15
Symbiosis as a General Principle in Eukaryotic Evolution Angela E. Douglas Department of Entomology, Cornell University, Ithaca, New York 14853 Correspondence: [email protected] Eukaryotes have evolved and diversified in the context of persistent colonization by non- pathogenic microorganisms. Various resident microorganisms provide a metabolic capabil- ity absent from the host, resulting in increased ecological amplitude and often evolutionary diversification of the host. Some microorganisms confer primary metabolic pathways, such as photosynthesis and cellulose degradation, and others expand the repertoire of secondary metabolism, including the synthesis of toxins that confer protection against natural enemies. A further route by which microorganisms affect host fitness arises from their modulation of the eukaryotic-signaling networksthat regulate growth, development, behavior, and other functions. These effects are not necessarily based on interactions beneficial to the host, but can be a consequence of either eukaryotic utilization of microbial products as cues or host– microbial conflict. By these routes, eukaryote–microbial interactions play an integral role in the function and evolutionary diversification of eukaryotes. E ukaryotes do not live alone. They bear living cells of bacteria (Eubacteria and Archaea), and often eukaryotic microorganisms, on their surfaces and internally without any apparent ill effect. Furthermore, there is now persuasive evidence that all extant eukaryotes are derived from an association with intracellular bacteria within the Rickettsiales that evolved into mi- tochondria (Williams et al. 2007), with the implication that this propensity to form persis- tent associations has very ancient evolutionary roots. In this respect, the eukaryotes are differ- ent from the bacteria, among which only a sub- set associate with eukaryotes, specifically mem- bers of about 11 of an estimated 52 phyla of Eubacteria (Sachs et al. 2011) and a tiny minor- ity of Archaea (Gill and Brinkman 2011). The current interest in the microbiota as- sociated with eukaryotes stems from key tech- nological advances for culture-independent analysis of microbial communities, especially high-throughput sequencing methods to iden- tify and quantify microorganisms (Caporaso et al. 2011; Zaneveld et al. 2011). The Human Microbiome Project (commonfund.nih.gov), MetaHIT (metahit.eu), and other initiatives are yielding unprecedented information on the taxonomic diversity and functional capabilities of microorganisms associated with humans, other animals, and also plants, fungi, and uni- cellular eukaryotes (the protists), as well as abi- otic habitats (Qin et al. 2010; Muegge et al. 2011; Human Microbiome Project 2012a; Lundberg et al. 2012; Bourne et al. 2013). Much of this Editors: Patrick J. Keeling and Eugene V. Koonin Additional Perspectives on The Origin and Evolution of Eukaryotes available at www.cshperspectives.org Copyright # 2014 Cold Spring Harbor Laboratory Press; all rights reserved; doi: 10.1101/cshperspect.a016113 Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113 1 on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/ Downloaded from

Symbiosis as a General Principle in Eukaryotic Evolution

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Symbiosis as a General Principle in EukaryoticEvolution

Angela E. Douglas

Department of Entomology, Cornell University, Ithaca, New York 14853

Correspondence: [email protected]

Eukaryotes have evolved and diversified in the context of persistent colonization by non-pathogenic microorganisms. Various resident microorganisms provide a metabolic capabil-ity absent from the host, resulting in increased ecological amplitude and often evolutionarydiversification of the host. Some microorganisms confer primary metabolic pathways, suchas photosynthesis and cellulose degradation, and others expand the repertoire of secondarymetabolism, including the synthesis of toxins that confer protection against natural enemies.A further route by which microorganisms affect host fitness arises from their modulation ofthe eukaryotic-signaling networks that regulate growth, development, behavior, and otherfunctions. These effects are not necessarily based on interactions beneficial to the host, butcan be a consequence of either eukaryotic utilization of microbial products as cues or host–microbial conflict. By these routes, eukaryote–microbial interactions play an integral role inthe function and evolutionary diversification of eukaryotes.

Eukaryotes do not live alone. They bear livingcells of bacteria (Eubacteria and Archaea),

and often eukaryotic microorganisms, on theirsurfaces and internally without any apparentill effect. Furthermore, there is now persuasiveevidence that all extant eukaryotes are derivedfrom an association with intracellular bacteriawithin the Rickettsiales that evolved into mi-tochondria (Williams et al. 2007), with theimplication that this propensity to form persis-tent associations has very ancient evolutionaryroots. In this respect, the eukaryotes are differ-ent from the bacteria, among which only a sub-set associate with eukaryotes, specifically mem-bers of about 11 of an estimated 52 phyla ofEubacteria (Sachs et al. 2011) and a tiny minor-ity of Archaea (Gill and Brinkman 2011).

The current interest in the microbiota as-sociated with eukaryotes stems from key tech-nological advances for culture-independentanalysis of microbial communities, especiallyhigh-throughput sequencing methods to iden-tify and quantify microorganisms (Caporasoet al. 2011; Zaneveld et al. 2011). The HumanMicrobiome Project (commonfund.nih.gov),MetaHIT (metahit.eu), and other initiativesare yielding unprecedented information on thetaxonomic diversity and functional capabilitiesof microorganisms associated with humans,other animals, and also plants, fungi, and uni-cellular eukaryotes (the protists), as well as abi-otic habitats (Qin et al. 2010; Muegge et al. 2011;Human Microbiome Project 2012a; Lundberget al. 2012; Bourne et al. 2013). Much of this

Editors: Patrick J. Keeling and Eugene V. Koonin

Additional Perspectives on The Origin and Evolution of Eukaryotes available at www.cshperspectives.org

Copyright # 2014 Cold Spring Harbor Laboratory Press; all rights reserved; doi: 10.1101/cshperspect.a016113

Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113

1

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

research has focused on the Eubacteria, but eu-karyotic members of the microbiota, especiallythe fungi, are increasingly being investigated(Iliev et al. 2012; Findley et al. 2013).

Although driven by technological change,these culture-independent studies of the micro-biota of humans and other eukaryotes are hav-ing profound consequences for our conceptualunderstanding. In particular, there is a growingappreciation that the germ theory of disease,which has played a crucial role in improvingpublic health and food production throughthe 20th century, has also led to the widespreadbut erroneous belief that all microorganismsassociated with animals and plants are patho-gens. This outmoded perception is increasinglybeing replaced by the recognition that eukary-otes are chronically infected with benign andbeneficial microorganisms, and that diseasecan result from disturbance to the compositionor activities of the microbiota (McFall-Ngaiet al. 2013; Stecher et al. 2013).

This article reviews the pervasive impact ofsymbiosis with microorganisms on the traitsof their eukaryotic hosts and the resultant con-sequences for the evolutionary history of eu-karyotes. For the great majority of associations,the effects of symbiosis can be attributed totwo types of interaction. The first interac-tion—“symbiosis as a source of novel capabili-ties”—is founded on metabolic or other traitspossessed by the microbial partner but not theeukaryotic host. By gaining access to these ca-pabilities, eukaryotes have repeatedly derivedenhanced nutrition, defense against natural en-emies, or other selectively important character-istics. The second interaction—“the symbioticbasis of health”—comprises the improved vigorand fitness that eukaryote hosts gain throughmicrobial modulation of multiple traits, includ-ing growth rates, immune function, nutrientallocation, and behavior, even though the effectscannot be ascribed to specific microbial capa-bilities absent from the host. There is increasingevidence that the health benefits of symbiosisare commonly a consequence of microbial mod-ulation of the signaling networks by which thegrowth and physiological function of eukaryotehosts are coordinated.

This article comprises three sections: thetwo types of interaction are considered in turn,with the key patterns and processes illustratedby specific examples from a range of symbiosesin animals, plants, and other eukaryotes; andthe concluding comments address some keyunanswered questions about symbiosis in eu-karyotes. This article does not review the fulldiversity of associations made in this article onthe general principles of symbiosis in eukary-otic evolution; interested readers are referredto Douglas (2010).

SYMBIOSIS AS A SOURCE OF NOVELCAPABILITIES

The Ancient Evolutionary Roots of SymbiosesFounded on Metabolism

Eukaryotes have repeatedly gained access tocomplex metabolic capabilities by forming sym-bioses with other organisms that possess thesefunctions. By acquiring the organism (not justthe genes), eukaryotes have gained not only thegenes coding specific enzymes, but also the ge-netic and cellular machinery for their regulatedexpression.

The capacity to acquire metabolic capabili-ties by symbiosis is evident in the common an-cestor of all extant eukaryotes and can be linkedto the metabolic limitations in the lineage giv-ing rise to the eukaryotes. In particular, thislineage lacked the capacity for autotrophiccarbon fixation and aerobic respiration; thesetraits had evolved in the Eubacteria, presumablymore recently than the common ancestor ofthe two groups. The association with the a-pro-teobacterial ancestor of mitochondria probablyevolved 1–2 billion years ago. All extant eukary-otes have mitochondria or are descended frommitochondriate ancestors, raising the possibili-ty that the acquisition of mitochondria definedthe evolutionary origin of eukaryotes (Embleyand Martin 2006). (We cannot, however, ex-clude the alternative scenario that the ancestorof mitochondria was acquired by bona fide eu-karyotic cells, with the subsequent extinction ofall amitochondriate eukaryotic lineages.) Fur-thermore, the mitochondria have undergone

A.E. Douglas

2 Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

substantial evolutionary diversification, includ-ing the transformation into hydrogenosomes,functionally distinct organelles that generateATP with the production of hydrogen and lackoxidative phosphorylation (Hjort et al. 2010).Hydrogenosomes have evolved multiple timesin secondarily anaerobic eukaryotes, includingvarious ciliate protists and the lorificeran ani-mals in anoxic marine sediments (van der Gie-zen et al. 2005; Danovaro et al. 2010). The evo-lutionary history of eukaryotic acquisition ofoxygenic photosynthesis by symbiosis with thecyanobacterial ancestor of chloroplasts is alsocomplex. Unlike mitochondria, which evolvedjust once, chloroplasts have evolved from twodifferent cyanobacterial groups (Keeling 2013).One cyanobacterial-derived chloroplast is alliedto Synechoccus and has been reported in the rhi-zarian amoeba Paulinella chromatophora (Na-kayama and Ishida 2009; Nowack and Grossman2012). The other lineage is in the very successfulArchaeplastids, including both algae (e.g., Rho-dophytes, Chlorophytes) and terrestrial plants,with the subsequent evolution of some algaeinto secondary, and even tertiary, chloroplastsin further groups of eukaryotes (Keeling 2013).

Without doubt, the evolutionary and eco-logical success of the eukaryotes is founded ontheir acquisition of aerobic respiration and pho-tosynthesis by symbiosis, with the transition ofbacterial symbionts to organelles. Without thesecapabilities, the eukaryotes as a group wouldhave been excluded from oxic habitats (occu-pied by the photosynthetic cyanobacteria andaerobic bacteria) and restricted to the low-ener-gy lifestyle dictated by hypoxia and anoxia. Thelocalization of electron transport chain in aero-bic respiration and photosynthesis to the intra-cellular compartments of mitochondria andplastids, respectively, provided additional evo-lutionary opportunities for eukaryotes. The re-striction of energy production to the organellarmembranes, separate from cell membrane func-tion, permitted large cell size and associated re-duction in the ratio of (cell membrane surfacearea)/(cell volume) without prejudicing energyproduction (Lane and Martin 2010). (Bacterialcells with the respiratory chain localized to thecell membrane are generally restricted to small

size with a high surface area:volume ratio, al-though some bacteria can attain relatively largecell sizes through various adaptations (Schulzand Jorgensen 2001). Furthermore, the bac-terial-derived organelles represent additionalsubcellular compartments with key roles in me-tabolism (e.g.,b-oxidation of fatty acids, steroidhormone synthesis) and signaling (e.g., apopto-sis, calcium signaling), thereby enhancing cellu-lar efficiency and capacity to mediate complexinteractions.

Symbioses Founded on Primary Metabolismof Microbial Symbionts

Symbiosis with microorganisms is not restrictedto associations involving eubacterial endosym-bionts that evolved in the distant ancestors ofmodern eukaryotes. Rather, the capacity to ac-quire bacterial or eukaryotic microorganisms isa persistent theme in the biology of the eukary-otes and a major factor shaping adaptation ofeukaryotes to different habitats and lifestyles, aswell as the evolutionary diversification of manygroups. Particularly striking examples are pro-vided by the acquisition of photosynthetic algae(eukaryotic microorganisms that are, them-selves, the product of symbiosis) by fungi, gen-erating the lichens that account for nearly half ofall Ascomycete fungal species, and by variousanimals, including the corals, whose capacityto form reefs in the photic zone is symbiosisdependent.

Furthermore, aerobic respiration and pho-tosynthesis are not the only complex metabolictraits acquired by eukaryotes through symbio-sis, although none of the microbial partnerscontributing these other traits are known tohave evolved into organelles. Nitrogen fixationand chemosynthesis are two metabolic capa-bilities that are apparently absent from the an-cestral eukaryote and have been acquired bymultiple eukaryotic groups by symbiosis. An-giosperm plants of the eurosid clade are partic-ularly predisposed to associate with nitrogen-fixing bacteria, notably the rhizobia (includingRhizobium and Bradyrhizobium in the a-pro-teobacteria) and the actinobacteria Frankia.The access to atmospheric nitrogen in these

Symbiosis in Eukaryotic Evolution

Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113 3

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

plants facilitates their utilization of nitrogen-poor soils, for example, in early succession com-munities, and is recognized as a contributoryfactor to the evolutionary diversification of le-gumes, with their large, nitrogen-rich seedsprotected by nitrogen-containing toxins, in-cluding alkaloids (Houlton et al. 2008; Corbyet al. 2011). Nitrogen-fixing symbioses have alsobeen reported in protists, for example, in ma-rine diatoms (Foster et al. 2011), and some an-imals, notably the wood-feeding shipwormsand some termites (Lechene et al. 2007; Desaiand Brune 2012).

Chemosynthesis (the capacity to fix carbondioxide using the energy derived from the oxi-dation of reduced substrates, such as hydrogen,hydrogen sulfide, and methane) is widely ex-ploited by animals. The habitats where theseassociations flourish are zones of oxic/anoxicmixing, including aquatic sediments, hydro-thermal vents, and hydrocarbon seeps. Manyanimals in these habitats bear chemosyntheticsymbionts on their body surface (e.g., annelidworms), in gills (e.g., bivalve mollusks), or in acentral tissue called the trophosome (in pogo-nophoran worms), and many display reducedcapacity for feeding and a digestive tract that ismuch reduced or absent (Petersen et al. 2011;Roeselers and Newton 2012).

Various eukaryotes have compounded themetabolic limitations of their eukaryotic inher-itance by the loss of further metabolic capabil-ities. Reduction of metabolic scope is particu-larly evident in the animals, which, as a group,lack the capacity to synthesize at least nine of the20 protein amino acids (the essential amino ac-ids) and various cofactors required for the func-tion of enzymes central to metabolism (e.g.,various vitamins). The arthropods have, addi-tionally, lost the capacity for sterol synthesis.Generally, these lost metabolic capabilities com-prise many reactions and tend to be energetical-ly demanding (Wagner 2005). Organisms withan ample dietary supply of amino acids or co-factors would gain no advantage from retainingthese capabilities (relaxed selection) and argu-ably the selective advantage of energetic effi-ciency from losing them. As predicted fromthese considerations, the sponges and choano-

flagellates (the most basal animals, and closestrelative of animals, respectively) feed on livingorganisms, especially bacteria, which provide anutritionally balanced diet.

Repeatedly in the evolution of animals, sym-biosis with microorganisms has enabled ani-mals to escape from the nutritional constraintof feeding on a nutritionally balanced diet.Many microbial symbionts of animals haveprobably evolved by the progressive delay in an-imal digestion of food-associated microorgan-isms, raising the evolutionary scenario of a shiftin the animal to a poor-nutrient diet coupled tothe transition of nutrient acquisition by diges-tion to biotrophic release from intact microbialcells. Two diets for which animals require nu-tritional support from microbial symbiontsare vertebrate blood and plant sap, as indicatedby the possession of microbial symbionts byall animals feeding on these diets through thelife cycle (Buchner 1965). For example, symbi-otic microorganisms are borne in the gut orspecialized cells of the obligately blood-feedingleeches, ticks, lice, and bedbugs, and they arebelieved to provide their animal hosts with Bvitamins. Plant sap feeding has also evolvedmultiple times, but only among insects of theorder Hemiptera (including whiteflies, aphids,cicadas, and planthoppers). These insects deriveessential amino acids, nutrients in short supplyin the plant phloem and xylem sap, and pos-sibly vitamins from their microbial partners(Douglas 2009).

Microbial symbioses have also played an im-portant role in animal utilization of structuralplant material, including lignocellulose. Her-bivory has evolved multiple times in the verte-brates and, where investigated, the extraction ofenergy from the ingested cellulose is dependenton cellulolysis by symbiotic bacteria in an an-aerobic portion of the gut, known as the fer-mentation chamber. The waste products of thebacterial fermentation are short-chain fatty ac-ids, such as acetic acid and butyric acid. Theseproducts are transported into the epithelial cellsof the animal and via the blood to other organs,where they are used as a substrate for aerobicrespiration (Karasov and Douglas 2013). Micro-bial degradation of ingested lignocellulose and

A.E. Douglas

4 Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

fermentation of cellulose are generally less im-portant for invertebrate, especially insect, her-bivores, but can make an important contribu-tion to the carbon economy of insects, such astermites, wood roaches, wood wasps, and somebeetles, that feed on wood products of very lownutritional content (Geib et al. 2008; Suen et al.2010; Adams et al. 2011). Multiple factors maycontribute to the greater contribution of mi-crobial symbionts to vertebrates than insectsfeeding on structural plant material. They in-clude the greater anatomical barriers to a fer-mentation chamber in a small, highly aerobicinsect than in a larger vertebrate with a predom-inantly anoxic gut lumen, and the widespreadoccurrence of intrinsic cellulases (i.e., of ani-mal origin) in invertebrates, including many in-sects, but apparently in no vertebrates (Davisonand Blaxter 2005; Calderon-Cortes et al. 2012).Consequently, microbial symbiosis is a generalprinciple for the evolution of herbivory in ver-tebrates, but its significance for invertebrate an-imals has to be assessed on a case-by-case basis.

Symbioses and Secondary Metabolism

Symbiotic microorganisms can also contributeto the secondary metabolism of their eukary-otic hosts. “Secondary metabolism” refers tometabolic reactions and pathways that do notcontribute directly to organismal growth andreproduction but that can be crucial to the fit-ness of the organism in the natural environ-ment. Secondary metabolism includes the syn-thesis of toxins, antibiotics, pheromones andallomones, and pigments, as well as the degra-dation of toxins and other secondary metabo-lites. The capacity of eukaryotes for secondarymetabolism varies widely, even among closelyrelated species, and the technical difficulties inworking with secondary metabolism has led tovarious failures to recognize the role of micro-organisms in the secondary chemistry ofsome eukaryotes, as well as some unsubstanti-ated claims of symbiotic secondary metabolism.

Two complementary approaches provide anexcellent test for the role of microbial symbiontsin secondary metabolism: identification of therelevant genes in the sequenced genome of the

microbial partner, and perfect correlation be-tween expression of the secondary metaboliccapability in the host and presence of the mi-crobial partner with these genes. This can beillustrated by the contribution of bacterial sym-bionts to the production of a biologically activeclass of secondary compounds, the polyketides.For example, the beetle Paederus bears bacte-ria of the genus Pseudomonas, which have thegenes for polyketide synthase and linked re-actions required for the production of the poly-ketide profile of these animals (Piel et al. 2004a),and elimination of the Pseudomonas in Paederusresults in loss of polyketide production (Kell-ner 2001). Similarly, many grasses are protectedfrom herbivory by alkaloids that are synthe-sized by clavicipitaceous fungal endophytesthat ramify through the plant tissues (Fletcherand Harvey 1981), and the genetic basis of thefungal synthesis of one key alkaloid loline by thefungal endophyte in the grass Lolium has beenestablished (Spiering et al. 2005). Secondarymetabolites synthesized by microbial symbiontshave also been shown to contribute to the pro-tection against predators and pathogens in mul-tiple benthic marine animals, including spong-es (Piel et al. 2004b), bryozoans (Sharp et al.2007), and tunicates (Kwan et al. 2012).

Just as secondary chemistry is very diverse,so are the routes by which eukaryotes havegained access to these often-complex biosyn-thetic pathways for secondary metabolite syn-thesis. This is illustrated by recent research onthe source of carotenoid pigments in animals.Carotenoids have antioxidant properties andare also used as pigments. Animals, generally,are unable to synthesize carotenoids, and manyderive carotenoids from their diet. However, thecarotenoid profile of some insects is indepen-dent of a dietary supply. In one group of in-sects, the whiteflies, the carotenoids are of sym-biotic origin, with the carotenoid biosynthesisgenes coded by its bacterial symbiont Portiera(Sloan and Moran 2013). In two other groups,the aphids and spider mites, the carotenoidbiosynthesis genes are encoded in the animalgenome, following independent lateral transfersfrom fungi (Moran and Jarvik 2010; Altinciceket al. 2012; Novakova and Moran 2012).

Symbiosis in Eukaryotic Evolution

Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113 5

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

There is one further aspect to secondarymetabolism that deserves careful consideration:the contribution of resident microorganismsto the detoxification of secondary metabolites.This is currently a “hot topic” in biomedicalresearch, in the light of evidence that the activ-ities of the gut microbiota can determine thehalf-life and metabolic fate of orally adminis-tered drugs, with clinically important impli-cations for drug efficacy and toxicity (Haiserand Turnbaugh 2012; Holmes et al. 2012; Haiseret al. 2013). An analogous issue has been ofpersistent concern in the discipline of herbi-vory for decades. Although herbivorous ani-mals generally possess a diverse array of detox-ifying enzymes, including cytochrome P450monooxygenases, glutathione S-transferases,and esterases (Despres et al. 2007), symbioticmicroorganisms have long been invoked to me-diate the detoxification of plant allelochemicals(Jones 1984; Berenbaum 1988; Dillon and Dil-lon 2004). Supportive evidence comes from afew systems. The microbiota in the foregut fer-mentation chamber (the rumen) of ruminantmammals has the potential to detoxify ingestedallelochemicals in the food, as illustrated by thereported capacity of the rumen microbiota ofindigenous Indonesian cattle to degrade thetoxic amino acid mimosine in the tropical le-gume Leucaena leucocephala (Cheeke 1994).Similarly, the capacity of reindeer to use rein-deer moss (a lichen, Cladonia rangiferina) hasbeen attributed to a bacterium Eubacteriumrangiferina in the reindeer rumen, which canmetabolize usnic acid, an abundant metabolitein the lichen that is toxic to most animals(Sundset et al. 2008, 2010). Plant allelochemicaldetoxification has also been implicated in cer-tain insect symbioses, notably bark-feeding bee-tles and leaf-cutting ants. The gut microbiota ofthe mountain pine beetle Dendroctonus ponder-osae include bacteria of the genera Pseudomo-nas, Serratia, Rahnella, and Burkholderia thatbear genes involved in the degradation of ter-penes, the principal defensive compounds ofthe trees infested by these beetles (Adams et al.2013). The leaf-cutting ant Acromyrmex echina-tior maintains, and feeds on, the symbiotic fun-gus Leucocoprinus gongylophorus in its nest. The

fungus genome has multiple genes for laccases(a type of phenoloxidase). Enzymatically ac-tive laccase enzyme is passed through the gutof ants feeding on the fungus and releasedonto plant material, where it can degrade plantcompounds, such as tannins and flavonoids (DeFine Licht et al. 2013).

THE SYMBIOTIC BASIS OF HEALTH

The function of all living organisms is orches-trated by signaling networks that coordinateprocesses within individual cells and, for mul-ticellular forms, also among cells, tissues, or-gans, and so on. It has long been known thatthese regulatory networks can be manipulatedby pathogens and parasites, with various con-sequences ranging from the reorganization ofthe cytoskeleton and intracellular traffickingto the restructuring of host growth patterns, re-productive schedules, and behavior (Gandonet al. 2002; Bhavsar et al. 2007; Hughes et al.2012). It is now becoming increasingly clearthat the nonpathogenic resident microorgan-isms can also modulate the signaling networksof their eukaryotic hosts and that these manip-ulations are generally advantageous for the host.Although the data are still fragmentary, variouslines of evidence are consistent with the hy-pothesis that eukaryotes derive health benefitsfrom symbiosis because their regulatory net-works are structured to function in the contextof interactions with the resident microbiota.

The health benefits of symbiosis can beillustrated by plant-growth-promoting rhizo-bacteria (PGPRs), including strains of Azo-spirillum, Bacillus subtilis, Pseudomonas putida,and Enterobacter cloacae. As the term suggests,PGPRs are associated with plant roots and pro-mote plant growth. The underlying processesare complex and variable, but very commonlyinvolve the capacity to synthesize signaling mol-ecules, including plant hormones, for example,the auxin indole-3-acetic acid (IAA), and vola-tiles, for example, 2,3-butanediol (Spaepenet al. 2007; Lugtenberg and Kamilova 2009;Roca et al. 2013). The best-studied interactionrelates to PGPR-derived IAA, which triggersincreased root branching and a higher densi-

A.E. Douglas

6 Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

ty of root hairs. This altered root morphologyenhances nutrient uptake from the soil, there-by promoting plant growth (Dobbelaere et al.1999; Steenhoudt and Vanderleyden 2000). Theimportant implication of these results is thatthe plant regulatory networks controlling thepattern of plant growth are not structured togenerate optimal growth of microbe-free plants,which is achieved only in the context of signalexchange with associated rhizosphere bacteria.

The IAA-producing PGPRs are an exampleof resident microorganisms that influence hostgrowth by the release of molecules that are alsoan integral part of the host-signaling network(in this case, a plant hormone). Other com-pounds produced by microbial symbionts arenot known elements of the host-signaling net-work but, nevertheless, act to modulate (ampli-fy or dampen) host signaling. One examplecomes from research on the association betweenthe Drosophila fruit fly and its gut microbiota.When the microorganisms are eliminated, theDrosophila display depressed development ratesand elevated levels of lipid and circulating glu-cose, similar to the phenotypic traits of flieswith impaired insulin signaling (Shin et al.2011). Flies infected with a mutant of a domi-nant bacterial symbiont, Acetobacter pomorum,that cannot produce the enzyme pyrroloquino-line quinone-dependent alcohol dehydrogenase(PQQ-ADH) also display these traits, togetherwith reduced expression of the genes for insu-lin-like peptides (dilp-3 and dilp-5). PQQ-ADHmediates the oxidation of ethanol to acetic acid.When acetic acid was supplied in the food, de-velopment rates, lipid and glucose contents, anddilp gene expression in flies bearing the mutantbacteria shifted to values found in conventionalflies. These data suggest that acetic acid pro-duced by wild-type A. pomorum stimulates in-sulin signaling in the fly. Although the processesby which acetic acid interacts with insulin sig-naling are unknown, the resultant speeding oflarval development is advantageous for the in-sect, which develops in rotting fruit and mustcomplete larval development before the fruitresources are exhausted. As with the IAA-pro-ducing PGPRs associated with plant roots, theacetic-acid-producing bacteria in fruit fly guts

promote host health and vigor by increasing theamplitude of host signaling.

The hyperlipidemia and hyperglycemiadisplayed by Drosophila containing mutant A.pomorum are reminiscent of the elevated lip-id and glucose levels in the laboratory mousetreated with antibiotics that alter the composi-tion of the gut microbiota (Cho et al. 2012). Thegut microbiota of the mouse can also be alteredby diet or mutation, especially of the mouseimmune system and nutrient signaling, withcorrelated phenotypic lesions, especially in nu-trient allocation and immune function (Mas-lowski and Mackay 2011; Claesson et al. 2012;Maynard et al. 2012). Transplant studies sug-gest that the altered microbial community like-ly contributes to the altered host phenotype.When introduced to germ-free control mice(reared on the standard diet/of wild-type geno-type), the recipients display similar deleteriousphenotypic traits (Vijay-Kumar et al. 2010;Smith et al. 2013). These various data sets re-inforce the growing appreciation that the com-position and activities of the resident micro-biota are important for host health, such thatperturbation of the microbiota can contributeto chronic ill health, a condition known as “dys-biosis.” The concept of dysbiosis, first coined acentury ago (Metchnikoff 1910), has gained rel-evance in the context of hypotheses put forwardto account for the increase in immunologicaland metabolic disease in humans, with reducedexposure of children to environmental micro-organisms (hygiene hypothesis) or to specificmembers of the human resident microbiota(disappearing microbiota hypothesis) as pos-sible drivers (Strachan 1989; Blaser and Falkow2009).

Given the fitness costs of dysbiosis, the sus-ceptibility of the regulatory circuits of animalsand plants to modulation by their resident mi-crobiota appears as a very real vulnerability.How is it that eukaryote-signaling networksare not insulated from the influence of residentmicroorganisms, but appear to be under jointhost–microbial control?

Two sets of processes may contribute to thesymbiotic basis of health. The first is that micro-organisms provide a reliable cue for current or

Symbiosis in Eukaryotic Evolution

Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113 7

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

future environmental conditions. Eukaryotesthat respond appropriately to the microbialproducts, that is, display regulated changes ingrowth or developmental patterns, behavior,and the like, would be at a selective advantage.For example, the profile of metabolic end-prod-ucts from bacteria associated with rotting fruitmay provide a reliable cue for the longevity ofthe fruit resource, enabling Drosophila larvae totitrate their developmental time to environmen-tal circumstance. Specifically, the amplified in-sulin production in response to Acetobacter-de-rived acetic acid increases developmental rates,and this may be highly adaptive in the naturalenvironment. Microbial products are used ascues by various eukaryotes. For example, a sul-fonolipid released from specific bacteria, Algo-riphagus machipongonensis, induces colony for-mation in choanoflagellates (Alegado et al.2012); N-acyl homoserine-lactones (quorum-sensing molecules) released from Vibrio bacteriapromote settling of motile zoospores of thegreen alga Enteromorpha (Joint et al. 2002);and compounds released from the bacteriumPseudoalteromonas associated with the substratein marine environments promote the settlingand metamorphosis of barnacle larvae (Had-field 2011).

I describe these various compounds as cuesbecause their production by microorganisms isindependent of the response of the eukaryotichost, that is, whether and how a eukaryote de-rives information from microbial compoundshas no effect on their production by the micro-organisms. If a microbe-derived compound is areliable cue, the signaling network of the eu-karyote may evolve to increasing dependenceon that cue through reduced responsiveness toother environmental factors, and this may resultin dependence on specific microbial productsfor sustained function of the signaling networkand, ultimately, host health.

The alternative basis for microbial impactson host-signaling networks derives from con-flict between the microorganisms and host.Conflict is inevitable because of incomplete se-lective overlap between a eukaryotic host and itsresident microbiota. A eukaryote is a nutrient-rich patch in which microorganisms tolerant of

the immune system and other defenses canproliferate, and it is also frequently a route fordispersal. Consequently, many microorganismshave no selective interest in the reproductiveoutput of their host. (Exceptionally, the variousmaternally inherited forms are in conflict withthe host over host sex ratio.) Thus, microorgan-isms in the animal gut may favor hyperphagiadespite its negative consequences for host fit-ness, modulate gut peristalsis rates to optimizetheir residence time at the expense of optimalnutrient absorption by the host, and dampenimmune responses to favor their persistence(de La Serre et al. 2010; Round et al. 2011; Mat-sumoto et al. 2012; Maynard et al. 2012).

Counterintuitively, host–microbial conflictover the amplitude of host-signaling pathwayscan lead to host dependence, but only where theprevalence of the interacting microorganismsin the host population is very high. This impor-tant effect was first shown by research on therelationship between a parasitic wasp, Asobaratabida, and the vertically transmitted bacteriumWolbachia. Several lines of evidence suggestthat Wolbachia induces oxidative stress and alinked dampening of apoptotic signaling, prob-ably as a result of disruption of iron metabo-lism (Kremer et al. 2009). As a likely conse-quence, elimination of Wolbachia by antibiotictreatment results in massive apoptosis (pro-grammed cell death) of the ovaries, leaving thewasp reproductively sterile. It has been arguedthat, over evolutionary time, the wasp apoptoticpathways have gained heightened responsive-ness, in compensation for the inhibitory manip-ulation by Wolbachia (Pannebakker et al. 2007).Importantly, this effect is constitutive (presum-ably because Wolbachia is always present undernatural conditions), with the consequence thatappropriate apoptotic signaling requires themanipulative intervention of the Wolbachia bac-teria. Analogous host compensation for micro-bial manipulation may contribute to the mul-tiple demonstrations of interactions betweenthe resident microbiota and the developmental,nutritional, neurological, and immunologicalhealth of animals (Stappenbeck et al. 2002;Smith et al. 2007; Bravo et al. 2011; Olszaket al. 2012). Additionally, host dependence on

A.E. Douglas

8 Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

microbial products for sustained health may beretained in eukaryotic lineages, where the con-flict is resolved because of the sheer complexityof signaling networks. Elimination of microbialmodulation of one pathway may be selectedagainst because the multiple, knock-on effectson linked signaling pathways are highly delete-rious. Ultimately, the symbiotic basis of hosthealth may only be explicable in the context ofthe deep evolutionary history of interactionsbetween eukaryotes and their resident micro-biota.

CONCLUDING COMMENTS

It has taken a long time for biologists gener-ally to appreciate the significance of associa-tions with symbiotic microorganisms for eu-karyotes. Despite the discovery of the dualnature of lichens (fungi and algae/cyanobacte-ria) and near-ubiquitous associations betweenplant roots and mycorrhizal fungi in the 1800s,symbioses were treated as mere curiosities ofnature through much of the last century. Theoverwhelming molecular evidence for bacterialorigins of mitochondria and plastids, togetherwith the realization that the resident microbio-ta is vital for the human health and productivi-ty of crops and livestock has, at last, placed sym-biosis in the mainstream of biology. And yet,multiple questions central to our understand-ing of the function and evolution of symbiosesremain.

Tremendous opportunities are being pro-vided by the latest sequencing technologies. Al-though currently used mostly to catalog thecomposition and activities of bacterial com-munities (Human Microbiome Project 2012b),these methods are starting to be applied to in-vestigate both the diversity of eukaryotic micro-bial symbionts and how microbial communitiesare structured (Costello et al. 2012). A key un-resolved question is whether the functionaltraits of resident microorganisms can be pre-dicted from their taxonomic composition.Perhaps such a relationship is frequently con-founded by microbe–microbe and microbe–host interactions, resulting in spatiotemporalvariation in the traits of microorganisms. Re-

cent advances in single-cell imaging and sin-gle-cell sequencing (Pamp et al. 2012; Perniceet al. 2012; Lasken 2013) offer unprecedentedopportunities to investigate such heterogeneity.

These ecological issues segue into long-standing evolutionary questions, particularlywhether the microbiota can influence the spe-ciation patterns and evolutionary diversifica-tion of their hosts (Brucker and Bordenstein2013). In certain instances, evolutionary changemay be facilitated by among-microbe transfer ofsymbiosis-related gene clusters via “symbiosisislands” (Finan 2002), equivalent to pathoge-nicity islands in pathogenic bacteria, or by thedisplacement of one microbial partner by an-other taxon with different traits (Koga et al.2013; Toju et al. 2013). With the dramaticallyimproving genomic technologies, it is becom-ing increasingly feasible to answer these funda-mental questions.

We should also recognize that most currentresearch is focused on a tiny subset of the di-versity of symbioses, notably a few phyla of Eu-bacteria associated with animals, especially hu-mans and laboratory mice. A major outstandingissue is the taxonomic and functional diversityof symbioses across the evolutionary radiationof eukaryotes. In particular, relatively little isknown about symbioses involving protists aseither host or symbiont, even though the pro-tists account for most of the evolutionary diver-sity of eukaryotes. Some of these associationsare apparently without parallel in animals orplants, for example, motility conferred on largeprotists by spirochete ectosymbionts (Clevelandand Grimstone 1964), and protection frompredators by extrusive structures associatedwith bacterial symbionts of the ciliate Euplodi-nium (Petroni et al. 2000). Reinvigorated re-search on symbioses involving protists has thepotential to expand and modify our concept ofthe general principles of symbiosis.

ACKNOWLEDGMENTS

This work is supported by NIH grant1R01GM095372-01, NSF grant BIO 1241099,and the Sarkaria Institute for Insect Physiologyand Toxicology.

Symbiosis in Eukaryotic Evolution

Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113 9

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

REFERENCES

Adams AS, Jordan MS, Adams SM, Suen G, Goodwin LA,Davenport KW, Currie CR, Raffa KF. 2011. Cellulose-de-grading bacteria associated with the invasive woodwaspSirex noctilio. ISME J 5: 1323–1331.

Adams AS, Aylward FO, Adams SM, Erbilgin N, AukemaBH, Currie CR, Suen G, Raffa KF. 2013. Mountain pinebeetles colonizing historical and naive host trees are as-sociated with a bacterial community highly enriched ingenes contributing to terpene metabolism. Appl EnvironMicrobiol 79: 3468–3475.

Alegado RA, Brown LW, Cao S, Dermenjian RK, Zuzow R,Fairclough SR, Clardy J, King N. 2012. Bacterial regula-tion of colony development in the closest living relativesof animals. eLife 1: e00013.

Altincicek B, Kovacs JL, Gerardo NM. 2012. Horizontallytransferred fungal carotenoid genes in the two-spottedspider mite Tetranychus urticae. Biol Lett 8: 253–257.

Berenbaum MR. 1988. Micro-organisms as mediators ofintertrophic and intratrophic interactions. In Novel as-pects of insect–plant interactions (ed. Barbosa P, Letour-neau DK), pp. 91–123. Wiley, New York.

Bhavsar AP, Guttman JA, Finlay BB. 2007. Manipulation ofhost-cell pathways by bacterial pathogens. Nature 449:827–834.

Blaser MJ, Falkow S. 2009. What are the consequences of thedisappearing human microbiota? Nat Rev Microbiol 7:887–894.

Bourne DG, Dennis PG, Uthicke S, Soo RM, Tyson GW,Webster N. 2013. Coral reef invertebrate microbiomescorrelate with the presence of photosymbionts. ISME J7: 1459.

Bravo JA, Forsythe P, Chew MV, Escaravage E, Savignac HM,Dinan TG, Bienenstock J, Cryan JF. 2011. Ingestion ofLactobacillus strain regulates emotional behavior andcentral GABA receptor expression in a mouse via thevagus nerve. Proc Natl Acad Sci 108: 16050–16055.

Brucker RM, Bordenstein SR. 2013. The hologenomic basisof speciation: Gut bacteria cause hybrid lethality in thegenus Nasonia. Science 341: 667–669.

Buchner P. 1965. Endosymbioses of animals with plant mi-croorganisms. Wiley, Chichester, UK.

Calderon-Cortes N, Quesada M, Watanabe H, Cano-Cama-cho H, Oyama K. 2012. Endogenous plant cell wall di-gestion: A key mechanism in insect evolution. Annu RevEcol Evol Syst 43: 45–71.

Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Loz-upone CA, Turnbaugh PJ, Fierer N, Knight R. 2011.Global patterns of 16S rRNA diversity at a depth of mil-lions of sequences per sample. Proc Natl Acad Sci 108:4516–4522.

Cheeke PR. 1994. A review of the functional and evolution-ary roles of the liver in the detoxification of poisonousplants, with special reference to pyrrolizidine alkaloids.Vet Hum Toxicol 36: 240–247.

Cho I, Yamanishi S, Cox L, Methe BA, Zavadil J, Li K, Gao Z,Mahana D, Raju K, Teitler I, et al. 2012. Antibiotics inearly life alter the murine colonic microbiome and adi-posity. Nature 488: 621–626.

Claesson MJ, Jeffery IB, Conde S, Power SE, O’Connor EM,Cusack S, Harris HM, Coakley M, Lakshminarayanan B,O’Sullivan O, et al. 2012. Gut microbiota compositioncorrelates with diet and health in the elderly. Nature 488:178–184.

Cleveland LR, Grimstone AV. 1964. The fine structure of theflagellate Mixotricha paradoxa and its associated micro-organisms. Proc Roy Soc London B 159: 668–686.

Corby HDL, Smith DL, Sprent JI. 2011. Size, structure andnitrogen content of seeds of Fabaceae in relation to nod-ulation. Bot J Lin Soc 167: 251–280.

Costello EK, Stagaman K, Dethlefsen L, Bohannan BJ, Rel-man DA. 2012. The application of ecological theory to-ward an understanding of the human microbiome. Sci-ence 336: 1255–1262.

Danovaro R, Dell’Anno A, Pusceddu A, Gambi C, Heiner I,Kristensen RM. 2010. The first metazoa living in perma-nently anoxic conditions. BMC Biol 8: 30.

Davison A, Blaxter M. 2005. Ancient origin of glycosyl hy-drolase family 9 cellulase genes. Mol Biol Evol 22: 1273–1284.

De Fine Licht HH, Schiott M, Rogowska-Wrzesinska A,Nygaard S, Roepstorff P, Boomsma JJ. 2013. Laccase de-toxification mediates the nutritional alliance betweenleaf-cutting ants and fungus-garden symbionts. ProcNatl Acad Sci 110: 583–587.

de La Serre CB, Ellis CL, Lee J, Hartman AL, Rutledge JC,Raybould HE. 2010. Propensity to high-fat diet-inducedobesity in rats is associated with changes in the gut micro-biota and gut inflammation. Am J Physiol 299: G440–G448.

Desai MS, Brune A. 2012. Bacteroidales ectosymbionts ofgut flagellates shape the nitrogen-fixing community indry-wood termites. ISME J 6: 1302–1313.

Despres L, David JP, Gallet C. 2007. The evolutionary ecol-ogy of insect resistance to plant chemicals. Trends EcolEvol 22: 298–307.

Dillon RJ, Dillon VM. 2004. The gut bacteria of insects:Nonpathogenic interactions. Annu Rev Entomol 49: 71–92.

Dobbelaere S, Croonenburghs A, Thys A, Vande Broek A,Vanderleyden J. 1999. Phytostimulatory effect of Azospi-rillum brasilense wild type and mutant strains altered inIAA production on wheat. Plant Soil 212: 155–164.

Douglas AE. 2009. The microbial dimension in insect nu-tritional ecology. Funct Ecol 23: 38–47.

Douglas AE. 2010. The symbiotic habit. Princeton UniversityPress, Princeton, NJ.

Embley TM, Martin W. 2006. Eukaryotic evolution, changesand challenges. Nature 440: 623–630.

Finan TM. 2002. Evolving insights: Symbiosis islands andhorizontal gene transfer. J Bac 184: 2855–2856.

Findley K, Oh J, Yang J, Conlan S, Deming C, Meyer JA,Schoenfeld D, Nomicos E, Park M, NIH IntramuralSequencing Center Comparative Sequencing Program,et al. 2013. Topographic diversity of fungal and bacterialcommunities in human skin. Nature 498: 367–370.

Fletcher LR, Harvey IC. 1981. An association of a Loliumendophyte with ryegrass staggers. New Zeal Vet J 29: 185–186.

A.E. Douglas

10 Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

Foster RA, Kuypers MM, Vagner T, Paerl RW, Musat N,Zehr JP. 2011. Nitrogen fixation and transfer inopen ocean diatom–cyanobacterial symbioses. ISME J5: 1484–1493.

Gandon S, Agnew P, Michalakis Y. 2002. Coevolution be-tween parasite virulence and host life-history traits. AmNat 160: 374–388.

Geib SM, Filley TR, Hatcher PG, Hoover K, Carlson JE, delMar Jimenez-Gasco M, Nakagawa-Izumi A, Sleighter RL,Tien M. 2008. Lignin degradation in wood-feeding in-sects. Proc Natl Acad Sci 105: 12932–12937.

Gill EE, Brinkman FS. 2011. The proportional lack ofarchaeal pathogens: Do viruses/phages hold the key?Bioessays 33: 248–254.

Hadfield MG. 2011. Biofilms and marine invertebrate lar-vae: What bacteria produce that larvae use to choosesettlement sites. Annu Rev Mar Sci 3: 453–470.

Haiser HJ, Turnbaugh PJ. 2012. Is it time for a metagenomicbasis of therapeutics? Science 336: 1253–1255.

Haiser HJ, Gootenberg DB, Chatman K, Sirasani G, BalskusEP, Turnbaugh PJ. 2013. Predicting and manipulatingcardiac drug inactivation by the human gut bacteriumEggerthella lenta. Science 341: 295–298.

Hjort K, Goldberg AV, Tsaousis AD, Hirt RP, Embley TM.2010. Diversity and reductive evolution of mitochondriaamong microbial eukaryotes. Philos Trans R Soc Lond BBiol Sci 365: 713–727.

Holmes E, Kinross J, Gibson GR, Burcelin R, Jia W, Petters-son S, Nicholson JK. 2012. Therapeutic modulation ofmicrobiota–host metabolic interactions. Sci Trans Med4: 136–137.

Houlton BZ, Wang YP, Vitousek PM, Field CB. 2008. Aunifying framework for dinitrogen fixation in the terres-trial biosphere. Nature 454: 327–330.

Hughes DP, Brodeur J, Thomas F. 2012. Host manipulationby parasites. Oxford University Press, Oxford.

Human Microbiome Project Consortium. 2012a. Structure,function and diversity of the healthy human micro-biome. Nature 486: 207–214.

Human Microbiome Project Consortium. 2012b. A frame-work for human microbiome research. Nature 486:215–221.

Iliev ID, Funari VA, Taylor KD, Nguyen Q, Reyes CN, StromSP, Brown J, Becker CA, Fleshner PR, Dubinsky M, et al.2012. Interactions between commensal fungi and theC-type lectin receptor Dectin-1 influence colitis. Science336: 1314–1317.

Joint I, Tait K, Callow ME, Callow JA, Milton D, WilliamsP, Camara M. 2002. Cell-to-cell communication acrossthe prokaryote–eukaryote boundary. Science 298: 1207.

Jones CG. 1984. Microorganisms as mediators of plant re-source exploitation by insect herbivores. In A new ecology:Novel approaches to interactive systems (ed. Price PW,et al.), pp. 53–99. Wiley, New York.

Karasov WH, Douglas AE. 2013. Comparative digestivephysiology. Compr Physiol 3: 741–783.

Keeling PJ. 2013. The number, speed, and impact of plastidendosymbioses in eukaryotic evolution. Annu Rev PlantBiol 64: 583–607.

Kellner RL. 2001. Suppression of pederin biosynthesisthrough antibiotic elimination of endosymbionts in Pae-derus sabaeus. J Insect Physiol 47: 475–483.

Koga R, Bennett GM, Cryan JR, Moran NA. 2013. Evolu-tionary replacement of obligate symbionts in an ancientand diverse insect lineage. Environ Microbiol 15: 2037–2081.

Kremer N, Voronin D, Charif D, Mavingui P, Mollereau B,Vavre F. 2009. Wolbachia interferes with ferritin expres-sion and iron metabolism in insects. PLoS Pathog 5:e1000630.

Kwan JC, Donia MS, Han AW, Hirose E, Haygood MG,Schmidt EW. 2012. Genome streamlining and chemicaldefense in a coral reef symbiosis. Proc Natl Acad Sci 109:20655–20660.

Lane N, Martin W. 2010. The energetics of genome com-plexity. Nature 467: 929–934.

Lasken RS. 2013. Single-cell sequencing in its prime. NatureBiotechnol 31: 211–212.

Lechene CP, Luyten Y, McMahon G, Distel DL. 2007.Quantitative imaging of nitrogen fixation by indi-vidual bacteria within animal cells. Science 317: 1563–1566.

Lugtenberg B, Kamilova F. 2009. Plant-growth-promotingrhizobacteria. Annu Rev Microbiol 63: 541–556.

Lundberg DS, Lebeis SL, Paredes SH, Yourstone S, Gehring J,Malfatti S, Tremblay J, Engelbrektson A, Kunin V, del RioTG, et al. 2012. Defining the core Arabidopsis thalianaroot microbiome. Nature 488: 86–90.

Maslowski KM, Mackay CR. 2011. Diet, gut microbiota andimmune responses. Nat Immunol 12: 5–9.

Matsumoto M, Ishige A, Yazawa Y, Kondo M, Muramatsu K,Watanabe K. 2012. Promotion of intestinal peristalsis byBifidobacterium spp. capable of hydrolysing sennosides inmice. PLoS ONE 7: e31700.

Maynard CL, Elson CO, Hatton RD, Weaver CT. 2012. Re-ciprocal interactions of the intestinal microbiota and im-mune system. Nature 489: 231–241.

McFall-Ngai M, Hadfield MG, Bosch TC, Carey HV, Doma-zet-Loso T, Douglas AE, Dubilier N, Eberl G, Fukami T,Gilbert SF, et al. 2013. Animals in a bacterial world, a newimperative for the life sciences. Proc Natl Acad Sci 110:3229–3236.

Metchnikoff E. 1910. The prolongation of life: Optimisticstudies. Heinemann, Portsmouth, NH.

Moran NA, Jarvik T. 2010. Lateral transfer of genes fromfungi underlies carotenoid production in aphids. Science328: 624–627.

Muegge BD, Kuczynski J, Knights D, Clemente JC, GonzalezA, Fontana L, Henrissat B, Knight R, Gordon JI. 2011.Diet drives convergence in gut microbiome functionsacross mammalian phylogeny and within humans. Sci-ence 332: 970–974.

Nakayama T, Ishida K. 2009. Another acquisition of a pri-mary photosynthetic organelle is underway in Paulinellachromatophora. Curr Biol 19: R284–R285.

Novakova E, Moran NA. 2012. Diversification of genesfor carotenoid biosynthesis in aphids following anancient transfer from a fungus. Mol Biol Evol 29: 313–323.

Symbiosis in Eukaryotic Evolution

Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113 11

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

Nowack EC, Grossman AR. 2012. Trafficking of protein intothe recently established photosynthetic organelles of Pau-linella chromatophora. Proc Natl Acad Sci 109: 5340–5345.

Olszak T, An D, Zeissig S, Vera MP, Richter J, Franke A,Glickman JN, Siebert R, Baron RM, Kasper DL, et al.2012. Microbial exposure during early life has persistenteffects on natural killer T cell function. Science 336: 489–493.

Pamp SJ, Harrington ED, Quake SR, Relman DA, BlaineyPC. 2012. Single-cell sequencing provides clues about thehost interactions of segmented filamentous bacteria(SFB). Genome Res 22: 1107–1119.

Pannebakker BA, Loppin B, Elemans CP, Humblot L, VavreF. 2007. Parasitic inhibition of cell death facilitates sym-biosis. Proc Natl Acad Sci 104: 213–215.

Pernice M, Meibom A, Van Den Heuvel A, Kopp C, Domart-Coulon I, Hoegh-Guldberg O, Dove S. 2012. A single-cellview of ammonium assimilation in coral–dinoflagellatesymbiosis. ISME J 6: 1314–1324.

Petersen JM, Zielinski FU, Pape T, Seifert R, Moraru C,Amann R, Hourdez S, Girguis PR, Wankel SD, Barbe V,et al. 2011. Hydrogen is an energy source for hydrother-mal vent symbioses. Nature 476: 176–180.

Petroni G, Spring S, Schleifer K-H, Verni F, Rosati G. 2000.Defensive extrusive ectosymbionts of Euplotidium (Cil-iophora) that contain microtubule-like structures arebacteria related to Verrucomicrobia. Proc Natl Acad Sci97: 1813–1817.

Piel J, Hofer I, Hui D. 2004a. Evidence for a symbiosis islandinvolved in horizontal acquisition of pederin biosyn-thetic capabilities by the bacterial symbiont of Paederusfuscipes beetles. J Bacteriol 186: 1280–1286.

Piel J, Hui D, Wen G, Butzke D, Platzer M, Fusetani N,Matsunaga S. 2004b. Antitumor polyketide biosynthesisby an uncultivated bacterial symbiont of the marinesponge Theonella swinhoei. Proc Natl Acad Sci 101:16222–16227.

Qin J, Li R, Raes J, Arumugam M, Burgdorf KS, ManichanhC, Nielsen T, Pons N, Levenez F, Yamada T, et al. 2010. Ahuman gut microbial gene catalogue established by meta-genomic sequencing. Nature 464: 59–65.

Roca A, Pizarro-Tobias P, Udaond Z, Fernandez M, MatillaMA, Molina-Henares MA, Molina L, Segura A, Duque E,Ramos JL. 2013. Analysis of the plant growth-promotingproperties encoded by the genome of the rhizobacteriumPseudomonas putida BIRD-1. Environ Microbiol 15: 780–794.

Roeselers G, Newton IL. 2012. On the evolutionary ecologyof symbioses between chemosynthetic bacteria and bi-valves. Appl Microbiol Biotechnol 94: 1–10.

Round JL, Lee SM, Li J, Tran G, Jabri B, Chatila TA, Maz-manian SK. 2011. The Toll-like receptor 2 pathway estab-lishes colonization by a commensal of the human micro-biota. Science 332: 974–977.

Sachs JL, Skophammer RG, Regus JU. 2011. Evolutionarytransitions in bacterial symbiosis. Proc Natl Acad Sci 108:10800–10807.

Schulz HN, Jorgensen BB. 2001. Big bacteria. Annu RevMicrobiol 55: 105–137.

Sharp KH, Davidson SK, Haygood MG. 2007. Localizationof “Candidatus Endobugula sertula” and the bryostatinsthroughout the life cycle of the bryozoan Bugula neritina.ISME J 1: 693–702.

Shin SC, Kim S-H, You H, Kim B, Kim AC, Lee K-A, YoonJ-H, Ryu J-H, Lee W-J. 2011. Drosophila microbiomemodulates host developmental and metabolic homeosta-sis via insulin signaling. Science 334: 670–674.

Sloan DB, Moran NA. 2013. The evolution of genomic in-stability in the obligate endosymbionts of whiteflies. Ge-nome Biol Evol 5: 783–793.

Smith K, McCoy KD, Macpherson AJ. 2007. Use of axe-nic animals in studying the adaptation of mammals totheir commensal intestinal microbiota. Sem Immunol 19:59–69.

Smith MI, Yatsunenko T, Manary MJ, Trehan I, MkakosyaR, Cheng J, Kau AL, Rich SS, Concannon P, Mycha-leckyj JC, et al. 2013. Gut microbiomes of Malawiantwin pairs discordant for kwashiorkor. Science 339:548–554.

Spaepen S, Versees W, Gocke D, Pohl M, Steyaert J, Vander-leyden J. 2007. Characterization of phenylpyruvate decar-boxylase, involved in auxin production of Azospirillumbrasilense. J Bacteriol 189: 7626–7633.

Spiering MJ, Moon CD, Wilkinson HH, Schardl CL. 2005.Gene clusters for insecticidal loline alkaloids in thegrass–endophytic fungus Neotyphodium uncinatum. Ge-netics 169: 1403–1414.

Stappenbeck TS, Hooper LV, Gordon JI. 2002. Develop-mental regulation of intestinal angiogenesis by indigen-ous microbes via Paneth cells. Proc Natl Acad Sci 99:15451–15455.

Stecher B, Maier L, Hardt WD. 2013. “Blooming” in the gut:How dysbiosis might contribute to pathogen evolution.Nat Rev Microbiol 11: 277–284.

Steenhoudt O, Vanderleyden J. 2000. Azospirillum, a free-living nitrogen-fixing bacterium closely associated withgrasses: Genetic, biochemical and ecological aspects.FEMS Microbiol Rev 24: 487–506.

Strachan DP. 1989. Hay fever, hygiene, and household size.Brit Med J 299: 1259–1260.

Suen G, Scott JJ, Aylward FO, Adams SM, Tringe SG, Pin-to-Tomas AA, Foster CE, Pauly M, Weimer PJ, Barry KW,et al. 2010. An insect herbivore microbiome withhigh plant biomass-degrading capacity. PLoS Genet 6:e1001129.

Sundset MA, Kohn A, Mathiesen SD, Praesteng KE.2008. Eubacterium rangiferina, a novel usnic acid-resis-tant bacterium from the reindeer rumen. Naturwiss 95:741–749.

Sundset MA, Barboza PS, Green TK, Folkow LP, Blix AS,Mathiesen SD. 2010. Microbial degradation of usnic acidin the reindeer rumen. Naturwiss 97: 273–278.

Toju H, Tanabe AS, Notsu Y, Sota T, Fukatsu T. 2013. Diver-sification of endosymbiosis: Replacements, co-speciationand promiscuity of bacteriocyte symbionts in weevils.ISME J 7: 1378–1390.

van der Giezen M, Tovar J, Clark CG. 2005. Mitochondrion-derived organelles in protists and fungi. Int Rev Cytol 244:175–225.

A.E. Douglas

12 Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

Vijay-Kumar M, Aitken JD, Carvalho FA, Cullender TC,Mwangi S, Srinivasan S, Sitaraman SV, Knight R, LeyRE, Gewirtz AT. 2010. Metabolic syndrome and alteredgut microbiota in mice lacking Toll-like receptor 5. Sci-ence 328: 228–231.

Wagner A. 2005. Energy constraints on the evolution of geneexpression. Mol Biol Evol 22: 1365–1374.

Williams KP, Sobral BW, Dickerman AW. 2007. A robust spe-cies tree for the a-proteobacteria. J Bac 189: 4578–4586.

Zaneveld JR, Parfrey LW, Van Treuren W, Lozupone C, Cle-mente JC, Knights D, Stombaugh J, Kuczynski J, KnightR. 2011. Combined phylogenetic and genomic approach-es for the high-throughput study of microbial habitatadaptation. Trends Microbiol 19: 472–482.

Symbiosis in Eukaryotic Evolution

Cite this article as Cold Spring Harb Perspect Biol 2014;6:a016113 13

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

2014; doi: 10.1101/cshperspect.a016113Cold Spring Harb Perspect Biol  Angela E. Douglas Symbiosis as a General Principle in Eukaryotic Evolution

Subject Collection The Origin and Evolution of Eukaryotes

FunctionEukaryotic Gen(om)e Architecture and Cellular The Persistent Contributions of RNA to

Jürgen Brosius

Mitochondrion Acquired?Eukaryotic Origins: How and When Was the

Anthony M. Poole and Simonetta Gribaldo

the Plant KingdomGreen Algae and the Origins of Multicellularity in

James G. Umen

Bacterial Influences on Animal OriginsRosanna A. Alegado and Nicole King

Phylogenomic PerspectiveThe Archaeal Legacy of Eukaryotes: A

Lionel Guy, Jimmy H. Saw and Thijs J.G. Ettemathe Eukaryotic Membrane-Trafficking SystemMissing Pieces of an Ancient Puzzle: Evolution of

Klute, et al.Alexander Schlacht, Emily K. Herman, Mary J.

OrganellesCytoskeleton in the Network of Eukaryotic Origin and Evolution of the Self-Organizing

Gáspár Jékely LifeIntracellular Coevolution and a Revised Tree ofOrigin of Eukaryotes and Cilia in the Light of The Neomuran Revolution and Phagotrophic

Thomas Cavalier-Smith

from Fossils and Molecular ClocksOn the Age of Eukaryotes: Evaluating Evidence

et al.Laura Eme, Susan C. Sharpe, Matthew W. Brown,

Consequence of Increased Cellular ComplexityProtein Targeting and Transport as a Necessary

Maik S. Sommer and Enrico Schleiff

SplicingOrigin of Spliceosomal Introns and Alternative

Manuel Irimia and Scott William Roy

How Natural a Kind Is ''Eukaryote?''W. Ford Doolittle

Eukaryotic Epigeneticsand Geochemistry on the Provenance ofImprints of Bacterial Biochemical Diversification Protein and DNA Modifications: Evolutionary

et al.L. Aravind, A. Maxwell Burroughs, Dapeng Zhang,

Insights from Photosynthetic EukaryotesEndosymbionts to the Eukaryotic Nucleus? What Was the Real Contribution of

David Moreira and Philippe Deschamps

http://cshperspectives.cshlp.org/cgi/collection/ For additional articles in this collection, see

Copyright © 2014 Cold Spring Harbor Laboratory Press; all rights reserved

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from

Phylogenomic PerspectiveThe Eukaryotic Tree of Life from a Global

Fabien BurkiComplex LifeBioenergetic Constraints on the Evolution of

Nick Lane

http://cshperspectives.cshlp.org/cgi/collection/ For additional articles in this collection, see

Copyright © 2014 Cold Spring Harbor Laboratory Press; all rights reserved

on March 17, 2022 - Published by Cold Spring Harbor Laboratory Press http://cshperspectives.cshlp.org/Downloaded from