how initial responses lead to longer-term effects of social interactions requires examining the changes in responses over time. Here we examined the effects of social interactions on transcriptomic signatures at two times, at the end of a 45-minute interaction and 4 hours later, in female Polistes fuscatus paper wasp foundresses. Female P . fuscatus have variable facial patterns that are used for visual individual recognition, so we separately examined the transcriptional dynamics in the optic lobe and the non-visual brain. Results demonstrate much stronger transcriptional responses to social interactions in the non-visual brain compared to the optic lobe. Differentially regulated genes in response to social interactions are enriched for memory-related transcripts. Comparisons between winners and losers of the encounters revealed similar overall transcriptional profiles at the end of an interaction, which significantly diverged over the course of 4 hours, with losers showing changes in expression levels of genes associated with aggression and reproduction in paper wasps. On nests, subordinate foundresses are less aggressive, do more foraging and lay fewer eggs compared to dominant foundresses and we find losers shift expression of many genes in the non-visual brain, including vitellogenin, related to aggression, worker behavior, and reproduction within hours of losing an encounter. These results highlight the early neurogenomic changes that likely contribute to behavioral and physiological effects of social status changes in a social insect.
Article
Full-text available
Biological big data are a massive amount of data generated from multi-omics experiments, such as genomics, transcriptomics, proteomics, metabolomics, phenomics, glycomics, epigenomics, and other omics. These data are used to study biological processes and to gain insights into how living systems work. It can also be used to develop new treatments for diseases and understand the causes of certain conditions. The storage and analysis of these data present several challenges owing to their sheer size and complexity. Storing these data efficiently requires a large amount of storage space and processing power. Furthermore, there are certain limitations in terms of the kind of insights that can be gained from multi-omics data because of their complexity. Despite these challenges, biological big data offers great potential for advancing our understanding of biology and developing new treatments for diseases. Big-data research is a rapidly growing field, with numerous applications. As the amount of data continues to increase, it is important to understand its storage, utility, limitations, and challenges. In this review article, various sources of big-data research and their storage capacities, limitations, and challenges are discussed. Factors affecting the data quality and accuracy have been reported. It will be helpful for researchers to understand the available big data in biology for their further utilization and integration into novel discovery.
Article
Full-text available
At present, there is a growing interest among researchers in studying the structure and function of the bee brain in relation to their cognitive behavior. The bee brain, despite its small size of approximately 1 million neurons, is known for its ability to facilitate effective communication and collaboration. Just like humans, the bee brain is also controlled by biogenic amines like dopamine, serotonin and tyramine, octopamine, and histamine. The honey bees communicate with each other by using a complex language called the “waggle dance”. Despite existing knowledge about the bee brain's neuroanatomy, there is still a need to understand which specific regions control cognition and social behavior in bees. This review aims to explore the different major parts of the bee brain and how each part contributes to modulating social behavior.
Article
Full-text available
Across evolutionary lineages, insects vary in social complexity, from those that exhibit extended parental care to those with elaborate divisions of labor. Here, we synthesize the sociogenomic resources from hundreds of species to describe common gene regulatory mechanisms in insects that regulate social organization across phylogeny and levels of social complexity. Different social phenotypes expressed by insects can be linked to the organization of co-expressing gene networks and features of the epigenetic landscape. Insect sociality also stems from processes like the emergence of parental care and the decoupling of ancestral genetic programs. One underexplored avenue is how variation in a group's social environment affects the gene expression of individuals. Additionally, an experimental reduction of gene expression would demonstrate how the activity of specific genes contributes to insect social phenotypes. While tissue specificity provides greater localization of the gene expression underlying social complexity, emerging transcriptomic analysis of insect brains at the cellular level provides even greater resolution to understand the molecular basis of social insect evolution.
Article
Full-text available
Here we report the genome sequence of the honeybee Apis mellifera, a key model for social behaviour and essential to global ecology through pollination. Compared with other sequenced insect genomes, the A. mellifera genome has high A+T and CpG contents, lacks major transposon families, evolves more slowly, and is more similar to vertebrates for circadian rhythm, RNA interference and DNA methylation genes, among others. Furthermore, A. mellifera has fewer genes for innate immunity, detoxification enzymes, cuticle-forming proteins and gustatory receptors, more genes for odorant receptors, and novel genes for nectar and pollen utilization, consistent with its ecology and social organization. Compared to Drosophila, genes in early developmental pathways differ in Apis, whereas similarities exist for functions that differ markedly, such as sex determination, brain function and behaviour. Population genetics suggests a novel African origin for the species A. mellifera and insights into whether Africanized bees spread throughout the New World via hybridization or displacement.
Article
The diversity of social behavior among birds and primates is surpassed only by members of the Hymenopteran insects, including bees, ants, and the genus Polistes, or paper-wasps. This volume combines incisive reviews and new, unpublished data in studies of paper-wasps, a large and varied group whose life patterns are often studied by biologists interested in social evolution. While this research is significant to the natural history of paper-wasps, it also applies to topics of general interest such as the evolution of cooperation, social parasitism, kin recognition, and the division of labor.
Chapter
A phenotype-centered view of evolution needs to start with a solid idea about the nature of the phenotype. This chapter and the next are devoted to two universal properties of phenotypes, plasticity, or responsiveness to environmental inputs; and modularity, or subdivision into semi-independent and dissociable parts (chapter 4). Of these two properties, plasticity is probably the more fundamental, for the ability to replicate, which distinguishes organic from inorganic nature, requires molecules which are interactive and precisely responsive— adaptively plastic. So plasticity must have been an early universal property of living things. The universality of modularity is a secondary, or “emergent” result of the universality of plasticity (see Wilczek, 2002, on emergent universality in physics). Any organism whose size, whether due to accretion or growth, is large enough to create internal environmental differences, such as those between the inner and the outer regions of a clump of material, has the potential for regional internal differentiation. As differentiation evolves to produce specialized parts and an internal division of labor, internal heterogeneity gives rise to conditional switches between developmental pathways. The result is a stucture characterized by somewhat discrete parts—modularity. Thus, given plasticity as a universal property of living matter, modularity follows. The present chapter describes some of the remarkable mechanisms of phenotypic plasticity. One reason to focus on mechanisms is to indicate the material basis for the evolution of plasticity, which is a product of concrete devices that are subject to genetic variation and selection. A cursory look at these mechanisms, however incomplete, by itself suggests the importance of plasticity in development and evolution, for the mechanisms of plasticity include some of the most ingenious and widely conserved creations of nature. Mechanisms of plasticity are further discussed in chapter 23, which describes how organisms assess environmental conditions when they adaptively switch between alternative developmental pathways. Phenotypic plasticity has already been defined as the ability of an organism to react to an environmental input with a change in form, state, movement, or rate of activity.
Article
The multiple independent origins of eusociality in the insect order Hymenoptera are clustered in only four of more than 80 families, and those four families are two pairs of closely related taxa in a single part of the order. Therefore, although ordinal-level characteristics can contribute to hymenopteran eusocial evolution, more important roles have been played by traits of infraordinal taxa that contain the eusocial forms. Many factors have been proposed and discussed, but assessments of traits' salience to eusocial evolution have heretofore not been joined to phylogenetics. In the present analysis, cladograms of superfamilies and families of Hymenoptera and of the family Vespidae are used to ordinate the appearance of traits that play roles in vespid eusociality. Proximity of traits' first appearance to the origin of eusocial Vespidae is taken as one measure of traits' salience to vespid eusocial evolution. Traits that subtend only eusocial taxa and that are uniquely associated with eusociality have foundations in more general traits that subtend more inclusive taxa. No single trait is uniquely causative of vespid eusocial evolution. High-salience traits that closely subtend vespid eusociality include nesting, oviposition into an empty nest cell, progressive provisioning of larvae, adult nourishment during larval provision malaxation, and inequitable food distribution among nestmates. The threshold characteristic of Polistes-grade eusociality is life-long alloparental brood care by first female offspring who remain, uninseminated, at their natal nest. Traits directly associated with occurrence of such workers are larva-adult trophallaxis, which can foster relatively low larval nourishment early in a colony cycle, and protogyny and direct larval development, which combine to yield restricted mating opportunities for female offspring that are the first to emerge in the colony cycle. Trait mapping suggests no role for asymmetry of relatedness due to haplodiploidy, but it suggests high salience for haplodiploidy as a mechanism enabling the production of all-female clutches of first offspring.
Book
A guide to using S environments to perform statistical analyses providing both an introduction to the use of S and a course in modern statistical methods. The emphasis is on presenting practical problems and full analyses of real data sets.
Book
Social behavior occurs in some of the smallest animals as well as some the largest, and the transition from solitary life to sociality is an unsolved evolutionary mystery. The Evolution of Social Wasps examines social behavior in a single lineage of insects, wasps of the family Vespidae. It presents empirical knowledge of social wasps from two approaches: one that focuses on phylogeny and life history; and one that focuses on individual ontogeny, colony development, and population dynamics. It also provides an extensive summary of the existing literature while demonstrating how it can be clouded by theory. This approach to the conflicting literature on sociality highlights how often repeated models can become fixed in the thinking of the scientific community. Instead, it presents a mechanistic scenario for the evolution of sociality in wasps that changes our perspective on kin selection, the paradigm that has dominated thinking about social evolution since the 1970s.
Article
Techniques involving whole-genome sequencing and whole-population sequencing (metagenomics) are beginning to revolutionize the study of ecology and evolution. This revolution is furthest advanced in the Bacteria and Archaea, and more sequence data are required for genomic ecology to be fully applied to the majority of eukaryotes. Recently developed next-generation sequencing technologies provide practical, massively parallel sequencing at lower cost and without the requirement for large, automated facilities, making genome and transcriptome sequencing and resequencing possible for more projects and more species. These sequencing methods include the 454 implementation of pyrosequencing, Solexa/Illumina reversible terminator technologies, polony sequencing and AB SOLiD. All of these methods use nanotechnology to generate hundreds of thousands of small sequence reads at one time. These technologies have the potential to bring the genomics revolution to whole populations, and to organisms such as endangered species or species of ecological and evolutionary interest. A future is now foreseeable where ecologists may resequence entire genomes from wild populations and perform population genetic studies at a genome, rather than gene, level. The new technologies for high throughput sequencing, their limitations and their applicability to evolutionary and environmental studies, are discussed in this review.