text
stringlengths
35
50.6k
An innovation is described for producing soy sauce koji mold spore inoculum by using inexpensive autoclavable plastic bags and reuseable plastic enclosures to make culture vessels. After growth, the spore mass could be dried and packaged in the same bag after removing the enclosure. Broken rice was used as the substrate for mold cultivation. Viable spore counts of 10 spores per g were obtained under optimal conditions. After drying at 50 degrees C for 6 h, the moisture content of the spore mass decreased from 35.22 to 6.32% with no significant effect on spore viability. The dry spores could be stored in the refrigerator or at room temperature for at least 3 months.
The activation of the preemergent herbicide 2-(2,4-dichlorophenoxy)ethyl sulfate (Crag herbicide) is initiated by soil microorganisms that are presumed to act by removing the ester sulfate group via some type of sulfatase enzyme. An enrichment technique with the herbicide as the sole source of sulfur led to the isolation of several pure cultures that could produce 2-(2,4-dichlorophenoxy)ethanol from the herbicide. One of these, a strain of Pseudomonas putida, was particularly active. Polyacrylamide gel zymograms of extracts of cells grown on nutrient broth showed the presence of three secondary and three primary alkylsulfatases. One of the latter enzymes was active toward Crag herbicide as well as sodium dodecyl sulfate. Maximum activity was obtained in the late-stationary phase of growth, and enzyme yields were not affected by either the presence or the absence of the herbicide in the growth medium. The enzyme was purified 2,670-fold to homogeneity by a combination of streptomycin sulfate treatment, heat treatment, and column chromatography on DEAE-cellulose, Sephacryl 200-S, and butyl agarose. The pure enzyme was tetrameric (molecular weight, 295,000) and most active at pH 6.0. Saturation kinetics with inhibition by excess substrate were observed for Crag herbicide and octyl sulfate. 2-Butox-yethyl sulfate was a relatively poor substrate, and dodecyltriethoxy sulfate was not hydrolyzed at all. Enzymatic hydrolysis of each substrate in the presence of H(2)O led to incorporation of O exclusively into SO(4) ions in all three cases. The Crag herbicide sulfatase therefore acts by cleaving the O-S bond of the C-O-S ester linkage, in contrast with other alkylsulfatases acting on long-chain alkyl sulfates.
Vanillic acid (4-hydroxy-3-methoxybenzoic acid) supported the anaerobic (nitrate respiration) but not the aerobic growth of Pseudomonas sp. strain PN-1. Cells grown anaerobically on vanillate oxidized vanillate, p-hydroxybenzoate, and protocatechuic acid (3,4-dihydroxybenzoic acid) with O(2) or nitrate. Veratric acid (3,4-dimethoxybenzoic acid) but not isovanillic acid (3-hydroxy-4-methoxybenzoic acid) induced cells for the oxic and anoxic utilization of vanillate, and protocatechuate was detected as an intermediate of vanillate breakdown under either condition. Aerobic catabolism of protocatechuate proceeded via 4,5-meta cleavage, whereas anaerobically it was probably dehydroxylated to benzoic acid. Formaldehyde was identified as a product of aerobic demethylation, indicating a monooxygenase mechanism, but was not detected during anaerobic demethylation. The aerobic and anaerobic systems had similar but not identical substrate specificities. Both utilized m-anisic acid (3-methoxybenzoic acid) and veratrate but not o- or p-anisate and isovanillate. Syringic acid (4-hydroxy-3,5-dimethoxybenzoic acid), 3-O-methylgallic acid (3-methoxy-4,5-dihydroxybenzoic acid), and 3,5-dimethoxybenzoic acid were attacked under either condition, and formaldehyde was liberated from these substrates in the presence of O(2). The anaerobic demethylating system but not the aerobic enzyme was also active upon guaiacol (2-methoxyphenol), ferulic acid (3-[4-hydroxy-3-methoxyphenyl]-2-propenoic acid), 3,4,5-trimethoxycinnamic acid (3-[3,4,5-trimethoxyphenyl]-2-propenoic acid), and 3,4,5-trimethoxybenzoic acid. The broad specificity of the anaerobic demethylation system suggests that it probably is significant in the degradation of lignoaromatic molecules in anaerobic environments.
We have developed a double-staining procedure for use with epifluorescence microscopy which allows the detection both of dividing cells and of ingested bacteria in food vacuoles of heterotrophic microprotozoa. Microprotozoan cells are stained sequentially with the DNA-specific fluorochrome DAPI (4',6-diami-dino-2-phenylindole) and the nonspecific protein stain fluorescein isothiocyanate. During microscopic examination, heterotrophic microprotozoan cells are first located with fluorescein isothiocyanate fluorescence and then epifluorescence filter sets are switched to permit inspection under DAPI fluorescence of the cell nuclei and of the contents of food vacuoles. Among in situ populations of estuarine microprotozoa sampled over a tidal cycle, we found from 2.2 to 5.2% of the heterotrophic cells in a recognizable stage of division (nuclei elongated or double). Batch culture growth experiments were also carried out both with natural populations and with two isolated species of estuarine microprotozoa. In these experiments, the frequency of dividing cells ranged from 1.2 to 3.8% and appeared to be negatively correlated with growth rate. Microprotozoan populations sampled in continental shelf waters off Savannah, Ga., had mean frequencies of dividing cells ranging from 2.0 to 5.0%. A large fraction of cells in heterotrophic microprotozoan populations (an average of 27.4 +/- 1.0% in estuarine water and of 30.1 +/- 4.8% in shelf water) had DAPI-stained inclusions, presumably recently ingested bacteria, in their food vacuoles.
Torulopsis petrophilum can synthesize either a glycolipid surfactant or a protein emulsifier depending on the substrate used. These compounds were not produced to facilitate the uptake of an insoluble carbon source. The glycolipids produced were identical to the mixture isolated from T. bombicola.
A bacterial consortium capable of sucrose degradation primarily to CH(4) and CO(2) was constructed, with acetate as the key methanogenic precursor. In addition, the effect of agar immobilization on the activity of the consortium was determined. The primary fermentative organism, Escherichia coli, produced acetate, formate, H(2), and CO(2) (known substrates for methanogens), as well as ethanol and lactate, compounds that are not substrates for methanogens. Oxidation of the nonmethanogenic substrates, lactate and ethanol, to acetate was mediated by the addition of Acetobacterium woodii and Desulfovibrio vulgaris. The methanogenic stage was accomplished by the addition of the acetophilic methanogen Methanosarcina barkeri and the hydrogenophilic methanogen Methanobacterium formicicum. Results of studies with low substrate concentrations (0.05 to 0.2% [wt/vol]), a growth-limiting medium, and the five-component consortium indicated efficient conversion (40%) of sucrose carbon to CH(4). Significant decreases in yields of CH(4) and rates of CH(4) production were observed if any component of the consortium was omitted. Approximately 70% of the CH(4) generated occurred via acetate. Agar-immobilized cells of the consortium exhibited yields of CH(4) and rates of CH(4) production from sucrose similar to those of nonimmobilized cells. The rate of CH(4) production decreased by 25% when cysteine was omitted from reaction conditions and by 40% when the immobilized consortium was stored for 1 week at 4 degrees C.
Two types of carbon sources-carbohydrate and vegetable oil-are necessary to obtain large yields of biosurfactant from Torulopsis bombicola ATCC 22214. Most of the surfactant is produced in the late exponential phase of growth. It is possible to grow the yeast on a single carbon source and then add the other type of substrate, after the exponential growth phase, and cause a burst of surfactant production. This product is a mixture of glycolipids. The maximum yield is 70 g liter, or 35% of the weight of the substrate used. An economic comparison demonstrated that this biosurfactant could be produced significantly more cheaply than any of the previously reported microbial surfactants.
The kinetics of mineralization of carbonaceous substrates has been explained by a deterministic model which is applicable to either growth or nongrowth conditions in soil. The mixed-order nature of the model does not require a priori decisions about reaction order, discontinuity period of lag or stationary phase, or correction for endogenous mineralization rates. The integrated equation is simpler than the integrated form of the Monod equation because of the following: (i) only two, rather than four, interdependent constants have to be determined by nonlinear regression analysis, (ii) substrate or product formation can be expressed explicitly as a function of time, (iii) biomass concentration does not have to be known, and (iv) the required initial estimate for the nonlinear regression analysis can be easily obtained from a linearized form rather than from an interval estimate of a differential equation. CO(2) evolution data from soil have been fitted to the model equation. All data except those from irradiated soil gave better fits by residual sum of squares (RSS) by assuming growth in soil was linear (RSS = 0.71) as opposed to exponential (RSS = 2.87). The underlying reasons for growth (exponential versus linear), no growth, and relative degradation rates of substrates are consistent with the basic mechanisms from which the model is derived.
Peptidase activity determinations involving native cells of Streptococcus cremoris and completely disrupted cell preparations, as well as experiments concerned with peptidase activity distribution among cell fractions obtained by a damage-restrictive removal of the cell wall and release of intracellular material, suggest the presence of peptidases with distinguishable locations. Alanyl, leucyl, and prolyl aminopeptidase activities are most likely located in the cell wall-membrane interface, showing no detectable association with the membrane. Lysyl aminopeptidase is present not only in this location, but also as an intracellular enzyme. Endopeptidase activity and glutamate aminopeptidase activity appear to be weakly associated with the membrane. The locations of these two peptidase activities, unlike those of the former aminopeptidase activities, impose a restriction on their expression. Results of experiments concerned with permeabilization of the membrane and findings regarding an effect of the local environment of the enzymes on their pH activity profiles are evaluated and considered as being indicative of the proposed location. The possible implications of these findings with respect to protein utilization during growth of the organism in milk are discussed.
The isolation and characterization of a Streptococcus lactis ML3 strain which possessed a recombinant lactose plasmid is described. The recombination events generating this plasmid occurred in vivo in a recombination-deficient strain and appeared to be mediated by transposition events. Restriction mapping revealed that the recombinant plasmid, pDA0307, contained a region of the lactose plasmid, pSK08, linked to another resident plasmid, pSK07. Copy number determinations indicated that the lac genes were present at approximately 20 copies per cell in pDA0307, whereas the lac genes are normally present at approximately 10 copies per cell in pSK08. The strain containing pDA0307 displayed a 21 to 54% increase in the expression of the Lac enzyme phospho-beta-d-galactosidase. However, the strain containing pDA0307 both grew and produced lactic acid in milk at rates identical to that of a strain containing pSK08. This result suggests that lac gene dosage of plasmid-linked lac genes was not limiting the rate at which these derivatives of S. lactis ML3 fermented milk.
Nitrapyrin inhibited growth, CH(4) oxidation, and NH(4) oxidation, but not the oxidation of CH(3)OH, HCHO, or HCOONa, by Methylosinus trichosporium OB3b, suggesting that nitrapyrin acts against the methane monooxygenase enzyme system. The inhibition of CH(4) oxidation could be reversed by repeated washing of nitrapyrin-inhibited cells, indicating that its effect is bacteriostatic. The addition of Cu did not release the inhibition. Methane oxidation was also inhibited by 6-chloro-2-picoline. These data suggest that the mode of action of nitrapyrin on M. trichosporium is different from that on chemoautotrophic NH(4) oxidizers or methanogens.
Conidia of domesticated yellow-green aspergilli from strains of Aspergillus oryzae (Ahlburg) Cohn and Aspergillus sojae Sakaguchi and Yamada ex Murakami, used in the preparation of koji inoculum, germinate approximately 3 h sooner than conidia produced by related wild species, Aspergillus flavus Link ex Fr. and Aspergillus parasiticus Speare. There was no consistent relationship between average conidium size and estimated 50% germination time. Germination trials were conducted on Czapek agar at 28 degrees C. A hypothesis is offered that, in the propagation of koji inoculum, selection has favored those individuals capable of rapid conidium germination and germ tube extension, attributes that enable them to gain the available substrate during intraspecific competition.
The gene responsible for the malolactic fermentation of wine was cloned from the bacterium Lactobacillus delbrueckii into Escherichia coli and the yeast Saccharomyces cerevisiae. This gene codes for the malolactic enzyme which catalyzes the conversion of l-malate to l-lactate. A genetically engineered yeast strain with this enzymatic capability would be of considerable value to winemakers. L. delbrueckii DNA was cloned in E. coli on the plasmid pBR322, and two E. coll clones able to convert l-malate to l-lactate were selected. Both clones contained the same 5-kilobase segment of L. delbrueckii DNA. The DNA segment was transferred to E. coli-yeast shuttle vectors, and gene expression was analyzed in both hosts by using enzymatic assays for l-lactate and l-malate. When grown nonaerobically for 5 days, E. coli cells harboring the malolactic gene converted about 10% of the l-malate in the medium to l-lactate. The best expression in S. cerevisiae was attained by transfer of the gene to a shuttle vector containing both a yeast 2-mum plasmid and yeast chromosomal origin of DNA replication. When yeast cells harboring this plasmid were grown nonaerobically for 5 days, ca. 1.0% of the l-malate present in the medium was converted to l-lactate. The L. delbrueckii controls grown under these same conditions converted about 25%. A laboratory yeast strain containing the cloned malolactic gene was used to make wine in a trial fermentation, and about 1.5% of the l-malate in the grape must was converted to l-lactate. Increased expression of the malolactic gene in wine yeast will be required for its use in winemaking. This will require an increased understanding of the factors governing the expression of this gene in yeasts.
Cells from glucose-limited chemostat cultures of Cytophaga johnsonae were subjected to a sudden relaxation of substrate limitation by injecting the cells into fresh batch cultures. Starvation experiments were carried out by injecting glucose-limited cells into batch cultures lacking glucose. Transient responses of biomass, glucose uptake and mineralization, ATP content, and viability on different agar media were monitored during these nutrient-shift experiments. Cells reacted differently depending on growth rate and time spent in the chemostat. Fast-growing cells showed an immediate adaptation to the new growth conditions, despite some initial overshoot reactions in ATP and uptake potential. In contrast, slowly growing cells and long-term-adapted cells showed extensive transient growth responses. Glucose uptake and mineralization potentials changed considerably during the transient growth phase before reaching new levels. During the starvation experiments, all cell types displayed a fast decrease in ATP, but the responses of the substrate uptake and mineralization potentials were strongly dependent upon the previous growth rate. Both potentials decreased rapidly in cells with high growth rates. On the other hand, cells with low growth rates maintained 80% of their uptake and mineralization potentials after 8 h of starvation. Thus, slowly growing cells are much better adapted for starvation than are fast-growing cells.
Specifically radiolabeled [C-lignin]lignocelluloses were prepared from the aquatic macrophytes Spartina alterniflora, Juncus roemerianus, Rhizophora mangle, and Carex walteriana by using [C]phenylalanine, [C]tyrosine, and [C]cinnamic acid as precursors. Specifically radiolabeled [C-polysaccharide]lignocelluloses were prepared by using [C]glucose as precursor. The rates of microbial degradation varied among [C-lignin]lignocelluloses labeled with different lignin precursors within the same plant species. To determine the causes of these differential rates, [C-lignin]lignocelluloses were thoroughly characterized for the distribution of radioactivity in nonlignin contaminants and within the lignin macromolecule. In herbaceous plants, significant amounts (8 to 24%) of radioactivity from [C]phenylalanine and [C]tyrosine were found associated with protein, although very little (3%) radioactivity from [C]cinnamic acid was associated with protein. Microbial degradation of radiolabeled protein resulted in overestimation of lignin degradation rates in lignocelluloses derived from herbaceous aquatic plants. Other differences in degradation rates among [C-lignin]lignocelluloses from the same plant species were attributable to differences in the amount of label being associated with ester-linked subunits of peripheral lignin. After acid hydrolysis of [C-polysaccharide]lignocelluloses, radioactivity was detected in several sugars, although most of the radioactivity was distributed between glucose and xylose. After 576 h of incubation with salt marsh sediments, 38% of the polysaccharide component and between 6 and 16% of the lignin component (depending on the precursor) of J. roemerianus lignocellulose was mineralized to CO(2); during the same incubation period, 30% of the polysaccharide component and between 12 and 18% of the lignin component of S. alterniflora lignocellulose was mineralized.
The kinetics of sulfate and acetate uptake was studied in the sulfate-reducing bacterium Desulfobacter postgatei (DSM 2034). Kinetic parameters (K(m) and V(max)) were estimated from substrate consumption curves by resting cell suspensions with [S]sulfate and [C]acetate. Both sulfate and acetate consumption followed Michaelis-Menten saturation kinetics. The half-saturation constant (K(m)) for acetate uptake was 70 muM with cells from either long-term sulfate- or long-term acetate-limited chemostat cultures. The average K(m) value for sulfate uptake by D. postgatei was about 200 muM. K(m) values for sulfate uptake did not differ significantly when determined with cells derived either from batch cultures or sulfate- or acetate-limited chemostat cultures. Acetate consumption was observed at acetate concentrations of </=1 muM, whereas sulfate uptake usually ceased at 5 to 20 muM. The results show that D. postgatei is not freely permeable to sulfate ions and further indicate that sulfate uptake is an energy-requiring process.
Fifteen Frankia strains from five different plant species were analyzed by two-dimensional polyacryl-amide gel electrophoresis to determine their relatedness by comparing the polypeptide patterns obtained. Three major subgroups (A, C, and D) were found in the Alnus-Comptonia-Myrica cross-inoculation group. An isolate from Purshia tridentata had a unique protein pattern and represents a distinct group of frankiae. Members of group A were isolated from root nodules of Alnus incana subsp. rugosa and Alnus viridis subsp. crispa. Group C organisms were from A. incana subsp. rugosa and Comptonia peregrina nodules, and group D organisms were from A. incana subsp. rugosa, A. viridis subsp. cripsa, and Myrica pensylvanica root nodules. Isolates from each gel group were obtained at several widely separated geographical locations. The results indicate that two-dimensional polyacrylamide gel electrophoresis is useful for identifying Frankia isolates.
The nutritional versatility of a vibrio-shaped, oxalate-utilizing isolate, strain NOX, obtained from tap water supplied with low concentrations of formate, glyoxylate, and oxalate, was determined by growth experiments with low-molecular-weight carbon compounds at high (grams per liter) and very low (micrograms per liter) concentrations. The organism, which was identified as a Spirillum species, appeared to be specialized in the utilization of a number of carboxylic acids. Yields of 2.9 x 10 CFU/mug of oxalate C and 1.2 x 10 CFU/mug of acetate C were obtained from growth experiments in tap water supplied with various low amounts of either oxalate or acetate. A substrate saturation constant of 0.64 muM oxalate was calculated for strain NOX from the relationship between growth rate and concentration of added oxalate. Maximum colony counts of strain NOX grown in ozonated water (dosages of 2.0 to 3.2 mg of O(3) per liter) were 15 to 20 times larger than the maximum colony counts of strain NOX grown in water before ozonation. Based on the nutritional requirements of strain NOX, it was concluded that carboxylic acids were produced by ozonation. Oxalate concentrations were calculated from the maximum colony counts of strain NOX grown in samples of ozonated water in which a non-oxalate-utilizing strain of Pseudomonas fluorescens had already reached maximum growth. The oxalate concentrations obtained by this procedure ranged from 130 to 220 mug of C/liter.
A thermostable amylase, possibly a beta-amylase from Thermoactinomyces sp. no. 2 isolated from soil, is reported. The enzyme was purified 36-fold by acetone precipitation, ion-exchange chromatography, and Sephadex G-200 gel filtration, and the molecular weight was estimated at 31,600. The enzyme was characterized by demonstration of optimum activity at 60 degrees C and pH 7 and by retention of 70% activity at 70 degrees C (30 min). It was stimulated by Mn and Fe but strongly inhibited by Hg. Maltose was the only detectable product of hydrolysis of starches and was quantitatively highest in plantain starch hydrolysate.
A mutant of Alternaria alternata excreted enhanced levels of carboxymethylcellulase and particularly beta-glucosidase when grown in cellulose liquid media. Both enzymes were purified two- to four-fold by ammonium sulfate precipitation and gel filtration, and the kinetic data showed K(m) values of 16.64 mg/ml of culture fluid for carboxymethylcellulase and 0.14 mM p-nitrophenyl-beta-d-glucoside and 0.81 mM cellobiose for beta-glucosidase at pH 5. Carboxymethylcellulase and extracellular beta-glucosidase functioned optimally at pH 5 to 6 and 4.5 to 5 and at temperatures of 55 to 60 and 70 to 75 degrees C, respectively. Both temperature optima and thermostability of beta-glucosidase were among the highest ever reported for the same enzyme excreted from cellulase and beta-glucosidase hyperproducing microorganisms.
To investigate the regulatory interactions of amino acid transport and incorporation, we determined the effects of dipeptides on amino acid uptake by bacteria in an estuary and a freshwater lake. Dipeptides noncompetitively inhibited net transport and incorporation of amino acids into macromolecules but had no effect on the ratio of respiration to incorporation. Nearly maximum inhibition occurred at peptide concentrations of <10 nM. In contrast, the initial uptake rate of glycyl-[C]phenylalanine was not affected by glycine or phenylalanine. Net amino acid transport appeared to be inhibited by the increased flux into the intracellular pools, whereas the incorporation of labeled monomers into macromolecules was isotopically diluted by the unlabeled amino acids resulting from intracellular hydrolysis of the dipeptide. Chloramphenicol, sodium azide, and dinitrophenol all inhibited the initial uptake rate of leucine and phenylalanine. These results suggest that in aquatic environments amino acids are taken up by active transport which is coupled closely to protein synthesis.
Because of the scarcity of literature on the successful use of serological methods for differentiation of Rhizobium meliloti isolates, the objectives of this study were to provide a rationale for selecting isolates to which antisera could be raised and to appraise the suitability of published methods of preparing R. meliloti antigens for the serological identification of field isolates. We used one-dimensional sodium dodecyl sulfate-polyacrylamide gel electrophoresis to develop protein profiles of eight field isolates and one commercial inoculant strain of R. meliloti in order to choose candidates that were either identical or distinctly different from each other for the production of antisera. The serological methods of tube agglutination and gel immunodiffusion complemented the sodium dodecyl sulfate-polyacrylamide gel electrophoresis method of identification. On the basis of their agglutination titers and gel immunodiffusion analysis, the isolates were placed in five serogroups which were identical to the groupings based on protein profiles. Antigenic characteristics of gel immunodiffusion antigens were influenced by the composition of the growth medium, sonication of whole-cell antigens, and the addition of Formalin. We recommend that careful attention be given to the effects of varying antigen preparation procedures when analyzing R. meliloti so that experimental protocols do not complicate the results. The wide range of homologous-antiserum titers observed for the nine isolates indicates different inherent degrees of immunogenicity of R. meliloti which cannot be predicted before serum production. The sodium dodecyl sulfate-polyacrylamide gel electrophoresis method is a useful tool for screening a collection of R. meliloti isolates to better ensure that strain-specific antisera representative of different types of organisms will be obtained.
Alginate lyases (EC 4.2.2.3) from two marine bacteria were isolated and partially characterized. A cell-bound lyase from isolate A3 had a molecular weight of approximately 100,000 and cleaved mannuronate blocks, apparently in an exo manner. A lyase recovered from the culture medium of isolate W3 was soluble in saturated ammonium sulfate, cleaved guluronate blocks, apparently in an endo manner, and had a molecular weight of 35,000. The thiobarbiturate test and urea-polyacrylamide gel electrophoresis were used to determine substrate specificity and mode of substrate cleavage by the enzymes.
Protoplasts of Claviceps purpurea were prepared by treatment of mycelium with a lytic mixture of snail gut enzyme and cellulase from Trichoderma viride. Such protoplasts could be efficiently lysed by Triton X-100 treatment at high osmotic pressure without Ca or Mg, allowing the release of intact vacuoles in high yields. Vacuoles obtained from cells grown in modified Vogel medium (vegetative-type cells not producing alkaloids) were isolated and purified by centrifugation from a 5% Ficoll 400 (wt/vol) phase into the interphase between two layers, one containing 0.25 M each of mannitol and sucrose, and one containing 0.5 M mannitol. Vacuoles derived from cells grown in a medium favoring ergot alkaloid synthesis (sclerotia-like cells) were isolated by gentle centrifugation of filtered protoplast lysates without addition of Ficoll 400. Biochemical analyses of the vacuole fraction isolated from either kind of cell revealed their function as compartments harboring several hydrolytic enzymes. However, the enrichment of free amino acids in vacuoles of sclerotia-like cells was less pronounced than that in vacuoles of vegetative-type cells, indicating a difference in metabolic compartmentation in the two types of cells.
Aceticlastic methanogens and other microbial groups were enumerated in a 58 degrees C laboratory-scale (3 liter) anaerobic digestor which was fed air-classified municipal refuse, a lignocellulosic waste (loading rate = 1.8 to 2.7 g of volatile solids per liter per day; retention time = 10 days). Two weeks after start-up, Methanosarcina sp. was present in high numbers (10 to 10 CFU/ml) and autofluorescent Methanosarcina-like clumps were abundant in sludge examined by using epifluorescence microscopy. After about 4 months of digestor operation, numbers of Methanosarcina sp. dropped 2 to 3 orders of magnitude and large numbers (most probable number = 10 to 10/ml) of a thermophilic aceticlastic methanogen morphologically resembing Methanothrix sp. were found. Methanothrix sp. had apparently displaced Methanosarcina sp. as the dominant aceticlastic methanogen in the digestor. During the period when Methanothrix sp. was apparently dominant, acetate concentrations varied between 0.3 and 1.5 mumol/ml during the daily feeding cycle, and acetate was the precursor of 63 to 66% of the methane produced during peak digestor methanogenesis. The apparent K(m) value obtained for methanogenesis from acetate, 0.3 mumol/ml, indicated that the aceticlastic methanogens were nearly saturated for substrate during most of the digestor cycle. CO(2)-reducing methanogens were capable of methanogenesis at rates more than 12 times greater than those usually found in the digestor. Added propionate (4.5 mumol/ml) was metabolized slowly by the digestor populations and slightly inhibited methanogenesis. Added n-butyrate, isobutyrate, or n-valerate (4.5 mumol/ml each) were broken down within 24 h. Isobutyrate was oxidized to acetate, a novel reaction possibly involving isomerization to n-butyrate. The rapid growth rate and versatile metabolism of Methanosarcina sp. make it a likely organism to be involved in start-up, whereas the low K(m) value of Methanothrix sp. for acetate may cause it to be favored in stable digestors operated with long retention times.
The betaine-stimulated differential synthesis of vitamin B(12), i.e., the increase in B(12) per increase in dry cell weight, by Pseudomonas denitrificans was inhibited by rifampin and chloramphenicol but not by benzylpenicillin and carbenicillin at concentrations of antibiotic that inhibit growth. The level of the first enzyme of corrin (and porphyrin) biosynthesis, delta-aminolevulinic acid synthetase, was decreased to a much greater degree by rifampin and chloramphenicol than by the penicillins. These data support the concept that betaine stimulation of B(12) synthesis is a result of its stimulation of synthesis of delta-aminolevulinic acid synthetase, a labile and presumably rate-limiting enzyme of corrin formation requiring continuous induction. In further support of this hypothesis, it was found that chloramphenicol immediately interfered with both vitamin B(12) and delta-aminolevulinic acid synthetase formation, no matter when it was added to the system.
Growth of Clostridium thermocellum in batch cultures was studied over a broad range of cellobiose concentrations. Cultures displayed important differences in their substrate metabolism as determined by the end product yields. Bacterial growth was severely limited when the initial cellobiose concentration was 0.2 (wt/vol), was maximal at substrate concentrations between 0.5 and 2.0%, and did not occur at 5.0% cellobiose. Ethanol accumulated maximally (38.3 mumol/10 cells) in cultures with an initial cellobiose concentration of 0.8%, whereas cultures in 2.0% cellobiose accumulated only 17.3 mumol, and substrate-limited cultures (0.2% cellobiose) accumulated little, if any, ethanol beyond that initially detected (8.3 mumol/10 cells). In a medium with 0.8% cellobiose, ethanol was produced at a constant rate of approximately 1.1 mumol/10 cells per h from late-logarithmic phase (16 h) of growth well into stationary phase (44 h). When ethanol was added exogenously at levels more than twice the maximum produced by the cultures themselves (0.5% [vol/vol]), neither the extent of growth (maximum Klett units, 150) nor the amounts of ethanol produced ( approximately 0.17%) by the culture was affected. The ratio of ethanol to acetate was highest (2.8) when cells were grown in 0.8% cellobiose and lowest (1.2) when cells were grown in 0.2% cellobiose.
Ruminococcus albus 8 was cultured with isolated alfalfa cell walls as the carbon source. The culture broth was assayed for muralytic enzyme activities. The effect, with respect to the production of such muralytic enzymes, of growing the microorganism on different carbon sources was also investigated. Also, the rates of solubilization and utilization by R. albus of individual alfalfa cell wall sugars during a 96-h growth period were examined.
A new acetotrophic marine methane-producing bacterium that was isolated from the methane-evolving sediments of a marine canyon is described. Exponential phase cultures grown with sodium acetate contained irregularly shaped cocci that aggregated in the early stationary phase and finally differentiated into communal cysts that released individual cocci when ruptured or transferred to fresh medium. The irregularly shaped cocci (1.9 +/- 0.2 mm in diameter) were gram negative and occurred singly or in pairs. Cells were nonmotile, but possessed a single fimbria-like structure. Micrographs of thin sections showed a monolayered cell wall approximately 10 nm thick that consisted of protein subunits. The cells in aggregates were separated by visible septation. The communal cysts contained several single cocci encased in a common envelope. An amorphous form of the communal cyst that had incomplete septation and internal membrane-like vesicles was also present in late exponential phase cultures. Sodium acetate, methanol, methylamine, dimethylamine, and trimethylamine were substrates for growth and methanogenesis; H(2)-CO(2) (80:20) and sodium formate were not. The optimal growth temperature was 35 to 40 degrees C. The optimal pH range was 6.5 to 7.0. Both NaCl and Mg were required for growth, with maximum growth rates at 0.2 M NaCl and 0.05 M MgSO(4). The DNA base composition was 41 +/- 1% guanine plus cytosine. Methanosarcina acetivorans is the proposed species. C2A is the type strain (DSM 2834, ATCC 35395).
The influence of carbon, nitrogen, and phosphate concentrations on growth and proteinase production by Pseudomonas fluorescens 32A was examined. In mineral salts medium containing dialyzed skim milk supernatant as an inducer, maximum growth was obtained at 1.0 and 2.5 mM orthophosphate at 20 and 5 degrees C, respectively. At both temperatures, 5 mM orthophosphate was required for maximum proteinase production, whereas significant inhibition was found at 10 mM. Orthophosphate was the only phosphate compound able to support growth. With sodium pyruvate as the carbon source, maximum enzyme synthesis was at 100 mM carbon at both temperatures. At both 20 and 5 degrees C maximum growth and enzyme production was found with 10 mM NH(4)Cl. A bioassay for available phosphate based on the growth of P. fluorescens 32A in phosphate-limited mineral salts medium showed that skim milk and skim milk supernatant contained 50 and 10 mM orthophosphate, respectively. Proteinase production in skim milk was 2.6- and 12-fold greater than that in optimal mineral salts medium at 20 and 5 degrees C, respectively. These results suggest that proteinase production in milk does not occur as a result of nutrient limitation and may be regulated in part by milk phosphates.
The levels of seven intermediary enzymes involved in acetate and butyrate formation from acetyl coenzyme A in the saccharolytic anaerobe Clostridium acetobutylicum were investigated as a function of time in solvent-producing batch fermentations. Phosphate acetyltransferase and acetate kinase, which are known to form acetate from acetyl coenzyme A, both showed a decrease in specific activity when the organism reached the solvent formation stage. The three consecutive enzymes thiolase, beta-hydroxybutyrylcoenzyme A dehydrogenase, and crotonase exhibited a coordinate expression and a maximal activity after growth had ceased. Only low levels of butyryl coenzyme A dehydrogenase activity were found. Phosphate butyryltransferase activity rapidly decreased after 20 h from 5 to 11 U/mg of protein to below the detection limit (1 mU/mg). Butyrate no longer can be formed, and the metabolic flux may be diverted to butanol. Butyrate kinase showed a 2.5- to 10-fold increase in specific activity after phosphate butyryltransferase activity no longer could be detected. These results suggest that the uptake of acetate and butyrate during solvent formation can not proceed via a complete reversal of the phosphate transferase and kinase reactions. The activities of all enzymes investigated as a function of time in vitro are much higher than the metabolic fluxes through them in vivo. This indicates that none of the maximal activities of the enzymes assayed is rate limiting in C. acetobutylicum.
An N-acetylgalactosamine-specific protein was purified from mycelial homogenates of the nematode-trapping fungus Arthrobotrys oligospora by using affinity chromatography. The molecular weight of the protein was estimated at 22,000 by its comparative mobility on sodium dodecyl sulfate-polyacrylamide slab gels. Pretreatment of nematodes with the purified protein reduced entrapment, indicating a role for the sugar-binding protein in recognition and capture of prey by the fungus.
Diel patterns of dissolved free amino acid (DFAA) concentration and microheterotrophic utilization were examined in the spring and fall of 1981 in euphotic waters from the base of the mixed layer off the southern California coast. The average depths of the isotherms sampled were 19.2 m for spring and 9.0 m for fall. Total DFAA levels were generally higher in the spring than in the fall, 18 to 66 nM and 14 to 20 nM, respectively. Two daily concentration maxima and minima were observed for total DFAAs as well as for most individual DFAAs. Maxima were usually measured in the mid-dark period and in the early afternoon; minima were typically observed in early morning and late afternoon. Bacterial cell numbers reached maximal values near midnight in both seasons. These increases coincided with one of the total DFAA maxima. The second total DFAA maximum occurred in early to midafternoon, during the time of maximum photosynthetic carbon production and rapid dissolved amino acid utilization. Microbial metabolism (incorporation plus respiration) of selected H-amino acids was 2.7 to 4.1 times greater during the daylight hours. DFAA turnover times, based on these metabolic measurements, ranged between 11 and 36 h for the amino acids tested, and rates were 1.7 to 3.7 times faster in the daylight hours than at night. DFAA distributions were related to primary production and chlorophyll a concentrations. Amino acids were estimated to represent 9 to 45% of the total phytoplankton exudate. Microheterotrophic utilization or production of total protein amino acids was estimated as 3.6 mug of C liter day in spring and 1.9 mug of C liter day in the fall. Assimilation efficiency for dissolved amino acids averaged 65% for marine microheterotrophs.
Alcoholic fermentation, growth, and glucoamylase production by 12 strains of Saccharomyces diastaticus were compared by using starch and dextrins as substrates. Haploid progeny produced from a rapidly fermenting strain, SD2, were used for hybridization with other S. diastaticus and Saccharomyces cerevisiae haploids. Alcoholic fermentation and enzyme production by hybrid diploids and their haploid parents were evaluated. Although the dosage of the STA or DEX (starch or dextrin fermentation) genes may enhance ethanol production, epistatic effects in certain strain combinations caused decreases in starch-fermenting activity. Both the nature of the starch or dextrin used and the fermentation medium pH had substantial effects on alcohol production. Commercial dextrin was not as good a substrate as dextrins prepared by digesting starch with alpha-amylase. Crude manioc starch digested by alpha-amylase was fermented directly by selected hybrids with almost 100% conversion efficiency. The manioc preparation contained adequate minerals and growth factors. This procedure should be suitable for direct commercial application in manioc-producing regions in Brazil and elsewhere. A rapidly fermenting haploid strain, SD2-A8, descended from strain SD2, contains two unlinked genes controlling formation of extracellular amylase. A convenient method for detecting these genes (STA genes) in replica plates containing large numbers of meiotic progeny was developed.
The effects of temperature, solvents, and cultural conditions on the fermentative physiology of an ethanol-tolerant (56 g/liter at 60 degrees C) and parent strain of Clostridium thermohydrosulfuricum were compared. An ethanol-tolerant mutant was selected by successive transfer of the parent strain into media with progressively higher ethanol concentrations. Physiological differences noted in the mutant included enhanced growth, tolerance to various solvents, and alterations in the substrate range and the fermentation end product ratio. Ethanol tolerance was temperature dependent in the mutant but not in the parent strain. The mutant grew with ethanol concentrations up to 8.0% (wt/vol) at 45 degrees C, but only up to 3.3% (wt/vol) at 68 degrees C. Low ethanol concentration (0.2 to 1.6% [wt/vol]) progressively inhibited the parent strain to where glucose was not fermented at 2.0% (wt/vol) ethanol. Both strains grew and produced alcohols on glucose complex medium at 60 degrees C in the presence of either 5% methanol or acetone, and these solvents when added at low concentration stimulated fermentative metabolism. The mutant produced ethanol at high concentrations and displayed an ethanol/glucose ratio (mole/mole) of 1.0 in media where initial ethanol concentrations were </=4.0% (wt/vol), whereas when ethanol concentration was changed from 0.1% to 1.6% (wt/vol), the ethanol/glucose ratio for the parent strain changed from 1.6 to 0.6. These data indicate that C. thermohydrosulfuricum strains are tolerant of solvents and that low ethanol tolerance is not a result of disruption of membrane fluidity or glycolytic enzyme activity.
The morphology and cellulases of Ruminococcus albus 8 were markedly affected by the inclusion of 3-phenylpropanoic acid (PPA) in a defined growth medium. PPA-grown bacteria produced substantial quantities of cell-bound cellulase, as well as a very high-molecular-weight extracellular enzyme and lesser amounts of two low-molecular-weight enzymes. PPA-deprived bacteria produced greater total amounts of cellulase, but all of it exists in soluble, low-molecular-weight forms. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed that the availability of PPA did not affect the kinds of proteins produced, but the distribution of two major proteins between cells and supernatant was PPA dependent. These two proteins (85 and 102 kilodaltons) were primarily associated with the cells of PPA-grown bacteria but were found chiefly in the supernatants of PPA-deprived cultures. Examination of thin sections of PPA-grown R. albus 8 by transmission electron microscopy showed a lobed ruthenium red-staining capsule surrounding the cell wall, as well as small vesicular structures (diameter, 0.05 to 0.06 mum) which appeared to aggregate into larger spherical units (diameter, 0.2 to 0.3 mum). In contrast, thin sections of PPA-deprived cells were devoid of vesicles and showed little or no capsule surrounding the cells.
Peat hydrolysate, a diluted acid-autoclaved extract of peat, was used as a substrate for the production of the extracellular polysaccharide pullulan by three strains of Aureobasidium pullulans, 140B, 142, and 2552. It was found that the addition of (NH(4))(2)SO(4) and K(2)HPO(4) as sources of nitrogen and phosphate, respectively, is not necessary for the polysaccharide production. The economically optimized culture medium for large-scale production of pullulan contains peat hydrolysate, 0.05% NaCl, 0.02% MgSO(4), and 0.01% antifoam FG-10. The initial pH of peat hydrolysate medium is adjusted to its optimum value of 6.0 with Ca(OH)(2). The total ingredient cost for the production of each kilogram of pullulan with optimized medium is only 1/10 of that with the nonoptimized medium. In this study, a zero cost for peat hydrolysate was assumed, since it is an effluent of the peat and peat processing industries.
The fermentation of cellodextrins by eight yeast species capable of fermenting cellobiose was monitored. Only two of these species, Torulopsis molischiana and T. wickerhamii, were able to ferment beta-glucosides with a degree of polymerization between one and six. These two species showed exocellular beta-glucosidase activity. Four other species were able to ferment cellotriose, and the last two species only fermented cellobiose. These latter six species produced a beta-glucosidase capable of attacking cellodextrins, but this enzyme was endocellular.
In free-living Rhizobium japonicum cultures, the stimulatory effect of CO(2) on nitrogenase (acetylene reduction) activity was mediated through ribulose bisphosphate carboxylase activity. Two mutant strains (CJ5 and CJ6) of R. japonicum defective in CO(2) fixation were isolated by mitomycin C treatment. No ribulose bisphosphate carboxylase activity could be detected in strain CJ6, but a low level of enzyme activity was present in strain CJ5. Mutant strain CJ5 also exhibited pleiotropic effects on carbon metabolism. The mutant strains possessed reduced levels of hydrogen uptake, formate dehydrogenase, and phosphoribulokinase activities, which indicated a regulatory relationship between these enzymes. The CO(2)-dependent stimulation of nitrogenase activity was not observed in the mutant strains. Both mutant strains nodulated soybean plants and fixed nitrogen at rates comparable to that of the wild-type strain.
Laboratory studies of methane formation in peat samples from an acid subarctic mire in Sweden indicated the presence of a low-temperature-adapted methanogenic flora. Enrichment culture studies with ethanol, acetate, hydrogen, or a combination of these as substrate for methane formation provided evidence for the existence of two different methanogenic populations in the peat: one, unaffected by hydrogen and using acetate, with a temperature optimum at 20 degrees C; the other, oxidizing hydrogen, with a temperature optimum at ca. 28 degrees C.
The role of anaerobic rumen fungi in in vitro forage fiber degradation was determined in a two forage x two inoculum source x five treatment factorial design. Forages used as substrates for rumen microorganisms were Coastal bermuda grass and alfalfa; inoculum sources were rumen fluid samples from a steer fed Coastal bermuda grass hay or alfalfa hay; treatments were whole rumen fluid (WRF), WRF plus streptomycin (0.2 mg/ml of rumen fluid) and penicillin (1.25 mg/ml of fluid), WRF plus cycloheximide (0.5 mg/ml of fluid), WRF plus streptomycin, penicillin, and cycloheximide, and McDougall buffer. Populations of fungi as shown by sporangial development were greater on bermuda grass leaves than on alfalfa leaflets regardless of inoculum source. However, endogenous fungal populations were greater from the alfalfa hay inoculum. Cycloheximide inhibited the fungi, whereas streptomycin and penicillin, which inhibit bacterial populations, resulted in an increase in numbers of sporangia in the alfalfa inoculum, suggesting an interaction between bacteria and fungi. Bacteria (i.e., WRF plus cycloheximide) were equal to the total population in degrading dry matter, neutral-detergent fiber (NDF), acid-detergent fiber (ADF), and cellulose for both inocula and both forages. Degradation of dry matter, NDF, ADF, and cellulose by anaerobic fungi (i.e., WRF plus streptomycin and penicillin) was less than that due to the total population or bacteria alone. However, NDF, ADF, and cellulose digestion was 1.3, 2.4, and 7.9 percentage units higher, respectively, for bermuda grass substrate with the alfalfa versus bermuda grass inoculum, suggesting a slight benefit by rumen fungi. No substantial loss of lignin (72% H(2)SO(4) method) occurred due to fungal degradation. The most active fiber-digesting population in the rumen was the bacteria, even when streptomycin and penicillin treatment resulted in an increase in rumen fungi over untreated WRF. The development of large numbers of sporangia on fiber may not indicate a substantial role as digesters of forage.
An alternative method for the conversion of cheese whey lactose into ethanol has been demonstrated. With the help of continuous-culture technology, a catabolite repression-resistant mutant of Saccharomyces cerevisiae completely fermented equimolar mixtures of glucose and galactose into ethanol. The first step in this process was a computer-controlled fed-batch operation based on the carbon dioxide evolution rate of the culture. In the absence of inhibitory ethanol concentrations, this step allowed us to obtain high biomass concentrations before continuous fermentation. The continuous anaerobic process successfully incorporated a cell-recycle system to optimize the fermentor productivity. Under conditions permitting a low residual sugar concentration (</=1%), maximum productivity (13.6 g liter h) was gained from 15% substrate in the continuous feed at a dilution rate of 0.2 h. Complete fermentation of highly concentrated feed solutions (20%) was also demonstrated, but only with greatly diminished fermentor productivity (5.5 g liter h).
Whole cells of Pseudomonas dacunhae containing l-aspartate beta-decarboxylase activity were immobilized by mixing a cell suspension with a liquid isocyanate-capped polyurethane prepolymer (Hypol; W. R. Grace & Co., Lexington, Mass.). The immobilized cell preparation was used to convert l-aspartic acid to l-alanine. Properties of the immobilized P. dacunhae cells containing aspartate beta-decarboxylase activity were investigated with batch reactors. Retention of enzyme activity was observed to be as much as 100% when cell lysis was allowed to occur before immobilization. The pH and temperature optima were determined to be 5.5 and 45 degrees C, respectively. Immobilized P. dacunhael-aspartate beta-decarboxylase activity was stabilized by the addition of 0.1 mM pyridoxal-5-phosphate and 0.1 mM alpha-ketoglutaric acid to a 1.7 M ammonium aspartate (pH 5.5) substrate solution. Under conditions of semicontinuous use in a batch reactor, a 2.5% loss in immobilized l-aspartate beta-decarboxylase activity was observed over a 31-day period.
Cells of Saccharomyces rouxii (a salt-tolerant yeast) were grown in the presence of two levels of NaCl, 0 and 15%. Mannans obtained from both the cell walls and culture filtrates (extracellular) were examined. Yields based on the dry weight of cells demonstrated that the levels of both cell wall and extracellular mannans were lower when cells were grown in the presence of 15% NaCl. However, the carbohydrate and protein contents of the mannan preparations were not altered. The cell wall mannans obtained from the two growth conditions had similar molecular weights, whereas the extracellular mannans had different molecular weight distributions. Structural analyses showed that the cell wall and extracellular mannans had similar structures. Both had an alpha1-6-linked backbone to which single mannose and mannobiose units were connected as side chains, predominantly by alpha1-2 linkages. The mannans also contained mannosyloligosaccharides, mannotriose, mannobiose, and mannose attached to protein through an O-glycosidic bond. The molecular structure of the cell wall mannans remained unchanged at both levels of NaCl. However, in the presence of 15% NaCl, the side chains consisting of a mannobiose unit were slightly reduced.
The response of methanogenesis and sulfate reduction to trimethylamine, choline, and glycine betaine was examined in surface sediments from the intertidal region of Lowes Cove, Maine. Addition of these substrates markedly stimulated methanogenesis in the presence of active sulfate reduction, whereas addition of other substrates, including glucose, acetate, and glycine, had no effect on methane production. Sulfate reduction was stimulated simultaneously with methanogenesis by the various quaternary amines and all other substrates examined. Incubation of exogenous trimethylamine, choline, or glycine betaine with either bromoethane sulfonic acid or sodium molybdate was used to establish pathways of degradation of the substrates. Methanogenesis dominated the metabolism of trimethylamine, although limited nonmethanogenic activity, perhaps by sulfate-reducing bacteria, was observed. Acetate was oxidized primarily by sulfate reducers. Both choline and glycine betaine were fermented stoichiometrically to acetate and trimethylamine; apparently, neither substrate could be utilized directly by methanogens or sulfate reducers, and the activities of fermenters, methanogens, and sulfate reducers were all required to effect complete mineralization. These observations support the hypothesis that the presence of quaternary amines can mediate the coexistence of sulfate reduction and methanogenesis in marine surface sediments; they also implicate methanogens in the nitrogen cycle of marine sediments containing quaternary amines.
Extracts prepared from non-solvent-producing cells of Clostridium acetobutylicum contained methyl viologen-linked hydrogenase activity (20 U/mg of protein at 37 degrees C) but did not display carbon monoxide dehydrogenase activity. CO addition readily inhibited the hydrogenase activity of cell extracts or of viable metabolizing cells. Increasing the partial pressure of CO (2 to 10%) in unshaken anaerobic culture tube headspaces significantly inhibited (90% inhibition at 10% CO) both growth and hydrogen production by C. acetobutylicum. Growth was not sensitive to low partial pressures of CO (i.e., up to 15%) in pH-controlled fermentors (pH 4.5) that were continuously gassed and mixed. CO addition dramatically altered the glucose fermentation balance of C. acetobutylicum by diverting carbon and electrons away from H(2), CO(2), acetate, and butyrate production and towards production of ethanol and butanol. The butanol concentration was increased from 65 to 106 mM and the butanol productivity (i.e., the ratio of butanol produced/total acids and solvents produced) was increased by 31% when glucose fermentations maintained at pH 4.5 were continuously gassed with 85% N(2)-15% CO versus N(2) alone. The results are discussed in terms of metabolic regulation of C. acetobutylicum saccharide fermentations to achieve maximal butanol or solvent yield.
Stable mixed continuous cultures of Pseudomonas sp. strain VM15C and Pseudomonas putida VM15A, the former of which produced a polyvinyl alcohol (PVA)-degrading enzyme and the latter of which produced an essential growth factor for PVA utilization by strain VM15C, were established with PVA as the sole source of carbon and energy with chemostat cultivation. A high extent of PVA degradation was achieved at dilution rates of less than 0.030/h. The predominant strain in the cultures was the primary metabolizer of PVA, strain VM15C. The growth supporter, strain VM15A, existed as a minor population, although its population was maintained at an almost constant level throughout a dilution region in which the VM15C population decreased markedly as the dilution rate was raised. A crude growth factor which was prepared from a culture supernatant of strain VM15A and increased the specific growth rate of strain VM15C with PVA in an axenic batch culture was also effective for enhancing the VM15C population and PVA degradation in the mixed continuous culture at a high dilution rate (0.064/h). This indicated that the growth-limiting substrate for strain VM15C in the mixed continuous culture is the growth factor produced by strain VM15A.
Experiments document the ability of two species of autotrophic methanogens to assimilate and utilize organic substrates as the nutrient sulfur or nitrogen source and as a carbon source during growth on H(2)-CO(2). Methanobacterium thermoautotrophicum strain DeltaH and the mesophilic species Methanobacterium sp. strain Ivanov grew with glutamine as the nitrogen source or cysteine as the sulfur source. M. thermoautotrophicum also utilized urea as the nitrogen source and as a carbon precursor for methane and cell synthesis. Methanobacterium sp. strain Ivanov grew with methionine as the sulfur source. The growth rate of two different Methanobacterium species was lower on an organic N or S source than on ammonium or sulfide. S and C tracer studies demonstrated that amino acid or urea assimilation correlated with time and amount of growth. The rate of [S]cysteine incorporation was similar in strain DeltaH (34 nmol h mg of cells) and strain Ivanov (23 nmol h mg of cells). However, the rate of [C]acetate incorporation was dramatically different (17 versus 208 nmol h mg of cells in strains DeltaH and Ivanov, respectively). [C]acetate accounted for 1.3 and 21.2% of the total cell carbon synthesized by strains DeltaH and Ivanov, respectively. Amino acids and urea were mainly assimilated into the cell protein fraction, but accounted for less than 2.0% of the total cell carbon synthesized. The data suggest that a biochemical-genetic approach to understanding cell carbon synthesis in methanogens is feasible; mutants that are auxotrophic for either acetate, glutamine, cysteine, or methionine are suggested as future targets for genetic studies.
A methanogenic consortium able to use 3-chlorobenzoic acid as its sole energy and carbon source was enriched from anaerobic sewage sludge. Seven bacteria were isolated from the consortium in mono- or coculture. They included: one dechlorinating bacterium (strain DCB-1), one benzoate-oxidizing bacterium (strain BZ-2), two butyrate-oxidizing bacteria (strains SF-1 and NSF-2), two H(2)-consuming methanogens (Methanospirillum hungatei PM-1 and Methanobacterium sp. strain PM-2), and a sulfate-reducing bacterium (Desulfovibrio sp. strain PS-1). The dechlorinating bacterium (DCB-1) was a gram-negative, obligate anaerobe with a unique "collar" surrounding the cell. A medium containing rumen fluid supported minimal growth; pyruvate was the only substrate found to increase growth. The bacterium had a generation time of 4 to 5 days. 3-Chlorobenzoate was dechlorinated stoichiometrically to benzoate, which accumulated in the medium; the rate of dechlorination was ca. 0.1 pmol bacterium day. The benzoate-oxidizing bacterium (BZ-2) was a gram-negative, obligate anaerobe and could only be grown as a syntroph. Benzoate was the only substrate observed to support growth, and, when grown in coculture with M. hungatei, it was fermented to acetate and CH(4). One butyrate-oxidizing bacterium (NSF-2) was a gram-negative, non-sporeforming, obligate anaerobe; the other (SF-1) was a gram-positive, sporeforming, obligate anaerobe. Both could only be grown as syntrophs. The substrates observed to support growth of both bacteria were butyrate, 2-dl-methylbutyrate, valerate, and caproate; isobutyrate supported growth of only the sporeforming bacterium (SF-1). Fermentation products were acetate and CH(4) (from butyrate, isobutyrate, or caproate) or acetate, propionate, and CH(4) (from 2-dl-methylbutyrate or valerate) when grown in coculture with M. hungatei. A mutualism among at least the dechlorinating, benzoate-oxidizing, and methane-forming members was apparently required for utilization of the 3-chlorobenzoate substrate.
Various basidiomycetes, ascomycetes, and deuteromycetes, grown in a sugar-rich liquid medium, were compared for laccase-producing ability and for the inducing effect of 2,5-xylidine on laccase production. Clear stimulation of the extracellular enzyme formation by xylidine was obtained in the cultures of Fomes annosus, Pholiota mutabilis, Pleurotus ostreatus, and Trametes versicolor, whereas Rhizoctonia praticola and Botrytis cinerea were not affected by the xylidine, and in the case of Podospora anserina a decrease in laccase activity was observed. The laccases were purified, and electrophoresis on polyacrylamide gels indicated a particular pattern for each laccase. The bands of the induced forms appeared only with basidiomycetes. The optimal pH of R. praticola laccase was in the neutral region, whereas the optima of all the other exolaccases were significantly lower (between pH 3.0 and 5.7). All laccases oxidized the methoxyphenolic acids under investigation, but there existed quantitative differences in oxidation efficiencies which depended on pH and on the nature (noninduced or induced) of the enzyme. The sensitivity of all enzymes to inhibitors did not differ considerably.
Fusarium tricinctum Fn-2B was grown on a rice substrate at room temperature (22 to 26 degrees C) for 2 weeks followed by growth at a low temperature (10 to 12 degrees C) for an additional 2 weeks. A total of 1.5 g of nivalenol and 80 mg of fusarenone-X were obtained from 2 kg of rice culture. The methods of production, extraction, and purification are described.
A mixed population, enriched and established in a defined medium, from a sewage sludge inoculum was capable of complete mineralization of 4-chlorobenzoate. An organism, identified as Arthrobacter sp., was isolated from the consortium and shown to be capable of utilizing 4-chlorobenzoate as the sole carbon and energy source in pure culture. This organism (strain TM-1), dehalogenated 4-chlorobenzoate as the initial step in the degradative pathway. The product, 4-hydroxybenzoate, was further metabolized via protocatechuate. The ability of strain TM-1 to degrade 4-chlorobenzoate in liquid medium at 25 degrees C was improved by the use of continuous culture and repeated sequential subculturing. Other chlorinated benzoates and the parent compound benzoate did not support growth of strain TM-1. An active cell extract was prepared and shown to dehalogenate 4-chloro-, 4-fluoro-, and 4-bromobenzoate. Dehalogenase activity had an optimum pH of 6.8 and an optimum temperature of 20 degrees C and was inhibited by dissolved oxygen and stimulated by manganese (Mn). Strain improvement resulted in an increase in the specific activity of the cell extract from 0.09 to 0.85 nmol of 4-hydroxybenzoate per min per mg of protein and a decrease in the doubling time of the organism from 50 to 1.6 h.
Multilocus isoenzyme electrophoresis was used to screen 47 field isolates of Yersinia ruckeri for electrophoretic variation at 15 enzyme loci. Only four electrophoretic types were observed, thus indicating that the genetic structure of Y. ruckeri is clonal. Forty-two isolates were of one electrophoretic type, a reflection of the low amount of genetic diversity extant in this species. Although sorbitol fermentation has been considered to be indicative of a second biotype, no significant gene frequency differences were found between the group of 20 isolates that readily used sorbitol as the sole carbon source and the group of 27 that did not.
The ratios of bicarbonate uptake to substrate oxidation were measured for three genera of nitrifying bacteria. The ratios for the two ammonium oxidizers tested were essentially the same; 0.0863 +/- 0.0055 and 0.0868 +/- 0.0091 mumol of bicarbonate were taken up per umol ammonium oxidized for Nitrosomonas europaea and a Nitrosospira strain, respectively. For Nitrobacter sp., a ratio of 0.0236 +/- 0.0013 mumol of bicarbonate taken up per umol of nitrite oxidized was obtained. Cells were grown in substrate-limited continuous culture and in batch culture, with generation times ranging between 16 and 189 h for the ammonium oxidizers and 18 and 69 h for Nitrobacter sp. All ratios appeared to be independent of growth rates and pH. However, short-term changes in substrate concentration and certain metabolic inhibitors significantly changed the efficiency of bicarbonate uptake. The significance of these results to the application of the nitrapyrin-sensitive bicarbonate uptake method for measuring nitrification rates in natural samples is discussed.
The cell-wall-associated proteolytic systems of several Streptococcus cremoris strains were analyzed by crossed immunoelectrophoresis. At least four immunologically different components of the proteolytic system's were found. One of these proteins was produced by all strains tested. The proteolytic activity of this enzyme was demonstrated with a zymogram staining technique which is based on the degradation of Coomassie-brilliant-blue-stainable casein. The crossed-immunoelectrophoresis patterns of the proteolytic systems of different S. cremoris strains indicated that each strain produces a characteristic combination of proteins. On the basis of these combinations, the different S. cremoris strains were classified into four groups.
For improved production of coenzyme A (CoA), a mutant of Brevibacterium ammoniagenes IFO127071 resistant to oxypantetheine, the corresponding oxygen analog of pantetheine, was obtained. In the mutant, activity of pantothenate kinase (EC 2.7.1.33), the first-step enzyme for the biosynthesis of CoA from pantothenic acid, l-cysteine, and ATP, was about threefold higher than that in the parent strain. As the main regulation mechanism of CoA biosynthesis in this bacterium is negative feedback inhibition of pantothenate kinase by CoA, the mutant is very useful as a catalyst for practical production of CoA. When added to culture broth of the mutant, pantothenate, l-cysteine, and AMP gave 9.3 mg of CoA per ml. With pantetheine and AMP, 11.5 mg of CoA per ml accumulated. These values were about threefold higher than those with the parent strain, and more than 70% of the added AMP was converted to CoA.
A novel thermophilic, coccoid methanogen isolated from nonthermal freshwater sediments is described. Hydrogen plus carbon dioxide and formate were substrates for methanogenesis, and methane production was stimulated by yeast extract, Casamino Acids, and tryptose. Growth also occurred autotrophically. Elevated levels of sodium chloride were not required for maximum growth and were inhibitory above 2%. The minimum doubling time occurred at 57 degrees C, and the upper and lower limits for methane production were 62 and 26 degrees C, respectively. The optimum pH for growth was between 7.0 and 7.5. Inhibitory antibiotics included metronidazole, anisomycin, chloramphenicol, and lasalocid. Colonies were circular, dark yellow, shiny, and convex with entire edges. Cells were 1 to 2.5 mum in diameter, nonmotile, occurring singly or in pairs, and fimbriated. Cells were lysed by pronase or trypsin digestion, glass-distilled water, and 1% sodium dodecyl sulfate. Electron micrographs of thin sections showed a monolayered cell wall ca. 20 nm thick. The DNA base ratio was 49.2 mol% guanine plus cytosine. The whole cell protein pattern differed from that of other named coccoid methanogens.
The alpha- and beta-glucosidase activity in natural samples can be readily measured during short incubation times (20 min) by using the artificial substrates 4-methylumbelliferyl-alpha-d-glucoside and 4-methylumbelliferyl-beta-d-glucoside. The apparent K(m) of both alpha- and beta-glucosidase for these respective substrates is 0.01 muM. The homologous disaccharides maltose and cellobiose competitively inhibit alpha- and beta-glucosidase, respectively. Absolute substrate specificity of the alpha- and beta-glucosidase is observed with respect to the configuration of carbon atoms 1 and 4. Enrichment cultures on either alpha- and beta-glucoside result in increasing activity of the corresponding glucosidase, both in absolute terms and with respect to the other glucosidase.
Bacterial populations attached to intestinal linings of shallow-water fish were compared to those free in the lumen for response to hydrostatic pressure and ability to degrade a variety of substrates. Results suggested that, unlike reports on gut-associated deep-sea bacteria, the two shallow-water populations were not significantly different in their pressure or substrate responsiveness.
The effect of environmental parameters on the growth and the tyrosine phenol-lyase content of Erwinia herbicola was investigated. On mineral medium containing glycerol, l-tyrosine increased the enzyme content 23-fold. When the l-tyrosine was also the carbon source, bacterial growth was 300 times greater than the basal level. On a rich medium, tyrosine phenol-lyase production was strongly dependent on pH and aeration. Catabolite repression and induction both probably control enzyme content.
The nitrogen content, distribution, and amino acid composition of protein material were determined in wood and lignin of Fagus sylvatica. The data indicated that the nitrogen originated from hydroxyproline-rich cell wall glycoprotein, about half of which may be bound to the lignin polymer. The implications for lignocellulose biodegradation are discussed.
Cellulase yields of 250 to 430 IU/g of cellulose were recorded in a new approach to solid-state fermentation of wheat straw with Trichoderma reesei QMY-1. This is an increase of ca. 72% compared with the yields (160 to 250 IU/g of cellulose) in liquid-state fermentation reported in the literature. High cellulase activity (16 to 17 IU/ml) per unit volume of enzyme broth and high yields of cellulases were attributed to the growth of T. reesei on a hemicellulose fraction during its first phase and then on a cellulose fraction of wheat straw during its later phase for cellulase production, as well as to the close contact of hyphae with the substrate in solid-state fermentation. The cellulase system obtained by the solid-state fermentation of wheat straw contained cellulases (17.2 IU/ml), beta-glucosidase (21.2 IU/ml), and xylanases (540 IU/ml). This cellulase system was capable of hydrolyzing 78 to 90% of delignified wheat straw (10% concentration) in 96 h, without the addition of complementary enzymes, beta-glucosidase, and xylanases.
Samples of peat from Pine Island and Brookston bogs in Minnesota were hydrolyzed with various concentrations of HCl or H(2)SO(4) solutions, before or after debituminization (an extraction process used to remove waxy materials, bitumens, from peat), to produce peak hydrolysates as growth substrates for Candida utilis. Hydrolysates were neutralized with concentrated NaOH solution to pH 3.5, 4.5, 5.0, 5.5, 6.0, and 7.0. The precipitated humates were removed by filtration. The resulting peat hydrolysates were amended with reagent-grade K(2)HPO(4), K(2)SO(4), and MgSO(4), 200, 100, and 50 mg per liter of peat hydrolysate, respectively. The debituminized peat produced more total nitrogen (TN) and reducing substances (RS) than the nondebituminized peat. Peat hydrolysates produced by HCl solutions contained slightly higher RS and TN than those produced by H(2)SO(4) solutions; however, the latter were better growth substrates than the former. The yield coefficients in both H(2)SO(4) and HCl hydrolysates initially decreased at 12 to 24 h and then increased gradually over the remaining incubation period (24 to 96 h). As TN and RS were increased, an increase in cell density, biomass, and productivity was observed. In contrast, a decrease in specific growth rate occurred as the RS and TN were increased. The generation time of C. utilis was affected by the concentrations of RS and TN. A peak substrate yield coefficient was found at pH 5.0 in HCl hydrolysates and at pH 6.0 to 6.5 in H(2)SO(4) hydrolysates. Good linear correlation coefficients were found between RS and biomass of C. utilis. The coefficients of correlation increased as the TN level in hydrolysates was increased.
Components from culture fluid and whole cells of Rhizobium trifolii were examined for effects on root hair morphology of white clover seedlings (Trifolium repens var. Ladino). Cell-free culture fluid, exopolysaccharides, supernatant fluid from the precipitation of the exopolysaccharides, capsular polysaccharides, lipopolysaccharides, and a protein fraction from culture fluids were assayed for morphogenetic effects on the root hairs of axenically grown clover seedlings. Crude fractions were chromatographed on Bio Gel A-5m (Bio-Rad Laboratories), and fractions collected were similarly assayed. Hexose, uronic acid, and protein concentrations were determined for all fractions assayed. Gel chromatography indicated the materials with deforming ability to be of high molecular weight (>10,000). For all fractions except exopolysaccharide, deforming ability was associated with a protein component. This suggested that two components were associated with deformation; both contained polysaccharides and one contained protein. Crude fractions differed in their ability to cause deformations and indicated the following relative ability (in decreasing order) to deform root hairs: cell-free culture fluid, capsular polysaccharides, protein from culture fluids, exopolysaccharide, and cell envelope. Lipopolysaccharides had no effect.
A metered blend of anaerobic-grade N(2), CO(2), and H(2)S gases was introduced into an illuminated, 800-ml liquid volume, continuously stirred tank reactor. The system, described as an anaerobic gas-to-liquid phase fed-batch reactor, was used to investigate the effects of H(2)S flow rate and light energy on the accumulation of oxidized sulfur compounds formed by the photoautotroph Chlorobium limicola forma thiosulfatophilum during growth. Elemental sulfur was formed and accumulated in stoichiometric quantities when light energy and H(2)S molar flow rate levels were optimally adjusted in the presence of nonlimiting CO(2). Deviation from the optimal H(2)S and light energy levels resulted in either oxidation of sulfur or complete inhibition of sulfide oxidation. Based on these observations, a model of sulfide and sulfur oxidases electrochemically coupled to the photosynthetic reaction center of Chlorobium spp. is presented. The dynamic deregulation of oxidative pathways may be a mechanism for supplying the photosynthetic reaction center with a continuous source of electrons during periods of varying light and substrate availability, as in pond ecosystems where Chlorobium spp. are found. Possible applications for a sulfide gas removal process are discussed.
Previous investigations have identified a quantitatively major intermediate of lignin degradation by Streptomyces viridosporus. The intermediate, a modified lignin polymer, acid-precipitable polymeric lignin (APPL), is released as a water-soluble catabolite and has been recovered in amounts equivalent to 30% of the lignin originally present in a corn stover lignocellulose substrate after degradation by this actinomycete. In the present work, APPLs were collected at various time intervals from cultures of two highly ligninolytic Streptomyces sp. strains, S. viridosporus T7A and S. badius 252, growing on corn stover lignocellulose. APPL production was measured over time, and the chemistry of APPLs produced by each organism after different time intervals was compared. Chemical characterizations included assays for lignin, carbohydrate, and ash contents, molecular weight distributions by gel permeation chromatography, and chemical degradation analyses by permanganate oxidation, acidolysis, and alkaline ester hydrolysis. Differences between the organisms were observed in the cultural conditions required for APPL production and in the time courses of APPL accumulation. S. viridosporus produced APPL in solid-state fermentation over a 6- to 8-week incubation period, whereas S. badius produced as much or more APPL, but only in liquid culture and over a 7- to 8-day incubation period. The chemistry of the APPLs produced also differed. S. viridosporus APPL was more lignin-like than that of S. badius and was slowly modified further over time, although no change in molecular weight distribution over time was observed. In contrast, S. badius APPL was less lignin-like and increased substantially in average molecular weight over time. Results indicated that differing mechanisms of lignin metabolism may exist in these two Streptomyces sp. strains. S. viridosporus APPL probably originates from the heart of the lignin and is released largely as the result of beta-ether cleavage and other oxidative reactions. S. badius APPL probably originates in the same manner; however, after release as a water-soluble catabolite, lower-molecular-weight intermediates of lignin degradation are repolymerized with APPL in a reaction catalyzed by an extracellular phenol oxidase. The chemical analyses and the presence of extracellular phenol oxidase in S. badius, but not in S. viridosporus, support this conclusion.
O-methylation of chloroguaiacols has been examined in a number of gram-positive and gram-negative bacteria to elucidate the effects of substrate concentration, growth conditions, and cell density. Substrate concentrations between 0.1 and 20.0 mg liter were used, and it was found that (i) yields of the O-methylated products were significantly higher at the lowest concentrations and (ii) rates of O-methylation were not linear functions of concentration. With 3,4,5-trichloroguaiacol, the nature of the metabolites also changed with concentration. During growth with a range of substrates, O-methylation of chloroguaiacols also took place. With vanillate, however, de-O-methylation occurred: the chlorocatechol formed from 4,5,6-trichloroguaiacol was successively O-methylated to 3,4,5-trichloroguaiacol and 3,4,5-trichloroveratrole, whereas that produced from 4,5-dichloroguaiacol was degraded without O-methylation. Effective O-methylation in nonproliferating suspensions occurred at cell densities as low as 10 cells ml, although both the yields and the rates were lower than in more dense cultures. By using disk assays, it was shown that, compared with their precursors, all of the O-methylated metabolites were virtually nontoxic to the strains examined. It is therefore proposed that O-methylation functions as a detoxification mechanism for cells exposed to chloroguaiacols and chlorophenols. In detail, significant differences were observed in the response of gram-positive and gram-negative cell strains to chloroguaiacols. It is concluded that bacterial O-methylation is to be expected in the natural environment subjected to discharge of chloroguaiacols.
The effects of air drying soil on denitrifying enzyme activity, denitrifier numbers, and rates of N gas loss from soil cores were measured. Only 29 and 16% of the initial denitrifying enzyme activity in fresh, near field capacity samples of Maury and Donerail soils, respectively, were lost after 7 days of air drying. The denitrifying activity of bacteria added to soil and activity recently formed in situ were not stable during drying. When dried and moist soil cores were irrigated, evolution of N gas began, and it maximized sooner in the dried cores. This suggests that the persistence of denitrifying enzymes permits accelerated denitrification when dried soils are remoistened. Enzyme activity increased significantly in these waterlogged cores, but fluctuations in enzyme activity were small compared with fluctuations in actual denitrification rate, and enzyme activities were always greater than denitrification rates. Apparent numbers of isolatable denitrifiers (most-probable-number counts) decreased more than enzyme activity as the soils were dried, but after the soils were rewetted, the extent of apparent growth was not consistently related to the magnitude of N loss. We hypothesize that activation-inactivation of existing enzymes by soil O(2) is of greater significance in transient denitrification events than is growth of denitrifiers or synthesis of new enzymes.
[C-lignin]lignocellulose was solubilized by alkaline heat treatment and separated into different molecular size fractions for use as the sole source of carbon in anaerobic enrichment cultures. This study is aimed at determining the fate of low-molecular-weight, polyaromatic lignin derivatives during anaerobic degradation. Gel permeation chromatography was used to preparatively separate the original C-lignin substrate into three component molecular size fractions, each of which was then fed to separate enrichment cultures. Biodegradability was assessed by monitoring total carbon dioxide and methane production, evolution of labeled gases, loss of C-activity from solution, and changes in gel permeation chromatographic elution patterns. Results indicated that the smaller the size of the molecular weight fraction, the more extensive the degradation to gaseous end products. In addition, up to 30% of the entire soluble lignin-derived carbon was anaerobically mineralized to carbon dioxide and methane.
Anaerobic enrichment cultures acclimated for 2 years to use a C-labeled, lignin-derived substrate with a molecular weight of 600 as a sole source of carbon were characterized by capillary and packed column gas chromatography. After acclimation, several of the active methanogenic consortia were inhibited with 2-bromoethanesulfonic acid, which suppressed methane formation and enhanced accumulation of a series of metabolic intermediates. Volatile fatty acids levels in 2-bromoethanesulfonic acid-amended cultures were 10 times greater than those in the uninhibited, methane-forming consortia with acetate as the predominant component. Furthermore, in the 2-bromoethanesulfonic acid-amended consortia, almost half of the original substrate carbon was metabolized to 10 monoaromatic compounds, with the most appreciable quantities accumulated as cinnamic, benzoic, caffeic, vanillic, and ferulic acids. 2-Bromoethanesulfonic acid seemed to effectively block CH(4) formation in the anaerobic food chain, resulting in the observed buildup of volatile fatty acids and monoaromatic intermediates. Neither fatty acids nor aromatic compounds were detected in the oligolignol substrate before its metabolism, suggesting that these anaerobic consortia have the ability to mediate the cleavage of the beta-aryl-ether bond, the most common intermonomeric linkage in lignin, with the subsequent release of the observed constituent aromatic monomers.
The influence of readily degradable, naturally occurring carbon substrates on the biodegradation of several monosubstitued phenols (m-cresol, m-aminophenol, p-chlorophenol) was examined. The natural substrate classes used were amino acids, carbohydrates, and fatty acids. Samples of the microbial community from Lake Michie, a mesotrophic reservoir, were adapted to different levels of representatives from each natural substrate class in chemostats. After an extended adaptation period, the ability of the microbial community to degrade the monosubstituted phenols was determined by using a radiolabeled substrate uptake and mineralization method. Several microbiological characteristics of the communities were also measured. Adaptation to increasing concentrations of amino acids, carbohydrates, or fatty acids enhanced the ability of the microbial community to degrade all three phenols. The stimulation was largest for m-cresol and m-aminophenol. The mechanism responsible for the enhancement of monosubstituted phenol metabolism was not clearly identified, but the observation that adaptation to amino acids also increased the biodegradation of glucose and, to a lesser extent, naphthalene suggests a general stimulation of microbial metabolism. This study demonstrates that prior exposure to labile, natural substrates can significantly enhance the ability of aquatic microbial communities to respond to xenobiotics.
A methodology for reoxygenation of in situ benthic chamber systems by enzymatic catalysis of hydrogen peroxide with catalase was developed. For a 10-liter benthic chamber, the injection of 1 ml of catalase suspension (26,000 U ml) followed by 10 ml of 0.5 M hydrogen peroxide solution resulted in complete reoxygenation within 2.5 min at 25 degrees C.
A methanogenic coccoid organism, Methanosarcina mazei LYC, was isolated from alkaline sediment obtained from an oil exploration drilling site. The isolate resembled M. mazei S-6 by exhibiting different morphophases during its normal growth cycle. It differed from M. mazei S-6 by undergoint a spontaneous shift from large, irregular aggregates of cells to small, individual, irregular, coccoid units. In batch cultures at pH 7.0, M. mazei LYC grew as aggregates during the early growth stage. As the batch culture began exponential growth, the cell aggregates spontaneously dispersed: the culture liquid became turbid, and myriads of tiny (diameter, 1 to 3 mum) coccoid units were observed under phase-contrast microscopy. Disaggregation apparently was accomplished by the production of an enzyme which hydrolyzed the heteropolysaccharide component of the cell wall; the enzyme was active on other Methanosarcina strains as well. Although the enzyme was active when tested at pH 6.0, it apparently was not produced at that pH: when strain LYC was grown at pH 6.0, only cell aggregates were present throughout batch growth. Individual coccoid cells of M. mazei LYC were sensitive to sodium dodecyl sulfate, but the large aggregates of cells were not. Strain LYC rapidly used H(2)-CO(2), in addition to methanol, and mono-, di-, and trimethylamine as methanogenic substrates; acetate was used very slowly. Its optimum growth temperature was 40 degrees C, and its optimum pH was 7.2.
Erwinia amylovora infected with bacteriophage ERA103 produced an enzyme which degraded the extracellular polysaccharide of noninfected cells. The depolymerase enzyme was purified 15-fold by a procedure which included ammonium sulfate precipitation, ultracentrifugation, CM-Sephadex batchwise separation, Sephadex G-50 column chromatography, and Sephacryl S-200 column chromatography. The enzyme had a molecular weight of approximately 21,000 and a pH optimum of 6.0. Activity was enhanced by supplements of 2-mercaptoethanol or dithiothreitol.
The laccase of the fungus Trametes versicolor was able to polymerize various halogen-, alkyl-, and alkoxy-substituted anilines, showing substrate specificity similar to that of horseradish peroxidase, whereas the laccase of Rhizoctonia praticola was active only with p-methoxyaniline. The substrate specificities of the enzymes were determined by using gas chromatography to measure the decrease in substrate concentration during incubation. With p-chloroaniline as the substrate, the peroxidase and the Trametes laccase showed maximum activity near pH 4.2. The transformation of this substrate gave rise to a number of oligomers, ranging from dimers to pentamers, as determined by mass spectrometry. The product profiles obtained by high-pressure liquid chromatography were similar for the two enzymes. A chemical reaction was observed between p-chloroaniline and an enzymatically formed dimer, resulting in the formation of a trimer. All three enzymes oxidized p-methoxyaniline to 2-amino-5-p-anisidinobenzoquinone di-p-methoxyphenylimine, but only the T. versicolor laccase and the peroxidase caused the formation of a pentamer (2,5-di-p-anisidinobenzoquinone di-p-methoxyphenylimine). Our results demonstrate that in addition to horseradish peroxidase, a T. versicolor laccase can also polymerize aniline derivatives.
Natural denitrification rates and activities of denitrifying enzymes were measured in an agricultural soil which had a 20-year past history of low pH (pH ca. 4) due to fertilization with acid-generating ammonium salts. The soil adjacent to this site had been limed and had a pH of ca. 6.0. Natural denitrification rates of these areas were of similar magnitude: 158 ng of N g of soil day for the acid soil and 390 ng of N g of soil day at the neutral site. Estimates of in situ denitrifying enzyme activity were higher in the neutral soil, but substantial enzyme activity was also detected in the acid soil. Rates of nitrous oxide reduction were very low, even when NO(3) and NO(2) were undetectable, and were ca. 400 times lower than the rates of N(2)O production from NO(3). Denitrification rates measured in slurries of the acid and neutral soil showed distinctly different pH optima (pH 3.9 and pH 6.3) which were near the pH values of the two soils. This suggests that an acid-tolerant denitrifying population had been selected during the 20-year period of low pH.
Nutritional and physical factors affecting the decomposition of [C]lignocellulose prepared from Douglas fir (Pseudotsuga menziesii) were examined by incubating the labeled substrate with homogenized surface wood scrapings obtained from a Douglas fir log in a Pacific Northwest stream. Incubations were conducted in distilled water, in stream water collected from four different sources, or in a defined mineral salts solution with or without supplemental N (KNO(3)). Decomposition rates of [C]lignocellulose, as measured by CO(2) evolution, were greater in each of the four filter-sterilized sources of stream water than in distilled water alone. Decomposition experiments conducted in stream water media with the addition of defined mineral salts demonstrated that [C]cellulose decomposition was stimulated 50% by the addition of either KNO(3) or KH(2)PO(4)/K(2)HPO(4) and further enhanced (167%) by a combination of both. In contrast, [C]lignin decomposition was stimulated (65%) only by the addition of both N and P. Decomposition of [C]lignocellulose was greatest when supplemental KNO(3) was supplied in concentrations of at least 10.0 mg of N liter but not increased further by higher concentrations. The decomposition of [C]lignocellulose increased as the incubation temperature was raised and NO(3)-N supplementation further increased these rates between three-and sevenfold over the range of temperatures examined (5 to 22 degrees C). Accumulation of NH(4) (2 to 4 mg of N liter) was always observed in culture filtrates of incubations which had been supplemented with KNO(3), the quantity being independent of NO(3) concentrations >/= 10 mg of N liter. The role of supplemental NO(3) in the decomposition of [C]lignocellulose is discussed in relation to wood decomposition and the low concentrations of N found in stream ecosystems of the Pacific Northwest.
Clostridium thermosulfurogenes, an anaerobic bacterium which ferments starch into ethanol at 62 degrees C, produced an active extracellular amylase and contained intracellular glucoamylase but not pullulanase activity. The extracellular amylase was purified 2.4-fold, and its general physicochemical and catalytic properties were examined. The extracellular amylase was characterized as a beta-amylase (1,4-alpha-d-glucan maltohydrolase) based on demonstration of exocleavage activity and the production of maltose with a beta-anomeric configuration from starch. The beta-amylase activity was stable and optimally active at 80 and 75 degrees C, respectively. The pH optimum for activity and the pH stability range was 5.5 to 6 and 3.5 to 6.5, respectively. The apparent [S](0.5V) and V(max) for beta-amylase activity on starch was 1 mg/ml and 60 U/mg of protein. Similar to described beta-amylase, the enzyme was inhibited by p-chloromercuribenzoate, Cu, and Hg; however, alpha- and beta-cyclodextrins were not competitive inhibitors. The beta-amylase was active and stable in the presence of air or 10% (vol/vol) ethanol. The beta-amylase and glucoamylase activities enabled the organism to actively ferment raw starch in the absence of significant pullulanase or alpha-amylase activity.
The coenzyme F(420) content of granular sludge grown on various substrates and substrate combinations was measured, and the potential of the sludge to form methane (maximum specific methane production rate) from hydrogen, formate, acetate, propionate, and ethanol was determined. The F(420) content varied between 55 nmol g of volatile suspended solids (VSS) for sludge grown on acetate and 796 nmol g of VSS for sludge grown on propionate. The best correlation was found between the F(420) content and the potential activity for methane formation from formate; almost no correlation, however, was found with acetate as the test substrate. The ratio between the potential methanogenic activities (qch(4)) of sludges grown on various substrates and their F(420) content was in general highest for formate (48.2 mumol of CH(4) mumol of F(420) min) and lowest for propionate (6.9 mumol of CH(4) mumol of F(420) min) as test substrates. However, acetate-grown granular sludge with acetate as test substrate showed the highest ratio, namely, 229 mumol of CH(4) mumol of F(420) min. The data presented indicate that the F(420) content of methanogenic consortia can be misleading for the assessment of their potential acetoclastic methanogenic activity.
Cell extracts of Clostridium thermohydrosulfuricum, an anaerobic bacterium which ferments starch into ethanol at 65 degrees C, contained both pullulanase and glucoamylase activities. The general physiochemical and catalytic properties of these enzyme activities were compared. Pullulanase and glucoamylase activities were stable and optimally active at 85 and 75 degrees C, respectively. The pH optima for activity and pH stability ranges were, respectively, 5.5 to 6 and 4.5 to 5.5 for pullulanase and 4 to 6 and 5 to 6 for glucoamylase. The apparent [S](0.5v) and V(max) for pullulanase activity on pullulan were 0.33 mg/ml and 2.6 U/mg of protein. The apparent [S](0.5v) and V(max) for glucoamylase activity on starch were of 0.41 mg/ml and 0.31 U/mg of protein. These enzymes were active and stable in the presence of air or 10% (vol/vol) ethanol. These enzyme activities allowed the organism to actively degrade raw starch into glucose in the absence of significant alpha-amylase activity.
An unusual Xanthobacter sp., capable of independent growth on cyclohexane as the sole source of carbon and energy, has been isolated from soil by using classical enrichment techniques. The mean generation time for growth on cyclohexane was 6 h. The microorganism showed a limited ability to utilize hydrocarbons, with only alicyclic hydrocarbons closely related to cyclohexane supporting growth. Ultrastructural studies indicated the presence of electron-transparent vesicles in the cyclohexane-grown Xanthobacter sp., but the presence of complex intracytoplasmic membranes could not be identified. A soluble inducible enzyme capable of oxidizing cyclohexane was identified in cell extracts. This enzyme had a pH optimum of 6.5, an absolute specificity for NADPH, and a stoichiometric requirement for molecular O(2) which was consistent with the formation of cyclohexanol. The enzyme showed no activity towards straight chain alkanes and only a limited activity towards unsaturated ring compounds. Enzymatic studies with cell extracts have indicated the main route of metabolism of cyclohexane by this Xanthobacter sp. to proceed via cyclohexane --> cyclohexanol --> cyclohexanone --> 1-oxa-2-oxocycloheptane (epsilon-caprolactone) --> 6-hydroxyhexanoate (6-hydroxycaproate) --> --> adipic acid. Alternative routes involving initial double hydroxylation of the cyclohexane ring may operate fortuituously but are unlikely to represent major pathways for the dissimilation of cyclohexane by this microorganism.
A mutant, USDA 206C, of Rhizobium fredii USDA 206 was obtained by passage on acridine plates. This mutant was cured of its 197-megadalton Sym plasmid but retained its symbiotic effectiveness. Multiple plasmid and chromosomally borne nif gene copies have previously been shown in R. fredii USDA 206. HindIII and EcoRI restriction enzyme digests of plasmid and total DNA showed that at least two nif gene copies are probably missing in USDA 206C. To compare the symbiotic effectiveness of USDA 206 and USDA 206C, plant tests were carried out. Statistically significant differences were obtained for nodule number, nodule mass, nitrogenase activity per plant, nitrogenase specific activity, and total plant dry weight. There was an apparent correlation between loss of Sym plasmidborne nif gene copies and reduction of overall symbiotic effectiveness. Delayed nodulation by strain USDA 206C relative to strain USDA 206 also indicated an association with the loss of plasmidborne nodulation functions and the reduced symbiotic effectiveness of strain USDA 206C.
PR toxin and eremofortin C are secondary metabolites of Penicillium roqueforti. The chemical structures of these two compounds are closely related to each other and differ only by an aldehyde and an alcohol group at the C-12 position. In an effort to better understand the biosynthesis of PR toxin, we discovered the enzyme of P. roqueforti that is responsible for the transformation of eremofortin C to PR toxin. The maximum activity of the enzyme in the culture medium was found to occur on day 13, which corresponded to the maximal production of PR toxin in the medium. The enzyme was isolated and purified from the culture medium and the mycelium of the fungus, respectively, through a procedure involving ammonium sulfate fractionation and DEAE-cellulose chromatography. The specific activity increased 20- and 8-fold, respectively, and the yield was 33.3 and 21.6%, respectively, for the enzyme from the medium and mycelium. The optimal pH for the enzyme reaction was ca. pH 5.6. The enzyme reaction was temperature dependent. The rates followed a linear time course when it catalyzed the transformation at 30 degrees C and decayed with time when reacted at higher temperatures. At 100 degrees C, the enzyme activity was completely lost. The K(m) and V(max) of the enzyme as determined at 30 degrees C were 0.02 mM and 4.0 mumol/min per mg, respectively. The molecular weight of the enzyme was estimated by gel filtration on a high-pressure liquid chromatography I-250 protein column to be ca. 40,000.
High-rate anaerobic digestion can be applied in upflow anaerobic sludge blanket reactors for the treatment of various wastewaters. In upflow anaerobic sludge blanket reactors, sludge retention time is increased by a natural immobilization mechanism (viz. the formation of a granular type of sludge). When this sludge is cultivated on acid-containing wastewater, the granules mainly consist of an acetoclastic methanogen resembling Methanothrix soehngenii. This organism grows either in rods or in long filaments. Attempts to cultivate a stable sludge consisting predominantly of Methanosarcina sp. on an acetate-propionate mixture as substrate by lowering the pH from 7.5 during the start-up to approximately 6 failed. After 140 days of continuous operation of the reactor a filamentous organism resembling Methanothrix soehngenii prevailed in the sludge. The specific methanogenic activity of this sludge on acetate-propionate was optimal at pH 6.6 to 6.8 and 7.0 to 7.2, respectively.
The production of pellets of the fungus Agaricus campestris NRRL 2334 was studied in submerged fermentation with peat extract as the main substrate source. Pellets up to 6 mm in diameter were obtained when the peat extract was diluted to reduce the concentration of growth inhibitors. Yeast extract and yeast extract plus glucose were the most effective nutrient supplements in the diluted peat extract media and stimulated the formation of large pellets which contained 44.4% crude protein, 2.8% fat, and 9% ash (dry weight basis). No solid supports were required for the growth of the pellets. The effects on the growth morphology of several dilution ratios of the peat extract, rates of agitation and aeration, and time were investigated.
By cloning the beta-galactosidase gene of Bacillus stearothermophilus IAM11001 (ATCC 8005) into Bacillus subtilis, enzyme production was enhanced 50 times. beta-Galactosidase could be purified to 80% homogeneity by incubating the cell extract of B. subtilis at 70 degrees C for 15 min, followed by centrifugation to remove the denatured proteins. Because of its heat stability and ease of production, beta-galactosidase is suitable for application in industrial processes.
Candida wickerhamii NRRL Y-2563 expressed beta-glucosidase activity (3 to 8 U/ml) constitutively when grown aerobically in complex medium containing either glycerol, succinate, xylose, galactose, or cellobiose as the carbon source. The addition of a high concentration of glucose (>75 g/liter) repressed beta-glucosidase expression (<0.3 U/ml); however, this yeast did produce beta-glucosidase when the initial glucose concentration was </=50 g/liter. When grown aerobically in medium containing glucose plus the above-listed carbon sources, diauxic utilization of the carbon source was observed and the expression of beta-glucosidase was glucose repressed. Surprisingly, glucose repression did not occur when the cells were grown anaerobically. When grown anaerobically in medium containing 100 g of glucose per liter, C. wickerhamii produced 6 to 9 U of enzyme per ml and did not demonstrate diauxic utilization of glucose-cellobiose mixtures. To our knowledge, this is the first report of apparent derepression of a glucose-repressed enzyme by anaerobiosis.
This study considered the possibility of using water expressed during the drying of fuel-grade peat as a substrate for microbial growth. Highly humified peat pressed for 2.5 min at 1.96 MPa produced water with a chemical oxygen demand of 690 mg/liter. Several biological compounds could be produced by using the organic matter in expressed peat water as a substrate. These included polymers such as chitosan, contained in the cell wall of Rhizopus arrhizus, and two extracellular polysaccharides, xanthan gum and pullulan, produced by Xanthomonas campestris and Aureobasidium pullulans, respectively. A very effective surfactant was produced by Bacillus subtilis grown in the expressed water. Small additions of nutrients to the peat pressate were necessary to obtain substantial yields of products. The addition of peptone, yeast extract, and glucose improved production of the various compounds. Biological treatment improved the quality of the expressed water to the extent that in an industrial process it could be returned to the environment.
The envelope proteins of 5 strains of the genus Desulfotomaculum and 12 strains of the genus Desulfovibrio were studied by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and immunoblotting. The Desulfovibrio strains exhibited a typical gram-negative cell envelope, whereas the cell envelope of Desulfotomaculum strains appeared to be gram-positive. A close relationship between strains of Desulfotomaculum nigrificans was observed. A comparison between different species of Desulfotomaculum revealed some degree of similarity between Desulfotomaculum nigrificans and Desulfotomaculum ruminis, whereas Desulfotomaculum orientis seemed unique. The strains of Desulfovibrio salexigens were quite different from the strains of the other species of Desulfovibrio. In two of the strains of Desulfovibrio desulfuricans, a species-specific antigen was observed. The strains of Desulfovibrio vulgaris, Desulfovibrio africanus, and Desulfovibrio gigas and one strain of Desulfovibrio desulfuricans exhibited a similar outer membrane protein profile and also showed very similar antigenic reactions.
Clostridium acetobutylicum NRRL B527 and ATCC 824 exhibited extracellular and cell-bound endoglucanase and cellobiase activities during growth in a chemically defined medium with cellobiose as the sole source of carbohydrate. For both strains, the endoglucanase was found to be mainly extracellular (70 to 90%) during growth in continuous or batch cultures with the pH maintained at 5.2, whereas the cellobiase was mainly cell associated (60 to 90%). During continuous cultivation of strain B527 with cellobiose as the limiting nutrient, maximum production of the endoglucanase and cellobiase occurred at pH values of 5.2 and 4.8, respectively. In the carbon-limited continuous cultures, strain 824 produced similar levels of endoglucanase, cellobiosidase, and cellobiase activities regardless of the carbon source used. However, in ammonium- or phosphate-limited cultures, with an excess of glucose, only 1/10 of the endoglucanase was produced, and neither cellobiosidase nor cellobiase activities were detectable. A crude extracellular enzyme preparation from strain B527 hydrolyzed carboxymethylcellulose and phosphoric acid-swollen cellulose readily and microcrystalline cellulose (A vicel) to a lesser extent. Glucose accounted for more than 90% of the reducing sugar produced by the hydrolysis of acid-swollen cellulose and Avicel. Strain B527 did not grow in medium with acid-swollen cellulose as the sole source of carbohydrate, although it grew readily on the products obtained by hydrolyzing the cellulose in vitro with a preparation of extracellular cellulase derived from the same organism.
Acetivibrio cellulolyticus extracellular cellulase extensively hydrolyzed crystalline celluloses such as Avicel (FMC Corp., Food and Pharmaceutical Products Div., Philadelphia, Pa.) but only if it was desalted and supplemented with Ca. The Ca effect was one of increased enzyme stability in the presence of the ion. Although preincubation of the cellulase complex at 40 degrees C for 5 h without added Ca had a negligible effect on endoglucanase activity or on the subseqent hydrolysis of amorphous cellulose, the capacity of the enzyme to hydrolyze crystalline cellulose was almost completely lost. Adsorption studies showed that 90% of the Avicel-solubilizing component of the total enzyme preparation bound to 2% Avicel at 40 degrees C. Under these conditions, only 15% of the endoglucanase and 25% of the protein present in the enzyme preparation adsorbed to the substrate. The protein profile of the bound enzyme, as analyzed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, was complex and distinctly different from the profile observed for total cellulase preparations. The specific activity of A. cellulolyticus cellulase with respect to Avicel hydrolysis was compared with that of commercially available Trichoderma reesei cellulase.
The proteolytic enzymes pronase, trypsin, and chymotrypsin and the surfactant Triton X-100 inhibited attachment of Vibrio proteolytica to the hydrophobic substratum polystyrene by >97%. These treatments had no effect on attachment to hydrophilic substrata such as glass or tissue culture dishes. Both pronase and Triton X-100 effected the removal of previously attached cells from polystyrene but not from hydrophilic surfaces. Removal of cells from polystyrene by pronase left material (which we have termed footprints) that stained with the protein-specific stain Hoechst 2495 but not with the DNA-specific stain Hoechst 33342. Pronase treatment also caused a significant decrease in cell surface hydrophobicity as determined by phase partitioning in hexane or petroleum ether. Collectively, these results imply the existence of separate mechanisms for the adhesion of V. proteolytica to hydrophilic and hydrophobic substrata and suggest a role for protein in the latter mechanism.
Carbon mineralization was examined in Lakehurst and Atsion sands collected from the New Jersey Pinelands and in Pahokee muck from the Everglades Agricultural Area. Objectives were (i) to estimate the carbon mineralization capacities of acidic, xeric Pinelands soils in the absence of exogenously supplied carbon substrate (nonamended carbon mineralization rate) and to compare these activities with those of agriculturally developed pahokee muck, and (ii) to measure the capacity for increased carbon mineralization in the soils after carbon amendment. In most cases, nonamended carbon mineralization rates were greater in samples of the acid- and moisture-stressed Pinelands soils than in Pahokee muck collected from a fallow (bare) field. Carbon amendment resulted in augmented catabolic activity in Pahokee muck samples, suggesting that the microbial community was carbon limited in this soil. With many of the substrates, no stimulation of the catabolic rate was detected after amendment of Pinelands soils. This was documented by the observation that amendment of Pahokee muck with an amino acid mixture, glucose, or acetate resulted in a 3.0-, 3.9-, or 10.5-fold stimulation of catabolic activity, respectively, for the added substrate. In contrast, amendment of the Pinelands soils resulted in increased amino acid and acetate catabolic rates in Lakehurst sand and increased acetate metabolism only in Atsion sand. Other activities were unchanged. The increased glucose respiration rates resulted from stimulation of existing microbial activity rather than from microbial proliferation since no change in the microbial growth rate, as estimated by the rate of incorporation of C-labeled acetate into cell membranes, occurred after glucose amendment of the soils. A stimulation of microbial growth rate was recorded with glucose-amended Lakehurst sand collected from the B horizon.
A model of growth and substrate utilization for ferrous-iron-oxidizing bacteria attached to the disks of a rotating biological contactor was developed and tested. The model describes attached bacterial growth as a saturation function in which the rate of substrate utilization is determined by a maximum substrate oxidation rate constant (P), a half-saturation constant (K(s)), and the concentration of substrate within the rotating biological contactor (S(1)). The maximum oxidation rate constant was proportional to flow rate, and the substrate concentration in the reactor varied with influent substrate concentration (S(0)). The model allowed the prediction of metabolic constants and included terms for both constant and growth-rate-dependent maintenance energies. Estimates for metabolic constants of the attached population of acidophilic, chemolithotrophic, iron-oxidizing bacteria limited by ferrous iron were: maximum specific growth rate (mu(max)), 1.14 h; half-saturation constant (K(s)) for ferrous iron, 54.9 mg/liter; constant maintenance energy coefficient (m(1)), 0.154 h; growth-rate-dependent maintenance energy coefficient (m'), 0.07 h; maximum yield (Y(g)), 0.063 mg of organic nitrogen per mg of Fe(II) oxidized.
Pathways of glucose catabolism, potentially operational in six strains of obligately aerobic, acidophilic bacteria, including Acidiphilium cryptum strain Lhet2, were investigated by short-term radiorespirometry and enzyme assays. Short-term radiorespirometry was conducted at pH 3.0 with specifically labeled [C]glucose. The high rate and yield of C-1 oxidized to CO(2) indicated that the Entner-Doudoroff, pentose phosphate, or both pathways were operational in all strains. Apparent nonequivalent yields of CO(2) from C-1 and estimated CO(2) from C-4 (C-1 > C-4) were suggestive of simultaneous glucose catabolism by both pathways in all strains tested. Variation in the relative contribution of the two pathways of glucose catabolism appears to account for observed strain differences. Calculation of the actual percent pathway participation was not feasible. Enzyme assays were completed with crude extracts of glucose-grown cells to substantiate the results obtained by radiorespirometry. The key enzymes of the pentose phosphate pathway (6-phosphogluconate dehydrogenase) and the Entner-Doudoroff pathway (2-keto-3-deoxy-6-phosphogluconate aldolase and 6-phosphogluconate dehydrase) were present in all strains examined (PW2, Lhet2, KLB, OP, and QBP). However, none of the strains exhibited detectable levels of the key enzyme of the Embden-Meyerhof-Parnas pathway, 6-phosphofructokinase. All strains contained glucose-6-phosphate dehydrogenase and fructose bisphosphate aldolase. The results of the enzyme study supported the contention that the pentose phosphate and Entner-Doudoroff pathways are operational for glucose catabolism in the acidophilic heterotrophs, and that the Embden-Meyerhof-Parnas pathway is apparently absent.
We compared the metabolism of methanol and acetate when Methanosarcina barkeri was grown in the presence and absence of Desulfovibrio vulgaris. The sulfate reducer was not able to utilize methanol or acetate as the electron donor for energy metabolism in pure culture, but was able to grow in coculture. Pure cultures of M. barkeri produced up to 10 mumol of H(2) per liter in the culture headspace during growth on acetate or methanol. In coculture with D. vulgaris, the gaseous H(2) concentration was </=2 mumol/liter. The fractions of CO(2) produced from [C]methanol and 2-[C]acetate increased from 0.26 and 0.16, respectively, in pure culture to 0.59 and 0.33, respectively, in coculture. Under these conditions, approximately 42% of the available electron equivalents derived from methanol or acetate were transferred and were utilized by D. vulgaris to reduce approximately 33 mumol of sulfate per 100 mumol of substrate consumed. As a direct consequence, methane formation in cocultures was two-thirds that observed in pure cultures. The addition of 5.0 mM sodium molybdate or exogenous H(2) decreased the effects of D. vulgaris on the metabolism of M. barkeri. An analysis of growth and carbon and electron flow patterns demonstrated that sulfate-dependent interspecies H(2) transfer from M. barkeri to D. vulgaris resulted in less methane production, increased CO(2) formation, and sulfide formation from substrates not directly utilized by the sulfate reducer as electron donors for energy metabolism and growth.
A system was developed for the rapid characterization of microbial pectic enzyme complexes and then tested on Erwinia chrysanthemi and Sclerotium rolfsii. Pectic enzymes in minute samples of crude culture filtrates were resolved by ultrathin-layer polyacrylamide gel isoelectric focusing and sodium dodecyl sulfate-polyacrylamide gel electrophoresis and then assayed with an ultrathin pectate-agarose overlay stained with ruthenium red. The simple procedure can be completed within 30 min after isoelectric focusing, can detect extremely low levels of pectate lyase (6.4 x 10 mumol of product per min), and is sufficiently sensitive to determine the pectate lyase isozyme profile of a single bacterial colony with a diameter of 4 mm. Pectate lyases and polygalacturonases can be distinguished by altering buffer conditions in the overlays. The assay system revealed additional isozymes not resolved by classical techniques and generally corroborated the previously published isoelectric points and molecular weights of the pectate lyase isozymes and exo-poly-alpha-d-galacturonosidase produced by E. chrysanthemi and the endopolygalacturonase and exopolygalacturonase produced by S. rolfsii.
A 75-kilobase plasmid from Bacillus thuringiensis var. kurstaki (HD-244) was associated with the k-73 type insecticidal crystal protein production by mating into B. cereus and subsequent curing of excess plasmids. This plasmid was partially digested with endonuclease R . Sau3A and the fragments were cloned into Escherichia coli (HB101) on vector pBR322. Candidate clones were screened for plasmid vectors which contained the expected insert size (at least 3 kilobases) and then with an enzyme-linked immunosorbent assay, using antisera prepared against electrophoretically purified, solubilized insecticidal crystal protein of 130,000 daltons. Several positive clones were isolated and were analyzed for expression, toxicity, and genetic content by restriction enzyme analysis. Electrophoretic transfer blots of proteins from a candidate E. coli clone, analyzed by enzyme-linked immunosorbent assay, demonstrated a predominant cross-reacting protein of about 140,000 daltons. Ouchterlony analysis also showed a single precipitin band. Extensive bioassays with Manduca sexta larvae revealed that the E. coli clones make toxin with a specific activity (50% lethal dose per microgram of cross-reacting protein) equivalent to that of the parental B. thuringiensis strain or a B. cereus trancipient carrying the toxin-encoding, 75-kilobase plasmid.
To better understand the ecological significance of pressure effects on bacteria in the abyssobenthic boundary layer, experimental suspensions of sediments and sinking particulates were prepared from samples collected in boxcore and bottom-moored sediment traps at two stations (depth, 4,470 and 4,850m) in the Demerara abyssal plain off the coast of Brazil. Replicate samples were incubated shipboard at 3 degrees C and at both atmospheric and deep-sea pressures (440 or 480 atm [4.46 x 10 or 4.86 x 10 kPa]) following the addition of [C]glutamic acid (<10 mug liter) or yeast extract (0.025%) and the antibiotic nalidixic acid (0.002%). In seven of the eight samples supplemented with isotope, a barophilic microbial response was detected, i.e., substrate incorporation and respiration were greater under in situ pressure than at 1 atm (101.3 kPa). In the remaining sample, prepared from a sediment trap warmed to 24 degrees C before recovery, pressure was observed to inhibit substrate utilization. Total bacterial counts by epifluorescence microscopy decreased with depth in each sediment core, as did utilization of glutamic acid. Significant percentages of the total bacterial populations in cold sediment trap samples (but not the prewarmed one or any boxcore sample) were abnormally enlarged and orange fluorescing after incubation with yeast extract and nalidixic acid under deep-sea conditions. Results indicated that in the deep sea, barophilic bacteria play a predominant role in the turnover of naturally low levels of glutamic acid, and the potential for intense microbial activity upon nutrient enrichment is more likely to occur in association with recently settled particulates, especially fecal pellets, than in buried sediments.
Of 20 strains of Clostridium spp. screened, 17 hydrolyzed larch wood xylan. Two strains of Clostridium acetobutylicum, NRRL B527 and ATCC 824, hydrolyzed xylan but failed to grow on solid media with larch xylan as the sole carbon source; however, strain ATCC 824 was subsequently found to grow on xylan under specified conditions in a chemostat. These two strains possessed cellulolytic activity and were therefore selected for further studies. In cellobiose-limited continuous cultures, strain NRRL B527 produced maximum xylanase activity at pH 5.2. Strain ATCC 824 produced higher xylanase, xylopyranosidase, and arabinofuranosidase activities in chemostat culture with xylose than with any other soluble carbon source as the limiting nutrient. The activities of these enzymes were markedly reduced when the cells were grown in the presence of excess glucose. The xylanase showed maximum activity at pH 5.8 to 6.0 and 65 degrees C. The enzyme was stable on the alkaline side of pH 5.2 but was unstable below this pH value. The extracellular xylanolytic activity from strain ATCC 824 hydrolyzed 12% of the larch wood xylan during a 24-h incubation period, yielding xylose, xylobiose, and xylotriose as the major hydrolysis products. Strain ATCC 824, after being induced to grow in batch culture in xylan medium supplemented with a low concentration of xylose, failed to grow reproducibly in unsupplemented xylan medium. A mutant obtained by mutagenesis with ethyl methanesulfonate was able to grow reproducibly in batch culture on xylan. Both the parent strain and the mutant were able to grow with xylan as the sole source of carbohydrate in continuous culture with the pH maintained at either 5.2 or 6.0. Under these conditions, the cells utilized approximately 50% of the xylan.