Open Access
How to translate text using browser tools
25 July 2009 Termites (Isoptera): Their Phylogeny, Classification, and Rise to Ecological Dominance
Michael S. Engel, David A. Grimaldi, Kumar Krishna
Author Affiliations +
Abstract

Like ants, termites are entirely eusocial and have profound ecological significance in the tropics. Following upon recent studies reporting more than a quarter of all known fossil termites, we present the first phylogeny of termite lineages using exemplar Cretaceous, Tertiary, and Recent taxa. Relationships among Recent families were largely unaffected by the addition of extinct taxa, but the analysis revealed extensive grades of stem-group taxa and the divergence of some modern families in the Cretaceous. Rhinotermitidae, Serritermitidae, and the “higher” termites (family Termitidae), which comprise 84% of the world termite species, diverged and radiated entirely in the Tertiary, corresponding to a significant increase in termite individuals in the fossil record. Radiation of the higher termites may have affected the formation of terrestrial carbon reserves like oil and coal. The higher classification of Isoptera is slightly revised based on the phylogenetic results. The following new taxa are proposed: Cratomastotermitidae, new family; Euisoptera, new clade; Archotermopsidae, new family; and Neoisoptera, new clade. In addition, the families Stolotermitidae, Stylotermitidae, and Archeorhinotermitidae are newly recognized or resurrected, and the families Termopsidae and Hodotermitidae are significantly restricted in composition.

INTRODUCTION

“Success” of a species or group of species is typically either ecological or evolutionary. Insects, in particular, are evolutionarily very successful because of their early origins in the Devonian and subsequent radiation into millions of species. It is only certain insects, however, that account for the remarkable overall ecological dominance of insects in terrestrial ecosystems, in terms of biomass and impact on biological communities, and chief among them are the social insects. This report is a phylogenetic synthesis of many recent discoveries of fossil termites, which is then used to consider the origins of termite ecological success.

Advanced sociality, or eusociality, involves overlapping generations of siblings that generally share a durable domicile in which groups of individuals specialize in tasks (i.e., castes), especially reproduction, foraging, and commonly defense and brood care (Wilson, 1971). A colony's efficiency in mobilizing its foragers and soldiers fosters the superior competitive ability of major eusocial groups like ants, vespid wasps, honeybees, and termites. For example, there are approximately 19,340 modern species of bees (Apoidea: Anthophila), but it is the large, complex colonies of honey bees (genus Apis) that easily outcompete the native social species, even leading to the latter's localized extirpation (e.g., Roubik et al., 1986; Sugden et al., 1996). Unlike bees, all of the 12,516 living species of ants and 2,958 living species of termites are eusocial (termite species numbers valid as of 11 March 2009: Krishna et al., in press). Basal lineages of both these groups generally live in small colonies of several dozen to several hundred individuals with less caste differentiation, and the most recently derived lineages (like army ants, leaf cutter ants, and mound-building termites) form massive colonies of over one million individuals with extreme caste differentiation. Termites in tropical and subtropical ecosystems are the major consumers of the most abundant biomolecule on land, cellulose, and its more inert form, lignocellulose. It is estimated, for example, that termites ingest 50%–100% of the dead plant biomass in tropical ecosystems (Bignell and Eggleton, 2000). Their abundance, like that of ants, frequently exceeds 1,000 individuals/m2 or 2,000 mg/m2, and it is estimated that gas excretion from termites and their nests contributes 2%–5% of the world's atmospheric methane (Sugimoto et al., 2000). The ecological impact of termites, even apart from the commercial damage they cause, is prodigious.

There has been intensive study of the relationships and fossil record of bees (Engel, 2001, 2004; Danforth et al., 2006; Michener, 2007; Ohl and Engel, 2007) and ants (Grimaldi et al., 1997; Grimaldi and Agosti, 2001; Dlussky and Rasnitsyn, 2002; Engel and Grimaldi, 2005; Moreau et al., 2006; Brady et al., 2006). Both of these groups appear to have originated in the late Early Cretaceous, ca. 100–120 Ma, with some modern subfamilies diverging in the Late Cretaceous. Though their evolutionary history has been studied less, it is now acknowledged that termites are highly modified, eusocial roaches (Cleveland et al., 1934; McKittrick, 1964; Lo et al., 2000; Deitz et al., 2003; Grimaldi and Engel, 2005; Klass and Meier, 2006), whose earliest fossils predate those of ants and bees by approximately 35 million years (Thorne et al., 2000; Engel et al., 2007a). Recent phylogenetic work on termites involves molecular and some morphologically based analyses of living species only (e.g., Kambhampati et al., 1996; Donovan et al., 2000; Thompson et al., 2000; Bitsch and Noirot, 2002; Klass and Meier, 2006; Inward et al., 2007a, 2007b; Legendre et al., 2008). Unique for any insect order, all early fossil termites have been classified into living families, particularly the Hodotermitidae, despite the fact that such fossils may be stem groups since they lack many derived features of living families.

Here we present the first analysis of relationships among fossil and living termite lineages, along with estimates of divergence times and ecological patterns of the major lineages. Recent studies have reported 18 new termites from the Cretaceous (Krishna and Grimaldi, 2000, 2003; Engel et al., 2007a; Grimaldi et al., 2008) and 38 species from the Tertiary (Nel and Bourguet, 2006; Wappler and Engel, 2006; Engel and Krishna, 2007a, 2007b; Engel et al., 2007b; Engel, 2008; Krishna and Grimaldi, 2009), comprising about one-quarter of all described fossil species. Nearly 80% of these species are preserved in amber, from the Early Cretaceous of Lebanon, the mid-Cretaceous of Myanmar and New Jersey, the Eocene of the Baltic Region and France, and the Miocene of Mexico and the Dominican Republic (deposits are reviewed in Rasnitsyn and Quicke, 2002; Grimaldi and Engel, 2005). The microscopic fidelity of preservation in amber allows uniquely detailed comparisons with living species, greatly facilitating phylogenetic analysis of extinct taxa. Insects in 110 Ma limestone from the Crato Formation of Brazil also have exceptional preservation, as mineralized replicas with cuticular microstructure and even some internal organs. As a result, Crato species preserved as series of specimens can be largely reconstructed (Grimaldi et al., 2008). Thus, now is an opportune time for deciphering nearly 140 million years of termite evolution.

Figure 1

Strict consensus cladogram with unambiguous character-state changes mapped. Character numbers appear above the branch, with the corresponding state beneath the branch. Branch “A” continued in figs. 2, 4, and 5. Chaeteessa (Mantodea), Panchlora (Blattaria), Periplaneta (Blattaria), and Cryptocercus (Blattaria) are the outgroup taxa.

i0003-0082-3650-1-1-f01.gif

MATERIALS AND METHODS

Termite specimens belonging to 38 exemplar living species of 36 genera, representing all seven traditionally recognized families and four subfamilies of the “higher” termites (family Termitidae), were dissected and directly examined; 38 extinct species in 25 genera were studied and compared to living taxa. The fossils ranged in age from Early Cretaceous (Berriasian, ca. 135 Ma, of Baissa, Siberia) to the Miocene (in amber from the Dominican Republic, ca. 17 Ma) (table 2). Species preserved as compressions or mineralized replicas were also included where bodies were preserved, not just isolated wings (we have omitted from the analysis taxa known only or largely from wings, such as Ulmeriella Meunier).

There were 108 morphological and biological attributes scored from imago, soldier, and worker castes where available (table 1 and appendix); outgroup taxa were other Dictyoptera, specifically the most basal mantis (Chaeteessa sp. Burmeister), the roaches Periplaneta Burmeister and Panchlora Burmeister, and the relict wood roaches (Cryptocercus spp. Scudder), the latter being the living sister group to termites. Within the matrix of 8,748 cells, 21% of the cells were coded as unknown (feature not observed in the fossils), and a further 9% as inapplicable (e.g., soldier characters in genera that lack this caste). Phylogenetic analysis employed NONA (Goloboff, 1997), using 500 replicates of the data set with random taxon addition and branch swapping on all shortest topologies. This yielded 2,088 equally most-parsimonious trees of 302 steps (C.I. 44, R.I. 86), the strict consensus of which resulted in 323 steps (unambiguous character changes mapped in figs. 1, 2, 4, 5). The relative abundance of termites through time was plotted as proportions (%) of the number of termite specimens to all insect specimens per fossil deposit (table 2); only deposits that yielded at least one termite specimen were plotted, in order to ensure the appropriate taphonomic conditions for termite preservation.

Figure 2

Continuation of figure 1 focusing on Hodotermitidae s.s., “Termopsidae” s.s., Archotermopsidae, n. fam., Stolotermitidae, stat. n., and various stem-group lineages. Branch “B” (Kalotermitidae + Neoisoptera) is depicted in fig. 4.

i0003-0082-3650-1-1-f02.gif

Figure 3

Strict consensus cladogram of all termite species studied superimposed on geological time scale, with graph of termite abundance.

i0003-0082-3650-1-1-f03.jpg

Figure 4

Continuation of figs. 12 focusing on Kalotermitidae. Branch “C” (Neoisoptera) in fig. 5.

i0003-0082-3650-1-1-f04.gif

Figure 5

Continuation of figs. 1, 2, and 4 focusing on Neoisoptera.

i0003-0082-3650-1-1-f05.gif

Table 1

Character and Character-State Descriptions The following list enumerates those characters and character states used in the analysis of relationships within Isoptera (refer to appendix 1 for codings). All characters were considered nonadditive and of equal weights

i0003-0082-3650-1-1-t01.gif

Table 2

Termite Abundance (% of all insect specimens) in Major Insect Deposits throughout the Cretaceous and Cenozoic

i0003-0082-3650-1-1-t02.gif

RESULTS AND DISCUSSION

Phylogeny

Analysis of the paleontological data resulted in a remarkably resolved topology for Isoptera (fig. 3). Structure of the consensus tree generally agrees with that from prior morphological and molecular studies (Kambhampati et al., 1996; Donovan et al., 2000; Thompson et al., 2000; Bitsch and Noirot, 2002; Inward et al., 2007a, 2007b), but depart in some respects from the recent molecular study of Legendre et al. (2008) (e.g., the relatively basal placement of Kalotermitidae). Legendre et al. (2008) had a dramatically reduced taxon sampling relative to other, more comprehensive treatments (e.g., Inward et al., 2007a).

In our analysis unequivocal relationships include Cryptocercus as the living sister group to the termites and Mastotermes Froggatt as the living sister group to all other termites—the lineage we are designating here as Euisoptera. Mastotermes darwiniensis Froggatt, from northern Australia and southern New Guinea, is the sole survivor of the formerly global Mastotermitidae (Thorne et al., 2000; Wappler and Engel, 2006), and it retains striking plesiomorphic features with roaches, such as laying its eggs in a vestigial pod or ootheca. Interestingly, the Mastotermitidae as it has historically been classified was recovered as monophyletic, despite opinion to the contrary (e.g., Jarzembowski, 1981). Cratomastotermes Bechly, from the Early Cretaceous Crato Formation of Brazil, was formerly placed in the Mastotermitidae (Bechly, 2007); in our analysis it is the most basal species of termite since it retains even more plesiomorphic features than Mastotermitidae (Grimaldi et al., 2008). This genus is accordingly placed in a new, extinct family, Cratomastotermitidae, new family (see Classification). Cretaceous fossils traditionally classified as Hodotermitidae, such as the extinct “genus” Meiatermes Lacasa-Ruiz and Martínez-Delclòs, actually comprise a grade of stem-group species that fall basal to the divergence of Termopsidae, true Hodotermitidae, and all other termites. Termopsidae stat. n. (see Classification) should be restricted to the Baltic amber species of Termposis Heer (Engel et al., 2007b). What we are designating as the true Hodotermitidae (see Classification) is a group of three genera of the “harvesters”—Anacanthotermes Jacobson, Hodotermes Hagen, Microhodotermes Sjöstedt—that feed on grasses in savanna and steppe biomes of Africa and Eurasia; monophyly of this group is confirmed by other analyses (Inward et al., 2007a, 2007b; Legendre et al., 2008). Three genera of highly disjunct wood feeders that are traditionally placed in the Termopsidae s.l.—Archotermopsis Desneux, Zootermopsis Emerson, and Hodotermopsis Holmgren—were not found to be monophyletic, contrary to other studies. Inclusion or exclusion of the fossils made no difference to the definition of “Termopsidae s.l.” except that Archotermopsis was sometimes found to be monophyletic. As such, the traditional concept of Termopsidae s.l. could not be supported and these genera are accordingly classified in the new family Archotermopsidae, new family (see Classification).

The austral disjuncts, Stolotermes Hagen and Porotermes Hagen, are sometimes placed within Termopsidae s.l. (e.g., Donovan et al., 2000; Thorne et al., 2000; Legendre et al., 2008) but should clearly be the separate family Stolotermitidae, stat. n. ( =  Stolotermitinae + Porotermitinae: see Classification), since they are the living sister group to Kalotermitidae plus higher termites in this study and in some molecular analyses (Inward et al., 2007a, 2007b). The divergence between Stolotermitidae and Kalotermitidae + higher termites was certainly in the Early Cretaceous, since there are three intermediate genera from the Early and mid-Cretaceous. The family Kalotermitidae, or dry-wood termites, is a cosmopolitan group of 457 living species, which has two known Cretaceous and three Tertiary stem-group taxa (only two of the latter were included in our study), with living species comprising a monophyletic group of probably Tertiary origin. Interestingly, a kalotermitid-like nest exists from the Late Cretaceous Javelina Formation of Texas (Rohr et al., 1986). Kalotermitidae is the sister group to an unequivocally monophyletic group we are calling the Neoisoptera, which is defined in part by the distinctive opening of the frontal gland called the fontanelle. The Neoisoptera is comprised of the Rhinotermitidae (13 living genera, 380 living species, all of which are wood feeders), the monotypic family Serritermitidae from Brazil, and the largest family, Termitidae. It also includes Archeorhinotermes rossi Krishna and Grimaldi in 100 Ma Burmese amber, the most derived termite from the Cretaceous, formerly placed in the Rhinotermitidae (Krishna and Grimaldi, 2003), but actually an extinct stem group to the rest of the Neoisoptera and here considered as the sole member of Archeorhinotermitidae, stat. n. (see Classification). Parastylotermes Snyder and Emerson and Stylotermes Holmgren and Holmgren (neither genus studied in prior analyses [Donovan et al., 2000; Inward et al., 2007a, 2007b; Legendre et al., 2008]) should be separated from Rhinotermitidae s. str., the latter doubtfully monophyletic (Donovan et al., 2000; Inward et al., 2007a, 2007b; Legendre et al., 2008) but clearly closely related to the Termitidae. These genera have been at times considered a separate family, as Stylotermitidae, stat. rev. (see Classification), and this status should be reinstated. The earliest rhinotermitids are Reticulitermes antiquus (Germar) and Heterotermes eocenicus Engel in Baltic amber (Engel et al., 2007b; Engel, 2008); the divergence of both families probably occurred in the Early Tertiary.

Classification

Several taxonomic changes are required in order to have the classification of termite families reflect our cladistic results. The revised, higher-level classification of Isoptera is outlined in table 3 (modified from Engel and Krishna, 2004a, 2004b, 2007c). The classification is that which is employed for (and will be further elaborated in) the forthcoming world catalog of Isoptera (Krishna et al., in press).

Table 3

Synonymical Hierarchical Classification of Isoptera (modified and updated from Engel and Krishna, 2004a, 2004b, 2007c)

i0003-0082-3650-1-1-t03.gif

Cratomastotermitidae, new family

Type genus

Cratomastotermes Bechly.

Diagnosis

Diagnosed by primitive retention of distinct cross veins, archedictyon between veins, arched humeral margin of forewing scale, large rectangular pronotum, pentamerous tarsi, and absence of ocelloids and fontanelle (refer also to Grimaldi et al., 2008).

Included genera

The family presently comprises a single genus, Cratomastotermes.

Family Termopsidae Holmgren, sensu novum

Termopsinae Holmgren, 1911: 35. Type genus: Termopsis Heer, 1849.

Comments

Since the time of Hagen (1858), the fossil Termopsis bremii Heer (1849) has been intricately linked to a group of otherwise plesiomorphic modern species. Over the intervening 150 years a few additional genera and several living species have been added to the group that became known as the Termopsinae and eventually Termopsidae (e.g., Emerson, 1933). Unfortunately, no singular, specialized (i.e., apomorphic) feature has truly united these taxa and the monophyly of the group has been suspect. Our analysis reveals that those fossils of the Tertiary genus Termopsis are, in fact, not related to the modern members otherwise classified in the family (namely Archotermopsis, Zootermopsis, and Hodotermopsis, below classified into a new family). In addition, those other fossil and living genera also traditionally classified in the family (e.g., Stolotermes and Porotermes of the Stolotermitinae and Porotermitinae, respectively) were similarly found to be unrelated to Termopsis, as well as unrelated to the aforementioned genera of Termopsinae. Accordingly, Termopsidae is here significantly restricted and considered to comprise only the genus Termopsis.

Archotermopsidae, new family

Type genus

Archotermopsis Desneux.

Diagnosis

The new family can be characterized by the following combination of attributes: absence of ocelloids and fontanelle, antennae with 22–27 articles, pronotum distinctly narrower than head, tarsi pentamerous (sometimes cryptically), fourth sternite with sole sternal gland, forewing scale overlapping hind-wing scale, humeral margin of scale flat, imago-worker mandibles with three marginal teeth (left side) and subsidiary tooth between apical and first marginal teeth (right side).

Included genera

Archotermopsis, Zootermopsis, Hodotermopsis, and tentatively †Parotermes Scudder. The Late Miocene genus †Gyatermes Engel and Gross (2009) may belong herein but must await more completely preserved material.

Family Hodotermitidae Desneux, sensu novum

Hodotermitini Desneux, 1904: 284. Type genus: Hodotermes Hagen, 1853.

Comments

Numerous genera of plesiomorphic fossil termites have been historically assigned to Hodotermitidae, leaving the impression that this group was once diverse in the past but has experienced significant extinction and that the modern taxa are relicts of this former diversity. In fact, our study demonstrates that none of the fossils assigned to Hodotermitidae can be considered actual hodotermitids. Instead, this assemblage represents a grade between several families and lineages of Isoptera. Hodotermitidae was recovered as a monophyletic group but strictly for the modern genera. We have accordingly restricted the sense of Hodotermitidae to those genera (listed below) and consider the remaining groups such as Cretotermitinae, Carinatermitinae, Lutetiatermitinae, and Caatingatermitinae (the latter two ill defined on teratologies and misinterpreted characters, respectively), among numerous other genera (e.g., Meiatermes; Melqartitermes Engel, Grimaldi, and Krishna; Mylacrotermes Engel, Grimaldi, and Krishna; Mariconitermes Fontes and Vulcano; Cratokalotermes Bechly) as incertae sedis among basal Euisoptera (fig. 3).

Included genera

Hodotermes, Anacanthotermes, and Microhodotermes.

Family Stolotermitidae Holmgren, status novus

Stolotermitinae Holmgren, 1910: 285. Type genus: Stolotermes Hagen, 1858.

Comments

The Stolotermitidae is here recognized to encompass the former subfamilies Stolotermitinae and Porotermitinae of Termopsidae s.l. For the moment the two subfamilies are retained despite each being monogeneric.

Included genera

Stolotermes (in Stolotermitinae) and Porotermes (in Porotermitinae).

Family Archeorhinotermitidae Krishna and Grimaldi, status novus

Archeorhinotermitinae Krishna and Grimaldi, 2003: 2. Type genus: Archeorhinotermes Krishna and Grimaldi, 2003.

Comments

Although previously classified as a primitive lineage in the Rhinotermitidae, affinities with this group are entirely plesiomorphic. As our analysis reveals, Archeorhinotermes is actually a stem group, basal to all Euisoptera. Accordingly we have removed the genus from Rhinotermitidae and elevated Archeorhinotermitinae to familial rank.

Included genera

The family includes only Archeorhinotermes at present.

Family Stylotermitidae Holmgren and Holmgren, status revivisco

Stylotermitinae Holmgren and Holmgren, 1917: 141. Type genus: Stylotermes Holmgren and Holmgren, 1917.

Comments

Stylotermes and the Tertiary genus Parastylotermes have historically been classified in the Rhinotermitidae. These genera are particularly distinctive in their possession of trimerous tarsi, a feature otherwise known only in Indotermes of the Termitidae. The significance of this tarsal reduction was used by some authors in the past to accord Stylotermitinae familial rank, in the same fashion that Indotermes was placed in a monogeneric family of its own (e.g., Roonwal, 1958). Herein we resurrect the familial status of the former. Although the trimerous condition of the tarsi is truly a distinctive synapomorphy for the group, the classificatory alteration is based on the fact that Stylotermitinae comprises a grade, along with Archeorhinotermitinae leading to Rhinotermitidae + Serritermitidae + Termitidae (fig. 3). As such, its inclusion within Rhinotermitidae renders the assemblage demonstrably paraphyletic.

Included genera

Stylotermes and Parastylotermes.

Ecology and Evolution

The circumtropical family Termitidae, or “higher termites,” comprises approximately 70% of all termite species and appears to be one of the most recent radiations of all insect groups that are ecologically significant. Monophyly of the Termitidae is well established; the family includes such familiar groups as the Macrotermitinae and Nasutitermitinae, some of which build huge mounds in grassland and scrub biomes; other nasute taxa build large arboreal nests of cartonlike, fecal material in tropical forests. The diets of Termitidae are extremely diverse, primitively being sound and rotting wood but also including humus, leaf litter, soil, grass, herbivore dung, and even the mycelia of a symbiotic fungus, Termitomyces R. Heim, that they cultivate in the nest like attine ants (e.g., Sands, 1969). The huge colonies and diverse diets of the Termitidae account for the overwhelming biomass of termites in tropical and subtropical environments. The earliest apparent termitid is an incomplete compression of an imago from the Oligocene of Brazil, ca. 30 Ma (Martins-Neto and Pesenti, 2006;, a putative termitid from the Bembridge Marls [Jarzembowski, 1980] is probably a rhinotermitid). The first diverse paleofaunas of Termitidae—more than 30 species—are in Miocene amber from the Dominican Republic (Krishna and Grimaldi, 2009) and Mexico, which are very similar to modern Neotropical faunas. This dramatic appearance is probably due to the fact that these ambers are the only major fossil insect deposits from the Neogene that were formed in the tropics. Were Termitidae abundant during the Eocene they should have been preserved in Baltic amber (Lutetian: ca. 45 Ma), since this deposit has yielded thus far most other living termite families (Engel et al., 2007b; Engel, 2008), as well as species belonging to a diversity of other tropical insect groups (Grimaldi and Engel, 2005). We estimate that Termitidae diverged from Rhinotermitidae sometime in the Early Paleogene (perhaps Late Paleocene or Early Eocene) and subsequently began its diversification in the latest Eocene (Priabonian-Bartonian: ca. 40 Ma) to Early Oligocene, continuing to radiate throughout the remainder of the Neogene and Quaternary.

Though Jurassic remains of Isoptera have not been found, Isoptera appear to have diverged from cryptocercid roaches in the Late Jurassic. This would make termites the oldest group of eusocial animals, predating the origins of ants by some 35 million years. Major geological and biotic events in the Cretaceous probably had little effect on termites, since basal divergences appear to have preceded the drift of Gondwanan continents and the angiosperm radiations. Unfortunately, the stratigraphic sampling is too poor in the Late Cretaceous and Paleocene to determine any effects of the end-Cretaceous extinctions. The Tertiary thermal maximum of the late Paleocene and Eocene, however, probably had a profound effect on termites, specifically on the global spread of Mastotermes and the radiations of living Kalotermitidae and Neoisoptera. The rapid spread of C4 grasslands in the Miocene (Jacobs et al., 1999) doubtless promoted a minor diversification of the harvesters and the explosive diversification of many Termitidae, such as the Macrotermitinae. A macrotermitine nest, in fact, is known from the Miocene of Chad (Duringer et al., 2006).

Throughout the Cretaceous and early Tertiary, including the mid-Eocene, termites represented less than 1% of all insect specimens in all fossil deposits (fig. 3). Their abundance rises in the late Eocene to approximately 2%, and then spikes from 5%–10% during the Miocene as both amber and compression fossils, to the present day (in copals, or subfossil resins) (table 2, fig. 3). This spike in abundance is due to the diversification of the Termitidae. The abundance of ants rises dramatically in the Eocene (Grimaldi and Agosti, 2001; Dlussky and Rasnitsyn, 2002), and ants are generally much more abundant in Tertiary insect deposits than are termites, probably because termites feed within the wood where they nest or they travel through tunnels from nest to food sources, so foragers are rarely exposed and imagoes are exposed only during brief nuptial flights. Interestingly, many termitids will forage in the open and these are concomitantly the most abundant termites in Dominican amber and copal.

Our analysis indicates the importance of including fossils in cladistic analyses rather than mapping their putative ages onto molecular-based trees for purposes of dating. The identity of stem groups is obscured in the latter method, resulting in overestimates for divergence times. In our analyses this overestimation is highlighted by the traditional taxonomic placement of Termopsis and most Cretaceous fossils as Termopsidae s.l. and Hodotermitidae, respectively (fig. 3). These taxa in fact represent either a grade to more nested termite lineages or stem groups to some modern families (fig. 3). Using such fossils to calibrate the basal nodes for Hodotermitidae or Termopsidae would result in significant overestimates of the ages of these groups. A failure to distinguish stem groups in an analysis probably explains prior overestimates of the age of other insect lineages (e.g., Moreau et al., 2006; Hunt et al., 2007).

Patterns in termite diversification are very similar to those of the ants (Grimaldi and Agosti, 2001): throughout the Cretaceous both groups were rare and consisted of basal lineages. The diversity and abundance of termites and ants spiked in the Tertiary when speciose groups that form large colonies with highly specialized castes (for ants, the subfamilies Dolichoderinae, Formicinae, and Myrmecinae) eclipsed the smaller colonies of more basal taxa—the concept of “dynastic succession” (Wilson and Hölldobler, 2005). For ants, approximately 70 million years passed from origin to ecological dominance; in termites, this period was 100 million years. Thus, eusociality per se does not result in ecological success, but living in very large colonies with extreme division of labor does. Why, then, did it take so long for large colonies to evolve? We suggest that social evolution is like any other highly adaptive feature, such as the evolution of flight in feathered theropods, and thus may take tens of millions of years to refine.

Similarly, the symbiosis of termites with intestinal protozoa or bacteria which aid their break down of lignocellulose does not alone explain their ecological success as basal termite lineages exhibit the same mutualistic relationship. While their critical role as carbon recyclers is made possible by this symbiosis, this association existed for tens of millions of years before their rise in abundance and diversity (fig. 3).

One question remains: how was wood decomposed in Mesozoic forests with few or no termites? Patterns of coal and oil deposition suggest that lignocellulose did not rapidly decompose prior to termites and the actions of fungi or other organisms at the time must have been either slow or negligible. Coal is formed from ombrotrophic, or waterlogged, peat (Scott, 1987). While there are some tropical peatlands, such as the coastal “moor” forests of western Borneo and southern Sumatra, these are dwarfed in area compared to the boreal peatlands of sphagnum and heaths that comprise some 3% of earth's land surface. Tropical ecosystems produce more biomass, but much greater plant detritus accumulates in boreal forests and peatlands, which is traditionally explained by lower boreal evapotranspiration and because seasonality limits decomposition (Scott, 1987). Indeed, boreal peat lands lay hundreds of miles north of the most northern termites in the genus Reticulitermes Holmgren (Rhinotermitidae). Maximum termite diversity is equatorial, and half of that diversity falls between 18° N and 30° S latitudes; by 48° N and S it is 1%–4% that of the diversity at the Equator (Eggleton, 2000). Tropical ecosystems where termites are most abundant and diverse have notoriously thin humus layers (Richards, 1996). This may explain why coals that were formed prior to the appearance of termites in the Paleozoic and Early Mesozoic, and in largely the same regions and habitats, decomposed less (i.e., contained significantly more vitrain) than Tertiary and modern peats (Shearer et al., 1995; Raymond et al., 2000), as well as the formation of some vast reservoirs of petroleum, like those in the Early Cretaceous Nubian sandstones of present-day Africa and the Middle East. While some Miocene coal formations are astonishingly thick (Shearer et al., 1995), these were formed in palaeoclimates that today would have very few or no termites.

Acknowledgments

We are grateful to V. Krishna, K.K. Magill, I.A. Hinojosa-Díaz, M. Knight, and S. Thurston for assistance during the various stages of this work, and to R. Scheffrahn and an anonymous reviewer for their constructive comments on the manuscript. Support was provided by NSF grants DEB-0542909 (to M.S. Engel), DEB-0542726 (to D.A. Grimaldi), DEB-9870097 (to K. Krishna and D.A. Grimaldi), and a Guggenheim Fellowship from the John Simon Guggenheim Memorial Foundation (to M.S. Engel). This is a contribution of the Division of Entomology, University of Kansas Natural History Museum.

REFERENCES

1.

D. Azar 2000. Les ambres mésozoïques du Liban. Ph.D. dissertation, University of Paris, Orsay,. 202. Google Scholar

2.

G. Bechly 2007. Isoptera: termites. In D. M. Martill, G. Bechly, and R. F. Loveridge . The Crato fossil beds of Brazil: window into an ancient world: 249–262. Cambridge Cambridge University Press. xvi+625. Google Scholar

3.

D. E. Bignell and P. Eggleton . 2000. Termites in ecosystems. In T. Abe, D. E. Bignell, and M. Higashi . Termites: evolution, sociality, symbioses, ecology: 363–387. Dordrecht Kluwer Academic Publishers. xxii+466. Google Scholar

4.

C. Bitsch and C. Noirot . 2002. Gut characters and phylogeny of the higher termites (Isoptera: Termitidae): a cladistic analysis. Annales de la Société Entomologique de France 38/3:201–210. Google Scholar

5.

S. G. Brady, T. R. Schultz, B. L. Fisher, and P. S. Ward . 2006. Evaluating alternative hypotheses for the early evolution and diversification of ants. Proceedings of the National Academy of Sciences of the United States of America 103/48:18172–18177. Google Scholar

6.

L. R. Cleveland, S. K. Hall, E. P. Sanders, and J. Collier . 1934. The wood-feeding roach Cryptocercus, its protozoa, and the symbiosis between protozoa and roach. Memoirs of the American Academy of Arts and Sciences 17:185–342. Google Scholar

7.

B. N. Danforth, S. Sipes, J. Fang, and S. G. Brady . 2006. The history of early bee diversification based on five genes plus morphology. Proceedings of the National Academy of Sciences of the United States of America 103/41:15118–15123. Google Scholar

8.

L. L. Deitz, C. Nalepa, and K-D. Klass . 2003. Phylogeny of the Dictyoptera re-examined (Insecta). Entomologische Abhandlungen 61:69–91. Google Scholar

9.

J. Desneux 1904. A propos de la phylogénie des Termitides. Annales de la Société Entomologique de Belgique 48/8:278–286. Google Scholar

10.

G. M. Dlussky and A. P. Rasnitsyn . 2002. Ants (Hymenoptera, Formicidae) of Formation Green River and some other middle Eocene deposits of North America. Russian Entomological Journal 11/4:411–436. Google Scholar

11.

S. E. Donovan, D. T. Jones, W. A. Sands, and P. Eggleton . 2000. Morphological phylogenetics of termites (Isoptera). Biological Journal of the Linnean Society 70/3:467–513. Google Scholar

12.

P. Duringer, M. Schuster, J. F. Genise, A. Likius, H. T. Mackaye, P. Vignaud, and M. Brunet . 2006. The first fossil fungus gardens of Isoptera: oldest evidence of symbiotic termite fungiculture (Miocene, Chad basin). Naturwissenschaften 93/12:610–615. Google Scholar

13.

P. Eggleton 2000. Global patterns of termite diversity. In T. Abe, D. E. Bignell, and M. Higashi . Termites: evolution, sociality, symbioses, ecology. Dordrecht Kluwer Academic Publishers. 25–51.xxii+466. Google Scholar

14.

A. E. Emerson 1933. A revision of the genera of fossil and Recent Termopsinae (Isoptera). University of California Publications in Entomology 6/6:165–196. Google Scholar

15.

M. S. Engel 2001. A monograph of the Baltic amber bees and evolution of the Apoidea (Hymenoptera). Bulletin of the American Museum of Natural History 259:1–192. Google Scholar

16.

M. S. Engel 2004. Geological history of the bees (Hymenoptera: Apoidea). Revista de Tecnologia e Ambiente 10/2:9–33. Google Scholar

17.

M. S. Engel 2008. Two new termites in Baltic amber (Isoptera). Journal of the Kansas Entomological Society 81/3:194–203. Google Scholar

18.

M. S. Engel and D. A. Grimaldi . 2005. Primitive new ants in Cretaceous amber from Myanmar, New Jersey, and Canada (Hymenoptera: Formicidae). American Museum Novitates 3485:1–23. Google Scholar

19.

M. S. Engel, D. A. Grimaldi, and K. Krishna . 2007a. Primitive termites from the Early Cretaceous of Asia (Isoptera). Stuttgarter Beiträge zur Naturkunde Serie B (Geologie und Paläontologie) 371:1–32. Google Scholar

20.

M. S. Engel, D. A. Grimaldi, and K. Krishna . 2007b. A synopsis of Baltic amber termites (Isoptera). Stuttgarter Beiträge zur Naturkunde Serie B (Geologie und Paläontologie) 372:1–20. Google Scholar

21.

M. S. Engel and M. Gross . 2009. A giant termite from the Late Miocene of Styria, Austria (Isoptera). Naturwissenschaften 96/2:289–295. Google Scholar

22.

M. S. Engel and K. Krishna . 2004a. Family-group names for termites (Isoptera). American Museum Novitates 3432:1–9. Google Scholar

23.

M. S. Engel and K. Krishna . 2004b. An overlooked family-group name for termites (Isoptera). Entomological News 115/3:168. Google Scholar

24.

M. S. Engel and K. Krishna . 2007a. Drywood termites in Dominican amber (Isoptera: Kalotermitidae). Beiträge zur Entomologie 57/2:263–275. Google Scholar

25.

M. S. Engel and K. Krishna . 2007b. New Dolichorhinotermes from Ecuador and in Mexican amber (Isoptera: Rhinotermitidae). American Museum Novitates 3592:1–8. Google Scholar

26.

M. S. Engel and K. Krishna . 2007c. Two overlooked family-group names for fossil termites (Isoptera: Mastotermitidae). Entomological News 118:105–106. Google Scholar

27.

P. A. Goloboff 1997. NoName (NONA), version 2.0. Program and documentation. Tucumán Fundación y Instituto Miguel Lillo. Google Scholar

28.

D. Grimaldi and D. Agosti . 2001. A formicine in New Jersey Cretaceous amber (Hymenoptera: Formicidae) and early evolution of the ants. Proceedings of the National Academy of Sciences of the United States of America 97/25:13678–13683. Google Scholar

29.

D. Grimaldi, D. Agosti, and J. M. Carpenter . 1997. New and rediscovered primitive ants (Hymenoptera: Formicidae) in Cretaceous amber from New Jersey, and their phylogenetic relationships. American Museum Novitates 3208:1–43. Google Scholar

30.

D. Grimaldi and M. S. Engel . 2005. Evolution of the insects. Cambridge Cambridge University Press. xv+755. Google Scholar

31.

D. A. Grimaldi, M. S. Engel, and K. Krishna . 2008. The species of Isoptera (Insecta) from the Early Cretaceous Crato Formation: a revision. American Museum Novitates 3626:1–30. Google Scholar

32.

D. Grimaldi, A. Shedrinsky, and T. P. Wampler . 2000. A remarkable deposit of fossiliferous amber from the Upper Cretaceous (Turonian) of New Jersey. In D. Grimaldi Studies on fossils in amber, with particular reference to the Cretaceous of New Jersey. Leiden Backhuys. 1–76.viii+498. Google Scholar

33.

H. A. Hagen 1853. Hr. Peters Berichtete über die von ihm gesammelten und von Hrn. Dr. Hermann Hagen bearbeiten Neuropteren aus Mosambique. Bericht über die zur Bekanntmachung geeigneten. Verhandlungen der Königlichen Preussischen Akademie der Wissenschaften zu Berlin 18:479–484. Google Scholar

34.

H. A. Hagen 1858. Monographie der Termiten. Linnaea Entomologica 12:1–342. Google Scholar

35.

O. Heer 1849. Die Insektenfauna der Tertiärgebilde von Oeningen und von Radoboj in Croatien. Zweiter Theil: Heuschrecken, Florfliegen, Aderflügler, Schmetterlinge und Fliegen. Leipzig Engelmann. iv+264. Google Scholar

36.

K. Holmgren and N. Holmgren . 1917. Report on a collection of termites from India. Memoirs of the Department of Agriculture of India 5:137–171. Google Scholar

37.

N. Holmgren 1910. Das System der Termiten. Zoologischer Anzeiger 35/9–10:284–286. Google Scholar

38.

N. Holmgren 1911. Termitenstudien: 2. Systematik der Termiten. Die Familien Mastotermitidae, Protermitidae und Mesotermitidae. Kungliga Svenska Vetenskapsakademiens Handlingar 46/6:1–86. Google Scholar

39.

T. Hunt, J. Bergsten, Z. Levkanicova, A. Papadopoulou, O. St John, et al . 2007. A comprehensive phylogeny of beetles reveals the evolutionary origins of a superradiation. Science 318/5858:1913–1916. Google Scholar

40.

D. J. G. Inward, G. Beccaloni, and P. Eggleton . 2007a. Death of an order: a comprehensive molecular phylogenetic study confirms that termites are eusocial cockroaches. Biology Letters 3/3:331–335. [Lo et al. (2007) provide a concise rebuttal to the synonymy of Isoptera proposed by these authors]. Google Scholar

41.

D. J. G. Inward, A. P. Vogler, and P. Eggleton . 2007b. A comprehensive phylogenetic analysis of termites (Isoptera) illuminates key aspects of their evolutionary biology. Molecular Phylogenetics and Evolution 44/3:953–967. Google Scholar

42.

B. F. Jacobs, J. D. Kingston, and L. L. Jacobs . 1999. The origin of grass-dominated ecosystems. Annals of the Missouri Botanical Garden 86:590–643. Google Scholar

43.

E. A. Jarzembowski 1977. Insect fossils from the wealden of the Weald. Proceedings of the Geologist's Association 87:443–447. Google Scholar

44.

E. A. Jarzembowski 1980. Fossil insects from the Bembridge Marls, Palaeogene of the Isle of Wight, southern England. Bulletin of the British Museum (Natural History) Geology 33/4:237–293. Google Scholar

45.

E. A. Jarzembowski 1981. An early Cretaceous termite from southern England (Isoptera: Hodotermitidae). Systematic Entomology 6:91–96. Google Scholar

46.

E. A. Jarzembowski 1984. Early Cretaceous insects from southern England. Modern Geology 9:71–93. Google Scholar

47.

E. A. Jarzembowski 1991. The weald clay of the Weald: report of 1988/89 field meetings. Proceedings of the Geologist's Association 102/2:83–92. Google Scholar

48.

S. Kambhampati, K. M. Kjer, and B. L. Thorne . 1996. Phylogenetic relationships among termite families based on DNA sequence of mitochondrial 16S ribosomal RNA gene. Insect Molecular Biology 5/4:229–238. Google Scholar

49.

K-D. Klass and R. Meier . 2006. A phylogenetic analysis of Dictyoptera (Insecta) based on morphological characters. Entomologische Abhandlungen 63/1–2:3–50. Google Scholar

50.

K. Krishna, M. S. Engel, and D. A. Grimaldi . In press. Termites of the world (Isoptera). Oxford Oxford University Press. Google Scholar

51.

K. Krishna and D. Grimaldi . 2000. A new subfamily, genus, and species of termite (Isoptera) from New Jersey Cretaceous amber. In D. Grimaldi Studies on fossils in Amber, with particular reference to the Cretaceous of New Jersey: 133–140. Leiden Backhuys. viii+498. Google Scholar

52.

K. Krishna and D. Grimaldi . 2003. The first Cretaceous Rhinotermitidae (Isoptera): a new species, genus, and subfamily in Burmese amber. American Museum Novitates 3390:1–10. Google Scholar

53.

K. Krishna and D. Grimaldi . 2009. Diverse Rhinotermitidae and Termitidae (Insecta: Isoptera) in Dominican amber. American Museum Novitates 3640:1–48. Google Scholar

54.

E. Krzemińska, W. Krzemiński, J-P. Haenni, and C. Dufour . 1992. Les fantômes de l'ambre: insectes fossiles dans l'ambre de la Baltique. Neuchâtel Musée d'Histoire Naturelle de Neuchâtel. 142. Google Scholar

55.

A. Lacasa-Ruiz and X. Martínez-Delclòs . 1986. Meiatermes: nuevo género fósil de insecto isóptero (Hodotermitidae) de las calizas Necomienses del Montsec (Provincia de Lérida, España). Lleida Institut d'Estudis Ilerdencs. 65. Google Scholar

56.

F. Legendre, M. F. Whiting, C. Bordereau, E. M. Cancello, T. A. Evans, and P. Grandcolas . 2008. The phylogeny of termites (Dictyoptera: Isoptera) based on mitochondrial and nuclear markers: implications for the evolution of the worker and pseudergate castes, and foraging behaviors. Molecular Phylogenetics and Evolution 48/2:615–627. Google Scholar

57.

N. Lo, M. S. Engel, S. Cameron, C. A. Nalepa, G. Tokuda, et al . 2007. Save Isoptera: a comment on Inward et al. Biology Letters 3/5:562–563. Google Scholar

58.

N. Lo, G. Tokuda, H. Watanabe, H. Rose, M. Slaytor, et al . 2000. Evidence from multiple gene sequences indicates that termites evolved from wood-feeding cockroaches. Current Biology 10/13:801–804. Google Scholar

59.

R. G. Martins-Neto and M. Pesenti . 2006. The first fossil Termitidae (Isoptera) from the Oligocene of South America: the Entre-Córregos Formation of the Aiuruoca Basin, Minas Gerais, Brazil. Journal of the Entomological Research Society 8/3:63–68. Google Scholar

60.

F. A. McKittrick 1964. Evolutionary studies of cockroaches. Cornell University Agricultural Experiment Station Memoir 389:1–197. Google Scholar

61.

C. D. Michener 2007. The bees of the world. [2nd ed.]. Baltimore, MD Johns Hopkins University Press. xvi+[i]+953. Google Scholar

62.

C. S. Moreau, C. D. Bell, R. Vila, S. B. Archibald, and N. E. Pierce . 2006. Phylogeny of the ants: diversification in the age of angiosperms. Science 312/5770:101–104. Google Scholar

63.

A. Nel and E. Bourguet . 2006. Termite [sic] of the Early Eocene amber of France (Isoptera: Mastotermitidae, Kalotermitidae). Neues Jahrbuch für Geologie und Paläontologie Monatshefte 2006/2:101–115. Google Scholar

64.

M. Ohl and M. S. Engel . 2007. Die Fossilgeschichte der Bienen und ihrer nächsten Verwandten (Hymenoptera: Apoidea). Denisia 20:687–700. Google Scholar

65.

E. E. Perkovsky, A. P. Rasnitsyn, A. P. Vlaskin, and M. V. Taraschuk . 2007. A comparative analysis of the Baltic and Rovno amber arthropod faunas: representative samples. African Invertebrates 48:229–245. Google Scholar

66.

V. Perrichot, D. Néraudeau, A. Nel, and G. de Ploëg . 2007. A reassessment of the Cretaceous amber deposits from France and their palaeontological significance. African Invertebrates 48:213–227. Google Scholar

67.

G. Poinar and R. Poinar . 1999. The amber forest: a reconstruction of a vanished world. Princeton, NJ Princeton University Press. xviii+239. Google Scholar

68.

J. Prokop 2003. Remarks on the palaeoenvironmental changes based on reviewed Tertiary insect associations from the Krušné hory piedmont basins and the České středohoří Mts in northwestern Bohemia (Czech Republic). Acta Zoologica Cracoviensia Supplement 46:329–344. Google Scholar

69.

A. P. Rasnitsyn 2008. Palaeontological Institute, Russian Academy of Sciences, Baissa Collections ( http://www.palaeoentomolog.ru/Collections/baissa.html). [accessed 28 August 2008]. Google Scholar

70.

A. P. Rasnitsyn and D. L. J. Quicke . 2002. History of insects. Dordrecht Kluwer Academic Publishers. xii+517. Google Scholar

71.

A. Raymond, P. Cutlip, and M. Sweet . 2000. Rates and processes of terrestrial nutrient cycling in the Paleozoic: the world before beetles, termites and flies. In W. Allmon and D. J. Bottjer . Evolutionary paleoecology: the ecological context of macroevolutionary change. New York Columbia University Press. 235–283.xi+357. Google Scholar

72.

P. W. Richards 1996. The tropical rain forest: an ecological study. [2nd ed.]. Cambridge Cambridge University Press. xxiii+575. Google Scholar

73.

D. M. Rohr, A. J. Boucot, J. Miller, and M. Abbott . 1986. Oldest termite nest from the Upper Cretaceous of west Texas. Geology 14:87–88. Google Scholar

74.

M. L. Roonwal 1958. Recent work on termite research in India (1947–57). Transactions of the Bose Research Institute 22:77–100. Google Scholar

75.

D. W. Roubik, J. E. Moreno, C. Vergara, and D. Wittmann . 1986. Sporadic food competition with the African honey bee: projected impact on neotropical social bees. Journal of Tropical Ecology 2:97–111. Google Scholar

76.

W. A. Sands 1969. The association of termites and fungi. In K. Krishna and F. M. Weesner . Biology of termites [Vol. 1]: 495–524. New York Academic Press. xiii+598. Google Scholar

77.

A. C. Scott 1987. Coal and coal-bearing strata: recent advances. Oxford Blackwell Science Publishing. vii+332. Google Scholar

78.

J. C. Shearer, T. A. Moore, and T. D. Demchuk . 1995. Delineation of the distinctive nature of Tertiary coal beds. International Journal of Coal Geology 28/2–4:71–98. Google Scholar

79.

E. Sontag 2003. Animal inclusions in a sample of unselected Baltic amber. Acta Zoologica Cracoviensia Supplement 46:431–440. Google Scholar

80.

E. A. Sugden, R. W. Thorp, and S. L. Buchmann . 1996. Honey bee-native bee competition in Australia: focal point for environmental change and apicultural response. Bee World 77:26–44. Google Scholar

81.

A. Sugimoto, D. E. Bignell, and J. A. MacDonald . 2000. Global impact of termites on the carbon cycle and atmospheric trace gases. In T. Abe, D. E. Bignell, and M. Higashi . Termites: evolution, sociality, symbioses, ecology: 409–435. Dordrecht Kluwer Academic Publishers. xxii+466. Google Scholar

82.

G. J. Thompson, O. Kitade, N. Lo, and R. H. Crozier . 2000. Phylogenetic evidence for a single, ancestral origin of a ‘true’ worker caste in termites. Journal of Evolutionary Biology 13/6:869–881. Google Scholar

83.

B. L. Thorne, D. Grimaldi, and K. Krishna . 2000. Early fossil history of the termites. In T. Abe, D. E. Bignell, and M. Higashi . Termites: evolution, sociality, symbioses, ecology: 77–93. Dordrecht Kluwer Academic Publishers. xxii+466. Google Scholar

84.

T. Wappler 2003. Systematik, Phylogenie, Taphonomie und Paläoökologie der Insekten aus dem Mittel-Eozän des Eckfelder Maares, Vulkaneifel. Clausthaler Geowissenschaften 2:1–241. Google Scholar

85.

T. Wappler and M. S. Engel . 2006. A new record of Mastotermes from the Eocene of Germany (Isoptera: Mastotermitidae). Journal of Paleontology 80/2:380–385. Google Scholar

86.

E. O. Wilson 1971. The insect societies. Cambridge, MA Belknap Press. x+584. Google Scholar

87.

E. O. Wilson and B. Hölldobler . 2005. Eusociality: origin and consequences. Proceedings of the National Academy of Sciences of the United States of America 102/38:13367–13371. Google Scholar

Appendices

APPENDIX

Data Matrix for Characters Described in Table 1 Below is provided the suite of character-state codings for characters used in the analysis of Isoptera relationships. Within the matrix “-” indicates inapplicable codings (e.g., soldier characters for species which lack this caste); “?” indicates unknown states; “*” indicates a subset polymorphism for character 73 with states 3, 4, 5, 6, and 7; while “$” indicates a subset polymorphism for character 92 with states 0 and 1

i0003-0082-3650-1-1-t04.gif
Michael S. Engel, David A. Grimaldi, and Kumar Krishna "Termites (Isoptera): Their Phylogeny, Classification, and Rise to Ecological Dominance," American Museum Novitates 2009(3650), 1-27, (25 July 2009). https://doi.org/10.1206/651.1
Published: 25 July 2009
Back to Top