ago-L- p“ 1" . ’3 1345556! - ~er .3: ‘ 1 r . l ”E.“ m m. a,“ u» ‘n .. ‘ n H. 3 I”... .; q r:~ r c. ' ,.. '2'. - N, ‘tfi'z -—-..m‘\ Wears 3003 37/ 7 7 M DR ”\ U B RARY '7 “336mm” Michigan State 3 University Physiological Ecology of Termite Gut Spirochetes presented by Joseph Rex Graber has been accepted towards fulfillment of the requirements for the Doctoral degree in Microbiologl jor Profe s Signature /2?9wgflm3r a Date MSU is an Manama AcdonlEqud Opportunity Institution PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJClRC/Dateouepes-sz PHYSIOLOGICAL ECOLOGY OF TERMITE GUT SPIROCHETES By Joseph Rex Graber A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHEOSOPHY Department of Microbiology and Molecular Genetics 2003 ABSTRACT PHYSIOLOGICAL ECOLOGY OF TERMITB GUT SPIROCHETES By Joseph Rex Graber T reponema Sp. strains ZAS-l and ZAS-2 were the first spirochetes to be isolated frOm termite hindguts and are also the first known spirochetal H2/COz-acetogens. The ZAS strains were examined for nutritional and physiological properties relevant to in situ growth and survival. In addition to using H2+C02 as growth substrates, both strains grew on a variety of organic compounds and were capable of mixotrophic growth (i.e. the simultaneous use of both H2 and organic substrates). Enzyme activities of the Wood/Ljungdahl pathway of acetogenesis were detected in both strains, whose H2 thresholds were within the range typical of acetogens (650 and 490 ppmv, respectively). Both strains were able to maintain growth in the presence of small amounts of O; (0.5%, vol/vol) and possessed enzyme activities which could mediate protection from 02. These results demonstrate that T reponema strains ZAS-1 and ZAS-2 are nutritionally versatile, perform acetogenesis by the Wood/Ljungdahl pathway, and are likely able to tolerate oxidative stress in the partially hypoxic hindgut environment. T reponema strain ZAS-9, an additional spirochete from Z. angusticollis hindguts, was not a homoacetogen, and in fact produced H; as a major product during sugar fermentation. This strain also differed from ZAS-1 and ZAS-2 in a variety of other properties, suggesting that strains ZAS-1 and ZAS-2 be assigned to a single new species of T reponema, whereas ZAS-9 should be considered a separate new Treponema species. Strains ZAS-1 and ZAS-2 had requirements for folate, and important cofactor in acetogenesis. On the notion that other termite gut microbes must supply folate in situ, heterotrophic organisms were isolated from the guts of the termite Zootermopsis angusticollis and screened for folate secretion. Two folate-secreting isolates (Serratia strain ZFX-1 and Lactococcus strain ZFX-2) were identified; both strains produced a compound that could replace the folate requirements of ZAS-1 and ZAS-2. The folate produced by both ZF X strains was identified as folinate. These results suggest that strains ZFX-1 and ZFX-2 are capable of supplying folinate to acetogenic spirochetes in the termite gut, and may also be important in providing folates to other members of the gut microbiota, as well as to the termite host. Despite the fact that methanogenesis is the more energetically favorable process, H2/COz-homoacetogens are the primary Hz-consumers in the guts of wood-feeding termites. To explain this observation, I hypothesized that gut methanogens are inhibited by pteridine compounds. The pteridine compound lumazine is an inhibitor of methanogenesis, and pteridines are known to be unusually prevalent in insect physiology. A variety of pteridines, as well as termite gut extracts, were tested for inhibition of cultures of the termite methanogen Methanobrevibacterfiliformis; only lumazine showed significant inhibitory activity. No lumazine was detected in termite hindguts by TLC and HPLC analysis. These results indicate that gut methanogens are not inhibited by pteridine compounds. The marginalization of methanogenesis in wood-feeding termites must be explained by other factors. ACKNOWLEDGEMENTS I would never have been able to complete the research presented in this dissertation without the support of my fiiends, family, and coworkers. First and foremost, my appreciation goes to my advisor, Dr. John Breznak, for the guidance and encouragement he has provided over the past five years and for teaching me what it means to truly be a scientist. I would also like to thank my fellow graduate students, especially Bradley Stevenson, Dan Buckley, Joel Klappenbach, Deborah Hogan, Joel Hashimoto, Keri Byzek, Merry Riley, Stephanie Eichorst, Kristin Huizinga, John Wertz, Dawn Greensides, Nikki Lebrasscur, and Kirsten Fertuck. Thanks to Kwi Kim for being a super tech and for looking out for all of the "kids" in the lab (myself included). I would also like to thank the members of my committee for their consistently solid advice: Dr. Mike Klug, Dr. C. A. Reddy, Dr. Mark Scriber, and especially Dr. Tom Schmidt who, owing to the "symbiotic" nature of the Schmidt and Breznak labs, has been like a second advisor to me. I am very grateful to the Department of Microbiology & Molecular Genetics for providing a superb educational environment and to the Center for Microbial Ecology for providing financial support that has allowed me to present my research at a number of scientific meetings and conferences. Finally, I would like to thank my mother, Faye, my father, Doug, and my brother, Jerome; whatever measure of success I've achieved personally and professionally is largely owed to them. iv TABLE OF CONTENTS LIST OF TABLES .......................................................................................................... vii LIST OF FIGURES ....................................................................................................... viii CHAPTER 1: INTRODUCTION .................................................................................... 1 Termite Gut Spirochetes: A Historical Perspective ................................................ 1 Nature & Function of the Termite Hindgut Community ........................................ 6 Physiochemical Gradients in the Termite Hindgut ............................................... 10 Acetogenesis vs. Methanogenesis in Wood-feeding Termites ............................. 12 The Spatial Resource Partitioning Hypothesis ..................................................... 17 Isolation of Termite Gut Spirochetes .................................................................... 20 Dissertation Research ........................................................................................... 25 References ............................................................................................................. 27 CHAPTER 2: PHYSIOLOGY AND NUTRITION OF T REPONEMA STRAINS ZAS-l AND ZAS-2, Hz/COz-ACETOGENIC SPIROCHETES FROM TERMITE HINDGUTS ..................................................................................................................... 37 Introduction ........................................................................................................... 37 Materials & Methods ............................................................................................ 38 Results ................................................................................................................... 42 Nutrition and growth of T reponema strains ZAS-1 and ZAS-2 ...................... 42 Mixotrophic growth ......................................................................................... 46 Hydrogen thresholds ........................................................................................ 46 Hz/Coz-acetogenic enzyme activity ................................................................ 49 Oxygen tolerance and oxidative stress enzymes ............................................. 49 Discussion ............................................................................................................. 53 References ............................................................................................................. 60 CHAPTER 3: TAXONOMY OF T REPONEMA STRAINS ZAS-l, ZAS-2, AND ZAS-9 ............................................................................................................................... 66 Introduction ........................................................................................................... 66 Materials & Methods ........................................................................................... 67 Results ................................................................................................................... 68 Morphology of T reponema strains ZAS-1, ZAS-2, and ZAS-9 ................ 68 Nutrition, growth, and fermentation products of strain ZAS-9 ................ 70 G + C content of DNA .............................................................................. 70 Discussion ............................................................................................................. 73 References ............................................................................................................. 76 CHAPTER 4: INTERSPECIES COFACTOR TRANSFER SUPPORTS THE FOLATE REQUIREMENT OF ACETOGENIC TERMITE GUT SPIROCHETES .............................................................................................................. 77 Introduction ........................................................................................................... 77 Materials & Methods ............................................................................................ 79 Results ................................................................................................................... 86 F olate requirements of T reponema strains ZAS-1 and ZAS-2 ................. 86 Isolation of folate-secreting bacteria from termite hindguts ..................... 88 Characterization of folate-secreting strains .............................................. 88 F olate production by Serratia str. ZFX-l and Lactococcus str. ZFX- ..... 94 Discussion ............................................................................................................. 97 References ........................................................................................................... 102 CHAPTER 5: ACETOGENESIS VS. METHANOGENESIS IN THE TERMITE HINDGUT: THE LUMAZINE HY POTHESIS ........................................................ 107 Introduction ........................................................................................................ 107 Materials & Methods .......................................................................................... 111 Results ................................................................................................................. 1 15 Methanogen inhibition assays ................................................................. 115 TLC of R. flavipes gut extracts ............................................................... 118 HPLC analysis of R. flavipes gut extracts .............................................. 118 Discussion ........................................................................................................... 120 CHAPTER 6: SUMMARY .......................................................................................... 129 vi LIST OF TABLES Table 1.1. Redox potentials, Gibbs free energy changes, and H2 thresholds for various anaerobic hydrogen consuming processes ........................................................................ 14 Table 2.1. Substrate utilization by Treponema strains ZAS-1 and ZAS-2 ...................... 44 Table 2.2. Cell yields and acetate production by Treponema strain ZAS-2 growing under unitrophic and mixotrophic conditions ............................................................................. 48 Table 2.3. Enzyme activities relevant to Hz/COz-acetogenesis in T reponema str. ZAS-l and ZAS-2 ......................................................................................................................... 51 Table 2.4. Oxygen and peroxide detoxifying enzyme activities in Treponema str. ZAS-1 and ZAS-2 ......................................................................................................................... 54 Table 3.1. Substrates Used as energy sources by T reponema strain ZAS-9 ................... 71 Table 3.2. Fermentation products of T reponema strain ZAS-9 ....................................... 72 Table 4.1. Substrates utilization by Serratia strain ZFX-l and Lactococcus strain ZFX-2 ................................................................................................................................ 93 Table 5.1. Thin layer chromatography of pteridines and R. flavipes gut extract ........... 119 vii LIST OF FIGURES Figure 1.1. Reticulitermes flavipes, a wood-feeding lower termite, pictured with a gut extracted from a separate individual. The volume of the hindgut is approximately 1 ul. Adapted from reference 18 ................................................................................................. 2 Figure 1.2. Phase contrast micrograph of diluted hindgut contents collected from the termite Reticulitermes flavipes. Spirochetes, including some attached to protozoa (P), are indicated by white arrows. Non-spirochetal prokaryotes are indicated by black arrows. Also labeled is an undigested wood particle (W). Adapted from reference 18 ................. 4 Figure 1.3. Carbon flow during cellulose digestion in wood-feeding termites. The thickness of arrows indicates the relative significance of the indicated microorganims in the hindgut .......................................................................................................................... 7 Figure 1.4. Radial profiles of oxygen (0) and hydrogen (0) in an agarose-embedded hindgut from Reticulitermes flavipes. The circular inset indicates the relative sizes of the oxic and anoxic zones in the gut (lefi half) and the hydrogen concentration gradient (right half). Adapted from reference 21 ..................................................................................... 11 Figure 1.5. In situ Gibb's free energies for methanogenesis (El) and acetogenesis (0) under termite hindgut conditions. Values were calculated at 25°C for pH 7.2 and the following substrate and product concentrations: acetate, 80 mM; methane, 1 mM; bicarbonate, 60 mM .......................................................................................................... 16 Figure 1.6. Diagram of the termite hindgut in radial cross-section, showing the spatial distributions of various microbial groups and gradients of 02 (blue) and H2 (red) .......... 19 Figure 1.7. Morphology of termite gut Treponema strain ZAS-l by phase contrast (a) and transmission electron microscopy of intact (c) and transverse sectioned (b) cells. The morphology of strain ZAS-2 is virtually identical to that of ZAS-1. Arrows indicate periplasmic flagella, whose insertion points are subterminal (c). Scale bars = 10 pm (a) and 0.1 pm (b, c) ............................................................................................................... 21 Figure 1.8. Phylogenetic tree inferred from 168 rDNA sequences of termite gut treponemes (strains ZAS-1, ZAS-2, and ZAS-9), representative known spirochetes, and spirochetal 16S rDNA clones generated directly from termite gut contents (T gut clones). A maximum likelihood technique (fastDNAml) was used to generate the tree. The vertical line on the right delimits a distinct subgroup (the "termite cluster”) within the genus Treponema. Scale bar represents units of evolutionary distance and is based on sequence divergence ......................................................................................................... 22 Figure 1.9. The Wood/Ljungdahl pathway of acetogenesis, highlighting enzymes of the methyl and carbonyl branches. THF=Tetrahydrofolate. Adapted from reference 36 ..... 24 viii Figure 2.1. H2 stimulation of chemoorganotrophic growth of Treponema strain ZAS-2. Substrates added to 2YACo medium under N2/CO2 head spaces (80/20, v/v) were: 10 mM maltose (O), 16 mM H2 (A) 10 mM maltose plus 16 mM H2 (Cl), and no added substrates (V) .................................................................................................................... 45 Figure 2.2. Mixotrophic growth of Treponema strain ZAS-2 on H2+CO2 and maltose. Cells were grown in bottles containing 100 ml of maltose supplemented 2YACo medium and 645 ml gas phase ........................................................................................................ 47 Figure 2.3. Determination of H2 threshold for replicate cultures (0 and Cl) of Treponema strain ZAS-2, growing under mixotrophic conditions (H2 + CO2 plus trimethoxybenzoate). H2 additions are indicated by arrows. The threshold value is indicated by a dashed line .............................................................................. 50 Figure 2.4. Effect of 02 on the growth of Treponema strain ZAS-2. Cultures were initially grown under 80% H2:20% CO2, with O2 additions made at 125 hours (indicated by arrow). The resulting concentration of O2 in the culture tube headspace (percentage by volume) was: 0 (CI), 0.5 (O), l (A), 2.5 (V), and 5 (O) .............................................. 52 Figure 3.1. Morphology of T reponema strains ZAS-2 (A & B) and ZAS-9 (C and D) by phase contrast (A & C) and transmission electron microscopy (B & D). Insets in panels A and C show cells of each strain, with wavelengths indicated by arrowed bars. Periplasmic flagella are indicated by arrows in panels B & D. Strain ZAS-1 is morphologically indistinguishable from strain ZAS-2. Bars, 5 pm (A and C), 2.5 pm (A and C insets), and 0.1 mm (B & D). Panel B is adapted from reference 4, and is presented here for comparison .......................................................................................... 69 Figure 4.1. Generalized structure of folate compounds (A) and pteridine portion of the reduced form of folate compounds, tetrahydrofolate (B). Individual derivatives vary in reduction state, C1 substitution (at N5 or N”), and number of glutamyl residues (generally from 1 to 9) ...................................................................................................... 78 Figure 4.2. Folate utilization by Treponema strains ZAS-l and ZAS-2. All folate compounds were provided at a final concentration of 500 ng/ml. The specific compounds were: folic acid (El), dihydrofolate (O), tetrahydrofolate (A), S-formyltetrahydrofolate (i.e. folinic acid) (V), and no addition (0) ....................................................................... 87 Figure 4.3. Folate secretion assay of strains ZFX-1 and ZFX-2. Putative folate secretion is indicated by the formation of satellite colonies of the folate-requiring bioassay organism, Enterococcus hirae, surrounding surface colonies of ZFX-l (inset A) and ZFX-2 (inset B) ................................................................................................................. 89 ix Figure 4.4. Phylogenetic trees inferred from 16S rDNA sequences (~1500 nt) of ZFX-1 (A), ZFX-2 (B), and related organisms. A maximum likelihood technique (fastDNAml) was used to generate the trees. The homologous sequence from Desulfovibrio senezii was used as an outgroup in tree A (not shown). The scale bars represent units of evolutionary distance and are based on sequence divergence .......................................... 91 Figure 4.5. Phase contrast micrographs of Serratia strain ZFX-1 (A) and Lactococcus strain ZFX-2 (B) ............................................................................................................... 92 Figure 4.6. HPLC chromatograms showing fluorescent (280/359 nm) compounds present in culture filtrates of strains ZFX-1 (A) and ZFX-2 (B) and for a standard mixture (C) of THF (peak 1, 1 ng), 5-CH3-TI-IF (peak 2, 1 ng). and S-HCO-THF (peak 3, 5 ng). In panels A & B, dashed and solid lines indicate samples taken at the time of inoculation and at the end of logarithmic growth, respectively. The peak eluting at 31 minutes in panel A could not be identified, but did not correlate to any folate standard ................... 95 Figure 4.7. Growth of Treponema strains ZAS-1 and ZAS-2 in folate-free medium 2YA supplemented with folinate (O) or culture filtrates of Serratia strain ZFX-l (O) or Lactococcus strain ZFX-2 (Cl). Culture filtrates were added at 10% (v/v) final concentration. F olinate was provided at 500 ng/ml final concentration. Negative controls (A) were supplemented with 10% uninoculated GM3 medium ....................................... 96 Figure 5.1. Generalized chemical structure of pterin compounds (A) and production of lumazine from pterin via enzymatic deamination (B) .................................................... 109 Figure 5.2. Methane production by growing cells of M. filr'formis following addition of lumazine (O), R. flavipes gut extract (A), and anoxic water (Cl). Lumazine was provided at a final concentration of 1 mM, and gut extract was added at 5% (vol/vol). The time of additions is indicated by the arrow ................................................................................. 116 Figure 5.3. Methane production by growing cells of M. filiformis following addition of biolumazine (<>), xantholumazine (A), and isoxantholumazine (V), all provided at 0.5 mM final concentrations. Additions of anoxic water (Cl) and 0.1 mM lumazine (0) were used as controls. The time of additions is indicated by the arrow .................................. 117 Chapter 1 Introduction Termite Gut Spirochetes: A Historical Perspective In 1877, Joseph Leidy published his first observations on the strikingly complex community of microbes inhabiting the intestines of the wood-feeding termite Reticulitermesflavipes (56) (Figure 1.1). Leidy initially sought to discover the food source of the insects by examining their gut contents. However, upon microscopic examination of fluid collected from hindguts of worker termites, he reports: "The brownish matter proved to be the semi—liquid food; but my astonishment was great to find it swarming with myriads of parasites, which indeed actually predominated over the real food in quantity. Repeated examination showed that all individuals harbored the same world of parasites wonderful in number, variety, and form. " Leidy reported detailed observations on a variety of protozoa inhabiting the gut fluid and also observed an abundant and diverse population of somewhat smaller organisms possessing a distinctive undulate morphology and rapid motility (Figure 1.2). Initially referred to by Leidy as Vibrio termitis, these organisms would subsequently be recognized as spirochetes (35). Spirochetes belong to an independent division (Spirochaetes) within the domain Bacteria. Nearly all representatives share a set of distinctive morphological features, including helical or undulate protoplasmic cylinder and periplasmic flagella located within the outer cell membrane (23, 25, 43). Spirochetes are generally 1 mm Hindgut \litlgut Figure 1.1. Reticulitermes flavipes, a wood-feeding lower termite, pictured with a gut extracted from a separate individual. The volume of the hindgut is approximately 1 ul. Adapted from reference 18. chemoheterotrophs, consuming carbohydrates, amino acids, long chain fatty acids, or long chain fatty alcohols as energy substrates (24). Despite these points of similarity, the various spirochetal species are physiologically diverse and capable of colonizing a wide range of habitats. This is illustrated by the differences between the free-living members of the genus Spirochaeta (5 7), most frequently found in aquatic or sedimentary environments, and the generally host-associated genus Treponema, which form relationships (ranging fi'om commensal to pathogenic) with a wide range of vertebrate and invertebrate animals (67). Termite hindguts harbor the most diverse and abundant population of spirochetes found in any known environment. Spirochetes can account for up to 50% of the prokaryotic cells observed in the small but densely packed termite hindgut (109-10ll total prokaryotic cells/ml) (74) (Figure 1.2). Termite gut spirochetes are morphologically diverse with cells ranging in length from 3 to 100 um and displaying a variety of wavelengths or body pitches (10); individual temrite species typically harbor from 12 to 15 visibly distinguishable morphotypes (94). These cells were readily identified as true spirochetes by the in situ observation of characteristic periplasmic flagella in tranmission electron micrographs (3, 9, 10). Spirochetes appear to colonize the central lumen of the hindgut almost exclusively and are only rarely observed among the dense biofilrn of organisms on the hindgut epithelial surface (9). In the lumen, spirochetes are observed both as free-swimming individuals and as epibionts attached to the surfaces of various gut protozoa (26, 83-85). In the case of the protozoa Mixotricha paradoxa, regularly arranged rows of spirochetes cover most of the cell surface. The coordinated motion of these epibionts propels the host cell in a unique motility symbiosis (26). Figure 1.2. Phase contrast micrograph of diluted hindgut contents collected from the termite Reticulitermes flavipes. Spirochetes, including some attached to protozoa (P), are indicated by white arrows. Non-spirochetal prokaryotes are indicated by black arrows. Also labeled is an undigested wood particle (W). Adapted fiom reference 18. Despite the conspicuous abundance of spirochetes in termite guts, attempts to isolate these organisms in pure culture met with repeated failure. To circumvent this problem, many researchers turned to cultivation independent molecular methods to survey the composition of the spirochetal community (1, 2, 70, 72, 74). By cloning and analyzing spirochetal 16S rDNA genes in community DNA collected from termite hindguts, these studies revealed that spirochetal morphological diversity in termite guts was paralleled by a high degree of phylogenetic diversity. Lilbum et a1. (59) undertook a comprehensive phylogenetic survey of spirochetal 16S rDNA sequences collected from seven termite species spanning five of the seven extant termite families. All spirochetal 16s rDNA sequences collected grouped within the genus Treponema. However, nearly all of these sequences formed a distinct cluster (the “termite cluster”) within the genus and had only limited similarity to the 16S sequences other treponemcs (S 91%). Curiously, the termite spirochetal sequences were most closely affiliated with Spirochaeta stenostrepta and Spirochaeta caldaria, two fi’ee-living spirochetes that nevertheless group with the neponemes on the basis of 16S rDNA sequences. In a single termite species (R. flavipes), it was estimated that 26 distinct spirochetal phylotypes were present, corresponding to at least 21 new species. FISH analysis of spirochetes in situ with a series of specific rRNA targeted probes resulted in labeling of both free-swimming spirochetes and protozoan epibionts, but individual phylotypes appeared to be restricted to one of the two life-styles (66). While these studies provided new information on spirochetal phylogeny, definitive information on the physiological properties of the termite gut spirochetes remained elusive. However, the observation that spirochetes were consistently present in seemingly healthy termites suggested a non-pathogenic or even beneficial relationship. This hypothesis was tested by Eutick et al. (40), using antibiotic treatments to selectively eliminate spirochetes from the guts of termites. While these treatments led to a reduction in termite vitality, it was impossible to determine if this was the direct result of spirochete elimination (as opposed to the loss of some other non-spirochetal member of the community) and if so, what specific benefits the host had derived from the spirochetal population. Until a representative could be isolated in pure culture, the role of spirochetes in the termite hindgut would remain a matter of speculation. Nature & Function of the Termite Hindgut Community Studies of the termite hindgut symbiotic system have most frequently focused on two problems posed by the utilization of wood as a primary food source by termites: First, how is this relatively refractory biopolymer converted to a metabolically useful source of carbon and energy? Second, how do termites thrive on a nutrient source containing low (0.05%) nitrogen? Accordingly, the roles of the hindgut microbial community in the digestion of lignocellulose and provision of carbon, energy, and nitrogen to their termite hosts evolved as central areas of research interest. Carbon flow in the hindgut of lower (i.e. evolutionarily basal termites comprising six of the seven termite families) wood-feeding termites is illustrated in Figure 1.3. Cellulose hydrolysis occurs by the combined actions of termite cellulases secreted fi'om the salivary glands (45, 96, 97) and cellulases of anaerobic protozoa (69, 98-100), the latter of which ferment the liberated glycosyl units within phagosomes by anaerobic protozoa residing within the hindgut: Cellulose Termite cellulases & Protozoa i Soluble Glycosides l Lactic Acid Bacteria (?) Protozoa & Protozoa & Other Gut Microbes Lactate Other Gut Microbes Unknown Acetate < H2 + 002 I Acetogenic Bacteria Termite Respiration Methanogenic Archaea C02 + H20 CH4 Figure 1.3. Carbon flow during cellulose digestion in wood-feeding termites. The thickness of arrows indicates the relative significance of the indicated microorganims in the hindgut. l.) [C6H1005]n + 3n H2O —-> 2n CH3COOH + 2n CO2 + 4n H2 The H2 and CO2 produced can serve as substrates for homoacetogenic bacteria or methanogenic archaea, which produce acetate or methane, respectively: 2.) 4 H2 + 2 CO2 —) CH3COOH + 2 H20 3.) 4H2+CO2—)CH4+2H2O Acetate is absorbed through the gut epithelium and oxidized to fuel the insect’s energy metabolism (42), while methane and unused H2 passively diffuses out of the gut. In the context of the symbiosis, methanogens could be viewed as parasites that consume reducing equivalents that could otherwise be directed towards acetate production. Acetate dominates the hindgut pool of volatile fatty acids, accounting for 94-98 mol% of all VFAs at concentrations ranging 58 to 81 mM (68). Oxidation of microbially produced acetate was estimated to satisfy between 71 and 100% of the host termite’s energy requirements (68). Given this observation, and the fact that methanogenesis is severely limited in the termite hindgut (see below), carbon flow in the hindgut was thought to be dominated by reactions 1 and 2, with bacterial acetogenesis accounting for up to 33% of total acetate production (15). However, recent studies using microinjection of radiolabelled substrates into intact guts have revealed that up to one-third of the total carbon flux passes through a previously unrecognized pool of lactate. The low standing concentration of lactate measured in the gut reflects the rapid turnover of this substrate to acetate (93). This observation has revised the estimated carbon flux accounted for by H2/CO2—acetogenesis to 10.5% of the total. Although the polysaccharide components of lignocellulose are very efficiently digested during passage though the hindgut (74-99% of cellulose, 65-87% hemicellulose) (38), high-molecular-weight core lignin does not appear to be significantly degraded in wood-feeding termites (16, 32, 38, 44). There is, however, evidence for some modification of lignin during gut passage via demethylation of the aromatic moieties of lignin sidechains (38), and a number of organisms capable of degrading aromatic lignin monomers have been isolated from termite hindguts (20, 51, 92). The contribution of these lignin-consuming activities to termite nutrition is currently unknown, but is likely to be relatively minor in comparison to cellulolytic activities. Hindgut microbes also play important roles in the nitrogen economy of termites (90). Reduction of N2 to ammonia (NH3) via the enzyme nitrogenase is a property unique to prokaryotes, and any N2 fixation measured in termites (either by the acetylene reduction assay or incorporation of l5N2 isotope) is thus attributable to bacteria or archaea. N2 fixation has been detected and quantified in a large number of different termite species (91), and a number of N2-fixing bacteria have been isolated from termites (41, 52, 77). A highly diverse set of nifH genes (encoding dinitrogenase reductase) have been detected in the hindguts a number of termite species (71, 73), and direct amplification of nif genes from community mRN A allowed an elegant assessment of which phylogenetic classes of nifl-I homologues were being expressed in the gut of the termite Neotermes koshunensis (65). In addition to providing newly fixed nitrogen to termites, members of the hindgut community act to prevent the loss of fixed nitrogen by recycling the nitrogen present in the excretory product uric acid. Potrikus and Breznak (78-81) demonstrated that uric acid is transported to the hindgut via the Malpighian tubules and isolated a variety of bacteria capable of anaerobically degrading uric acid with the liberation of ammonia (which is presumably reabsorbed by the termite). Physiochemical Gradients in the Termite Hindgut Early studies of redox conditions existing within termite hindguts indicated an effectively anoxic environment (5, 6). While this result was in keeping with the presence of strict anaerobes in the hindgut, such as cellulolytic protozoa and methanoarchaea, a number of other observations suggested a more complex oxygen status. In cultivation studies of gut bacteria, a large number of the isolates were observed to be aerotolerant, facultatively aerobic (39, 87), or even strictly aerobic (92). Moreover, the ready mineralization of lignin monomers fed to termites was shown to be dependent on hindgut bacteria and the presence of O2, and a variety of bacteria isolated fiom hindguts were shown to be capable of O2-dependent in vitro degradation of aromatic monomers of lignin (20, 51, 92). An elegant series of studies by Brune and coworkers (19, 21, 37) addressed the issue of O2 in termite guts by making fine scale microelectrode measurements of the in situ physiochemical characteristics of termite guts. These studies revealed a highly structured environment with distinct radial gradients of H2 and 02 concentration (Figure 1.4). 02 partial pressures were highest near the gut epithelium and rapidly decreased to non-detectable levels within 20-200 pm of the periphery. Conversely, hydrogen partial pressure was very high (approx. 50 mbar, or 50,000 ppmv) in the central luminal portion of the gut, and decreased steeply towards the gut periphery. Two main zones of H2 consumption were observed; one in the center of the lumen, and a second directly at the gut epithelium. The hindguts of both lower and higher termites had pH values of around 7. These results provided convincing new evidence that as much as 60% of the 10 Partial pressure of gases (m bar) 60 l 80 i l J Depth (um) 1200 1500 ? Agarose Gut Figure 1.4. Radial profiles of oxygen (0) and hydrogen (0) in an agarose- embedded hindgut from Reticulitermes flavipes. The circular inset indicates the relative sizes of the oxic and anoxic zones in the gut (left half) and the hydrogen concentration gradient (right half). Adapted from reference 21. 11 gut volume is in fact hypoxic in character and also demonstrated that the luminal region of termite hindguts is among the most H2-rich environments observed in nature (62). Anaerobiosis in the gut lumen appears to be maintained largely by respiratory activity of gut bacteria, and elimination of these organisms with antibiotic treatments renders the gut fully oxic (95). This observation fits well with the epithelial zone of O2 consumption observed by Brune et al. (19) (Figure 1.4). High rates of O2-dependent acetate oxidation have been observed in the gut (93), suggesting that acetate produced in the central lumen may be one of the major substrates of O2-consuming epithelial biofilrn. While this acetate consumption would represent a loss of oxidizable substrate for the host, acetate oxidizers (and other O2-consuming bacteria) could be viewed as playing an essential symbiotic role in maintaining conditions of anoxia in the lumen, which favors the continued homoacetic fermentation of cellulose. Acetogenesis vs. Methanogenesis in Wood-feeding Termites A significant role for H2-consuming, CO2-reducing homoacetogens in termite hindguts was first suggested by the dominance of acetate in the hindgut VFA pool and the observation of relatively low rates of H2 and CH4 emission from termites (68). It was also observed that H2 and CH4 emission rates were significantly increased in termites fed antibacterial drugs (68). These observations suggested that H2 consumption in the hindgut was dominated by bacterial homoacetogens. This hypothesis was confirmed by the demonstration of H2-dependent fixation of 14CO2 to acetate by termite gut homogenates and the inhibition of this activity by antibacterial drugs (11). Brauman et al. (7) found that acetogens out-process methanogens as H2-consumers in the majority of 12 wood-feeding termite species examined (although the opposite was true in termites from other feeding guilds, such as the soil-feeders). This situation, while unusual, is highly beneficial to the termite due to the significant contribution of acetogens to termite nutrition (15, 93). The dominance of acetogenesis over methanogenesis in the termite hindgut is a puzzling situation. In anoxic habitats, various metabolic classes of organisms (sulfate- reducing bacteria, methanogens, H2/CO2-acetogens, etc) compete for reducing potential, usually in the form of H2 produced by ferrnentative organisms. For a given electron donor (in this case, H2) thermodynamics suggests that the group of organisms using the most exergonic terminal electron-accepting metabolic process (i.e. the process with the most positive redox potential, E°') should dominate other competitors, as implied by the equation: AG°' = -n F AE°' [n = number of e' transferred, F = Faraday constant, AE°' = E°' redox compound - E°' (H+/H2) ]. This correlates to progressively lower minimum thresholds values of H2 (33). The minimum H2 threshold is the H2 concentration at which an organism's metabolic reaction is still sufficiently exergonic to produce a useful amount of energy. This amount, referred to as the critical Gibbs free energy (AG), is the amount of energy required to translocate 1 mol of protons through a full energized cell membrane, which in turn correlates to the minimum energy quanta sufficient to drive the synthesis of ATP (approximately -23 kJ/mol substrate, or 1/3 of an ATP) (30, 31, 33). Thus, sulfate-reducing bacteria (average E°’= -217 mV) are able to lower the H2 concentration to a point at which methanogens (average E°’= -238) can no longer derive useful energy from their metabolism. Table 1.1 provides a list of Gibbs fiee energy 13 Table 1.1. Redox potentials, Gibbs free energy changes, and H2 thresholds for various anaerobic hydrogen consuming processes”. E°' b AG°'/reaction H2 Thresholdc Metabolic Process: (mV) (kJ) (ppmv) Acetogenesis: 4 H2 + 2 HCO;' + H+ -> CH3COO' + 4 H20 -279 -104.4 260 - 950 Methanogenesis: 4 H2 + HCO3' + H+ - CH4 + 3 H20 -238 -135.6 25 - 100 Sulfate Reduction: 4H2+SO.2'+H*-> HS'+4H2O -217 452.0 8-16 ‘Adapted from values presented in reference (33). l’E°' of the couples accepting electrons fiom H2 calculated relative to the redox potential 0sz (414 mV) cRange of H2 threshold values observed for representatives of each of the metabolic types grown in pure culture. 14 values, redox potentials, and H2 thresholds of some typical groups of hydrogen consuming anaerobes. This model of competition is referred to as the threshold model (29, 31, 86), and is generally in good agreement with actual H2 concentrations measured in natural habitats in which the dominant H2-consuming process was known (27, 49, 62, 63). Methanogens, as would be predicted, dominate as the terminal electron sink organisms in habitats in which CO2/I-ICO3' is the main terminal electron acceptor, including gastrointestinal systems such as bovine rumen (4, 13, 34, 64, 82). In the termite hindgut, CO2 appears to be the primary electron acceptor. Under these conditions, methanogens should out-process acetogens due to the more positive E°’ value of the CO2/CH4 redox couple, which equates to a more exergonic reaction and lower H2 threshold (see Table 1). This is certainly the case for most other gastrointestinal systems, including the bovine rumen (4, 13, 34, 64, 82). Figure 1.5 shows free energy values of acetogenesis and methanogenesis calculated for pH levels and concentrations of acetate, bicarbonate, and methane typical of the termite gut environment and are calculated across the range of radial hydrogen concentrations measured in the termite hindgut. While methanogenesis is clearly the more favorable reaction at all H2 concentrations, acetogens appear to effectively dominate H2 consumption in the termite hindgut. A variety of homoacetogenic bacteria have been isolated from several species of termites, including members of the genera Sporomusa, Acetonema, and Clostridium (12, 47, 48). These organisms, however, did not seem to colonize the hindgut at population densities sufficient to support measured rates of H2/CO2 acetogenesis (15) and molecular . phylogenetic studies of hindgut microbial communities of other termite species have never detected phylotypes closely related to the isolated termite acetogens (70, 72). 15 -120- -110- E .. g -100- o m .. 9 2 -90- 3 2 '80“ (>6 . (D -70- <1 . E 17, -60- s. 1 -50e -40- '30 l I I I l I I I I I 1o 20 30 4o 50 Hydrogen partial pressure (mbar) Figure 1.5. In situ Gibb's free energies for methanogenesis (El) and acetogenesis (0) under termite hindgut conditions. Values were calculated at 25°C for pH 7.2 and the following substrate and product concentrations: acetate, 80 mM; methane, 1 mM; bicarbonate, 60 mM. 16 Furthermore, the H2 thresholds of the isolated acetogens (251-380 ppmv) were significantly higher than those of the cultured termite methanogens (36-45 ppmv) (17). In coculture competitions for limited H2, the termite acetogen Sporomusa termitida was consistently dominated by the termite methanogen Methanobrevibacter cuticularis (17). Given these observations, it was considered doubtful that the cultured termite acetogens represented the dominant acetogens in the termite hindgut. It is unknown what factors in the termite gut favor acetogenesis over the more energetically favorable process of methanogenesis. There are other habitats in which acetogens dominate methanogens, including mildly acidic (76) or carbon limited (46) freshwater sediments. Aceto genesis is also favored over methanogenesis at low temperatures (28, 88, 89). None of these conditions, however, would appear to be relevant to the termite hindgut habitat. Mixotrophic growth by the simultaneous utilization of organic substrates and H2 has also been proposed to increase energy yields and competitive ability of acetogens (50, 60, 61), and mixotrophic growth has been demonstrated in the termite acetogen Sporomusa termitida (14). Hydrogen thresholds of acetogens growing mixotrophically are indeed lower than those measured under unitrophic growth, but are still significantly higher than those characteristic of methanogenesis (75), making the significance of mixotrophy in acetogen/methanogen competitions somewhat debatable. The Spatial Resource Partitioning Hypothesis The puzzling dominance of acetogenesis over methanogenesis in the termite hindgut has led to speculation that direct competition for H2 does not actually occur between the 17 two groups. This was first suggested by the localization of methanogens in the hindgut of the termite R. flavipes: F420 autofluorescent cells (a diagnostic characteristic of methanogens) were rarely observed in the luminal gut fluid and instead appeared to colonize the gut epithelium almost exclusively (53, 54). Considered in light of the physiochemical gradient data of Ebert and Brune (37) and Brune et al. (19), this observation suggests an alternative model of H2 consumption in termite guts referred to as the spatial resource partitioning hypothesis (22, 37). In this model, methanogens and acetogens occupy distinct physiochemical niches within the hindgut, precluding any direct competition for H2 (see Figure 1.6). Methanogens appear to be restricted to the region of lowest hydrogen concentration on the gut epithelium, where they may be additionally inhibited by inwardly diffirsing 02. If acetogens were to inhabit the central lumen, they would have access to hydrogen produced by gut protozoa at levels 100 to 1000 fold higher than their minimum thresholds (37). This would be consistent with the two regions of H2 consumption (one in central lumen, and one near the gut epithelium) observed in the lumen by Ebert and Brrme (37) (Figure 1.4). As promising as this hypothesis may seem, it is not without shortcomings. First, it offers no explanation for the counterintuitive restriction of methanogens to the gut wall, where substrate H2 is scarce and inhibitory O2 is abundant. Also, in freshly prepared gut homogenates (in which all microbial spatial associations have been effectively disrupted), H2-dependent acetogenesis from CO2 still consistently dominates methanogenesis (11). This would imply that other, as yet unknown, factors also contribute to the dominance of 18 Bacteria Cellulolytic Protozoans Figure 1.6. Diagram of the termite hindgut in radial cross-section, showing the spatial distributions of various microbial groups and gradients of 02 and H2. homoacetogens over methanogens as H2-consumers in the hindguts of wood-feeding termites. Isolation of Termite Gut Spirochetes In 1999, years of periodic but persistent cultivation efforts were finally rewarded with the first successful enrichment and isolation of spirochetes fi'orn the hindgut of the dampwood termite Zootermopsis angusticollis (55). A number of factors contributed to the success of these enrichments: the medium was low in fermentable carbon sources, was supplemented with rumen fluid, and was incubated under an anoxic H2/CO2 atmosphere. The enrichment medium was also supplemented with rifamycin and phosphomycin (antibiotics to which spirochetes are resistant) and bromoethanosulfonate (a selective inhibitor of methanogens). The development of this enrichment medium was the result of patient microscopic observation of differential spirochetal growth in a variety of exploratory media of varying constitutions over 2-3 months (owing to the slow growth of the termite gut spirochetes) (17, 18). Of the dozen initially isolated spirochete strains, two (strains ZAS-1 and ZAS-2) grew to sufficient densities to permit physiological studies (55). The two strains are similar in size (0.2 pm by 3 to 7 urns) and display the characteristic morphological features of spirochetes (i.e. undulate cell shape and periplasmic flagella) (Figure 1.7). The 168 rDNA sequences of ZAS-1 and ZAS-2 are 98% similar to each other and, like other termite gut spirochetes, group within the so-called "termite cluster" of the genus T reponema (Figure 1.8). The most similar l6S rDNA sequences among cultivated spirochetes were those of S. stenostrepta and S. caldaria (92-93% similarity). Both 20 Figure 1.7. Morphology of termite gut Treponema strain ZAS-1 by phase contrast (a) and transmission electron microscopy of intact (c) and transverse sectioned (b) cells. The morphology of strain ZAS-2 is virtually identical to that of ZAS-1. Arrows indicate periplasmic flagella, whose insertion points are subterminal (c). Scale bars = 10 pm (a) and 0.1 pm (b, c). 21 T Gut Clone RFS3 ' T Gut Clone RF825 ZAS-1 ZAS-2 T Gut Clone ZASBQ 'Tennite ZAS-9 Cluster" T Gut Clone NL-1 Spirochaeta stenostrepta Spirochaeta caldan'a Treponema pallidum Treponema denticola Spirochaeta zuelzerae I I Treponema succinifaciens Treponema bryantii —-I Treponema pectinovorum —I Spirochaeta halophila Spirochaeta aurantia Bomelia burgdorfen’ I Cn'stispira pectinis I I Leptonema illini , Leptospira biflexa I Brachyspira hyodysenten’ae I Brevinema andersonii I Escherichia coli 0.1 0 Figure 1.8. Phylogenetic tree inferred from 168 rDNA sequences of termite gut treponemes (strains ZAS-1, ZAS-2, and ZAS-9), representative known spirochetes, and spirochetal 16$ rDNA clones generated directly from termite gut contents (T gut clones). A maximum likelihood technique (fastDNAml) was used to generate the tree. The vertical line on the right delimits a distinct subgroup (the "termite cluster“) within the genus Treponema. Scale bar represents units of evolutionary distance and is based on sequence divergence. 22 strains required anoxic conditions and provision of yeast autolysate and an ll-cofactor mixture in the growth media (55). The most interesting property of ZAS-1 and ZAS-2, however, was their ability to carry out H2/CO2 acetogenesis. This capability was first suggested by the consumption of headspace H2 and CO2 in culture tubes with the coincident production of stoichiometric amounts of acetate as the sole end product. Subsequently, it was shown that pure cultures of strains ZAS-1 and ZAS-2 incorporated l4CO2 into both C atoms of acetate. To further test for the presence of the characteristic Wood/Ljungdahl pathway of acetogenesis (Figure 1.9) (36), both strains were successfully assayed for the presence of three characteristic enzyme activities (carbon monoxide dehydrogenase, formate dehydrogenase, and hydrogenase) of the pathway. While the possibility that termite hindgut spirochetes were acetogens had been hypothesized many years ago (8), the confirmation of this metabolic activity in ZAS-1 and ZAS-2 was still something of a surprise, as no other spirochete had ever been shown to be capable of H2/CO2 acetogenesis. Given their abundance in the hindgut microbial community, the observation of acetogenesis in the ZAS strains suggests that spirochetes may be the dominant acetogens in the termite hindgut. This hypothesis is supported by the primarily luminal localization of spirochetes, which places them among (or attached to) H2-producing protozoa and within the luminal zone of H2-consumption observed by Ebert et al. (37). This interpretation, however, should be viewed with a note of caution. It is presently unknown what percentage of spirochetes in the termite hindgut are actually acetogens and whether acetogenic spirochetes exist in termites other than Z. angusticollis. An 23 Methyl Branch Carbonyl Branch C02 002 Formate Dehydrogenase 21H] HCOOH Formyl-THF Synthetase THF’ ATP 2 [H] \I ADP HCO-THF CO Dehydrogenase Methenyl-THF Cyclohydrolase 4 [H] Methylene-THF Dehydrogenase Methylene-THF Reductase CH3-THF Methyltranserase CO Dehydrogenase v CH3-corrinoid [CO] CoA CH3-CO-S-CoA - - +cell carbon CoA Pi CH3-CO-P ADP ATP CH3COOH Figure 1.9. The Wood/Ljungdahl pathway of acetogenesis, highlighting enzymes of the methyl and carbonyl branches. THF=Tetrahydrofolate. Adapted from reference 36. 24 additional spirochete subsequently isolated from Z. angusticollis, T reponema strain ZAS- 9 (more thoroughly described in Chapter 3), is not capable of H2/CO2-acetogenesis. This observation suggests that the morphological and phylogenetic diversity observed in termite-associated spirochetes is likely paralleled by a similarly high degree of physiological diversity. In addition to their role in the provision of acetate to their hosts, termite spirochetes are likely to perform a variety of other functions in the termite hindgut. In keeping with this hypothesis, strain ZAS-9 was recently determined to be capable of N2 fixation (58), another property not previously observed in spirochetes. Strains ZAS-1, ZAS-2, and ZAS-9 each possess two distinct nrfl-I homologs, several of which were nearly identical to nifH sequences known to be expressed in termite hindguts (58, 65). Nitrogen fixation was unambiguously demonstrated in ZAS-9 by the acetylene reduction assay, by the fixation of 15N2 into cellular 15N, and by growth with N2 as the primary nitrogen source. ZAS-1 and ZAS-2, however, showed only low levels of nitrogenase activity and no enhancement of nitrogen limited growth by the provision of N2. These results strongly suggest that distinct populations of spirochetes make significant contributions to host nutrition both in terms of carbon (via acetogenesis) and nitrogen provision. Dissertation Research The research presented in this dissertation focuses on the physiological ecology of termite gut spirochetes. Chapter 2 presents a more thorough physiological characterization of 25 strains ZAS-l and ZAS-2, with special attention focused on those properties relevant to survival within the termite hindgut and their role as symbiotic homoacetogens. Chapter 3 provides further description of strain ZAS-9, explores the taxonomic and genomic properties of all three ZAS strains, and proposes Latinate species epithets. Chapter 4 examines the requirement of strains ZAS-1 and ZAS-2 for the vitamin folate (a critical cofactor in acetogenesis) and the role of other members of the hindgut microbiota in providing this factor in situ. Chapter 5 revisits the issue of acetogenesis vs. methanogenesis in termite hindguts, and tests an alternative to the spatial resource partitioning hypothesis. Finally, chapter 6 summarizes the main conclusions of these studies. 26 References Berchtold, M., W. Ludwig, and H. Konig. 1994. 16S rDNA sequence and phylogenetic position of an uncultivated spirochete fi'om the hindgut of the termite Mastotermes darwiniensis Froggatt. FEMS Microbiology Letters 123:269-273. Berchtold, M., and H. Konig. 1996. Phylogenetic analysis and in situ identification of uncultivated spirochetes from the hindgut of the termite Mastotermes darwiniensis. Syst Appl Microbiol 19:66-73. Bermudes, D., D. Chase, and L. Margulis. 1988. Morphology as a basis for taxonomy of large spirochetes symbiotic in wood-eating cockroaches and termites. Int J Syst Bacteriol 38:291-302. Bernalier, A., M. Lelait, V. Rochet, J.-P. Grivet, G. R. Gibson, and M. Durand. 1996. Acetogenesis from H2 and CO2 by methane and non-methane- producing human colonic bacterial communities. F ES Microbial. Ecol. 19:193- 202. Bignell, D. E. 1980. Determination of pH and oxygen status in the guts of lower and higher termites. J Insect Physiol 26: 183-188. Bignell, D. E. 1984. Direct potentiometric determination of redox potentials of the gut contents in the termites Zootermopsis nevadensis and Cubitermes severus and in three other arthropods. J Insect Physiol 30: 169-174. Brauman, A., M. D. Kane, M. Labat, and J. A. Breznak. 1992. Genesis of acetate and methane by gut bacteria of nutritionally diverse termites. Nature 257:1384-1387. Breznak, J. A. 1973. Biology of non-pathogenic, host-associated spirochetes. CRC Crit Rev Microbiol 2:457-489. Breznak, J. A., and H. S. Pankratz. 1977. In situ morphology of gut microbiota of wood-eating termites [Reticulitermesflavipes (Kollar) and Coptotermes formosanus Shiraki). Appl Environ Microbiol 33:406-426. 27 10. ll. 12. 13. 14. 15. 16. 17. 18. 19. 20. Breznak, J. A. 1984. Hindgut spirochetes of termites and Cryptocercus punculatus, p. 67-70. In N. R. Krieg and J. G. Holt (ed.), Bergey's Manual of Systematic Bacteriology, vol. 1. Williams & Witkins, Baltimore. Breznak, J. A., and J. M. Switzer. 1986. Acetate synthesis fi'om H2 plus CO2 by termite gut microbes. Appl Environ Microbi0152:623-630. Breznak, J. A., J. M. Switzer. and H. J. Seitz. 1988. Sporomusa termitida sp. nov., an H2/CO2-utilizing acetogen isolated from termites. Arch Microbiol 150:282-288. Breznak, J. A., and M. D. Kane. 1990. Microbial H2/CO2 acetogenesis in animal guts: nature and nutritional significance. FEMS Microbiol. Rev.:309-314. Breznak, J. A., and J. S. Blum. 1991. Mixotrophy in the termite gut acetogen, Sporomusa termitida. Arch Microbiol 156: 105-1 10. Breznak, J. A. 1994. Acetogenesis fiom carbon dioxide in termite guts. In H. L. Drake (ed.), Acetogenesis. Chapman & Hall, New York. Bro-alt, J. A., and A. Brune. 1994. Role of microorganisms in the digestion of lignocellulose by termites. Annu Rev Entomol 39:453-487. Breznak, J. A. 2000. Ecology of prokaryotic microbes in guts of wood- and litter- feeding termites, p. 209-231. In T. Abe, D. E. Bignell, and M. Higashi (ed.), Termites: Evolution, Sociality, Symbioses, Ecology. Kluwer Academic, Dordrecht/Norvvell, MA. Breznak, J. A., and J. R. Leadbetter. 2002. Termite Gut Spirochetes. In M. Dworkin (ed.), The Prokaryotes: an evolving electronic resource for the microbiological community. Springer-Verlag, New York. Brune, A., D. Emerson, and J. A. Bro-alt. 1995. The termite gut microflora as an oxygen sink: rrricroelectrode determination of oxygen and pH gradients in guts of lower and higher termites. Appl Environ Microbiol 61 :2681-2687. Brune, A., E. Miambi, and J. A. Breznak. 1995. Roles of oxygen and the intestinal microflora in the metabolism of lignin-derived phenylpropanoids and 28 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. other monoaromatic compounds by termites. Appl Environ Microbiol 61 :2688- 2695. Brune, A. 1998. Termite guts: the world's smallest bioreactors. Trends Biotechnol 16: 16-21 . Brune, A., and M. Friedrich. 2000. Microecology of the termite gut: structure and function on a microscale. Curr Opin Microbiol 3:263-269. Canale-Parola, E. 197 7. Physiology and evolution of Spirochetes. Bacteriol Rev 41:181-204. Canale-Parola, E. 1984. The Spirochetes, p. 38-70. In N. R. Krieg and J. G. Holt (ed.), Bergey's Manual of Systematic Bacteriology, 9th ed, vol. 1. Williams & Wilkins, Baltimore. Charon, N. W., E. P. Greenberg, M. B. H. Koopman, and R. J. Limberger. 1992. Spirochete Chemotaxis, Motility, and the Structure of the Spirochetal Periplasmic Flagella. Res Microbiol 143:597-603. Cleveland, L. R., and A. V. Grimstone. 1964. The fine structure of the flagellate Mixotricha paradoxa and its associated microorganisms. Procedings of the Royal Society of London Series B 159:668-86. Conrad, R., B. Schink, and T. J. Phelps. 1986. Thermodynamics of H2- consuming and H2 producing metabolic reactions in diverse methanogenic environments under in situ conditions. FEMS Microbial. Ecol. 38:353-360. Conrad, R., and B. Wetter. 1990. Influence of temperature on energetics of hydrogen metabolism in homoacetogenic, methanogenic, and other anaerobic bacteria. Arch Microbiol 155:94-98. Conrad, R. 1995. Soil microbial processes involved in production and consumption of atmospheric trace gases, p. 207-250, Advances in Microbial Ecology, Vol 14, vol. 14. Conrad, R. 1996. Soil microorganisms as controllers of atmospheric trace gases (H2, CO, CH4, OCS, N20, and NO). Microbiol Rev 60:609-640. 29 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. Conrad, R. 1999. Contribution of hydrogen to methane production and control of hydrogen concentrations in methanogenic soils and sediments. FEMS Microbiol Ecol 28:193-202. Cookson, L. J. 1987. I“C-Lignin degradation by three Australian termite species. Wood Sci.Technol. 21:11-25. Cord-Ruwisch, R., H. Seitz, and R. Conrad. 1988. The capacity of hydrogenotrophic anaerobic bacteria to compete for traces of hydrogen depends on the redox potential of the terminal electron acceptor. Arch. Microbiol. 149:350—7. De Graeve, K. G., J. P. Grivet, M. Durand, P. Beaumatin, C. Cordelet, G. Hannequart, and D. Demeyer. 1994. Competition between reductive acetogenesis and methanogenesis in the pig large-intestinal flora. J. Appl. Biotechnol. 76:55-61. Dobell, C. 1912. Research on the spirochetes and related organisms. Archiv firr protistenkunde 26:119-230. Drake, H. L. 1994. Acetogenesis, acetogenic bacteria, and the acetyl-CoA "Wood/Ljundahl" pathway: Past and current perspectives, p. 1-60. In H. L. Drake (ed.), Acetogenesis. Chapman and Hall, New York. Ebert, A., and A. Brune. 1997. Hydrogen concentration profiles at the oxic- anoxic interface: a microsensor study of the hingut of the wood-feeding lower termite Reticulitermesflavipes. Appl Environ Microbiol 63:4039-4046. Esenther, G. R., and T. K. Kirk. 1974. Catabolism of aspen sapwood in Reticulitermesflavipes. Ann Entomol Soc Am 67:989-991. Eutick, M. L., R. W. O'Brien, and M. Slaytor. 1978. Bacteria from the gus of Australian termites. Appl Environ Microbiol 35:823-828. Eutick, M. L., P. Veivers, R. W. O'Brien, and M. Slaytor. 1978. Dependence of the higher termite, Nasutitermes exitiosus and the lower termite, Coptotermes lacteus on their gut flora. J Insect Physiol 24:363-368. 30 41. 42. 43. 45. 46. 47. 48. 49. 50. French, J. R. J., G. L. Turner, and J. F. Bradbury. 1976. Nitrogen fixation by bacteria from the hindguts of termites. J Gen Microbiol 95:202-206. Hogan, M. E., M. Slaytor, and R. W. Obrien. 1985. Transport of volatile fatty acids across the hindgut of the cockroach Panesthia cribrata Saussure and the termite, Mastotermes darwiniensis Froggatt. J Insect Physiol 31:587-591. Holt, S. C. 1978. Anatomy and chemistry of spirochetes. Microbiology Reviews 42:114-160. Hyodo, F., J. Azuma, and T. Abe. 1999. Estimation of the effect of passage through the gut of a lower termite, Coptotermesformosanus Shiraki, on lignin by solid-state CP/MAS 13C NMR. Holzforschung 53:244-246. Inoue, T., K. Murashima, J. I. Azuma, A. Sugimoto, and M. Slaytor. 1997. Cellulose and xylan utilisation in the lower termite Reticulitermes speratus. J Insect Physiol 43:235-242. Jones, J. G., and B. M. Simon. 1985. Interaction of acetogens and methanogens in anaerobic freshwater sediments. AEM 49:944-948. Kane, M. D., A. Brauman, and J. A. Bro-alt. 1991. Clostridium mayombei Sp- Nov, an H2/CO2 acetogenic bacterium fi’om the gut of the African soil-feeding termite, Cubitermes speciosus. Arch Microbiol 156:99-104. Kane, M. D., and J. A. Breznak. 1991. Acetonema longum Gen-Nov Sp-Nov, an H2/CO2 acetogenic bacterium from the termite, Pterotermes occidentis. Arch Microbiol 156:91-98. Kramer, H., and R. Conrad. 1993. Measurement of dissolved H2 concentrations in methanogenic environments with a gas-diffusion probe. FEMS Microbiol Ecol 12:149-158. Krumholz, L. R., S. H. Harris, S. T. Tay, and J. M. Sulflita. 1999. Characterization of two subsurface H2-utilizing bacteria, Desulfomicrobium hypogeium sp. nov. and Acetobacterium psammolithicum sp. nov., and their ecological roles. Appl Environ Microbiol 65:2300-2306. 31 51. 52. 53. S4. 55. 56. 57. 58. 59. 60. Kuhnigk, T., E.-M. Borst, A. Ritter, P. Kampfer, A. Graf, H. Hertel, and H. Konig. 1994. Degradation of lignin monomers by the hingut flora of xylophagous termites. Syst Appl Microbiol 17:76-85. Kuhnigk, T., J. Branke, D. Krekeler, H. Cypionka, and H. Konig. 1996. A feasible role of sulfate-reducing bacteria in the termite gut. Syst Appl Microbiol 19:139-149. Leadbetter, J. R., and J. A. Breznak. 1996. Physiological ecology of Methanobrevibacter cuticularis sp. nov. and Methanobrevibacter curvatus sp. nov., isolated from the hindgut of the termite Reticulitermesflavipes. Appl Environ Microbiol 62:3620-3631. Leadbetter, J. R., L. D. Crosby, and J. A. Breznak. 1998. Methanobrevibacter filiformis sp. nov., a filamentous methanogen fi'om termite hindguts. Arch Microbiol 169:287-292. Leadbetter, J. R., T. M. Schmidt, J. R. Graber, and J. A. Bro-alt. 1999. Acetogenesis fi'om H2 plus CO2 by spirochetes from termite guts. Science 283:686-689. Leidy, J. 1874-1881. The parasites of the termites. J. Acad. Nat. Sci. (Phil) 8:425-447. Leschine, S., B. J. Faster, and E. Canale-Parola 2002, posting date. Free- Living Saccharolytic Spirochetes: The Genus Spirochaeta. The prokaryotes: an evolving electronic resource for the microbiological community. Dworkin, M. (ed.). [Online.] Lilburn, T. C., K. S. Kim, N. E. Ostrom, K. R. Byzek, J. R. Leadbetter, and J. A. Breznak. 2001. Nitrogen fixation by symbiotic and fi'ee-living spirochetes. Science 292:2495-2498. Lilburn, T. G., T. M. Schmidt, and J. A. Breznak. 1999. The phylogenetic diversity of termite gut spirochetes. Environ. Microbiol. 1:331-345. Liu, S., and J. M. Sulflita. 1993. H2-CO2 dependent anaerobic O-demethylation activity in subsurface sediments and by an isolated bacterium. Appl Environ Microbiol 59:1325-1331. 32 61. 62. 63. 64. 65. 66. 67. 68. 69. Liu, S., and J. M. Sulflita. 1995. H2 as an energy source for mixotrophic acetogenesis by the reduction of CO2 and Syringate by Acetobacterium woodii and Eubacterium limosum. Curr Microbiol 31 :245-250. Lovely, D. R., and S. Goodwin. 1988. Hydrogen concentration as an indicator of the predominant terminal electron-accepting process in aquatic sediments. Geochimicaet Cosmochimica Acta 52:2993-3003. Lovley, D. R., D. F. Dwyer, and M. J. King. 1982. Kinetic analysis of competition between sulfate reducers and methanogens for hydrogen in sediments. Appl Environ Microbiol 43:1373-1379. Morvan, B., F. Bonnemoy, G. Fonty, and P. Gouet. 1996. Quantitative determination of H2-utilizing acetogenic and sulfate reducing bacteria and methanogenic archaea fiom the digestive tract of different mammals. Curr. Microbiol. 32:129-133. Noda, S., M. Ohkuma, R. Usami, K. Horikoshi, and T. Kudo. 1999. Culture- independent characterization of a gene responsible for nitrogen fixation in the symbiotic microbial community in the gut of the termite Neotermes koshunensis. Appl Environ Microbiol 65:4935-4942. Noda, S., M. Ohkuma, A. Yamada, Y. Hongoh, and T. Kudo. 2003. Phylogenetic position and in situ identification of ectosymbiotic spirochetes on protists in the termite gut. Appl Environ Microbiol 69:625-633. Norris, S. J., B. J. Paster, A. Moter, and U. B. Gobe12002, posting date. The Genus Treponema. The prokaryotes: an evolving electronic resource for the microbiological community. Dworkin, M (ed.). [Online.] Odelson, D. A., and J. A. Breznak. 1983. Volatile fatty acid production by the hindgut microbiota of xylophagous termites. Appl Environ Microbiol 45: 1602- 161 3. Odelson, D. A., and J. A. Breznak. 1985. Nutrition and growth characteristics of Trichomitopsis termopsidis, a cellulolytic protozoan from termites. Appl Environ Microbiol 49:614-621. 33 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. Ohknma, M., and T. Kudo. 1996. Phylogenetic diversity of the intestinal bacterial community in the termite Reticulitermes speratus. Appl Environ Microbiol 62:461-468. Ohkuma, M., S. Noda, R. Usami, K. Horikoshi, and T. Kudo. 1996. Diversity of nitrogen fixation genes in the symbiotic intestinal microflora of the termite Reticulitermes speratus. Appl Environ Microbiol 62:2747-2752. Ohkuma, M., and T. Kudo. 1998. Phylogenetic analysis of the symbiotic intestinal microflora of the termite Cryptotermes domesticus. F ems Microbiology Letters 164:389-395. Ohkuma, M., S. Noda, and T. Kudo. 1999. Phylogenetic diversity in the nitrogen fixation genes in the symbiotic microbial community in the guts of diverse termites. Appl Environ Microbiol 65:4926—4934. Paster, B. J., F. E. Dewhirst, S. M. Cooke, V. Fussing, L. K. Poulsen, and J. A. Breznak. 1996. Phylogeny of not-yet-cultured spirochetes from termite guts. Appl Environ Microbiol 62:347-352. Peters, V., P. H. J anssen, and R. Conrad. 1998. Efficiency of hydrogen utilization during unitrophic and mixotrophic growth of Acetobacterium woodii on hydrogen and lactate in the chemostat. FEMS Microbiol Ecol 26:317-324. Phelps, T. J., and J. G. Zeikus. 1984. Influence of pH on terminal carbon metabolism in anoxic sediments from a mildly acidic lake. Appl Environ Microbiol 48: 1088-1095. Potrikus, C. J., and J. A. Breznak. 1977. Nitrogen-fixing Enterobacter aglomerans isolated fi‘om guts of wood-eating termites. Appl Environ Microbiol 33:392-399. Potrikus, C. J., and J. A. Breznak. 1980. Uric acid degrading bacteria in guts of termites [Reticulitermesflavipes (Kollar)]. Appl Environ Microbiol 40:117-124. Potrikus, C. J., and J. A. Breznak. 1980. Anaerobic degradation of uric acid by gut bacteria of termites. Appl Environ Microbiol 40: 125-132. 34 80. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. Potrikus, C. J., and J. A. Breznak. 1980. Uric acid in wood-eating termites. Insect Biochemistry 10:19-&. Potrikus, C. J., and J. A. Breznak. 1981. Gut bacteria recycle uric acid nitrogen in termites - a strategy for nutrient conservation. Proc Natl Acad Sci U S A 78:4601-4605. Prins, R. A., and A. Lankhorst. 1977. Sythesis of acetate from CO2 in the cecum of some rodents. FEMS Microbiol. Letters 1:255-258. Radek, R., K. Hausmann, and A. Breunig. 1992. Ectobiotic and endocytobiotic bacteria associated with the termite flagellate Joenia annectens. Acta Protozoologica 31 :93-107. Radek, R., J. Rosel, and K. Hausmann. 1996. Light and electron microscopic study of the bacterial adhesion to termite flagellates applying lectin cytochemistry. Protoplasma 193:105-122. Rosel, J., R. Radek, and K. Hausmann. 1996. Ultrastructure of the trichomonad flagellate Stephanonympha nelumbium. J Eukaryot Microbiol 43:505-511. Schink, B. 1997 . Energetics of syntrophic cooperation in methanogenic degradation. Microbiol. Molec. Bio. Rev. 61:262-280. Schultz, J. E., and J. A. Breznak. 1978. Heterotrophic bacteria present in hindguts of wood-eating termites [Reticulitermesflavipes (Kollar)]. Appl Environ Microbiol 35:930-936. Schulz, S., and R. Conrad. 1996. Influence of temperature on pathways to methane production in the permanently cold profundal sediment of Lake Constance. FEMS Microbiol Ecol 20: 1-14. Schulz, S., H. Matsuyama, and R. Conrad. 1997. Temperature dependence of methane production from different precursors in a profundal sediment (Lake Constance). FEMS Microbiol Ecol 22:207-213. Slaytor, M., and D. J. Chappell. 1994. Nitrogen metabolism in termites. Comp Biochem Physiol 107:1-10. 35 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. Tayasu, I., A. Sugimoto, E. Wada, and T. Abe. 1994. Xylophagous termites depending on atmospheric nitrogen. Naturwissenschafien 81:229-231. Tholen, A., B. Schink, and A. Brune. 1997. The gut microflora of Reticulitermes flavipes, its relation to oxygen, and evidence for oxygen-dependent acetogenesis by the most abundant Enterococcus sp. FEMS Microbiol Ecol 24: 137-149. Tholen, A., and A. Brune. 2000. Impact of oxygen on metabolic fluxes and in situ rates of reductive acetogenesis in the hindgut of the wood-feeding termite Reticulitermesflavipes. Environ. Microbiol. 2:436-449. To, L. P., L. Margulis, D. Chase, and W. L. Nutting. 1980. The symbiotic community of the Sonoran desert termite: Pterotermes occidentis. Biosystems 13:109-137. Veivers, P. C., R. W. Obrien, and M. Slaytor. 1982. Role of bacteria in maintaining the redox potential in the hindgut of termites and preventing entry of foreign bacteria. J Insect Physiol 28:947-951. Watanabe, H., H. N akamura, G. Tokuda, I. Yamaoka, A. M. Scrivener, and H. Noda. 1997. Site of secretion and properties of the endogenous endobeta-1,4- glucanase components from Reticulitermes speratus (Kolbe), a Japanese subterranean termite. Insect Biochem Mol Biol 27 :305-313. Watanabe, H., H. Noda, G. Tokuda, and N. Lo. 1998. A cellulase gene of termite origin. Nature 394:330-331. Yamin, M. A. 1978. Axenic cultivation of the cellulolytic flagellate T richomitopsis termopsidis (Cleveland) from the termite Zootermopsis. J Protozool 25:535-538. Yamin, M. A. 1980. Cellulose metabolism by the termite flagellate T richomitopsis termopsidis. Appl Environ Microbiol 39:859-863. Yamin, M. A. 1981. Cellulose metabolism by the flagellate Trichonympha from a termite is independent of endosymbiotic bacteria. Science 21 1:58-59. 36 Chapter 2 Physiology and Nutrition of Treponema strains ZAS-1 and ZAS-2, H2/CO2-Acetogenic Spirochetes from Termite Hindguts Introduction Spirochetes are among the most abundant microbial groups in termite hindguts, accounting for up to one-half of the prokaryotic community (43). For more than a century, however, our knowledge of these organisms was largely limited to sporadic reports of their presence in various termite species, their morphological diversity, and their physical association with termite gut protozoa (8). Although elimination of spirochetes from the termite gut led to a decrease in termite survivorship (21), specific roles of spirochetes and factors contributing to their abundance in the hindgut have remained obscure. Over the past ten years, our understanding of termite hindgut spirochetes has advanced dramatically. Cultivation-independent molecular approaches revealed that they group within the genus T reponema and that the large majority of 16S rDNA clones form a phylogenetically discrete cluster [the “termite c1uster”] within this genus (35). These studies also revealed a striking degree of phylogenetic diversity amongst the termite gut treponemes, with as many as 21 distinct species occurring within a single termite host species (35). A few years ago, the first pure cultures of these organisms were isolated in our laboratory and were found to possess metabolic capabilities hitherto unknown in the Spirochaetes division of the Bacteria, including acetogenesis fi'om H2 + CO2 (32) and N2 37 fixation (34). Both of these processes are unique to prokaryotes, and have been demonstrated to be important in the provision of carbon, nitrogen, and energy to termites (6, 47). Acetogenesis plays a particularly prominent role in termite metabolism: 71- 100% of the host’s energy requirements can be met by oxidation of acetate produced by hindgut microbes, and it has been estimated that from 10.5% to 33% of this acetate production is attributable to H2/CO2 acetogenesis (6, 42, 52). The availability of pure cultures of termite gut spirochetes has allowed the exploration of properties relevant to their growth and survival in situ. In this paper, we report on the nutritional, physiological and biochemical properties of Treponema strains ZAS-1 and ZAS-2, H2/CO2-acetogenic spirochetes isolated from hindguts of the California dampwood termite, Zootermopsis angusticollis (Hagen) (32). Additional information regarding the taxonomy, nomenclature, and genomic properties of these strains is included in chapter 3. Material & Methods Media and cultivation methods Routine cultivation of strains ZAS-1 and ZAS-2 was in butyl rubber stoppered tubes or bottles containing ca. one-filth their volume of medium 2YACo (32) under an atmosphere of 80% H2 / 20% CO2 (v/v). Medium 2YACo consisted of a mixture of inorganic salts, vitamins, cofactors, and 2% (v/v) yeast autolysate (equiv. to ca. 2.2 mg dry solids/m1). The medium was buffered by inclusion of 70 mM NaHCO3 and 10mM 3- N-[morpholino] propanesulfonic acid (MOPS) and reduced with dithiothreitol (DTT) (1 mM final conc.). The pH of the medium prior to inoculation was 7 .2. Unless otherwise 38 noted, cultures were grown at 30°C on a reciprocal shaker (50 oscillations per min.) with vessels held in a horizontal position. Nutritional and growth studies The ability of commercial yeast extracts to replace yeast autolysate in 2YACo medium was tested by using the following products at 6 mg/ml final concentration: Tastone nos. 900, 154, and 310; Amberex nos.1003 and 695; and Amberferm nos. 5925, 5902, and 5021 (Red Star Bioproducts, Juneau, Wisconsin). Cofactor requirements were evaluated by testing for an attenuation of growth in 2YACo medium lacking one of the eleven cofactors present in the cofactor stock solution (32). Candidate compounds were then tested in reciprocal experiments in which they alone were incorporated into medium 2YACo instead of the cofactor mixture. Substrate utilization studies were performed with cells growing under an atmosphere of N2/CO2 (80/20, v/v). For ZAS-1, medium 2YACo was modified to ' contain Amberferm 5902 (4 mg/ml) in place of yeast autolysate. Unmodified medium 2YACo was used for ZAS-2, with a small amount H2 (16 mM) added to the N2/CO2 headspace to aid in the initiation of growth (see below). H2 was not added to cultures grown on methoxylated aromatics. An increase in cell yield (>20%) in the presence of a substrate, as compared to its absence, was taken to indicate utilization of the substrate as an energy source. Cell growth was determined by measuring the optical density (OD) of cultures at 600 nm with a Milton Roy Spectronic 20 colorimeter. OD readings were converted to cell number by reference to a standard curve relating these quantities. Substrate carbon balances were determined under conditions of substrate-limited growth. 39 Carbon recoveries for methoxylated aromatic compounds were calculated based on acetate production expected fiom demethylation of the aromatic (R) substrate according to the following equation: R(-OCH3)n + 0.5n C02 -> R (-OH)n + 0.75n CH3COOH + 0.5n H20 Organic acid production was determined by using a high performance liquid chromatograph (HPLC) with refractive index detection (4). Aromatic compounds were analyzed using a Beckrnan model 127 HPLC equipped with a model 168 photodiode array detector and an Alltech Lichrosorb RP-18 column (250 x 4.6 mm, 10 um particle size). The mobile phase was 0.1% phosphoric acid with a methanol gradient, linearly increasing from 48 to 55% in 30 minutes. The flow rate was 1.5 mein. To test for mixotrophic growth, strain ZAS-2 was grown in 750 ml bottles containing 100 ml 2YACo media with 2 mM maltose and a 650 ml headspace composed of 20% H2, 20% CO2, 60% N2 (vol/vol). Consumption of H2 was followed by gas chromatography (5). Maltose consumption and organic acid production were followed by using the anthrone assay (1) and HPLC analysis (above), respectively. Determination of hydrogen thresholds Hydrogen thresholds were determined as described by Lovley (37). In brief, cultures were grown under H2/C02 (80/20, v/v) in medium 2YACo unmodified (ZAS-2) or modified to contain 4 mg/ml Amberferm 5902 in place of yeast autolysate (ZAS-1). When cultures reached mid-log phase (at which point further growth of both strains was strictly dependent on the presence of H2), the headspace was replaced with N2/CO2 (80:20, v/v), followed by the introduction of a small amount of H2 (ca. 6000 ppmv). The 40 basal level to which this H2 was consumed (i.e. the H2 threshold) was determined through three cycles of H2 addition and consumption for each culture. H2 was monitored by using a Trace Analytical RGA2 gas chromatograph equipped with a RGD2 trace gas detection unit. Enzyme Assays Cells from mid-log phase cultures (ODooo values of 0.15 to 0.3) were harvested by centrifugation (16,000 x g for 10 min) and resuspended at 10x their original concentration in appropriate assay buffer (as cited below) containing 10 mM ethylenediaminetetraacetic acid (EDTA), 1 mM phenylrnethanesulfonyl fluoride (PMSF), and leupeptin (10 rig/ml) to inhibit proteases. While held at 4°C, cells were disrupted by sonication 3 times for 30 s each with a Branson Model 450 sonifier (power setting of 5; 50% duty cycle) equipped with a stepped micro-tip. The resulting crude cell extracts were assayed for enzyme activities. Formyltetrahydrofolate (formyl-THF) synthetase, methenyl-THF cyclohydrolase and methylene-THF dehydrogenase (41) and methylene-THF reductase (39) were assayed as described. THF and THF derivatives used in the assays were obtained from Schircks Laboratories (J ona, Switzerland). Catalase was assayed by measuring the rate of decrease in the A240 of H202 (2). Oxidase and peroxidase activities coupled to the oxidation of NADH or NADPH were assayed as described by Stanton (50). Superoxide dismutase was assayed by the xanthine/xanthine oxidase-cytochrome c reduction method (22). Protein content of cell extracts was measured by the Lowry assay (38). 41 Absorbance measurements were made using a Perkin-Elmer Lambda 14 UVN IS spectrophotometer. Oxygen Tolerance Cells growing under anoxic conditions were tested for the ability to maintain growth after the addition of various concentrations of O2 to the headspace. Replicate cultures (n = 3) were grown under H2/CO2 with oscillation (above) in 18 mm anaerobe tubes containing Sml of medium 2YACo modified by the inclusion of 10 mM maltose, but containing no D'I'I‘ reducing agent. When cells reached mid-log phase, the headspace was balanced to atmospheric pressure with H2/CO2, and sterile 02 was injected to a final headspace concentration of 0.5%, 1%, 2.5%, or 5% (v/v). Cultures were then reincubated, and firrther growth was monitored as described above. Results Nutrition and growth of Treponema strains ZAS-1 and ZAS-2 Cells of both strains grew in medium 2YACo under H2+CO2 within an initial pH range of 6.5 to 7.8, with an optimum at pH 7.2. No growth was observed in media with an initial pH 56.0 or 28.0. Both strains grew within a temperature range of 23°C to 32°C, with an optimum at 30°C. No growth occurred at 4°C, or 34°C. Under optimum conditions (H2/CO2 + organic substrates), the shortest doubling time of cells was 24 hrs (ZAS-1) and 29 hrs (ZAS-2). 42 A variety of commercial yeast extracts (6 mg/ml final concentration) were tested for their ability to replace yeast autolysate in medium 2YACo. Tastone 900, Amberferm 5925, and Amberferm 5902 could replace yeast autolysate for growth of strain ZAS-1, whereas Tastone 154, Tastone 310, Amberex 1003, and Amberex 695 were suitable replacements for growth of strain ZAS-2. No single product was effective for both organisms. Growth rates and cell yields of ZAS-1 and ZAS-2 in media containing commercial yeast extracts were similar to those measured in media 2YACo. Besides H2 + C02, a variety of hexoses, pentoses, and disaccharides were also used as energy sources by both strains and were fermented homoacetogenically (Table 1). Curiously, however, when grown on organic substrates under N2/CO2, strain ZAS-2 displayed prolonged lag phases (272 h) prior to the initiation of growth. The provision of small amounts of H2 (16 mM) or the use of larger inocula (> 5% v/v) eliminated such lags (Figure 2.1). Strain ZAS-2 was additionally able to utilize methoxylated aromatic compounds (syringate, ferulate, vanillate, and trimethoxybenzoate) as energy sources when supplied at 52.5 mM (higher concentrations inhibited growth). Cell doubling times were typically greater than 100 hours on these substrates, and acetate production was consistent with demethylation of the compounds (i.e. the aromatic ring did not appear to be cleaved) (Table 1). This was confirmed for trimethoxybenzoate, which was quantitatively converted to gallic acid and acetate. Neither strain was capable of growth on other C1 compounds or methyl group donors tested (methanol, formats, C0, betaine, or choline). Relatively low concentrations of CO were inhibitory to both strains; addition of 1% CO (v/v) to the headspace of actively growing cultures resulted in the immediate cessation of growth. 43 5.80508 005 28555 050880 05 .8 850350805 805 508095 850388 058000 8 50009. .NOQNE 05 88000 00 b08000 88020 50>C0T£ 8 .2308 00 505800.. .858 88555 .8 0w8=00.~0mo .2833 505 05 wet—00— 080508 00$ 8:508 oo<>m 505508 8 850805 08500 85 538m .8.“ 50500.80 05 85 538w 0088358580 .8 085580 .588 5088.555 053 0050800.. 858 85 020$... 8888800 .8.“ 05: 5080008 8 85 G3 00:05.80 8 503838 .3888 00? «Comm 8 0885 53 .8 530..» 8w 8.5 .8850 .555 .3 508:8 8: 0.53 088:0 580 .0883 .080bw .05>:.Sa .0880— .00 .0885 85508 .08an 8.: .3 508058 0.8 508:8 8: :5 50505 008835 .948 on: 5 85 G48 8 00 .983 5880800 050805 50538508 05 05 8085 28m .8 85880080 85 a 5 50288 0.53 0008038 :5 m2 - v.80 - .20.? .8 228; m _ m - .38 - .20.: .8 2238 8 - exam - .20.? .3. 85:80 5 - .88 - .2: ._ .8 3880885085 - R - $8 .3. .20.? 828:8 8 R :3 $3 .20.: .03.... 8932 2: 3 saw as .25; .052 as? - 0N - 50 .3. .2me 08:58... 2 mm as $3 .20.: .250 3882 K mm :3 $3 .20.? .2085 88:8 we 3 .80 :8 page .208 .09.: «-08 705 ”-98 708 ~08 708 .8825 35%: mun—uh. Mam—£509 Sewn—€250”— 83oo< .— na\m=oo “:3? ~02 N-m20%) in the presence of a substrate, as compared to its absence, was taken to indicate utilization of the substrate as an energy source. Cell growth was determined by measuring the optical density (OD) of cultures at 600 nm on a Milton Roy Spectronic 20 calorimeter. OD readings were converted to cell number by reference to a standard curve relating these quantities. Substrate-product carbon balances were determined under conditions of substrate-limited growth. Organic acid production was determined by using high performance liquid chromatograph (HPLC) with refractive index detection (2). Ethanol production was enzymatically measured (ethanol assay kit 332-A; Sigrna- Aldrich, St. Louis, Missouri). Production of H2 was determined by gas chromatography (3). 67 Genomic G-l-C Content Determination The guanosine plus cytosine (G+C) content of genomic DNA was calculated fi'om the apparent ratios of deoxyguanosine and deoxythymidine (dGuo/dThd), which were determined by using the HPLC method of Mesbah et al. (6), with the following modifications: genomic DNA was harvested from mid-log phase cultures by using a Qiagen genomic DNA isolation kit with a 500/G purification column (Qiagen Inc., Chatsworth, California). A Shimadzu HPLC equipped with a model SPD-lOA UV/V IS detector (Shimadzu Corp., Kyoto, Japan) and an Alltech Alltirna C18 column (250 x 4.6 mm, 5 pm particle size) (Alltech Assoc., Deerfield, Illinois) were used. Data was analyzed by using the Shimadzu 82 Start (v. 7.1) software package. Results Morphology of Treponema strains ZAS-1, ZAS-2, and ZAS-9 The morphology of strains ZAS-9 and ZAS-2 is depicted in Figure 3.1. Cells of ZAS-2 were 3 to 7 pm in length and had a wavelength of 2.3 :l: 0.2 um (11 = 15). Strain ZAS-1 was virtually indistinguishable fi'am ZAS-2 on a morphological basis. In contrast, cells of ZAS-9 were longer (10 to 12 pm in length by 0.2 to 0.3 pm in width) than cells of ZAS-1 and ZAS-2 and displayed a shorter wavelength [1.1 :l: 0.1 pm (n = 15)], giving the cells a more tightly coiled appearance. 68 Figure 3.1. Morphology of Treponema strains ZAS-2 (A & B) and ZAS-9 (C and D) by phase contrast (A & C) and transmission electron microscopy (B & D). Insets in panels A and C show cells of each strain, with wavelengths indicated by arrowed bars. Periplasmic flagella are indicated by arrows in panels B & D. Strain ZAS-1 is morphologically indistinguishable from strain ZAS-2. Bars. 5 um (A and C), 2.5 pm (A and C insets), and 0.1 pm (B & D). Panel B is adapted from reference 4. and is presented here for comparison. 69 Nutrition, Growth, and Fermentation Products of strain ZAS-9 Energy sources used by ZAS-9 for growth are presented in Table 3.1. Carbohydrates were the only growth-supportive substrates; no growth occurred in the presence of organic acids or the amino acid glycine. ZAS-9 was also incapable of using H2 (+ CO2) as an energy sources, and no enhancement of growth (in terms of rate or cell yield) was observed when H2 was provided along with a fermentable carbohydrate (data not shown). ZAS-9 did not grow in medium 2YACo in the absence of added growth substrates. Major products of maltose fermentation by ZAS-9 were acetate, adrenal, H2, and CO2 (Table 3.2). The carbon and electron recoveries (90% and 88%, respectively), and the lack of any other detectable products in HPLC profiles, suggested that these were the only fermentation products of strain ZAS-9. G+C Content of DNA Genomic G+C content of the three ZAS strains were similar: ZAS-1, 51.04 :1: 0.09 mol%; ZAS-2, 50.87 :t 0.17 mol%; ZAS-9, 50.02 i 0.11 mol% (n=6 for each strain). G+C content was also measured for the DNA of the free-living spirochete Spirochaeta aurantia J 1 (65.60 :t 0.05 mol%) for comparison. As a control, the G+ C content was determined for Myxococcus xanthus DK1622, a strain previously assayed by the same method (6); the value determined here (67.41 :t 0.07 mol%) was in good agreement with that reported in the previous study (67.55 i 0.02 mol %). For all strains examined, peaks corresponding to the four unmodified deoxynucleosides (dGua, dThy, dAde, and dCyt) were the only peaks observed in HPLC chromatograms. 7O Table 3.1. Substrates Used as Energy Sources by Treponema strain ZAS-9 Substratea Doubling Time (h) Yield (cells/ml) Glucose 35 3.1x10“ Fructose 37 2.6x108 Ribose 42 1.4x108 Xylose 35 2.8x108 Maltose 47 5.6x108 Cellobiase 45 4.8x108 “All substrates were provided at a final concentration of 5 mM except for H2 (see methods). H2, mannitol, arabinose, sucrose, trehalose, glycine, lactate, pyruvate, and uric acid were not utilized. No growth was observed in medium 2YACo in the absence of added substrates. 71 Table 3.2. Fermentation Products of T reponema strain ZAS-9 Product mmol per 100 mmol maltose“ Acetate 280 Ethanol 80 H2 520 C02 360” “Carbon recovery = 90%; electron recovery = 88%. bAssumed to be equal to the sum of acetate plus ethanol 72 Discussion As reported by Leadbetter et al. (4) and in Chapter 2, ZAS-1 and ZAS-2 are both homoacetogens capable of growth on H2 + CO2 as well as a variety of organic compounds. The only significant differences observed between the two strains have been the ability of ZAS-2 to grow on methoxylated aromatic compounds (albeit very slowly) and its apparent reliance on H2 as a growth stimulant. ZAS-1 and ZAS-2 are otherwise highly similar to each other in terms of morphology, physiology, 16S rRNA sequence, and genomic G+C content. Given these results, it seems reasonable to conclude that ZAS-1 and ZAS-2 represent two strains of a single species. By contrast, strain ZAS-9 differs substantially from ZAS-1 and ZAS-2 in a number of taxonomically relevant properties, including cell morphology and its capacity for nitrogen fixation and N2-dependant growth. Unlike ZAS-1 and ZAS-2, ZAS-9 is not a homoacetogen, and does not consume H2 during growth. In fact, H2 is a major product of fermentation by ZAS-9. ZAS-9’s non-homoacetogenic nature of ZAS-9 is also supported by the fact that cells possess a hydrogenase activity (1.15 U/mg protein), but no formate dehydrogenase or CO dehydrogenase (J. A. Breznak, personal communication). The demonstration of H2 evolution by ZAS-9 and H2/C02-acetogenesis by ZAS-1 and ZAS-2 suggests that interspecies H2 transfer between spirochetes may be an important component of H2 turnover in the termite hindgut. In addition to the genomic G+C content data reported here, other genomic properties of the ZAS strains have been the subject of studies by other members of our laboratory. As with G+C content, genome sizes determined by pulsed field gel electrophoresis of macrorestricted genomic DNA were fairly similar between the three 73 strains: ZAS-1, 3.46 Mb; ZAS-2, 3.84 Mb; ZAS-9, 3.9 Mb (Brendan Keough & Kwi Kim, personal communication). Each strain also possessed two copies of the small subunit rRNA-encoding gene (Kwi Kim, personal communication). The S. aurantia J1 genomic G + C content determined here (65.6 mol%, as measured by direct HPLC analysis of deoxynucleosides) was in relatively good agreement with value previously reported by Breznak and Canale-Parola (1) for this organism as measured by the equilibrium density centrifugation (pCsCl) method (64.5 mol%). Both values, however, were significantly higher than the G+C content inferred by the thermal denaturation (Tm) method (60.4 mol%) in the same study, in which a discrepancy of approximately 4 to 6 mol% G+C content between the pCsCl and Tm methods was also reported for the spirochetes S. stenostrepta and S. Iitoralis (1). While the basis for this discrepancy is unlmown, the results of this study suggest that genomic G+C contents for free-living spirochetes are more accurately measured by direct HPLC quantification of deoxynucleosides or density centrifugation than by thermal denaturation. In summary, T reponema strains ZAS-1 and ZAS-2 are sufficiently similar to be considered two strains of a single new species, whereas the morphological and physiological differences displayed by strain ZAS-9 support the assignment of this strain to a separate new species. Latinate species epithets for these organisms are currently under consideration. These results also further our knowledge of diversity in terrnite- associated spirochetes; the previously reported morphological and phylogenetic diversity of these organisms appears to be paralleled by physiological diversity (of an as yet unknown extent). Our current understanding of the ZAS strains suggests that the primary 74 contribution of ZAS-l and ZAS-2 to termite nutrition is through H2/CO2-acetogenesis, while ZAS-9 is likely to be of greater importance in the provision of fixed nitrogen to their hosts. Acknowledgements Thanks to Kristina Stredwick for providing cultures of Myxococcus xanthus DK1622 and to Kwi Kim, Brendan Keough, and Dr. John Breznak for sharing results of their own ongoing studies. 75 References Breznak, J. A., and E. Canale-Parola. 1975. Morphology and physiology of Spirochaeta aurantia strains isolated from aquatic habitats. Arch Microbial 105:1-12. Breznak, J. A., and J. M. Switzer. 1986. Acetate synthesis from H2 plus CO2 by termite gut microbes. Appl Environ Microbiol 52:623-630. Breznak, J. A., and J. S. Blum. 1991. Mixotrophy in the termite gut acetogen, Sporomusa termitida. Arch Microbiol 156:105-110. Leadbetter, J. R., T. M. Schmidt, J. R. Graber, and J. A. Breznak. 1999. Acetogenesis fiom H2 plus CO2 by spirochetes from termite guts. Science 283:686-689. Lilburn, T. C., K. S. Kim, N. E. Ostrom, K. R. Byzek, J. R. Leadbetter, and J. A. Breznak. 2001. Nitrogen fixation by symbiotic and flee-living spirochetes. Science 292:2495-2498. Mesbah, M., U. Premachandran, and W. B. Whitman. 1989. Precise measurement of the G + C content of deoxyribonucleic acid by high performance liquid chromatography. Int J Syst Bacterial 39:159-167. 76 Chapter 4 Interspecies Cofactor Transfer Supports the Folate Requirement of Homoacetogenic Termite Gut Spirochetes Introduction Folate is found in all living cells, playing a central metabolic role as a carrier of one-carbon (C1) groups. In particular, the synthesis of purines, pyrimidines, and several amino acids require folate cofactors (9). “Folate” is actually a generic term referring to folic acid (i.e. pteroylpolyglutarnate) compounds, all of which are composed of a pteridine ring, a p-aminobenzoate moiety, and one or more glutamate residues (Figure 4.1). Individual folate derivatives vary in reduction state, nature and position of C1 substitution, and degree of polyglutamation. Intracellular folates generally occur at the biologically active tetrahydrofolate (TI-IF) level of reduction, but are otherwise diverse in form and highly variable between organisms (8). While the majority of animals require a dietary source of folates, plants and bacteria are generally capable of de novo folate synthesis. In Chapter 2, T reponema strains ZAS-1 and ZAS-2 were shown to have a strict requirement for folate. This was particularly surprising for homoacetogenic organisms, since three enzymes on the methyl-forming branch of the Wood/Ljungdahl pathway of acetogenesis require THF as a cofactor (27). Other homoacetogens appear to be capable of synthesizing folate, and a number of acetogens grown in defined media have no 77 A 0H ? —NH k.“ N/ NH2 Pterin Ring p-Aminobenzoate B. on )\ I H \ H NH2 N m H COOH Glutamate Figure 4.1. Generalized structure of folate compounds (A) and the pteridine portion of the reduced form of folate compounds, tetrahydrofolate (B). Individual derivatives vary in reduction state, C1 substitution (at N5 or N10), and number of glutamyl residues (generally from 1 to 9). 78 apparent folate requirements (12, 26, 39). The absence of the ability to synthesize folate, however, is not uncommon in host-associated organisms, including the spirochetes T reponema bryantii (42) and T. phagedenis (43). In considering potential sources of folate compounds for the ZAS strains in the termite hindgut, it was hypothesized that the most likely candidates would be other members of the gut microbial community. Folate secretion has been observed in members of diverse bacterial genera, including Bacillus, Pseudomonas, Aeromonas, Serratia (l 8), Propionibacterium (l6), and Bifidobacterium (10). Moreover, production of folate by bacteria has long been thought to play an important role animal nutrition (37) and has been clearly demonstrated in a number of gastrointestinal systems (6, 7, 22, 29, 37,38) Research presented in this chapter examines the folate requirements of Treponema strains ZAS-1 and ZAS-2 in greater detail and identifies folate-secreting bacteria from the hindgut of the termite Zootermopsis angusticollis (the original source of the ZAS strains) that are likely to provide folate to the ZAS strains in vivo. Material and Methods Termites Specimens of Zootermopsis angusticollis were collected in the San Gabriel Mountains of southern California and maintained in the laboratory on Pinus ponderosa stump material in polypropylene containers. Specimens were used within one month of their collection. 79 Folate Compounds Folic acid, dihydrofolate (DHF), tetrahydrofolate (THF), 5-methyltetrahydrofolate (5- CH3-THF), 5-10-methylenetetrahydrofolate (5-10-CH2-THF), 5-formyltetrahydrofolate (5-HCO-THF; folinate), lO-formylfolate (IO-HCO-folate), and di- and triglutamate forms of folic acid were obtained fi'om Schirck’s Laboratories (Jana, Switzerland). Standard solutions of folate compounds were prepared as described in Kdnings et al. (21) and contained 10 mM 2-mercaptoethanol (final concentration) and 2% Na - ascorbate (w/v) to prevent the oxidation of reduced folate compounds. All preparative and analytical procedures involving folate compounds or culture filtrates were carried out under reduced lighting conditions to minimize photochemical degradation. Media and Cultivation Methods Routine growth of Treponema strains ZAS-1 and ZAS-2 was as described in Chapter 2, with the following exception: medium 2YACo was modified to medium 2YAFo by replacing the ll-cofactor mixture (25) with 500 ng/ml folinate (final concentration). Medium 2YA was identical to 2YAFo, but lacked folinate. Culture media for folate-excreting strains (ZFX strains) varied according to experimental conditions. Medium GMl was adapted from a bifidobacterial enrichment medium TPY (2) and consisted of (in grams per liter): glucose, 15; tryptone (Difco), 10; phytone (BBL), 5; yeast extract (Difco), 2.5; cysteine - HCl, 0.5; K2HP04, 2; MgCl2 - 6 H20, 0.5; ZnSO4 - 7 H2O, 0.25; CaCl2, 0.15; FCC13, 0.001; Tween 80 (1 ml/L). Plates for anoxic incubations were supplemented with sterile PdCl2 (30 mg/L final concentration) after autoclaving. Medium GM2 contained (g/l): NaCl, 1.0; KCl, 0.5; MgCl2-6H2O, 0.4; 80 CaCl2-2H2O, 0.1;NH4C1, 0.3; KH2PO4, 0.2; Na2SO4, 0.15; NaHCO3, 5.8; 3-N- [morpholino] propanesulfonic acid (MOPS) (10 mM final concentration) and 0.5% (v/v) yeast autolysate. Trace element solution SL1 1, selenite-tungstate solution, 7-vitamin solution, and vitamin-B12 solution were added as described by Widdel and Bak (50). Complex medium GM3 contained 0.2% yeast extract, 0.2% peptone, and 20 mM glucose. Semi-defined medium GM4 contained (g/l): NaCl, 1.0; KCl, 0.5; MgC12-6H20, 0.4; CaCl2-2H2O, 0.1; NH4C1, 0.3; KH2PO4, 0.2; Na2SO4, 0.15; NaI-ICO3, 5.8; casamino acids (Difco), 5; MOPS (10 mM). Afier autoclaving, medium GM4 was supplemented with (g/l): glutathione, 0.01; asparagine, 0.04; glutarnine, 0.04; uracil, 0.02; adenine, 0.01; guanine, 0.01; glucose (20 mM final concentration) and trace element and vitamin solutions described in Van Neil et al. (47). Routine cultivation of ZF X strains was on plates of Reinforced Clostridial Medium (Difco) containing 2% agar at 30°C. All described were at an initial pH of 7.2-7.4. Enrichment, Enumeration, and Isolation of Termite Hindgut Heterotrophs Worker larvae of Z. angusticollis were degutted with forceps, removing any attached midgut segments fiom hindguts. Two extracted hindguts were placed in 5 ml of dithiothreitol (DTT)-reduced buffered salt solution (24) and homogenized while held in an anoxic glove box (3). Enumerations were performed by preparing a serial 10-fold dilution series of gut homogenate in tubes of anoxic ZFX medium. Aliquots (0.1 ml) of the resulting dilutions were spread in triplicate on plates of GMl medium containing 2% agar and incubated under anoxic (95% N2: 5% H2), hypoxic (98.5% N2: 1.5% 02), or oxic (air) conditions at 25°C. Colonies on the high dilution plates were enumerated, then 81 grouped by colony and cell morphology. and representatives were selected for folate- secretion bioassays. Folate Secretion Biaassays The folate secretion bioassay was a modification of the folate diffusion bioassay described by Hewitt and Vincent (15). The folate-requiring bioassay organism, Enterococcus hirae was obtained from the American Type Culture Collection (ATCC 8043). Liquid cultures of E. hirae were grown overnight in brain heart infusion (Difco) at 37°C, and 0.25 ml of the culture was transferred to 25 ml of 50% strength AOAC folate assay medium (AOAC-F A) (Difco) and was again incubated at 37°C overnight. Twenty ml of the resultant culture was added to 250 ml AOAC-F A medium containing 2% agar, which had been autoclaved and cooled to 50°C. This suspension was dispensed into Petri plates and used for bioassays. To test for folate secretion, colonies of interest were picked and patched onto the assay plates and incubated overnight under oxic, hypoxia, or anoxic conditions. Formation of satellite E. hirae colonies around colonies of the test organism was taken to indicate secretion of a putative folate compound(s). Bifidobacterium infantis strain S12 (ATCC 15697), a known folate secretar (10), was used as a positive control. Strains yielding a positive test result were designated ZFX strains, streaked for isolation, and subjected to further characterization. Nucleotide Sequence Analysis of 16s rDNA The 16s rDNA gene was PCR amplified from putative folate-secreting strains using primers 8f(5'-AGAGTTTGATCCTGGCTCAG-3') and 1492r (5'- 82 GGTTACCTTGTTACGACT‘T-3'). Each 100-ul PCR reaction mixture contained the primers (30 pM each), deoxynucleoside triphosphates (Boehringer-Mannheim; 50 11M each), MgCl2 (2 mM), 2.5 U of T aq polymerase (Gibco), 10 pl of PCR buffer (supplied with enzyme) and a small amount of colony material. The reactions were performed with a Gene Amp model 9600 thermocycler (Perkin-Elmer). PCR amplifications consisted of a 3-min hold at 95°C, followed by 30 cycles consisting of 30 s at 95°C, 30 s at 55°C, and 45 s at 72°C. After thermocycling, an additional extension was performed for 10 min at 72°C. Positive controls contained E. coli genomic DNA and negative controls contained no template. After amplification, PCR products were purified using a QIAquick PCR purification kit (Qiagen, Valencia, CA). and subjected to restriction fragment length polymorphism (RF LP) analysis to eliminate redundancy (23). Each amplirner was digested with RsaI and HpaII (New England Biolabs) and the resulting fiagments were resolved on 2.5% NuSieve gel (FMC BioProducts) to yield a restriction fiagment length polymorphism (RF LP) pattern. One representative of each RFLP banding pattern was selected for sequencing. Nearly full length (~1500 nt) 16S rDNA sequences were determined using an ABI PRISM 3100 Genetic Analyzer and overlapping eubacterial sequencing primers: 8f, 339f, 515f, 700f, 776f, 9341', 1100f, 337r, 531r, 6851‘, 1100r, and 1492r. Phylogenetic analysis was performed using the Ribosomal Database Project sequence database (28) and the ARB software package (www.biol.chemie.tu- muenchen.de). Sequences were aligned using the ARB automatic aligner, followed by manual correction of ambiguous regions. A maximum likelihood method (fastDNAml) was used to generate phylogenetic trees (34). 83 Nutritional and growth studies Substrate utilization was assessed by using medium GM2 as a basal medium to which test substrates were added. Media were held in rubber stoppered 18 mm anaerobe tubes (Bellco Glass; Vineland, NJ) under atmospheres of 100% N2 or in foam-stoppered nephlometry flasks under air. Incubations were at 30°C. An increase in cell yield (>20%) in the presence of a substrate, as compared to its absence, was taken to indicate utilization of the substrate as an energy source. Cell yield was determined by measuring the optical density (OD) of cultures at 600 nm with a Milton Roy Spectronic 20 calorimeter. Soluble metabolic products were detected by high performance liquid chromatography (HPLC) with refractive index detection (3), and H2 production was measured by gas chromatogaphy (4). To prepare culture filtrates of ZF X strains, sub-samples of stationary phase cultures gown in medium GM2 were supplemented with 10 mM 2-mercaptoethanol (final concentration) and 2% Na - ascorbate (w/v) and adjusted to pH 7. Samples were centrifuged at 10,000 g for 10 minutes, and supematants were degassed, flushed with N2, passed through a 0.2 pm syringe filter, and stored under N2 at -80°C until analysis. Production of folate compounds supporting the gowth of Treponema strains ZAS-1 and ZAS-2 was assessed by replacing folinate in medium 2YAFo with 10% (v/v) ZFX culture filtrates. Cultures of ZAS-1 and ZAS-2 that served as inocula in these experiments were gown in folinate-free 2YA medium through two transfers to reduce residual folinate carried over with the inoculum. 84 Identification of Folate Compounds Identification of unknown folate compounds was done by comparing HPLC retention times with those of folate standards. The HPLC protocol used was a modification of the procedure described by Pheiffer et al. (35). Folates were separated by using a Shimadzu HPLC system (Shimadzu Corp., Kyoto, Japan) equipped with an Alltech Alltima C13 column (250 x 4.6 mm, 5 pm particle size) (Alltech Assoc., Deerfield, Illinois). Gradient elution was performed with acetonitrile and 33 mM phosphoric acid (pH 2.3) was performed as follows: acetonitrile was held at 5% for the first 9 minutes, then was raised linearly to 7% over the next 13 min, then raised linearly to 16% for the next 9 min, held at 16% for 14 nrin, then raised to 25% over two min, held at 25% for 5 min, and finally decreased to 5% acetonitrile within 3 min. The flow rate was 0.8 mllmin, and sample injection voltune was 20 pl. Separated folate compounds were monitored by using a Shimadzu model SPD-lOA UVN IS detector set at 280 nm and a model RF-lOA fluorescence detector set 280 nm excitation and 356 emission wavelengths for reduced folate compounds and 360/460nm for 10-HCO-folate (20). Data was analyzed with the Shimadzu EZ Start (v. 7.1) software package. F olate concentrations reported for culture filtrates are corrected for folates carried over in culture inocula. Rat plasma conjugase (RPC) was used to remove excess glutamate residues (i.e. more than a single glutamate) from folate compounds, as these interfere with HPLC analysis. RPC was prepared as described in Pfeiffer et al. (35) using rat plasma obtained from Pel-Freez Biologicals (Rogers, AR). Activity of the crude RPC preparation was 85 confirmed as described previously (35). ZFX strains were gown in medium GM4, and filtrates were collected as described above. RPC preparation was added to filtrates at 5% (v/v), and the mixture was incubated at 37°C for one hour, heated to 100°C for 5 min, and cooled an ice for five minutes. Samples were centrifuged at 10,000 x g for 10 minutes at 5°C. Supematants were collected and stored at -80°C until analysis. Results Folate Requirements of Treponema strains ZAS-1 and ZAS-2 Treponema strains ZAS-1 and ZAS-2 were tested for their ability to use folates in the three major states of reduction (folic acid, DHF, and THF) and folinate (5-HCO-THF) (Figure 4.2). ZAS-2 was able to use all of the folates tested, whereas ZAS-1 gew only when provided with THF or folinate. Growth of ZAS-1 appeared to be inhibited in the presence of folate and dihydrofolate compared to the limited gowth seen in the negative control cultures. Addition of 500 ng/ml folinate to these “inhibited” cultures eliminates this effect (data not shown), suggesting that the relatively large pools of folate and dihydrofolate may bind to (and thus rendered inaccessible) the small amounts of gowth supportive folinate carried over with the inoculum (36). Provision of p-aminobenzoate, glutamate, various pteridines (pterin, 6-hydroxymethlypterin, pteroic acid), and the pterin precursor guanosine 5’-triphosphate (GTP) did not alleviate the requirement of either strain for preformed folates. 86 0.8 : ZAS-1 fit 0 '5'0'100'150 200'250' Time (hrs) Figure 4.2. Folate utilization by 'li’eponema strains ZAS-1 and ZAS-2. All folate compounds were provided at a final concentration of 500 ng/ml. The specific compounds were: folic acid (El), dihydrofolate (O), tetrahydrofolate (A), 5-formyltetrahydrofolate (i.e. folinic acid) (V), and no addition (0). 87 Isolation of Folate-Secreting Bacteria from Termite Hindguts Colonies derived from the highest dilutions of Z. angusticollis gut homogenates were sorted on the basis of morphological characteristics (both colonial and microscopic) and incubation condition (oxic, anoxic, or hypoxic). A total of 32 colonies were then bioassayed for folate secretion by using Enterococcus hirae, an organism with a strict folate requirement. Eighteen of the screened colonies supported the gowth of satellite colonies of E. hirae in the folate-free bioassay medium, implying secretion of a putative folate into the surrounding medium (Figure 4.3). These colonies were designated “ZFX” strains. Bifidobacterium infantis str. S12, a known folate secretar (10), was used as a positive control organism in bioassays. The diameter of the zone of E. hirae satellite colonies surrounding ZF X colonies was similar to that observed surrounding B. infantis 812 colonies. Of the colonies yielding negative results, thirteen failed to gow on the assay plates (perhaps indicating a requirement for folate in these isolates) and one gew but did not support the gowth of E. hirae. Colonies yielding positive results were streaked for isolation and subjected to further characterization. Characterization of Folate-Secreting Strains 16S rDNA genes were PCR amplified from genomic DNA of the eighteen putatively folate-secreting isolates using bacteria specific primers. Analysis of the eighteen clones revealed two distinct RF LP banding patterns. Nearly complete 16S rDNA sequences were obtained for one isolate representing each RFLP pattern, desigrated strains ZFX-1 and ZFX-2. Phylogenetic analysis of these sequences indicated that strain ZFX-1 gouped with the genus Serratia, whereas ZFX-2 gouped within the 88 ('35-- W "i\ / Bioassay Organism Figure 4.3. Folate secretion assay of strains ZFX-1 and ZFX-2. Putative folate secretion is indicated by the formation of satellite colonies of the folate-requiring bioassay organism, Enterococcus hirae, surrounding surface colonies of ZFX-1 (inset A) and ZFX-2 (inset B). 89 genus Lactococcus (Figure 4.4). ZFX-1 had 99.7% 16S rRNA sequence similarity to Serratia grimesii and 99.1% similarity of Serratia liquefaciens. Strain ZFX-2 was 99.9% similar in 168 rRNA sequence to Lactococcus lactis subsp. lactis. All ZFX-1 and ZFX-2 type isolates gew well on plates incubated under oxic, anoxic, and hypoxic conditions. Cells of all strains corresponding to the ZFX-1 RFLP pattern were motile gam-negative rods approximately 0.6-0.8 x 1.5-2 pm in size, and all ZFX-2 type isolates were non-motile coccoid cells approximately 0.5-0.8 x 0.9-1.1 pm in size (Figure 4.5). Taken together with the RF LP data, these results suggest that all folate- secreting isolates obtained under the three enrichment conditions were similar, if not identical, to representative strains ZFX-1 or ZFX-2. Mean numbers of strains ZFX-1 and ZFX-2 occurring per Z. angusticollis gut were estimated to be 5.1 i 0.6 x 105 and 5.4 i 0.7 x 105 (n = 3) respectively. A variety of hexoses, pentoses, and disaccharides were used as gowth substrates by both strains during gowth under aerobic conditions; ZFX-1 was additionally capable of using organic acids (Table 4.1). During anaerobic gowth on glucose, strain ZFX-1 produced a mixture of succinate, acetate, ethanol, 2-3 butanediol, and H2. No organic products were detected during aerobic gowth of ZFX-l on glucose. Under both oxic and anoxic conditions, ZFX-2 produced lactate as the sole end product fi'om glucose. Lactate production was stoichiometrically consistent with glucose consumption, and the fermentation pattern was not altered during gowth in a complex medium (GM3). 90 Serratia plymuthica Aranicola proteolyticus Serratia proteomaculans Serratia fonticola Serratia grimesii Serratia quuefaciens ZFX-1 r——[7Aphid S-Symbiants Serratia fican'a Serratia entomophila _l -— Serratia odorifera Serratia marcesens 0.10 B ZFX-2 Lactococcus lactis subsp. lactis Lactococcus lactis subsp. cremorum Lactococcus lactis TmL05 Lactococcus garvieae Lactococcus plantarum l— Streptococcus infantan’us Streptococcus bovis Streptococcus equi Streptococcus sanguinis l l— Streptococcus pyogenes Streptococcus canis ll ,, Escherichia coli 0.10 Figure 4.4. Phylogenetic trees inferred from 16S rDNA sequences (~1500 nt) of ZFX-1 (A), ZFX-2 (B), and related organisms. A maximum likelihood technique (fastDNAml) was used to generate the trees. The homologous sequence from Desulfovibrio senezii was used as an outgroup in tree A (not shown). The scale bars represent units of evolutionary distance and are based on sequence divergence. 91 . I A o -.- \ 1 -~ ‘ \ ’ ’ 5pm B I t a! 0 ° r. '° r- 1. 0. ’Q 3 r. 5pm Figure 4.5. Phase contrast micrographs of Serratia strain ZFX-1 (A) and Lactococcus strain ZFX-2 (B). 92 Table 4.1. Substrate Utilization by Serratia strain ZFX-1 and Lactococcus strain ZFX-2 Strain ZFX-l Strain ZFX-2 Substratea Doubling Time (h) Doubling Time (h) Glucose 1 .30 0.73 Mannitol 1.45 0.73 Ribose 2.65 0.81 Xylose 1.92 0.82 Maltose 1 .22 0.75 Cellobiose n.u. 0.85 Trehalose 1.11 0.81 Acetate 1.85 n.u. Pyruvate 1 .58 n.u. Lactate 1.67 n.u. Fumarate 1.59 n.u. Malate 1.50 n.u. Succinate l . 1 8 n.u. 'All substrates were provided at a final concentration of 10 mM except for methoxylated aromatic compounds (2mM). Substrates tested but not utilized are indicated by n.u. Arabinose, glycine, formate, methanol, ferulate, vanillate, syringate, and trimethoxybenzoate were not utilized by either strain. 93 Folate Production by Serratia strain ZFX-2 and Lactococcus strain ZFX-1 To characterize the putative folate compound(s) secreted by the ZF X strains, cultures were grown in anoxic semi-defined medium GM4. Culture filtrates were collected at the time of inoculation (i.e. 0 hrs.) and at the end of logarithmic growth (9 hrs. for ZFX-1, 7 hrs. for ZFX-2). Reducing agents (2-mercaptoethanol and Na-ascorbate) were immediately added to protect folates from oxidative damage, and filtrates were then treated with rat plasma conjugase (RPC) to remove excess glutamate residues prior to analysis. Good resolution of folate standards was observed in HPLC analysis using fluorescence detection for reduced folates and UV absorbance for unreduced forms (see methods and Figure 4.6). A distinct peak with a retention time identical to that of authentic folinate (i.e. 5-HCO-THF) was observed in culture filtrates of both ZFX strains (Figure 4.6). Concentrations of folinate in culture fluids of ZFX-l and ZFX-2 at the end of growth were 146 ng/ml and 117 ng/ml, respectively (corrected for folinate carried over in inocula). An additional fluorescent compound with a retention time of 31 minutes was produced during growth of ZFX-1; the identity of this compound is unknown, but it did not correspond to any standard folate compound tested under the three detection methods. No other peaks corresponding to folates were observed using any of the three detection methods, suggesting that folinate is the sole folate compound secreted by strains ZFX-1 and ZFX-2 at significant levels. Provision of 10% culture filtrates (v/v) of strains ZFX-l and ZFX-2 supported the growth of Treponema strains ZAS-1 and ZAS-2 in the absence of added folate compounds (Figure 4.7). The filtrates used in these experiments were not treated with 94 15‘ A ’>‘ é a) 10- O C 0) O (I) 9 .3 5‘ Ll. O‘N‘ .7 20 2'5 3'0 35 4o 15- B 9 E, 8 10 C Q) 0 (D 9 s 5* LI. o~”"“' 20 25 30 35 4o 15* C 2 S‘ E 8 10- c 1 8 3 (I) § 2 5‘ LL 0: 20 25 3'0 35 40 Time (min) Figure 4.6. HPLC chromatograms showing fluorescent (280/359 nm) compounds present in culture filtrates of strains ZFX-1 (A) and ZFX-2 (B) and for a standard mixture (C) of THF (peak 1, 1 ng), 5-CH3-THF (peak 2, 1 ng), and 5-HCO-THF (peak 3, 5 ng). In panels A & B, dashed and solid lines indicate samples taken at the time of inoculation and at the end of logarithmic growth, respectively. The peak eluting at 31 minutes in panel A could not be identified, but did not correlate to any folate standard. 95 0.8 3 ZAS-1 If. 095 0.01- ' ifrfl ' r ' 'I'I I'I'I'I 0 50 100 150 200 250300350400 450 500 0.8 - ZAS-2 A L44 0,01. . .. . . o ' 5'0 '160'1éo'zoo'zéo'aoo'356466456560 Time (hrs) Figure 4.7. Growth of Treponema strains ZAS-1 and ZAS-2 in folate free media supplemented with folinate (O), cultures filtrates of Serratia strain ZFX-1 (O) and Lactococcus strain ZFX-2 (El). Culture filtrates were added at 10% (v/v) final concentration. Folinate was provided at 500 ng/ml final concentration. Negative controls (A) were supplemented with 10% uninoculated GM3 medium. 96 RPC, indicating that ZAS-1 and ZAS-2 are capable of using the folinate compounds secreted by the ZF X strains in their native states of polyglutamation. Provision of ZFX culture filtrates did not replace the requirement of either spirochete for yeast autolysate in medium 2YACo, and addition of ZAS culture filtrates did not result in any discernible enhancement of folate production by either ZF X strain (date not shown). Discussion The observed folate requirement of Treponema stains ZAS-1 and ZAS-2 in vitro suggested that a source of folate compounds must exist within the termite hindgut. Given the critical nature of this cofactor in the homoacetogenic metabolism of these organisms [(27); Chapter 2] and the importance of C02-reductive acetogenesis in termite nutrition (5, 45), determination of the source of folates in situ was deemed important to a better understanding of the physiological ecology of the ZAS strains and the termite hindgut symbiotic system as a whole. There are several potential sources of folate in termite hindguts. Since plants are capable of folate synthesis, wood ingested by termites is one possibility. However, the primarily structural role of lignocellulose in plants suggests that this material is likely be relatively low in folate, since this cofactor is normally involved in active biosynthetic processes. The termite host itself is also potential source, although the ability of termites to synthesize folate compounds is unknown. De novo synthesis of folate has been demonstrated in mosquito muscle tissue (19, 48, 49), but the majority of insects require a dietary source of folates. In any event, it seems unlikely that the gut microbial 97 community would have significant access to folate pools existing within host cells. As such, the most likely source of folates would be other microbes inhabiting the hindgut. F olate synthesis is a common capability in bacteria, and secretion of folate compounds has been demonstrated in a variety of bacterial species, including members of the genera Bacillus, Pseudomonas, Aeromonas, Serratia (l8), Propionibacterium (16), and Bifidobacterium (10). Bacterial folate production has been observed in a variety of gastrointestinal systems (7, 29, 37), and host uptake of bacterially produced folates has been unambiguously demonstrated in rat and human gastrointestinal systems (6, 38). The presence of folate-requiring bacteria in the bovine rumen (14, 41, 42) and observation of THF production in rumen enrichment cultures (41) suggests that extensive folate cross- feeding also occurs in the bovine rumen. Folate secretion by Z. angusticollis gut isolates Serratia strain ZFX-1 and Lactococcus strain ZFX-2 was demonstrated by both microbiological assay with E. hirae and by direct detection of folate compounds in ZF X culture filtrates. Folate concentrations in ZF X culture fluids were within the range observed for other folate secreting bacteria, which secrete folates to concentrations 20—160 ng/ml in culture fluids (10, 16, 18). Levels of folate production similar to those of ZFX-2 (100 ng/ml) have been reported in Lactococcus lactis strain M61363. In contrast to ZFX-2 though, 90% of the folate produced by strain M61363 is accumulated intracellularly rather than being secreted into the environment (17). Levels of folate production and secretion also vary widely between individual species and strains of propionibacteria and bifidobacteria (10, 16). It has been hypothesized that strains showing the highest levels of folate secretion 98 lack feed-back inhibition mechanisms controlling folate release, a trait which may be selected for in densely colonized gastrointestinal environments (10). Although no Serratia strains have been identified in recent culture collections (11, 40, 44) or molecular studies (32, 33) of Reticulitermes sp. termite gut bacteria, several strains identified as pathogenic Serratia marcescens have been isolated from Z. angusticollis (13). However, the termites from which strain ZFX-1 was isolated appeared to be in good health, and no subsequent signs of disease were observed in other members of the same colony. These observations, and the presence of a relatively large population of strain ZFX-1 in the gut (~5 x105 cells per gut), suggests that ZFX-1 is more likely to be a beneficial member of the hindgut microbiota. Strain ZFX-1 showed 96% sequence similarity to the 16S rRNAs of a number of secondary symbionts (S-symbionts) of aphids (46) (Figure 4.4). The role of S-symbionts in aphids is currently unknown, but the primary aphid symbiont Buchnera has been implicated in the provision of the vitamin riboflavin to their hosts (30). The observation of folate secretion by Serratia strain ZFX- 1 raises the possibility that the related aphid S-symbionts could also be involved in the provision of vitamins to their hosts. Lactic acid bacteria (LAB), including Lactococcus species similar to strain ZFX- 2, are frequently cultured from termite hindguts (1 , 11, 40, 44). The population levels determined for strain ZFX-2 in the Z. angmticollis hindgut are consistent with numbers of LAB observed in other termite species. LAB are estimated to represent approximately 3% of the total gut bacterial count in the termite Reticulitennesflavipes (44). Lactococci isolated from the guts of other termites exhibit a heterolactic fermentation under anoxic conditions (producing lactate plus smaller amounts of ethanol, acetate, and formate) and 99 shift their fermentation pattern to favor mostly acetate production under oxic conditions (1, 44). In contrast, strain ZFX-2 performs a homolactic fermentation of glucose under both oxic and anoxic conditions. The lack of some medium component required for acetate formation seems unlikely, since medium GM2 is similar in composition to the culture medium used in previous studies of lactococci (l, 44), and growth on a more complex media (GM3) did not result in any change in the fermentation pattern of ZFX-2. The inability of strain ZFX-2 to shifi to the more energetically favorable fermentation pattern favoring acetate production could indicate a lesion in a component of the pyruvate dehydrogenase complex. Given the wide distribution of LAB among diverse termite species (1), it is possible that these organisms are significant folate providers in many hindgut systems. Considering to high degree of variability in folate secretion observed among closely related strains of other folate-secreting bacteria though, it would be necessary to evaluate folate secretion in strains isolated from diverse termites to determine if this is a common property of termite-associated Lactococcus species. The relatively high levels of folinate secretion observed in strains ZFX-1 and ZFX-2 suggests that these organisms represent the major source of folate for Treponema strains ZAS-1 and ZAS-2 in the Z. angusticollis hindgut. In situ confirmation of this hypothesis, however, presents a considerable technical challenge due to the very small volume of extracellular fluid in termite hindguts (0.27-3.2 pl) (31) and the likelihood that secreted folates are efficiently absorbed by folate-requiring members of the dense gut microbial community. In mammalian gastrointestinal systems, investigators have tracked the incorporation of radiolabelled p-aminobenzoate into folate compounds, demonstrating in situ bacterial production and host-uptake of folates (6, 38). This method could prove 100 useful in confirming bacterial folate synthesis in termite hindguts and determining the in situ role of folate-secretors in providing this vitamin to their termite hosts, as well as to other members of termite hindgut microbial communities. Acknowledgments I wish to thank to Dr. Jesse Gregory (University of Florida) and Dr. Tsunenobu Tarnura (University of Alabama) for helpful comments on folate assay techniques. Thanks also to Dr. Jared Leadbetter (California Technical Institute) for providing the Z. angusticollis termites used in this study. 101 10. References Bauer, S., A. Tholen, J. Overmann, and A. Brune. 2000. Characterization of abundance and diversity of lactic acid bacteria in the hindgut of wood- and soil- feeding termites by molecular and culture-dependent techniques. Arch Microbiol 173:126-137. Biavati, B., and P. Mattarelli. 2001. The Family Bifidobacteriaceae. In M. Dworkin (ed.), The Prokaryotes: an Evolving Electronic Resource. Springer- Verlag, New York. Breznak, J. A., and J. M. Switzer. 1986. Acetate synthesis from H2 plus C02 by termite gut microbes. Appl Environ Microbi0152:623-630. Breznak, J. A., and J. S. Blum. 1991. Mixotrophy in the termite gut acetogen, Sporomusa termitida. Arch Microbiol 156:105-110. Breznak, J. A. 1994. Acetogenesis from carbon dioxide in termite guts. In H. L. Drake (ed.), Acetogenesis. Chapman & Hall, New York. Camila, E., J. Zimmerman, J. B. Mason, B. Golner, R. Russell, J. Selhub, and I. H. Rosenberg. 1996. Folate synthesized by bacteria in the human upper small intestine is assimilated by the host. Gastroenterol 110:991-998. Coates, M. E., J. E. Ford, and G. F. Harrison. 1968. Intestinal synthesis of vitamins of the B complex in chicks. Br J Nutr 22:493-500. Cossins, E. A. 1984. Folates in biological materials, p. 1-60. In R. L. Blakely and S. J. Benkovic (ed.), Folates and Pterins, vol. 1. John Wiley & Sons, New York. Cossins, E. A. 2000. The fascinating world of folate and one-carbon metabolism. Can. J. Bot-Rev. Can. Bot. 78:691-708. Deguchi, Y., T. Morishita, and M. Mutai. 1985. Comparative studies on synthesis of water-soluble vitamins among human species of Bifidobacteria. Agric. Biol. Chem. 49:13-19. 102 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. Eutick, M. L., R. W. O'Brien, and M. Slaytor. 1978. Bacteria from the gus of Australian termites. Appl Environ Microbiol 35:823-828. Gentbner, B. R. S., C. L. Davis, and M. P. Bryant. 1981. Features of rumen and sewage sludge strains of Eubacterium Iimosum, a methanol-utilizing and Hz-Coz- utilizing species. Appl Environ Microbiol 42:12-19. Grimont, P. A. D., and F. Grimont. 1978. The genus Serratia. Annu Rev Microbiol 32:221-248. Herbeck, J. L., and M. P. Bryant. 1974. Nutritional requirements of the intestinal anaerobe Ruminococcus bromii. App1.Microbiol. 28:1018-1022. Hewitt, W., and S. Vincent. 1989. Theory and Application of Microbiological Assay. Academic Press, San Diego, CA. Hugenholtz, J., J. Hunik, H. Santos, and E. Smid. 2002. Nutraceutical production by propionibacteria. Lait 82:103-112. Hugenholtz, J., W. Sybesma, M. N. Groot, W. Wisselink, V. Ladero, K. Burgess, D. van Sinderen, J. C. Piard, G. Eggink, E. J. Smld, G. Savoy, F. Senna, T. Jansen, P. Hols, and M. Kleerebezem. 2002. Metabolic engineering of lactic acid bacteria for the production of nutraceuticals. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology 82:217-235. Iwai, K., M. Kobasbi, and H. Fujisawa. 1970. Occurence of Crithidia factors and folic acid in various bacteria. J Bacteriol 104:197-201. Jaffe, J. J., J. J. McCormack, E. Meymarian, and H. M. Doremus. 1977. Comparative activity and properties of lactate dehydrogenase, xanthine dehydrogenase, and dihydrofolate reductase in normal and Brugia pahangi infected Aedes aegypti. J. Parasitol. 63:547-553. Kariluoto, M. S., L. T. Vahteristo, and V. I. Piironen. 2001. Applicability of microbiological assay and affinity chromatography purification followed by high- perfonnance liquid chromatography (HPLC) in studying folate contents in rye. J Sci Food Agric 81:938-942. 103 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. Konings, E. J. M. 1999. A validated liquid chromatographic method for determining folates in vegetables, milk powder, liver, and flour. J AOAC Int 82:119-127. Krause, L. J., C. W. Forsberg, and D. L. O'Connor. 1996. Feeding human milk to rats increases Bifidobacterium in the cecum and colon which correlates with enhanced folate status. J Nutr 126:1505-1511. Laguerre, G., M. Allard, F. Revoy, and N. Amarger. 1994. Rapid identification of rhizobia by restriction fragment length polymorphism analysis of PCR- arnplified 16S rRNA genes. Appl Environ Microbiol 60:56-63. Leadbetter, J. R., and J. A. Breznak. 1996. Physiological ecology of Methanobrevibacter cuticularis sp. nov. and Methanobrevibacter curvams sp. nov., isolated from the hindgut of the termite Reticulitennesflavipes. Appl Environ Microbiol 62:3620-3631. Leadbetter, J. R., T. M. Schmidt, J. R. Graber, and J. A. Breznak. 1999. Acetogenesis from H2 plus C02 by spirochetes from termite guts. Science 283:686-689. Leigh, J. A., F. Mayer, and R. S. Wolfe. 1981. Acetogenium kivui, a new thermophilic hydrogen-oxidizing, acetogenic bacterium. Arch Microbiol 129:275- 280. Ljungdahl, L. G. 1984. Other firnctions of folates, p. 555-579. In R. L. B. a. S. J. Benkovic (ed.), Folates and Pterins, vol. 1. John Wiley and Sons, New York. Maidak, B. L., G. J. Olsen, N. Larsen, R. Overbeek, M. J. McCaughey, and C. R. Woese. 1997. The RDP (Ribosomal Database Project). Nucleic Acids Res 25:109-110. Miller, H. T., and T. D. Lackey. 1963. Intestinal synthesis of folic acid in monoflora chicks. J Nutr 80:236-242. Nakabachi, A., and H. Ishikawa. 1999. Provision of riboflavin to the host aphid, Acyrthosiphon pisum, by endosymbiotic bacteria, Buchnera. J Insect Physiol 45:1-6. 104 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. Odelson, D. A., and J. A. Breznak. 1983. Volatile fatty acid production by the hindgut microbiota of xylophagous termites. Appl Environ Microbiol 45:1602- 1613. Obkuma, M., and T. Kudo. 1996. Phylogenetic diversity of the intestinal bacterial community in the termite Reticulitermes speratus. Appl Environ Microbiol 62:461-468. Obkuma, M., and T. Kudo. 1998. Phylogenetic analysis of the symbiotic intestinal microflora of the termite Cryptotermes domesticus. Ferns Microbiology Letters 164:389-395. Olsen, G. J., H. Matsuda, R. Hagstrom, and R. Overbeek. 1994. FasDNArnl: a tool for construction of phylogenetic trees of DNA sequences using maximum likelihood. Comput. Appl. Biosci. 10:41-48. Pfeiffer, C. M., L. M. Rogers, and J. F. Gregory. 1997. Determination of folate in cereal-grain food products using trienzyme extraction and combined affinity and reversed-phase liquid chromatography. J Agric Food Chem 45:407-413. Pfleiderer, W. 1985. Chemistry of naturally occurring pterins, p. 43-114. In R. L. a. B. Blakely, S. J. (ed.), Folates and Pterins, vol. 2. John Wiley & Sons, New York. Rodriguez, M. S. 1978. A conspectus of research on folacin requirements of man. J Nutr 108:1987-2075. Rong, N., J. Selhub, B. R. Goldin, and I. H. Rosenberg. 1991. Bacterially synthesized folate in rat large intestine is incorporated into host tissue folyl polyglutarnates. J Nutr 121:1955-1959. Schink, B. 1984. Clostridium magnum sp. nov., a non-autotrophic homoacetogenic bacterium. Arch Microbiol 137 :250-255. Schultz, J. E., and J. A. Breznak. 1978. Heterotrophic bacteria present in hindguts of wood-eating termites [Reticulitermesflavipes (Kollar)]. Appl Environ Microbiol 35:930-936. 105 41. 42. 43. 45. 46. 47. 48. 49. 50. Slyter, L. L., and J. M. Weaver. 1977. Tetrahydrofolate and other growth requirements of certain strains of Ruminococcusflavefaciens. Appl Environ Microbiol 33:363-369. Stanton, T. B., and E. Canale-Parola. 1980. Treponema bryantii sp nov, a rumen spirochete that interacts with cellulolytic bacteria. Arch Microbiol 127:145-156. Steinman, H. G., V. I. Oyama, and H. O. Schulze. 1954. Carbon dioxide, cocarboxylase, citrovorum factor, and coenzyme A as essential growth factors for a saccharolytic T reponeme. J Biol Chem 211:327-335. Tholen, A., B. Schink, and A. Brune. 1997. The gut microflora of Reticulitermes flavipes, its relation to oxygen. and evidence for oxygen-dependent acetogenesis by the most abundant Enterococcus sp. FEMS Microbiol Ecol 24:137-149. Tholen, A., and A. Brune. 2000. Impact of oxygen on metabolic fluxes and in situ rates of reductive acetogenesis in the hindgut of the wood-feeding termite Reticulitermesflavipes. Environ. Microbiol. 2:436-449. Unterman, B. M., P. Baumann, and D. L. McClean. 1989. Pea aphid symbiont relations established by analysis of 16S rRNAs. J Bacteriol 171 :2970-2974. van Niel, E. W. J., and B. Hahn-Hagerdal. 1999. Nutrient requirements of lactococci in defined growth media. Appl Microbiol Biotechnol 52:617-627. Venters, D. 1971. The mode of systemic insecticidal action of sulphaquinoxaline against mosquitoes. Trans. R. Soc. Trop. Med. Hyg. 65:24-25. Venters, D. 1972. Folate metabolism in the muscles of Aedes aegypti. Insect Biochemistry 2:153-160. Widdel, F., and F. Bak. 1992. Gram-negative mesophilic sulfate-reducing bacteria, p. 3358-3378. In A. Balows, H. G. Truper, M. Dworkin, W. Harder, and K.-H. Schleifer (ed.), The Prokaryotes, 2nd ed, vol. 4. Springer-Verlag, New York. 106 Chapter 5 Acetogenesis vs. Methanogenesis in the Termite Hindgut: The Lumazine Hypothesis Introduction The puzzling dominance of acetogenesis over methanogenesis in guts of wood- feeding termites was discussed in Chapter 1. To summarize, bacterial H2/C02-acetogens and archaea] methanogens are both present in termite hindguts, but the former clearly out-process the latter as Hz-consumers, as indicated by: i) the low rates of H2 or methane emission measured in most wood-feeding termites (3, 35); ii) the increased rate of methane emission observed after feeding termites antibacterial drugs (35); and iii) the preferential Hz-dependent reduction of 14CO; to acetate, rather than methane, in termite gut homogenates (5). The dominance of H2/COz-acetogenesis in termite hindguts is surprising, since methanogens utilize a more energetically favorable form of metabolism and are the dominant Hz-consumers in most anoxic habitats in which C02/HC03' is the primary terminal e' acceptor (2, 7, 13, 33, 37). While unusual, the dominance of termite C02-reductive acetogenesis as an electron sink is highly beneficial to termites: acetate production in the hindgut fuels the majority of the insect’s energy metabolism (35), whereas CH4 emission represents a loss of otherwise useful carbon and energy. Explaining this peculiar situation is important in understanding the functional ecology of the termite hindgut community. 107 The spatial resource partitioning (SRP) hypothesis (10, 15) discussed in Chapter 1 offers the best current explanation for the dominance of H2/CO2-acetogenesis in termite hindguts. In this model, methanogens and acetogens occupy distinct physiocherrrical niches within the hindgut and hence do not directly compete for H2. In most termites, methanogens appear to be restricted to the region of lowest hydrogen concentration on the gut epithelium (22, 23), where they may be additionally inhibited by inwardly diffusing O2 (9). In contrast, homoacetogens such as Treponema strains ZAS-1 and ZAS-2 colonize the anoxic lumen of the hindgut, where hydrogen production by gut protozoa and other luminal microbes maintains H2 at levels 10 to 100 fold higher than the threshold for H2/CO2-acetogenesis (15). Although the SRP hypothesis helps to reconcile the prevalence of acetogenesis in termite hindguts, it offers no explanation for the restriction of methanogens to the gut epithelium, nor does it explain why acetogens continue to dominate H2-consumption in termite gut homogenates, wherein the spatial distribution of microbes is homogenous and all members should have equal access to H2 (3, 5). While the spatial separation of acetogens and methanogens undoubtedly contributes to the dominance of acetogenesis as the primary H2-sink, it seems likely that some other factor(s) also affects this outcome. It is possible that methanogens are in some way prevented from colonizing the H2 rich gut lumen, perhaps due to the presence of some inhibitory compound. Pteridines are a class of fluorescent compounds possessing a pyrimidine-pyrazine ring structure. Most naturally occurring pteridine compounds are 2-amino-4-hydroxy derivatives and are collectively referred to as pterins (Figure 5.1A). Pterins are involved in a variety of physiological functions, and all living organisms appear to contain at least 108 R1 NH2 R2 N Pterin N N / I \ Deaminase N / \ L / —- I NH2 N N HO \ N N Pterin Lumazine Figure 5.1. Generalized chemical structure of pterin compounds (A) and production of lumazine from pterin via enzymatic deamination (B). 109 trace amounts of them. Insects, however, contain higher concentrations of pterins than most organisms and seem to have developed more varied physiological firnctions for these compounds (49). Insects are capable of synthesizing a wide variety of pteridine compounds (19, 46, 47), which play significant roles in pigmentation (1, 29, 31, 40), nitrogen storage and excretion (11, 18), and development (20, 30). Although there have been no studies of pteridines in termites, these compounds could play a significant role in the competition between hindgut acetogens and methanogens: the pteridine compound lumazine has been demonstrated to be a potent inhibitor of methanogenesis (34). Lumazines are usually formed by the enzymatic deamination of a pterin (Figure 5.13) (38). Lumazine compounds (1, 16, 17, 36) and pterin deaminase activities (17, 43-45) have been detected in a wide range of insect species, and various bacteria have also been shown to possess pterin deaminase activities (26-28, 41). Nagar-Anthal et al. (34) demonstrated that relatively low concentrations of lumazine (0.09-0.6 mM) completely inhibited growth and methanogenesis in phylogenetically diverse methanogens, whereas much higher concentrations had no apparent inhibitory effects on non-methanogenic organisms, including H2/CO2- homoaceto gens. The compound tested in that study, lumazine (2, 4-pteridinedione), is the dearninated derivative of unsubstituted pterin; other lumazine and pteridine compounds were not examined for anti-methanogenic properties. The observation that lumazine inhibited methanogenesis, in conjunction with the known importance of pterin metabolism in insects, raises the possibility that lumazine (or other pteridine compounds) function as inhibitors of methanogenesis in hindguts of wood-feeding termites. In this model, the termite host would excrete lumazine (or related 110 pterins, which could be subsequently deaminated by gut bacteria) into the hindgut. The presence of this compound in the extracellular gut fluid would act to inhibit methanogens in the H2 rich gut lumen, leaving homoacetogens as the primary H2-consuming group. This would also potentially explain the restriction of methanogens to the gut epithelium, since organisms living in biofilm communities are typically less susceptible to the action of inhibitory compounds (12, 48). In this chapter, the lumazine hypothesis is investigated in three ways: i) the susceptibility of the termite gut methanogen Methanobrevibacterfilrfannis to inhibition by lumazine, other pteridine derivatives, and termite gut extracts is evaluated; ii) the acetogenic termite gut spirochetes, Treponema strains ZAS-1 and ZAS-2 are screened for excretion of pteridines and the ability to deaminate exogenous pterins; and iii) gut extracts of the termite Reticulitermesflavipes are screened for the presence of lumazine and other pteridine derivatives. Material and Methods Organisms, Media, and Cultivation Conditions T reponema strains ZAS-1 and ZAS-2 and Methanobrevibacterfilzfonnis (DSM 11501) were grown in medium 2YACo (24) as described in Chapter 2, with the following modification: the cofactor mixture was either omitted (M. filiformis) or replaced with 500 ng/ml folinate (strains ZAS-1 and ZAS-2). All strains were cultivated under anoxic conditions with atmospheres containing 80% H2 and 20% C02. 111 Termite Gut Extract Preparation Reticulitermesflavipes worker termites were collected in Dansville, Mich. as described in Chapter 4 and used within 24 hours. Termites were chilled to 4°C and degutted by using forceps (4). To estimate the concentration of R. flavipes gut extract that would yield measurable amounts of pteridine compounds, termite biomass was compared with that of Pyrrhocoris aptems, an insect in which pteridine concentrations have been quantified (36). A total of 300 extracted guts were pooled in a glass tissue homogenizer containing 1.5 ml of 50 mM Na2CO3/NaHC03 buffer (pH 10). The extraction of pteridine compounds fiom gut homogenates was a modification of the procedure described in Melber et a1. (30). Homogenates were heated to 100°C for one minute and immediately cooled on ice, followed by the addition of 5 m1 of chloroform. Suspensions were mixed by vortexing and centrifuged for 10 min at 3000 x g. The upper aqueous phase (containing pterins) was removed and added to 0.2 ml of chloroform, mixed, and centrifuged for 5 min at 10,000 x g. The upper aqueous phase was collected and stored at -80°C until analysis. All preparations were carried out under reduced lighting conditions. Treponema strain ZAS-l and ZAS-2 Pterin Deaminase Assays Pterin compounds (0.1 mM, final concentration) were added to cultures of ZAS-1 and ZAS-2 at the time of inoculation. At the onset of stationary phase, culture filtrates were obtained by centrifuging cells for 10 min at 10,000 x g and collecting the supematants, which were then neutralized, degassed, and passed through a 0.2 um pore 112 size syringe filter into a pre-sterilized sealed tube filled with N2. Pterin deaminase activity was assessed by TLC analysis of filtrates (see below) to detect any shifts in Rf value or coloration consistent with conversion of the pterin compound to its lumazine derivative. Methanogen inhibition assays The effects of various pteridine compounds and R. flavipes gut extracts on M. filiformis were tested in replicate (n=2-4) liquid cultures. Methane production was assessed by gas chromatography as described previously (35). Pteridines, gut extracts, or ZAS strain culture filtrates were added to cultures after the onset of growth (as determined by optical density readings) and methane production. Termite gut extracts were added to culture tubes at 5% (vol/vol) final concentration, a dose based on comparison with potentially inhibitory pteridine concentrations measured in P. apterus (see above). Pteridines were provided at final concentrations of 0.5 mM (lumazines) and 0.2 mM (pterins) unless otherwise noted. ZAS culture filtrates were provided at a final concentration of 10% (vol/vol). Sterile anoxic water was added to negative control tubes. At the time of additions, culture tubes were wrapped in aluminum foil to protect compounds fi'om photochemical degradation. Subsequent assessments of culture growth were based on methane production alone. All pteridine compounds were obtained from Schirck’s Laboratories (Jona, Switzerland). 113 Thin Layer Chromatography As an initial screen for the presence of pteridine compounds, R. flavipes gut homogenates were analyzed by thin layer chromatography (TLC) (30). Gut homogenate (10 pl) was applied to 20 x 20 cm TLC plates of CEL 400-10 0.1 mm microcrystalline cellulose (Machery-Nagle; Diiren, Germany). Chromatograrns were developed by using an ascending solvent [isopropanolz 2% (w/v) NH3-acetate (1:1)] in a sealed tank lined with Whatrnan no. 1 filter paper. For two-dimensional chromatography, plates were developed in the first dimension as described above, allowed to dry, and developed in the second dimension using 3% NH4C1. Fluorescent compounds on TLC plates were visualized under a long wavelength (366 nm) U.V. lamp (Black-Ray, Ultra-Violet Products, San Gabriel, CA). The ratio of distance traveled by the solute to that of the mobile phase (Rf value) and visual appearance of fluorescant spots in gut extracts were compared to those of standard pteridine compounds (500 uM concentrations applied as 10 ul spots). HPLC Analysis Preparative one-dimensional TLC was performed prior to HPLC analysis of fluorescent compounds in gut extract. A total of 50 ul of R. flavipes gut extract was applied to a TLC plate and developed as described above. Each fluorescent spot was scraped from the glass backing with a razor blade and resuspended in 1 ml of 10 mM phosphoric acid buffer (pH 3.2) containing 4% methanol. The resulting suspension was agitated for 5 min, centrifuged for 15 min at 12,000 x g to sediment cellulose particles, and the supernatant was collected and used for HPLC analysis. 114 The HPLC system and coltunn used for analysis was described in Chapter 4. The HPLC protocol was a modification of the method described by Porcar et al. (36). The mobile phase consisted of 10 mM potassium phosphate buffer (pH 3.2) with 0.5% methanol (v/v) pumped at a flow rate of 1 mein. The injection volume was 20 pl, and each sample was analyzed by fluorescence detection at the following wavelengths (excitation/emission): 335/440 nm, 350/410 nm, and 365/490 nm. Results Methanogen Inhibition Assays Lumazine, other pteridine compounds, and gut extracts of Reticulitennesflavipes were tested for the ability to inhibit methanogenesis by Methanobrevibacterfilifonnis, a methanogen previously isolated from the gut of R. flavipes (23). The addition of lumazine to actively growing of cultures of M. filiformis resulted in a substantial reduction in the rate of methane emission; this effect was consistently observed at final lumazine concentrations ranging from 0.1 to 1 mM (Figures 5.2 and 5.3). In contrast, no inhibition of M. filzformis was observed following the addition of 5% (v/v) R. flavipes gut extract (Figure 5.2) or extracts of whole R. flavipes worker termites (data not shown). Addition of alternatively substituted lumazine derivatives also yielded no significant reduction in CH. emission (Figure 5.3). A variety of pterin compounds commonly observed in insects (pterin, biopterin, neopterin, leucopterin, xanthopterin, and isoxanthopterin) added to M. filiformis cultures at near saturating concentrations (0.2 mM) also had no inhibitory effects on methane emission. Analogous experiments using culture filtrates of T reponema strains ZAS-1 and ZAS-2 provided at 10% (vol/vol) final 115 0.063 '- 0.054- 0.045- 0.036" o.027- l Methane (mmol) 0.018“ 0.009- l ' l ' l ' 1 fl 1 u ‘ l o 20 4o 60 80 100 120 Time (h) Figure 5.2. Methane production by growing cells of M. filiformis following addition of lumazine (0), Fl. flavipes gut extract (A), and anoxic water (D). Lumazine was provided at a final concentration of 1 mM, and gut extract was added at 5% (vol/vol). The time of additions is indicated by the arrow. 116 0.081 " 0.072 ‘- 0.063 ‘ 0.0544 0.045 ‘ Methane (mmol) 0.036 -' 0.027 '- 0.018 " 0.009 * V V I I I I I I l 20 4o 60 80 100 120 140 Time (h) Figure 5.3. Methane production by growing cells of M. filiformis following addition of biolumazine (O), xantholumazine (A), and isoxantholumazine (V), all provided at 0.5 mM final concentrations. Additions of anoxic water (El) and 0.1 mM lumazine (0) were used as controls. The time of additions is indicated by the arrow. 117 concentration also had no effect on growth or methanogenesis by M. filiformis (data not shown). Thin Layer Chromatography of R. flavipes Gut Extracts TLC analysis of R. flavipes gut extracts revealed six fluorescent spots which were designated A through F (Table 5.1). Two-dimensional TLC separations of gut extract showed no additional spots, suggesting that the six spots represented the only fluorescent compounds present at detectable levels. A variety of pteridine compounds (both pterins and lumazine derivatives) commonly found in insects were also screened. Of the six spots detected in R. flavipes gut extracts, five (A-E) were not similar in Rf value or coloration to any screened pteridine. Spot F, however, was similar to leucopterin in both properties (Table 5.1). To assay T reponema strains ZAS-l and ZAS-2 for pterin deaminase activity, the pterins listed in table 5.1 were added to ZAS-1 and ZAS-2 cultures at the onset of growth and culture filtrates were harvested at the end of lag phase. No detectable shifts in Rf value or coloration were observed in any of the compounds, indicating a lack of pterin deaminase activity. TLC analysis of late log phase and stationary phase culture filtrates of Treponema strains ZAS-l and ZAS-2 showed no detectable production of fluorescent compounds during growth. HPLC Analysis of R. flavipes Gut Extracts To remove potentially interfering compounds present in R. flavipes gut extracts, the extracts were pre-purified by TLC prior to HPLC analysis. This purification was first tested by applying a total of 2.5 nmol of lumazine to TLC plates, developing the plate as 118 Table 5.1. Thin Layer Chromatography of Pteridines and R. flavipes Gut Extract Compound Rf Value Color Gut Extract Fluorescent Spots”: Spot A 0.99 Whitish-Green Spot B 0.85 Blue-Green Spot C 0.66 Blue-Green Spot D 0.56 Blue-White Spot E 0.35 Light Green Spot F 0.13 Green Pteridine Comp_ound Standards : Lumazine 0.52 Light Green Xantholumazine 0.27 Blue-Green Isoxantholumazine 0.31 Dark Purple Pterin 0.50 Blue Xanthopterin 0.26 Green Isoxanthopterin 0.24 Purple Biopterin 0.58 Blue-Purple Neopterin 0.50 Blue-Purple Leucopterin 0.14 Green ‘Gut extract and 0.5 mM pteridine standard solutions were applied to TLC plates as 10 ul spots. 119 described above, and extracting the resultant spot from material excised from the plate. Recovery of lumazine from this material (as determined by HPLC analysis) was greater than 90%. HPLC analysis of pterin and lumazine standards (listed in table 5.1) resulted in good linearity of response for micro- to nanomolar concentrations of all compounds tested. The six fluorescent compounds present in R. flavipes gut extracts were subjected to HPLC analysis with fluorescence detection. Five of the compounds (A, B, C, D, and E) showed no detectable peaks at any of the three pteridine-specific fluorescence excitation/emission wavelength combinations tested (data not shown). Spots C and E showed U.V. absorbance at 250 nm, suggesting that what had been interpreted as fluorescence on TLC plates may have actually been U.V. absorbance. Compound F yielded a single fluorescent peak with similar retention time to that of standard leucopterin. The amount of leucopterin present in gut extract equates to a concentration of 0.04 nmol per R. flavipes gut. If all of the detected leucopterin was present in gut fluid and assuming an approximate volume of 0.27 ul fluid per gut (35), leucopterin concentration in gut fluid would be 0.15 mM. Discussion The results of this study demonstrate that neither lumazine nor other pteridine compounds are likely to play a significant role in the inhibition of termite gut methanogens. Of the ten pteridines tested, only unsubstituted lumazine (2, 4- pteridinedione) showed any significant inhibition of the termite methanogen M. filiformis. While lumazine has been reported in the fruit fly Drosophila melanogaster (21, 39) and 120 the honeybee Apis mellifica (14), this compound was not present at detectable levels in gut extracts of R. flavipes by either TLC or HPLC analysis. The only pteridine compound observed in gut extracts was leucopterin, which was shown to be non- inhibitory to M. filiformis at higher concentrations (200 uM) than could have been present in R. flavipes gut fluid (150 uM at maximum). A more likely explanation for the presence of leucopterin in the gut would be as a nitrogen excretory product, since this is one of roles played by pterin compounds in other insects (49). It remains possible that lumazine (or pteridines) are present in the temrite hindgut at concentrations below the detection limits of the assays used in this study, but it is doubtful that such low levels of these compounds would have any significant activity against methanogens in the gut. In a series of independent experiments conducted by Dr. John Breznak, filter paper containing the methanogen-inhibitor bromoethanosulfonate (BES) or lumazine (5 and 500 uM final concentrations, respectively), were fed to R. flavipes worker termites. Methane emission rates measured in live termites fed BES were significantly reduced, whereas lumazine-fed termites actually showed a substantial increase in rates of methane emission (John Breznak, personal communication). The cause of this outcome is as yet unknown, but could be explained by either the degradation of ingested lumazine by the termite or gut microbiota (possibly to methanogenic precursors) or by the presence of a population of gut methanogens more resistant to lumazine inhibition than M. fihformis. Regardless of the reason, the failure of exogenous lumazine to inhibit methanogens in situ further indicates that this compound is unlikely be significantly involved in methanogen inhibition in the termite hindgut. 121 At this time, the failure of termite gut methanogens to dominate H2 consumption in the termite hindgut remains somewhat enigmatic. While it is possible that some non- pteridine inhibitor of methanogens is present in the gut lumen, the abolition of methanogenesis observed in the BBS-feeding studies [(32) and those discussed above] suggests that colonization of the gut epithelium affords methanogens no special protection from inhibitory substances. Also, incubation of live termites (32) and agarose embedded termite guts (42) under H2-enriched atmospheres results in an increase in methane emission, suggesting that substrate limitation is the dominant limiting factor for the population of epithelial methanogens. This inference is further supported by the observation of significantly higher rates of methane emission in Zootermopsis angusticollis than in other wood-feeding termites (3). Trichomonad flagellates containing endosymbiotic methanogens colonize the gut of Z. angusticollis (25), and by living within these protozoa, intracellular methanogens apparently escape the restriction of others methanogens to the gut wall and gain increased access to H2 produced by cellulose degradation. Taken together, these observations suggest that the spatial resource partitioning hypothesis articulated by Brune and coworkers (10, 15) remains the best current explanation for the dominance of H2/CO2-acetogens as H2 consumers in the hindguts of most wood- and grass-feeding termites. Since no termite gut H2/CO2-acetogen has yet displayed any property which would make them unusually competitive with methanogens [(6, 8), and Chapter 2], it would be predicted that successful colonization of the gut lumen by methanogenic archaea would result in a methanogenically dominated H2-flow, equating to a significant loss of energy for the termite host. This makes the 122 counterintuitive restriction of methanogens to the gut epithelium all the more puzzling. This unusual situation may yet be explained by considering other factors specific to growth and survival in termite hindguts, which could include avoidance of grazing by gut protozoa, avoiding being washed out of the gut, or a reliance on maintaining a close physical association with some other member of the epithelial microflora. Determining which (if any) of these factors plays a role in influencing the localization of methanogens in the hindgut remains a critical unresolved issue in understanding the functional ecology of the termite hindgut symbiotic system. 123 References Bel, Y., M. Porcar, R. Sacha, V. Nemec, and J. Ferre. 1997. Analysis of pteridines in Pyrrhocoris apterus (L) (Heteroptera, Pyrrhocoridae) during development and in body-color mutants. Arch Insect Biochem Physiol 34:83-98. Bernalier, A., M. Lelait, V. Rocbet, J.-P. Grivet, G. R. Gibson, and M. Durand. 1996. Acetogenesis fiom H2 and CO2 by methane and non-methane- producing human colonic bacterial communities. FES Microbial. Ecol. 19:193- 202. Brauman, A., M. D. Kane, M. Labat, and J. A. Breznak. 1992. Genesis of acetate and methane by gut bacteria of nutritionally diverse termites. Nature 257:1384-1387. Breznak, J. A., and H. S. Pankratz. 1977. In situ morphology of gut microbiota of wood-eating termites [Reticulitermesflavipes (Kollar) and Coptotermes formosanus Shiraki). Appl Environ Microbiol 33:406-426. Breznak, J. A., and J. M. Switzer. 1986. Acetate synthesis from H2 plus CO2 by termite gut microbes. Appl Environ Microbiol 52:623-630. Breznak, J. A., J. M. Switzer, and H. J. Seitz. 1988. Sporomusa termitida sp. nov., an H2/CO2-utilizing acetogen isolated from termites. Arch Microbiol 150:282-288. Breznak, J. A., and M. D. Kane. 1990. Microbial H2/CO2 acetogenesis in animal guts: nature and nutritional significance. FEMS Microbiol. Rev.:309-3l4. Breznak, J. A. 2000. Ecology of prokaryotic microbes in guts of wood- and litter- feeding termites, p. 209-231. In T. Abe, D. E. Bignell, and M. Higashi (ed.), Termites: Evolution, Sociality, Symbioses, Ecology. Kluwer Academic, Dordrecht/Norwell, MA. Brune, A., D. Emerson, and J. A. Breznak. 1995. The termite gut microflora as an oxygen sink: microelectrode determination of oxygen and pH gradients in guts of lower and higher termites. Appl Environ Microbiol 61 :2681-2687. 124 10. 11. 12. l3. 14. 15. 16. 17. 18. 19. 20. Brune, A., and M. Friedrich. 2000. Microecology of the termite gut: structure and function on a microscale. Curr Opin Microbial 3:263-269. Cochran, D. G. 1975. Excretion in insects, p. 179-281. In D. J. C. a. B. A. Kelby (ed.), Insect Biochemistry and Function. John Wiley and Sons, New York. Costerton, J. W., Z. Lewandowski, D. E. Caldwell, D. R. Korber, and H. M. Lappinscott. 1995. Microbial Biofihns. Annu Rev Microbiol 49:711-745. De Graeve, K. G., J. P. Grivet, M. Durand, P. Beaumatin, C. Cordelet, G. Hannequart, and D. Demeyer. 1994. Competition between reductive acetogenesis and methanogenesis in the pig large-intestinal flora. J. Appl. Biotechnol. 76:55-61. Dustmann, J. H. 1971. Pteridine bei der honigbiene Apis met/{flea isolierung neuer lumazine. Hoppe Seylers Z Physiol Chem 352:1599-1600. Ebert, A., and A. Brune. 1997. Hydrogen concentration profiles at the oxic- anoxic interface: a nricrosensor study of the hingut of the wood-feeding lower termite Reticulitermesflavipes. Appl Environ Microbiol 63:4039-4046. Goto, M., M. Konishi, K. Sugiura, and M. Tsusue. 1966. The structure of a yellow pigment from the mutant lemon of Bombyx mori. Bull Chem Soc Jpn 39:929-932. Gyure, W. L. 1974. Catabolism of isoxanthopterin during the development of the silkworm, Bombyx mori. Insect Biochemistry 4:303-312. Harmsen, R. 1966. The excretory role of pteridines in insects. Journal of Experimental Biology 45:1-13. Harmsen, R. 1969. The effect of atmospheric oxygen pressure on the biosynthesis of simple pteridines in the pierid butterflys. J Insect Physiol 15:2239- 2244. Hudson, B. W., A. H. Bartel, and R. Craig. 1959. Pteridines in milkweed bug, Oncopeltus fasciatus (Dallas) H. Quantitative determination of pteridine content of tissues during growth. J Insect Physiol 3:63-73. 125 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. Katah, S., T. Sueaka, and S. Yamada. 1967. Neopterin in the sepia mutant of adult Drosophila melanogaster. Insect Biochemistry 10:119-123. Leadbetter, J. R., and J. A. Breznak. 1996. Physiological ecology of Methanobrevibacter cuticularis sp. nov. and Methanobrevibacter curvatus sp. nov., isolated from the hindgut of the termite Reticulitennesflavipes. Appl Environ Microbiol 62:3620—363 1. Leadbetter, J. R., L. D. Crosby, and J. A. Breznak. 1998. Methanobrevibacter filiformis sp. nov., a filamentous methanogen from termite hindguts. Arch Microbiol 169:287-292. Leadbetter, J. R., T. M. Schmidt, J. R. Graber, and J. A. Bro-ak. 1999. Acetogenesis from H2 plus CO2 by spirochetes from termite guts. Science 283:686-689. Lee, M. J., P. J. Schreurs, A. C. Messer, and S. H. Zinder. 1987. Association of methanogenic bacteria with flagellated protozoa fi'om a termite hindgut. Curr Microbial 15:337-341. Levenberg, B., and O. Hayaishi. 1959. A bacterial pterin deaminase. J Biol Chem 234:955-961. Levy, C., and W. S. McNutt. 1962. The biological transformation of xanthopterin by a bacterium isolated from soil. Biochemistry 1:1161-1170. McNutt, W. S. 1962. The metabolism of isoxanthopterin by Alcaligenesfaecalis. J Biol Chem 238:1116-1121. Melber, C., and G. H. Schmidt. 1992. Identification of fluorescent compounds in certain species of Dysdercus and some of their mutants (Heteroptera, Pyrrhocoridae). Comp Biochem Physiol 101:115-133. Melber, C., and G. H. Schmidt. 1994. Quantitative variations in the pteridines during the postembryonic development of Dysdercusspecies (Heteroptera, Pyrrhocoridae). Comp Biochem Physiol 108:79-94. 126 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. Melber, C., and G. H. Schmidt. 1997. Body colouration related to the deposition of pteridines in the epidermis and other organs of Dysdercus species (Insecta; Heteroptera: Pyrrhocoridae). Comp Biochem Physiol 1 16:17-28. Messer, A. C., and M. J. Lee. 1989. Effect of chemical treatments on methane emission by the hindgut rrricrobiota in the termite Zootermopsis angusticollis. Microb Ecol 18:275-284. Marvan, B., F. Bonnemoy, G. Fonty, and P. Gouet. 1996. Quantitative determination of H2-utilizing acetogenic and sulfate reducing bacteria and methanogenic archaea from the digestive tract of different mammals. Curr. Microbial. 32:129-133. NagarAntbal, K. R., V. E. Warrell, R. Teal, and D. P. Nagle. 1996. The pterin lumazine inhibits growth of methanogens and methane formation. Arch Microbiol 166:136-140. Odelson, D. A., and J. A. Breznak. 1983. Volatile fatty acid production by the hindgut microbiota of xylophagous termites. Appl Environ Microbiol 45:1602- 1 61 3. Porcar, M., Y. Bel, R. Sacha, V. Nemec, and J. Ferre. 1996. Identification of pteridines in the firebug, Pyrrhocoris apterus (L) (Heteroptera, Pyrrhocoridae) by high-performance liquid chromatography. J Chromatogr A 724:193-197. Prins, R. A., and A. Lankhorst. 1977. Sythesis of acetate from CO2 in the cecum of some rodents. FEMS Microbiol. Letters 1:255-258. Rembold, H. 1985. Catabolism of pterins, p. 155-177. In R. L. B. a. S. J. Benkovic (ed.), Folates and Pterins, vol. 2. John Wiley and Sons, New York. Sagiura, K., and M. Gata. 1967. Biosythesis of pteridines in Drosophila melanogaster. Biochem Biophys Res Commun 28:687-691. Tabata, H., T. Hasegawa, M. N akagasbi, S. Takikawa, and M. Tsuse. 1996. Occurence of biopterin in the wings of Morpho butterflies. Experientia 52:85-87. Takikawa, S., C. Kitayama-yokokawa, and M. Tsusue. 1979. Pterin deanrinase from Bacillus megaterium. J Biochem (Tokyo) 85:785-790. 127 42. 43. 45. 46. 47. 48. 49. Tholen, A., and A. Brune. 2000. Impact of oxygen on metabolic fluxes and in situ rates of reductive acetogenesis in the hindgut of the wood-feeding termite Reticulitermesflavipes. Environ. Microbiol. 2:436-449. Tsusue, M. 1967. Occurence of sepiapterin deaminase in the silkworm, Bombyx mori. Experientia 23:1 16-1 17. Tsusue, M. 1971. Studies on sepiapterin deaminase fi'om the silkworm, Bombyx mori. J Biochem (Tokyo) 69:781-788. Tsusue, M., and T. Mazda. 1977. Occurrence of sepiapterin deaminase in the normal type silkworm, Bombyx mori. Experientia 33:854-855. Watt, W. B. 1967. Pteridine biosynthesis in the butterfly Colias eurytheme. J Biol Chem 242:565-572. Watt, W. B. 1972. Xanthine dehydrogenase and pteridine metabolism in Colias butterflys. J Biol Chem 247:1445-1451. Whiteley, M., M. G. Bangera, R. E. Bumgarner, M. R. Parsek, G. M. Teitzel, S. Lory, and E. P. Greenberg. 2001. Gene expression in Pseudomonas aeruginosa biofilrns. Nature 413:860-864. Zeigler, 1., and R. Harmsen. 1969. The biology of pteridines in insects. Adv. Insect. Physiol. 6:139-207. 128 Chapter 6 Summary This dissertation has investigated of the physiological properties of the termite gut spirochetes T reponema strains ZAS-l , ZAS-2 (3), and ZAS-9 (5) and their relationships with other members of the termite hindgut microbial community. Although over a hundred years of research had yielded a substantial amount of data on the abundance, diversity, and distribution of termite gut spirochetes, their physiological properties and firnctional roles in situ were poorly understood, largely due to a lack of any cultured representatives. The availability of pure cultures of Treponema strains ZAS-l, ZAS-2, and ZAS-9, the first spirochetes to be isolated fiom the guts of wood-feeding termites, has finally made in vitro physiological studies possible. Gaining further understanding of the physiology, nutrition, and growth requirements of these organisms in vitro has resulted in the formulation of new hypotheses regarding their in situ roles and relationships with other members of the termite gut microbial community. The research presented in Chapter 2 focused on the physiological properties relevant to the in situ growth and survival of Treponema strains ZAS-1 and ZAS-2, H2/CO2-homoacetogenic termite gut spirochetes. Strains ZAS-1 and ZAS-2 proved to be nutritionally versatile and were capable of mixotrophic utilization of H2 and organic substrates, including various carbohydrates and (in the case of ZAS-2) methoxylated aromatic compounds. These results suggest that the ZAS strains are likely to contribute to termite nutrition via acetogenesis from fermentation and demethylation of organic 129 substrates in situ, as well as from H2 and CO2 produced by other members of the gut microbiota. ZAS-l and ZAS-2 were shown to be similar to other homoacetogens in possessing enzyme activities of the Wood/Ljungdahl pathway and displaying H2 thresholds within the range of typical of known acetogens. In comparison to other homoacetogens however, ZAS-l and ZAS-2 were less efficient in terms of energy conservation; this may represent an adaptation to a symbiotic lifestyle which prevents excessive cell growth (which could be detrimental to the host) while still allowing production of acetate to fuel termite energy metabolism. Both strains had requirements for folate, a curious observation considering the importance of this cofactor in acetogenesis. Although ZAS-l and ZAS-2 are strict anaerobes, both strains are capable of some degree of O2 tolerance and detoxification, which is likely an adaptation to the partially hypoxic conditions encountered by spirochetes in termite hindguts. Taken together, these findings demonstrate that Treponema strains ZAS-1 and ZAS-l are well adapted to life in the termite hindgut and suggests that their relationship to their termite hosts is beneficial in nature. Chapter 3 provided a more thorough description of Treponema strain ZAS-9, which differed significantly fi'om ZAS-l and ZAS-2 in terms of morphology and metabolism. ZAS-9 is not a H2/CO2-homoacetogen and in fact produced H2 as a product of sugar fermentation. This raises the possibility that interspecies H2 transfer between spirochetes may be an important component of H2 turnover in the termite hindgut. In terms of taxonomy, the results suggest that ZAS-1 and ZAS-2 should be regarded as two strains of single new species in the genus Treponema, whereas strain ZAS-9 should be considered a separate new species of Treponema distinct fi'om that accommodating ZAS- 130 1 and ZAS-2 (Latinate species epithets are currently under consideration). The physiological differences observed among the three strains of termite gut spirochetes currently in pure culture likely offers only an introductory glimpse into the functional diversity of these organisms. Future cultivation efforts should continue to yield new insights into the physiological capabilities of termite gut spirochetes. The observation of folate requirements in Treponema strains ZAS-l and ZAS-2 in Chapter 2 resulted in an investigation of the in situ source of this cofactor, which was presented in Chapter 4. Since folate-secreting bacteria had been observed in a variety of other gastrointestinal systems(1, 2, 7), it was hypothesized that other members of the termite gut microbial community were providers of folate. Consistent with this hypothesis, two folate-secreting strains were isolated from guts of the termite Zootermopsis angusticollis, Serratia strain ZFX-1 and Lactococcus strain ZFX-2. Provision of culture filtrates of ZFX-1 and ZFX-2 supported the growth of Treponema strains ZAS-l and ZAS-2 in the absence of added folates. The folate compound produced by both ZF X strains was determined to be folinate (5-HCO-tetrahydrofolate), which was demonstrated to be a growth supportive form of the cofactor for ZAS-l and ZAS-2. These results suggest that folate secreting bacteria such as strains ZFX-1 and ZFX-2 are the most likely providers of folates to the ZAS strains in situ. The ZF X strains may also be important in providing folate to other members of the gut microbial community to the termite host. Finally, Chapter 5 tested the hypothesis that methanogenic archaea are inhibited by the presence of pteridine compounds in the termite hindgut, allowing H2/CO2- homoacetogens to act the primary H2-consumers in the guts of wood-feeding termites. 131 The pteridine compound lumazine has been shown to be inhibitory to methanogenesis (6), and pteridine compounds are typically present in insects at relatively high levels (as compared to other animals) (8). The presence of inhibitory pteridines in the lumen of the termite hindgut would help to explain the counterintuitive restriction of methanogens to the region of lowest H2 concentration in the gut, the epithelium. However, lumazine was the only pteridine which had any inhibitory effect on the termite methanogen Methanobrevibacterfilyformis, and this compound was not detected in gut extracts of the termite Reticulitermesflavipes. It was therefore deemed doubtful that pteridines play a significant role in the inhibition of methanogenesis in the R. flavipes hindgut, and the localization of termite methanogens to the gut epithelium must be explained by other factors. In concluding this dissertation, it seems appropriate to revisit the pioneering work of Dr. Joseph Leidy (4). In 1881, after years of microscopic observations describing the diverse inhabitants of termite hindguts, Leidy makes the following comment: "Termites... are so common, easily obtained and preserved alive, and their parasites are so exceedingly numerous, constant in their occurrence, and curious, that once the fact becomes sufi'iciently known, the insects will become subjects to illustrate at once the infinity of life and the wonders that are revealed by the microscope. " Over one hundred years later, the termite hindgut continues to fascinate microbiologists, and it seems safe to conclude that it will continue to do for at least another hundred years to come. 132 References Camila, E., J. Zimmerman, J. B. Mason, B. Golner, R. Russell, J. Selhub, and I. H. Rosenberg. 1996. Folate synthesized by bacteria in the human upper small intestine is assimilated by the host. Gastroenterol 110:991-998. Krause, L. J., C. W. Forsberg, and D. L. O'Connor. 1996. Feeding human milk to rats increases Bifidobacterium in the cecum and colon which correlates with enhanced folate status. J Nutr 126:1505-1511. Leadbetter, J. R., T. M. Schmidt, J. R. Graber, and J. A. Breznak. 1999. Acetogenesis fiom H2 plus CO2 by spirochetes fi'om termite guts. Science 283:686-689. Leidy, J. 1874-1881. The parasites of the termites. J. Acad. Nat. Sci. (Phi1.) 8:425-447. Lilburn, T. C., K. S. Kim, N. E. Ostrom, K. R. Byzek, J. R. Leadbetter, and J. A. Breznak. 2001. Nitrogen fixation by symbiotic and fi'ee-living spirochetes. Science 292:2495-2498. NagarAnthal, K. R., V. E. Warrell, R. Teal, and D. P. N agle. 1996. The pterin lumazine inhibits growth of methanogens and methane formation. Arch Microbiol 166:136-140. Slyter, L. L., and J. M. Weaver. 1977. Tetrahydrofolate and other growth requirements of certain strains of Ruminococcusflavefaciens. Appl Environ Microbiol 33:363-369. Zeigler, 1., and R. Harmsen. 1969. The biology of pteridines in insects. Adv. Insect. Physiol. 6:139-207. 133 IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII llélijliliflsliizilflil'i1111(1)1| 144