首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
3.
We redemonstrate that SwrA is essential for swarming motility in Bacillus subtilis, and we reassert that laboratory strains of B. subtilis do not swarm. Additionally, we find that a number of other genes, previously reported to be required for swarming in laboratory strains, are dispensable for robust swarming motility in an undomesticated strain. We attribute discrepancies in the literature to a lack of reproducible standard experimental conditions, selection for spontaneous swarming suppressors, inadvertent genetic linkage to swarming mutations, and auxotrophy.Many species of bacteria are capable of flagellum-mediated swimming motility in liquid broth. Of those species, a subset is also capable of a related, but genetically separable, form of flagellum-mediated surface movement called swarming motility (17). Examples of swarming-proficient species include Proteus mirabilis, Vibrio parahaemolyticus, Serratia marcescens, Escherichia coli, Salmonella enterica, and Bacillus subtilis (1, 15, 16, 20, 28). In general, swarming requires a surfactant or wetting agent to reduce surface tension, an increase in flagellar number per cell, and other genetic features that are distinct from swimming (7, 14).There is confusion in the literature concerning the genetic requirements of the swarming phenotype of B. subtilis. It is generally accepted that the ancestral undomesticated strain B. subtilis 3610 exhibits robust swarming motility (18, 20, 33). Swarming motility of strain 3610 requires the production of a secreted surfactant, called surfactin (6, 20), to reduce surface tension and permit surface spreading, and it also requires the protein SwrA to activate flagellar biosynthesis gene expression and increase the number of flagella on the cell surface (5, 20). Some reports claim that domesticated derivatives of 3610, such as the commonly used laboratory strain 168, are also swarming proficient (10, 18, 19, 24). Strain 168, however, is defective in both surfactin production (9, 25) and SwrA (5, 21, 31), and thus, swarming 168 strains challenge the genetic definition of swarming motility. Our lab has never observed swarming in laboratory strains, and here we investigated swarming motility in a reportedly swarming-proficient 168 strain.We obtained a reportedly swarming-proficient 168 strain (13) (generous gift of Simone Séror, Orsay University, Paris-Sud, France) (Table (Table1)1) and compared its swarming phenotype to that of 3610 under our standard conditions (20). Swarm plates were prepared one day prior to use with 25 ml of LB medium (10 g Bacto tryptone, 5 g Bacto yeast extract, 5 g NaCl per liter) fortified with 0.7% Bacto agar. To minimize water on the agar surface and thus minimize the potentially confounding influence of swimming motility, plates were dried 20 min prior to inoculation and 10 min postinoculation open-faced in a laminar flow hood. For qualitative swarm assays, plates were centrally inoculated with cells from a freshly grown overnight colony using a sterile stick. For quantitative swarm expansion assays, 1 ml of cells grown to mid-exponential phase (optical density at 600 nm [OD600], 0.5) was resuspended in PBS buffer (8 g NaCl, 0.2 g KCl, 1.44 g Na2HPO4, 0.24 g KH2PO4 per liter, pH 7.0) containing 0.5% India ink (Higgins) to an OD600 of 10 and centrally spotted (10 μl). Swarm expansion was measured at 0.5-h intervals along a transect on the plate. Plates were incubated at 37°C in 20 to 30% humidity. Whereas strain 3610 was swarming proficient, strain 168 (Orsay) was swarming deficient (Fig. (Fig.1A).1A). Thus, strain 168 (Orsay) appeared to behave similarly to all other laboratory strains we have tested previously (20, 21).Open in a separate windowFIG. 1.Swarming motility on LB and B media. In qualitative plate images, colonized agar appears white and uncolonized agar appears black on LB and B media, as indicated. Swarming cells colonize a larger surface area than nonswarming cells. All strains are derivatives of strain 3610 unless otherwise indicated. Bar, 2 cm. (A) Quantitative swarm expansion assays on solid medium and growth in liquid medium of the indicated strains on LB medium (closed symbols) and on B medium (open symbols). To indicate variability in a particular experiment, we have reproduced the quantitative swarm expansion assay of strain 3610 on LB and B media with error bars in Fig. S5 in the supplemental material. (B) Quantitative swarm expansion assays on LB (closed symbols) and B (open symbols) media. The following strains were used: DS3337 (sfp), DS2415 (swrA), DS5106 (168 swrA+), DS5758 (168 sfp+), and DS5759 (168 swrA+ sfp+). In all assays, B medium was made according to reference 2 except for strain DS5759, for which B medium was supplemented with 780 μM threonine to compensate for thrC auxotrophy. (C) Swarm plates of the indicated strains on LB medium made with equal parts peptone instead of tryptone. (D) Quantitative swarm expansion assays of the indicated 3610-derived mutant strains on LB medium (closed symbols) and on B medium (open symbols). The following strains were used: DS72 (yvzB), DS2268 (epr), DS3903 (phrC), DS4978 (rapC), DS4979 (oppD), DS2509 (swrB), and DS3649 (degU). All points are averages for three replicates.

TABLE 1.

Strains
StrainGenotypea
168trpC2 swrA sfp (13)
3610Wild type
DS72yvzB::tet (21)
DS2268epr::kan
DS2415ΔswrA
DS2509ΔswrB
DS3337sfp::mls
DS3649ΔdegU
DS3903phrC::spec
DS4978rapC::spec
DS4979oppD::kan
DS5106168 trpC2 swrA sfp amyE::PswrA-swrA cat
DS5758168 trpC2 swrA sfp amyE::sfp+ cat
DS5759168 trpC2 swrA sfp amyE::PswrA-swrA cat thrC::sfp+ mls
Open in a separate windowaAll strains are in the 3610 genetic background unless otherwise indicated.We next explored the genetic basis for the swarming defect we observed in strain 168 (Orsay). As with other laboratory strains, colonies of strain 168 (Orsay) failed to produce the transparent ring normally indicative of surfactin production, due to a mutation of the gene sfp (25). Complementation with the wild-type sfp gene in 168 was sufficient to restore surfactin production but was insufficient to restore swarming motility (Fig. (Fig.1B)1B) (20). Laboratory strains also fail to swarm because of a loss-of-function frameshift mutation in the gene encoding SwrA (5, 21). Sequencing of the swrA gene confirmed that strain 168 (Orsay) contained the frameshift mutation, but introduction of a swrA complementation construct at an ectopic site in the chromosome (amyE::PswrA-swrA) was also insufficient to restore swarming motility (Fig. (Fig.1B).1B). Swarming motility was fully rescued, however, when sfp and swrA were simultaneously complemented in the 168 strain (Fig. (Fig.1B)1B) or when the swrA frameshift mutation was repaired in spontaneous suppressors isolated from 168 complemented with sfp alone (see Fig. S1 in the supplemental material). Furthermore, mutation of either sfp or swrA in the 3610 genetic background abolished swarming (Fig. (Fig.1B).1B). We conclude that Sfp and SwrA are necessary for swarming. We further conclude that, with respect to swarming motility, strain 168 (Orsay) is genetically no different from any other laboratory strain we have tested, as it fails to swarm due to simultaneous defects in Sfp and SwrA (21). We infer that the apparent swarming observed in some laboratory strains is not due to genetic differences but rather due to differences in experimental conditions.In our swarming assays, we take steps to minimize surface water. In some cases of the reported swarming of strain 168, plates were poured 1 h before use, dried for 5 min, and incubated at 60 to 70% humidity (13). When 0.7% agar LB plates were freshly poured and not dried, we noticed that toothpick inoculation of the cells disturbed the agar surface and caused a pool of water to well forth from the agar (see Fig. S2 in the supplemental material). Pools of water emerged even when the plates were dried for 5 or 10 min prior to inoculation, but water did not emerge when the plates were dried for 15 min or longer (see Fig. S2 in the supplemental material). The colony size of strain 168 was proportional to the amount of water extracted from the agar, but the cells did not exhibit swarming motility (see Fig. S2 in the supplemental material). We conclude that excess water was not sufficient to promote swarming of the laboratory strain. Nonetheless, we recommend drying plates for 20 min prior to inoculation to minimize any contribution of swimming motility to apparent surface migration.Another difference in experimental conditions may concern the nutritional content of the medium. Some labs have tested swarming motility on LB medium in which tryptone was replaced by an equal amount of peptone (13). We reproduced the “LB” medium containing peptone and found that whereas strain 3610 was swarming proficient, strain 168 was swarming deficient (Fig. (Fig.1C).1C). Thus, the peptone substitution did not promote swarming in lab strains.Some labs have also reported swarming of laboratory strains on a defined medium called B medium [15 mM (NH4)2SO4, 8 mM MgSO4·7H2O, 27 mM KCl, 7 mM sodium citrate·H2O, 50 mM Tris·HCl (pH 7.5), 2 mM CaCl2·2H2O, 1 μM FeSO4·7H2O, 10 μM MnSO4·4H2O, 0.6 mM KH2PO4, 4.5 mM glutamic acid, 860 μM lysine, 780 μM tryptophan, and 0.5% glucose) (2, 13, 18, 19). In our hands, 3610 was swarming proficient on B medium, but strain 168 was swarming deficient (Fig. (Fig.1A).1A). We conclude that altering medium composition was insufficient to promote swarming of laboratory strains. Furthermore, mutation of either sfp or swrA rendered strain 3610 nonswarming on B medium, and complementation of sfp and swrA restored B medium swarming to strain 168 (Fig. (Fig.1B).1B). We conclude that the genetic requirements for swarming are the same for both LB and B medium.On undefined rich LB medium, strain 3610 swarmed rapidly as a featureless monolayer, whereas on defined B medium, it swarmed in a branched dendritic pattern (18, 20) (Fig. (Fig.1A).1A). In addition, the growth rate of 3610 in liquid B medium and swarm rate on solid B medium were both reduced fivefold relative to comparable assays with LB (Table (Table2),2), suggesting that the rate of swarming and the rate of growth were related. To further explore the connection between growth rate and swarming rate, we performed swarm expansion assays at lower temperatures. At 30°C, the growth rate in LB broth was reduced 2.5-fold relative to 37°C, and the swarming rate on LB agar was reduced 2.5-fold as well (Table (Table2;2; also, see Fig. S3 in the supplemental material). We conclude that swarming rate is correlated with growth rate. We infer that differences in growth may account for differences in swarm patterns (11). We note that regardless of the medium composition or the growth rate, the duration of the lag prior to swarming initiation was relatively constant.

TABLE 2.

Growth rates and swarm ratesa
MediumTemp (°C)Swarm rate (mm/h)Growth rate (generations/h)Reduction inb:
Swarm rateGrowth rate
LB37153.511
LB3061.42.52.5
B3730.855
Open in a separate windowaStrain 3610 was used to generate all data.bRelative to cells cultured in LB at 37°C (standard conditions).Ultimately we were unable to reproduce swarming in laboratory strains, and we reassert that laboratory strains are defective for swarming-motility. It is difficult to explain reports of swarming-proficient laboratory strains, because these cells are defective for both surfactin and swrA. Thus, the apparent swarming of strain 168 must be due to poorly reproducible environmental factors and/or selection for genetic revertants.  相似文献   

4.
The biological, serological, and genomic characterization of a paramyxovirus recently isolated from rockhopper penguins (Eudyptes chrysocome) suggested that this virus represented a new avian paramyxovirus (APMV) group, APMV10. This penguin virus resembled other APMVs by electron microscopy; however, its viral hemagglutination (HA) activity was not inhibited by antisera against any of the nine defined APMV serotypes. In addition, antiserum generated against this penguin virus did not inhibit the HA of representative viruses of the other APMV serotypes. Sequence data produced using random priming methods revealed a genomic structure typical of APMV. Phylogenetic evaluation of coding regions revealed that amino acid sequences of all six proteins were most closely related to APMV2 and APMV8. The calculation of evolutionary distances among proteins and distances at the nucleotide level confirmed that APMV2, APMV8, and the penguin virus all were sufficiently divergent from each other to be considered different serotypes. We propose that this isolate, named APMV10/penguin/Falkland Islands/324/2007, be the prototype virus for APMV10. Because of the known problems associated with serology, such as antiserum cross-reactivity and one-way immunogenicity, in addition to the reliance on the immune response to a single protein, the hemagglutinin-neuraminidase, as the sole base for viral classification, we suggest the need for new classification guidelines that incorporate genome sequence comparisons.Viruses from the Paramyxoviridae family have caused disease in humans and animals for centuries. Over the last 40 years, many paramyxoviruses isolated from animals and people have been newly described (16, 17, 22, 29, 31, 32, 36, 42, 44, 46, 49, 58, 59, 62-64). Viruses from this family are pleomorphic, enveloped, single-stranded, nonsegmented, negative-sense RNA viruses that demonstrate serological cross-reactivity with other paramyxoviruses related to them (30, 46). The subfamily Paramyxovirinae is divided into five genera: Respirovirus, Morbillivirus, Rubulavirus, Henipavirus, and Avulavirus (30). The Avulavirus genus contains nine distinct avian paramyxovirus (APMV) serotypes (Table (Table1),1), and information on the discovery of each has been reported elsewhere (4, 6, 7, 9, 12, 34, 41, 50, 51, 60, 68).

TABLE 1.

Characteristics of prototype viruses APMV1 to APMV9 and the penguin virus
StrainHostDiseaseDistributionFusion cleavagecGI accession no.
APMV1/Newcastle disease virus>250 speciesHigh mortalityWorldwideGRRQKRF45511218
InapparentWorldwideGGRQGRLa11545722
APMV2/Chicken/CA/Yucaipa/1956Turkey, chickens, psittacines, rails, passerinesDecrease in egg production and respiratory diseaseWorldwideDKPASRF169144527
APMV3/Turkey/WI/1968TurkeyMild respiratory disease and moderate egg decreaseWorldwidePRPSGRLa209484147
APMV3/Parakeet/Netherlands/449/1975Psittacines, passerines, flamingosNeurological, enteric, and respiratory diseaseWorldwideARPRGRLa171472314
APMV4/Duck/Hong Kong/D3/1975Duck, geese, chickensNone knownWorldwideVDIQPRF210076708
APMV5/Budgerigar/Japan/Kunitachi/1974Budgerigars, lorikeetsHigh mortality, enteric diseaseJapan, United Kingdom, AustraliaGKRKKRFa290563909
APMV6/Duck/Hong Kong/199/1977Ducks, geese, turkeysMild respiratory disease and increased mortality in turkeysWorldwidePAPEPRLb15081567
APMV7/Dove/TN/4/1975Pigeons, doves, turkeysMild respiratory disease in turkeysUnited States, England, JapanTLPSSRF224979458
APMV8/Goose/DE/1053/1976Ducks, geeseNone knownUnited States, JapanTYPQTRLa226343050
APMV9/Duck/NY/22/1978DucksNone knownWorldwideRIREGRIa217068693
APMV10/Penguin/Falkland Islands/324/2007Rockhopper penguinsNone KnownFalkland IslandsDKPSQRIa300432141
Open in a separate windowaRequires the addition of an exogenous protease.bProtease requirement depends on the isolate examined.cPutative.Six of these serotypes were classified in the latter half of the 1970s, when the most reliable assay available to classify paramyxoviruses was the hemagglutination inhibition (HI) assay (61). However, there are multiple problems associated with the use of serology, including the inability to classify some APMVs by comparing them to the sera of the nine defined APMVs alone (2, 8). In addition, one-way antigenicity and cross-reactivity between different serotypes have been documented for many years (4, 5, 14, 25, 29, 33, 34, 41, 51, 52, 60). The ability of APMVs, like other viruses, to show antigenic drift as it evolves over time (37, 43, 54) and the wide use and availability of precise molecular methods, such as PCR and genome sequencing, demonstrate the need for a more practical classification system.The genetic diversity of APMVs is still largely unexplored, as hundreds of avian species have never been surveyed for the presence of viruses that do not cause significant signs of disease or are not economically important. The emergence of H5N1 highly pathogenic avian influenza (HPAI) virus as the cause of the largest outbreak of a virulent virus in poultry in the past 100 years has spurred the development of surveillance programs to better understand the ecology of avian influenza (AI) viruses in aquatic birds around the globe, and in some instances it has provided opportunities for observing other viruses in wild bird populations (15, 53). In 2007, as part of a seabird health surveillance program in the Falkland Islands (Islas Malvinas), oral and cloacal swabs and serum were collected from rockhopper penguins (Eudyptes chrysocome) and environmental/fecal swab pools were collected from other seabirds.While AI virus has not yet been isolated from penguins in the sub-Antarctic and Antarctic areas, there have been two reports of serum antibodies positive to H7 and H10 from the Adélie species (11, 40). Rare isolations of APMV1, both virulent (45) and of low virulence (8), have been reported from Antarctic penguins. Sera positive for APMV1 and AMPV2 have also been reported (21, 24, 38, 40, 53). Since 1981, paramyxoviruses have been isolated from king penguins (Aptenodytes patagonicus), royal penguins (Eudyptes schlegeli), and Adélie penguins (Pygoscelis adeliae) from Antarctica and little blue penguins (Eudyptula minor) from Australia that cannot be identified as belonging to APMV1 to -9 and have not yet been classified (8, 11, 38-40). The morphology, biological and genomic characteristics, and antigenic relatedness of an APMV recently isolated from multiple penguin colonies on the Falkland Islands are reported here. Evidence that the virus belongs to a new serotype (APMV10) and a demonstration of the advantages of a whole genome system of analysis based on random sequencing followed by comparison of genetic distances are presented. Only after all APMVs are reported and classified will epidemiological information be known as to how the viruses are moving and spreading as the birds travel and interact with other avian species.  相似文献   

5.
Degenerate primers for the [FeFe] hydrogenase (hydA) were developed and used in PCRs to examine hydA in microbial mats that inhabit saltern evaporative ponds in Guerrero Negro (GN), Mexico. A diversity of deduced HydA was discovered that revealed unique variants, which may reflect adaptation to the environmental conditions present in GN.Hydrogen (H2) is an important intermediate in the decomposition of organic matter in anaerobic environments and is the basis for many syntrophic interactions that occur in the food chain such as those between acetogens and methanogens and/or sulfate reducers. Little is known concerning the potential role of fermentative bacteria and/or H2 cycling in phototrophic microbial mat systems, such as the 6-cm-thick and extremely diverse microbial mats inhabiting the saline ponds at the Exportadora de Sal SA saltern in Guerrero Negro (GN), Baja California Sur, Mexico. Recent 16S rRNA gene surveys in GN revealed an abundance of bacteria in the upper 2 mm of the mat that, based on their phylogenetic affiliation, are thought to harbor fermentative metabolisms (6, 8, 13, 16, 20, 28). Similarly, previous studies have shown that the flux of H2 from the surface of the phototrophic mats present in GN is ∼150-fold higher during the night than during the day (9). Together, these results suggest the potential for fermentative metabolisms in H2 and carbon cycling in the GN saline mat environment.The [FeFe] hydrogenase has a central role in primary and secondary substrate fermentations, catalyzing the oxidation of excess reducing equivalents coupled to the reduction of protons, yielding H2 (2, 17, 29, 30). The [FeFe] hydrogenase is only known to occur in anaerobic bacteria and a small number of green algae (23, 29, 30), making this a useful biomarker for examining the diversity and distribution of this functional class of organism. Eighty-two putative [FeFe] hydrogenase large-subunit protein sequences (HydA) present in the GenBank database that represent the known diversity of putative HydA were screened for the L1 ([FLI]TSC[C/S]P[GAS]W[VIQH]) and L2 ([IVLF]MPCx[ASRD]K[KQ]xE) (conserved residues are in bold and underlined, and bracketed positions indicate “semiconserved” residues at that position) signature sequence motifs (17, 30). Degenerate PCR primers, corresponding to positions 272 to 279 (AD[M/L]TIMEE) and positions 420 to 427 (TGGVMEAA) in the Clostridium pasteurianum protein sequence, were designed for use in specific gene amplifications.Mat core samples (1 by 6 cm) were collected at 2:00 pm from pond 4 (35°C, pH 8.0, and 80 ppt salt) near pond 5 at the Exportadora de Sal saltworks, GN, on 13 February 2005. Mat cores were sectioned in 1-mm increments on site and immediately flash frozen in liquid nitrogen as previously described (16). Genomic DNA was extracted by bead beating in the presence of phenol-chloroform-isoamyl alcohol and sodium dodecyl sulfate as previously described (7) and was quantified by using the High DNA Mass Ladder (Invitrogen, Carlsbad, CA). Approximately 500-bp fragments of hydA were amplified in triplicate using primers FeFe-272F (5′-GCHGAYMTBACHATWATGGARGA-3′, where H = A, C, T; Y = C, T; M = A, C; B = C, G, T; W = A, T; R = A, G; 432-fold degeneracy) and FeFe-427R (5′-GCNGCYTCCATDACDCCDCCNGT-3′, where N = A, C, T, G; Y = C, T; D = A, G, T; 864-fold degeneracy) and 35 cycles of PCR as previously described (4). An empirically determined annealing temperature of 56.5°C, a MgCl2 concentration of 1.5 mM, and a primer concentration of 1 μM for each forward and reverse primer were utilized in 50-μl PCR mixtures containing 10 ng of genomic DNA as the template. Equal 40-μl volumes of three replicate PCR products were pooled, purified using the Promega Wizard purification kit (Madison, WI), quantified using the Low Mass DNA Ladder (Invitrogen), cloned using the pGEM Easy Vector System (Promega), and sequenced by using the M13F-M13R primer pair as previously described (5).Deduced amino acid sequences were screened for the presence of the L1 and L2 HydA sequence motifs as described above. ClustalX (version 2.0.8) (15) was employed to align inferred amino acid sequences and to create distance matrices for use in identifying and clustering operational taxonomic units (OTUs) with DOTUR (26). Calculations of Shannon diversity and Chao1 and Ace1 deduced amino acid sequence richness were completed with DOTUR by using the furthest-neighbor algorithm and a precision of 0.01. The phylogenetic position of putative HydA was assessed with MRBAYES (10, 25). Tree topologies were sampled every 500 generations for 2,000,000 generations (burnin = 1,000,000) by using the WAG evolutionary model with fixed amino acid frequencies and gamma-shaped rate variation with a proportion of invariable sites as recommended by ProtTest (1). The Saccharomyces cerevisiae Narf protein, a homolog of HydA, was used as the outgroup. The phylogenetic tree was projected using TreeView (version 1.6.6) (21).An unexpectedly diverse assemblage of putative HydA variants was present in the top 1 mm of the GN microbial mat, corresponding to the photic zone. At a sequence identity threshold (SIT) of 100%, the predicted Shannon diversity index was 3.84 and the mean Chao1 HydA richness was 187 unique OTUs (see Fig. S1a in the supplemental material). When a 99% SIT was applied, the predicted Chao1 richness estimate decreased 56% to 82 unique OTUs and the Shannon diversity index decreased slightly to 3.55 (data not shown). A further decrease in the SIT from 99 to 85% resulted in a slight decrease (23%) in the Chao1 predicted richness and a modest decrease in the Shannon diversity index from 3.55 to 3.28, suggesting that the majority of the putative HydA OTUs present in the top 1 mm of the mat is a result of recent evolution or speciation.The putative HydA variants from GN resolved into 42 OTUs which clustered into one of seven distinct phylogenetic clusters (Fig. (Fig.1A),1A), although clusters 5 and 7 could not be adequately resolved. Intriguingly, the mean identity scores determined from sequence alignment matrices for members of each of the seven sequence clusters recovered from GN, when compared to sequences in the GenBank database, were very low (47.5 to 73.2%) (Table (Table1),1), suggesting that the putative HydA sequences recovered from GN represent novel sequence space.Open in a separate windowFIG. 1.(A) Phylogenetic tree based on deduced putative HydA amino acid sequences from GN and reference HydA sequences. Nodes with posterior probabilities of <50% were collapsed in this analysis (posterior probabilities of 100 are denoted by asterisks). Bar, one substitution per 10 sites. (B) Partial-length deduced amino acid sequences of environmental and reference deduced HydA sequences illustrating the phylogenetic coherence of novel substitutions and insertions in and upstream of L1 sequence motif. Cluster designations (C1 to C7) correspond to those presented in Table Table11.

TABLE 1.

Phylogeny of partial-length putative HydA sequence clusters
Sequence clusteraNo. of clones in clusterMaximal intracluster sequence divergence (%)bMost closely related sequencecPhylumdMean % sequence identity (range)e
C12249.4Moorella thermoacetica ATCC 39073Firmicutes61.8 (56.4-67.3)
C21641.6Opitutus terrae PB90-1Verrucomicrobia64.3 (56.3-71.1)
C3628.8Bacteroides thetaiotaomicron VPI-5482Bacteroidetes72.1 (69.8-73.2)
C4938.6Alkaliphilus oremlandii OhILAsFirmicutes65.1 (61.4-66.4)
C5338.6Moorella thermoacetica ATCC 39073Firmicutes61.4 (NA)
C6751.8Heliobacillus mobilisFirmicutes50.3 (47.5-54.6)
C7232.9Thermoanaerobacterium saccharolyticum JW/SL-YS485Firmicutes61.2 (58.5-64.2)
Open in a separate windowaSequence clusters correspond to those presented in Fig. Fig.11.bMaximum intracluster sequence divergence among sequences within a cluster as calculated by the ClustalX sequence identity matrix following alignment using the Gonnet protein weight matrix (gap opening penalty of 13, gap extension penalty of 0.05).cThe most closely related sequence is defined as the taxon with the HydA sequence most closely related to that of members within the cluster as presented in Fig. Fig.11.dPhylum designation of most closely related HydA sequence as defined in Fig. Fig.11.eSequence identity means and ranges were calculated by using the ClustalX sequence identity matrix as described in footnote b. NA, not available.The highest proportion of putative HydA sequences recovered from the top 1-mm transect of the GN microbial mat were related to acetogens most closely affiliated with the Firmicutes (66.2% of the clones) and the Verrucomicrobia (24.6%), with a lower proportion related to the Bacteroidetes (9.2%). These observations support those of previous studies that identified the highest abundance of 16S rRNA genes related to fermentative/acetogenic organisms within the phyla Firmicutes, Verrucomicrobia, and Bacteroidetes in the upper few millimeters of microbial mat (16), coinciding with the mat transect where nighttime H2 concentrations were also elevated (9). The recovery of putative HydA sequences in the photic zone of the GN microbial mat is consistent with the results of previous studies conducted in the phototrophic mats in Yellowstone National Park, WY, which documented the potential for the secondary fermentation of cyanobacterial primary fermentation products in the top photic layers of the mat at night (3). Importantly, cyanobacterial fermentation may also contribute to the production of H2 in the mats during periods of anoxia (9).Deduced amino acid sequences from GN had significantly lower (P = <0.01 for both GRAVY and aliphatic indices) hydropathy indices than nonhalotolerant HydA sequences from GenBank (see Table S1 in the supplemental material), a feature which may reflect molecular adaptation to salinity. These observations are consistent with the results of previous investigations of deduced amino acid sequences from a variety of halophilic organisms that indicate hydrophobicity indices lower than those of their nonhalophilic counterparts, which is hypothesized to reduce the effects of salt-driven protein misfolding and/or aggregation (11, 12, 14, 24, 27, 31). In addition, many of the putative HydA sequences recovered from GN exhibited previously unobserved substitutions in the L1 motif and novel insertion domains upstream from the L1 motif (Fig. (Fig.1B).1B). Residues within L1 are involved in the coordination of the oxygen-labile [4Fe-4S] subcluster of the H cluster of HydA (18, 19, 22). Substitutions in this region may have implications for the redox properties of this [4Fe-4S] cluster and thus may be important in conferring stability in the presence of oxygen (23).The discovery of a rich assemblage of putative HydA variants in the top 1 mm of the GN microbial mat underscores the utility of these primers in examining the natural diversity of putative HydA and fermentative organisms in complex microbial ecosystems. The composition of putative HydA sequences from the top 1 mm of the microbial mat suggests adaption to the conditions present in the evaporative salterns in GN.  相似文献   

6.
The human gut microbe Bacteroides fragilis can alter the expression of its surface molecules, such as capsular polysaccharides and SusC/SusD family outer membrane proteins, through reversible DNA inversions. We demonstrate here that DNA inversions at 12 invertible regions, including three gene clusters for SusC/SusD family proteins, were controlled by a single tyrosine site-specific recombinase (Tsr0667) encoded by BF0667 in B. fragilis strain YCH46. Genetic disruption of BF0667 diminished or attenuated shufflon-type DNA inversions at all three susC/susD genes clusters, as well as simple DNA inversions at nine other loci, most of which colocalized with susC/susD family genes. The inverted repeat sequences found within the Tsr0667-regulated invertible regions shared the consensus motif sequence AGTYYYN4GDACT. Tsr0667 specifically mediated the DNA inversions of 10 of the 12 regions, even under an Escherichia coli background when the invertible regions were exposed to BF0667 in E. coli cells. Thus, Tsr0667 is an additional globally acting DNA invertase in B. fragilis, which probably involves the selective expression of SusC/SusD family outer membrane proteins.The human gut harbors an abundant and diverse microbiota. Bacteroides is one of the most abundant genera of human gut microflora (10, 17, 20), and the biological activities of Bacteroides species are deeply integrated into human physiology through nutrient degradation, the production of short-chain fatty acids, or immunomodulatory molecules (11-14, 24). Recent genomic analyses of Bacteroides have revealed that the bacteria possess redundant abilities not only to bind and degrade otherwise indigestible dietary polysaccharides but also to produce vast arrays of capsular polysaccharide (5, 19, 38, 39). These functional redundancies have been established by the extensive duplication of various genes that encode molecules such as glycosylhydrolases, glycosyltransferases, and outer membrane proteins of the SusC/SusD family (starch utilization system) known to be involved in polysaccharide recognition and transport (7, 27, 28, 30). It has been assumed that these functional redundancies of Bacteroides contribute to the stability of the gut ecosystem (3, 21, 23, 32, 39).Another characteristic feature common in Bacteroides species is that the expression of some of the genes is altered in an on-off manner by reversible DNA inversions at gene promoters or within the protein-coding regions (5, 9, 19, 38, 39). These phase-variable phenotypes are associated with surface architectures such as capsular polysaccharides and SusC/SusD family proteins (5, 6, 16, 19). Our previous genomic analyses of Bacteroides fragilis strain YCH46 revealed that it contained as many as 31 invertible regions in its chromosome (19). These invertible regions can be grouped into six classes according to the internal motif sequences within inverted repeat sequences (IRs) (Table (Table1).1). The DNA inversions within these regions are thought to be controlled by site-specific DNA invertases specific to each class. B. fragilis strain YCH46 contains 33 tyrosine site-specific recombinases (Tsr) genes and four serine site-specific recombinases (Ssr) genes. Generally, DNA invertases mediate DNA inversions at adjacent regions, such as FimB and FimE, that flip their immediate downstream promoters to generate a phase-variable phenotype of type I pili in Escherichia coli (15). B. fragilis is unique in that this anaerobe possesses not only locally acting DNA invertases but also globally acting DNA invertases that mediate DNA inversions at distant loci (8, 29). It has been reported that B. fragilis possesses at least two types of master DNA invertase that regulate DNA inversions at multiple loci simultaneously (8, 29). One is Mpi, an Ssr that mediates the on-off switching of 13 promoter regions (corresponding to class I regions in B. fragilis strain YCH46), including seven promoter regions for capsular polysaccharide biosynthesis in B. fragilis strain NCTC9343 (8). The other master DNA invertase is Tsr19, a Tsr that regulates DNA inversions at two distantly located promoter regions (corresponding to class IV regions in B. fragilis strain YCH46) associated with the large encapsulation phenotype (6, 26, 29). The invertible regions contain specific consensus motifs within the IRs corresponding to each DNA invertase and constitute a regulatory unit. We designated this type of regulatory unit as an “inverlon,” which consists of at least two invertible regions controlled by a single master DNA invertase.

TABLE 1.

Classification of the invertible regions in B. fragilis strain YCH46 based on internal motif sequences within IRs
ClassaNo.Consensus motif sequencesbMaster DNA invertase genecRegulated genesSource or reference(s)
I14ARACGTWCGTBF2765 (mpi)Capsular polysaccharide biosynthesis genes8
II10AGTTC{N5}GAACTBF0667susC/susD paralogsThis study
III3GTTAC{N7}GTAACBF3038, BF4033, BF4283Putative outer membrane protein genes36
IV2TACTTANTAGGTAANAGAABF2766Extracellular polysacharide biosynthesis genes6, 26, 29
V1TCTGCAAAGNCTTTGCAGABF0667susC/susD paralogsThis study
VI1ACTAAGTTCTATCGGBF0667susC/susD paralogsThis study
Open in a separate windowaOur previous classification of the invertible regions identified in B. fragilis strain YCH46 genome (19).bConsensus motif sequences found within IRs are shown. R = A or G, W = A or T, and N = A, G, C, or T.cThe gene identifications in B. fragilis strain YCH46 genome are shown.Our previous studies indicated that an additional inverlon other than the Mpi- and Tsr19-regulated inverlons is present in B. fragilis, based on the finding that at least 10 invertible regions (corresponding to class II regions in B. fragilis strain YCH46) contain a particular consensus motif sequence (AAGTTCN5GAACTT) within their IRs (19) but do not appear to colocalize with a DNA invertase gene. The majority of the class II regions were associated with the selective switching of a particular set of susC/susD family genes. Since the SusC/SusD family of outer membrane proteins play an important role in polysaccharide utilization by Bacteroides (3, 23, 32), the inverlon associated with the phase variation of SusC/SusD family proteins would likely be involved in the survival of this anaerobe in the distal gut.In the present study, we sought to identify the DNA invertase regulating the additional inverlon in B. fragilis. Our results indicated that the Tsr encoded by BF0667 is a master DNA invertase for this inverlon (designated the Tsr0667-inverlon) in B. fragilis.  相似文献   

7.
Arthrobacter sp. strain JBH1 was isolated from nitroglycerin-contaminated soil by selective enrichment. Detection of transient intermediates and simultaneous adaptation studies with potential intermediates indicated that the degradation pathway involves the conversion of nitroglycerin to glycerol via 1,2-dinitroglycerin and 1-mononitroglycerin, with concomitant release of nitrite. Glycerol then serves as the source of carbon and energy.Nitroglycerin (NG) is manufactured widely for use as an explosive and a pharmaceutical vasodilator. It has been found as a contaminant in soil and groundwater (7, 9). Due to NG''s health effects as well as its highly explosive nature, NG contamination in soils and groundwater poses a concern that requires remedial action (3). Natural attenuation and in situ bioremediation have been used for remediation in soils contaminated with certain other explosives (16), but the mineralization of NG in soil and groundwater has not been reported.To date, no pure cultures able to grow on NG as the sole carbon, energy, and nitrogen source have been isolated. Accashian et al. (1) observed growth associated with the degradation of NG under aerobic conditions by a mixed culture originating from activated sludge. The use of NG as a source of nitrogen has been studied in mixed and pure cultures during growth on alternative sources of carbon and energy (3, 9, 11, 20). Under such conditions, NG undergoes a sequential denitration pathway in which NG is transformed to 1,2-dinitroglycerin (1,2DNG) or 1,3DNG followed by 1-mononitroglycerin (1MNG) or 2MNG and then glycerol, under both aerobic and anaerobic conditions (3, 6, 9, 11, 20), and the enzymes involved in denitration have been characterized in some detail (4, 8, 15, 21). Pure cultures capable of completely denitrating NG as a source of nitrogen when provided additional sources of carbon include Bacillus thuringiensis/cereus and Enterobacter agglomerans (11) and a Rhodococcus species (8, 9). Cultures capable of incomplete denitration to MNG in the presence of additional carbon sources were identified as Pseudomonas putida, Pseudomonas fluorescens (4), an Arthobacter species, a Klebsiella species (8, 9), and Agrobacterium radiobacter (20).Here we describe the isolation of bacteria able to degrade NG as the sole source of carbon, nitrogen, and energy. The inoculum for selective enrichment was soil historically contaminated with NG obtained at a facility that formerly manufactured explosives located in the northeastern United States. The enrichment medium consisted of minimal medium prepared as previously described (17) supplemented with NG (0.26 mM), which was synthesized as previously described (18). During enrichment, samples of the inoculum (optical density at 600 nm [OD600] ∼ 0.03) were diluted 1/16 in fresh enrichment medium every 2 to 3 weeks. Isolates were obtained by dilution to extinction in NG-supplemented minimal medium. Cultures were grown under aerobic conditions in minimal medium at pH 7.2 and 23°C or in tryptic soy agar (TSA; 1/4 strength).Early stages of enrichment cultures required extended incubation with lag phases of over 200 h and exhibited slow degradation of NG (less than 1 μmol substrate/mg protein/h). After a number of transfers over 8 months, the degradation rates increased substantially (2.2 μmol substrate/mg protein/h). A pure culture capable of growth on NG was identified based on 16S rRNA gene analysis (504 bp) as an Arthrobacter species with 99.5% similarity to Arthrobacter pascens (GenBank accession no. GU246730). Purity of the cultures was confirmed microscopically and by formation of a single colony type on TSA plates. 16S gene sequencing and identification were done by MIDI Labs (Newark, DE) and SeqWright DNA Technology Services (Houston, TX). The Arthrobacter cells stained primarily as Gram-negative rods with a small number of Gram-positive cocci (data not shown); Gram variability is also a characteristic of the closely related Arthrobacter globiformis (2, 19). The optimum growth temperature is 30°C, and the optimum pH is 7.2. Higher pH values were not investigated because NG begins to undergo hydrolysis above pH 7.5 (data not shown). The isolated culture can grow on glycerol, acetate, succinate, citrate, and lactate, with nitrite as the nitrogen source. Previous authors described an Arthrobacter species able to use NG as a nitrogen source in the presence of additional sources of carbon. However, only dinitroesters were formed, and complete mineralization was not achieved (9).To determine the degradation pathway, cultures of the isolated strain (5 ml of inoculum grown on NG to an OD600 of 0.3) were grown in minimal medium (100 ml) supplemented with NG at a final concentration of 0.27 mM. Inoculated bottles and abiotic controls were continuously mixed, and NG, 1,2DNG, 1,3DNG, 1MNG, 2MNG, nitrite, nitrate, CO2, total protein, and optical density were measured at appropriate intervals. Nitroesters were analyzed with an Agilent high-performance liquid chromatograph (HPLC) equipped with an LC-18 column (250 by 4.6 mm, 5 μm; Supelco) and a UV detector at a wavelength of 214 nm (13). Methanol-water (50%, vol/vol) was used as the mobile phase at a flow rate of 1 ml/min. Nitrite and nitrate were analyzed with an ion chromatograph (IC) equipped with an IonPac AS14A anion-exchange column (Dionex, CA) at a flow rate of 1 ml/min. Carbon dioxide production was measured with a Micro Oxymax respirometer (Columbus Instruments, OH), and total protein was quantified using the Micro BCA protein assay kit (Pierce Biotechnology, IL) according to manufacturer''s instructions. During the degradation of NG the 1,2DNG concentration was relatively high at 46 and 72 h (Fig. (Fig.1).1). 1,3DNG, detected only at time zero, resulted from trace impurities in the NG stock solution. Trace amounts of 1MNG appeared transiently, and trace amounts of 2MNG accumulated and did not disappear. Traces of nitrite at time zero were from the inoculum. The concentration of NG in the abiotic control did not change during the experiment (data not shown).Open in a separate windowFIG. 1.Growth of strain JBH1 on NG. ×, NG; ▵, 1,2DNG; ⋄, 1MNG; □, 2MNG; ○, protein.Results from the experiment described above were used to calculate nitrogen and carbon mass balances (Tables (Tables11 and and2).2). Nitrogen content in protein was approximated using the formula C5H7O2N (14). Because all nitrogen was accounted for throughout, we conclude that the only nitrogen-containing intermediate compounds are 1,2DNG and 1MNG, which is consistent with previous studies (6, 9, 20). The fact that most of the nitrogen was released as nitrite is consistent with previous reports of denitration catalyzed by reductase enzymes (4, 8, 21). The minor amounts of nitrate observed could be from abiotic hydrolysis (5, 12) or from oxidation of nitrite. Cultures supplemented with glycerol or other carbon sources assimilated all of the nitrite (data not shown).

TABLE 1.

Nitrogen mass balance
Time (h)% of total initial nitrogen by mass recovered ina:
Total recovery (%)
1MNG2MNG1,2DNG1,3DNGNGProteinNitriteNitrate
0NDbND0.9 ± 0.70.8 ± 0.682 ± 5.20.8 ± 0.214 ± 0.70.8 ± 0.3100 ± 5.3
460.1 ± 0.00.8 ± 0.27.9 ± 0.4ND35 ± 3.62.0 ± 0.549 ± 1.11.7 ± 0.096 ± 4.2
720.1 ± 0.00.9 ± 0.24.3 ± 4.2ND5.0 ± 0.43.3 ± 0.281 ± 4.23.9 ± 1.998 ± 6.8
94ND0.6 ± 0.4NDND0.6 ± 0.43.2 ± 0.095 ± 102.6 ± 1.6102 ± 10
Open in a separate windowaData represent averages of four replicates ± standard deviations.bND, not detected.

TABLE 2.

Carbon mass balance
Time (h)% of total initial carbon by mass recovered in:
Total recovery (%)
1MNGa2MNGa1,2DNGa1,3DNGaNGaProteinaCO2b
0NDcND1.6 ± 1.21.9 ± 0.492 ± 5.84.4 ± 0.9100 ± 8.4
460.5 ± 0.22.6 ± 0.613 ± 0.7ND39 ± 3.913 ± 3.028 ± 5.796 ± 14.1
720.4 ± 0.02.9 ± 0.77.3 ± 7.0ND5.6 ± 0.422 ± 1.259 ± 8.397 ± 17.6
94ND2.8 ± 0.3NDND0.8 ± 0.518 ± 0.371 ± 4.593 ± 5.6
Open in a separate windowaData represent averages of four replicates ± standard deviations.bData represent averages of duplicates ± standard deviations.cND, not detected.In a separate experiment cells grown on NG were added to minimal media containing 1,3DNG, 1,2DNG, 1MNG, or 2MNG and degradation over time was measured. 1,2DNG, 1,3DNG, and 1MNG were degraded at rates of 6.5, 3.8, and 8 μmol substrate/mg protein/hour. No degradation of 2MNG was detected (after 250 h), which indicates that 2MNG is not an intermediate in a productive degradation pathway. Because 1,3DNG was not observed at any point during the degradation of NG and its degradation rate is approximately one-half the degradation rate of 1,2DNG, it also seems not to be part of the main NG degradation pathway used by Arthrobacter sp. strain JBH1. The above observations indicate that the degradation pathway involves a sequential denitration of NG to 1,2DNG, 1MNG, and then glycerol, which serves as the source of carbon and energy (Fig. (Fig.2).2). The productive degradation pathway differs from that observed by previous authors using both mixed (1, 3, 6) and pure cultures (4, 9, 11, 20), in which both 1,3- and 1,2DNG were intermediates during NG transformation. Additionally, in previous studies both MNG isomers were produced regardless of the ratio of 1,2DNG to 1,3DNG (3, 4, 6, 9, 20). Our results indicate that the enzymes involved in denitration of NG in strain JBH1 are highly specific and catalyze sequential denitrations that do not involve 1,3DNG or 2MNG. Determination of how the specificity avoids misrouting of intermediates will require purification and characterization of the enzyme(s) involved.Open in a separate windowFIG. 2.Proposed NG degradation pathway.Mass balances of carbon and nitrogen were used to determine the following stoichiometric equation that describes NG mineralization by Arthrobacter sp. strain JBH1: 0.26C3H5(ONO2)3 + 0.33O2 → 0.03C5H7O2N + 0.63CO2 + 0.75NO2 + 0.75H+ + 0.17H2O. The result indicates that most of the NG molecule is being used for energy. The biomass yield is relatively low (0.057 mg protein/mg NG), with an fs (fraction of reducing equivalents of electron donor used for protein synthesis) of 0.36 (10), which is low compared to the aerobic degradation of other compounds by pure cultures, for which fs ranges between 0.4 and 0.6 (10, 14). The results are consistent with the requirement for relatively large amounts of energy during the initiation of the degradation mechanism (each denitration probably requires 1 mole of NADH or NADPH [21]).Although NG degradation rates were optimal at pH 7.2, they were still substantial at values as low as 5.1. The results suggest that NG degradation is possible even at low pH values typical of the subsurface at sites where explosives were formerly manufactured or sites where nitrite production lowers the pH.NG concentrations above 0.5 mM are inhibitory, but degradation was still observed at 1.2 mM (data not shown). The finding that NG can be inhibitory to bacteria at concentrations that are well below the solubility of the compound is consistent with those of Accashian et al. (1) for a mixed culture.The ability of Arthrobacter sp. strain JBH1 to grow on NG as the carbon and nitrogen source provides the basis for a shift in potential strategies for natural attenuation and bioremediation of NG at contaminated sites. The apparent specificity of the denitration steps raises interesting questions about the evolution of the pathway.  相似文献   

8.
9.
Par-1 is an evolutionarily conserved protein kinase required for polarity in worms, flies, frogs, and mammals. The mammalian Par-1 family consists of four members. Knockout studies of mice implicate Par-1b/MARK2/EMK in regulating fertility, immune homeostasis, learning, and memory as well as adiposity, insulin hypersensitivity, and glucose metabolism. Here, we report phenotypes of mice null for a second family member (Par-1a/MARK3/C-TAK1) that exhibit increased energy expenditure, reduced adiposity with unaltered glucose handling, and normal insulin sensitivity. Knockout mice were protected against high-fat diet-induced obesity and displayed attenuated weight gain, complete resistance to hepatic steatosis, and improved glucose handling with decreased insulin secretion. Overnight starvation led to complete hepatic glycogen depletion, associated hypoketotic hypoglycemia, increased hepatocellular autophagy, and increased glycogen synthase levels in Par-1a−/− but not in control or Par-1b−/− mice. The intercrossing of Par-1a−/− with Par-1b−/− mice revealed that at least one of the four alleles is necessary for embryonic survival. The severity of phenotypes followed a rank order, whereby the loss of one Par-1b allele in Par-1a−/− mice conveyed milder phenotypes than the loss of one Par-1a allele in Par-1b−/− mice. Thus, although Par-1a and Par-1b can compensate for one another during embryogenesis, their individual disruption gives rise to distinct metabolic phenotypes in adult mice.Cellular polarity is a fundamental principle in biology (6, 36, 62). The prototypical protein kinase originally identified as a regulator of polarity was termed partitioning defective (Par-1) due to early embryonic defects in Caenorhabditis elegans (52). Subsequent studies revealed that Par-1 is required for cellular polarity in worms, flies, frogs, and mammals (4, 17, 58, 63, 65, 71, 89). An integral role for Par-1 kinases in multiple signaling pathways has also been established, and although not formally addressed, multifunctionality for individual Par-1 family members is implied in reviews of the list of recognized upstream regulators and downstream substrates (Table (Table1).1). Interestingly, for many Par-1 substrates the phosphorylated residues generate 14-3-3 binding sites (25, 28, 37, 50, 59, 61, 68, 69, 78, 95, 101, 103). 14-3-3 binding in turn modulates both nuclear/cytoplasmic as well as cytoplasmic/membrane shuttling of target proteins, thus allowing Par-1 activity to establish intracellular spatial organization (15, 101). The phosphorylation of Par-1 itself promotes 14-3-3 binding, thereby regulating its subcellular localization (37, 59, 101).

TABLE 1.

Multifunctionality of Par-1 polarity kinase pathwaysa
Regulator or substrateFunctionReference(s)
Regulators (upstream function)
    LKB1Wnt signaling, Peutz-Jeghers syndrome, insulin signal transduction, pattern formation2, 63, 93
    TAO1MEK3/p38 stress-responsive mitogen-activated protein kinase (MAPK) pathway46
    MARKKNerve growth factor signaling in neurite development and differentiation98
    aPKCCa2+/DAG-independent signal transduction, cell polarity, glucose metabolism14, 37, 40, 45, 59, 75, 95
    nPKC/PKDDAG-dependent, Ca2+-independent signal transduction (GPCR)101
    PAR-3/PAR-6/aPKC(−); regulates Par-1, assembly of microtubules, axon-dendrite specification19
    GSK3β(−); tau phosphorylation, Alzheimer''s dementia, energy metabolism, body patterning54, 97
    Pim-1 oncogene(−); G2/M checkpoint, effector of cytokine signaling and Jak/STAT(3/5)5
    CaMKI(−); Ca2+-dependent signal transduction, neuronal differentiation99
Substrates (downstream function)
    Cdc25CRegulation of mitotic entry by activation of the cdc2-cyclin B complex25, 72, 78, 103
    Class II HDACControl of gene expression and master regulator of subcellular trafficking28, 50
    CRTC2/TORC2Gluconeogenesis regulator via LKB1/AMPK/TORC2 signaling, PPARγ1a coactivator49
    Dlg/PSD-95Synaptogenesis and neuromuscular junction, tumor suppressor (102)104
    DisheveledWnt signaling, translocation of Dsh from cytoplasmic vesicles to cortex73, 94
    KSR1Regulation of the Ras-MAPK pathway68, 69
    MAP2/4/TAUDynamic instability (67, 83) of microtubules, Alzheimer''s dementia (30)11, 31-33, 47, 70, 96
    Mib/NotchMind bomb (Mib degradation and repression of Notch signaling results in neurogenesis)57, 74, 81
    Par3/OSKAR/LglCytoplasmic protein segregation, cell polarity, and asymmetric cell division7, 10
    Pkp2Desmosome assembly and organization; nuclear shuttling68, 69
    PTPH1Linkage between Ser/Thr and Tyr phosphorylation-dependent signaling103
    Rab11-FIPRegulation of endocytosis (23), trafficking of E-cadherin (64)34
Open in a separate windowaLKB1 also is known as Par-4; MARKK also is known as Ste20-like; (−), inhibitory/negative regulation has been shown; GPCR, G protein-coupled receptors. MARKK is highly homologous to TAO-1 (thousand-and-one amino acid kinase) (46).The mammalian Par-1 family contains four members (Table (Table2).2). Physiological functions of the Par-1b kinase have been studied using targeted gene knockout approaches in mice (9, 44). Two independently derived mouse lines null for Par-1b have implicated this protein kinase in diverse physiological processes, including fertility (9), immune system homeostasis (44), learning and memory (86), the positioning of nuclei in pancreatic beta cells (35, 38), and growth and metabolism (43).

TABLE 2.

Terminology and localization of mammalian Par-1 family members
SynonymsaSubcellular localization
Par-1a, MARK3, C-TAK1, p78/KP78, 1600015G02Rik, A430080F22Rik, Emk2, ETK-1, KIAA4230, mKIAA1860, mKIAA4230, M80359Basolateralb/apicalc
Par-1b, EMK, MARK2, AU024026, mKIAA4207Basolateral
Par1c, MARK1Basolateral
Par1d, MARK4, MARKL1Not asymmetricd
Open in a separate windowaPar should not to be confused with protease-activated receptor 1 (PAR1 [29]); C-TAK1, Cdc twenty-five C-associated kinase 1; MARK, microtubule affinity regulating kinase; MARKL, MAP/microtubule affinity-regulating kinase-like 1.bBasolateral to a lesser degree than Par-1b (37).cHuman KP78 is asymmetrically localized to the apical surface of epithelial cells (76).dVariant that does not show asymmetric localization in epithelial cells when overexpressed (95).Beyond Par-1b, most information regarding the cell biological functions of the Par-1 kinases comes from studies of Par-1a. Specifically, Par-1a has been implicated in pancreatic (76) and hepatocarcinogenesis (51), as well as colorectal tumors (77), hippocampal function (100), CagA (Helicobacter pylori)-associated epithelial cell polarity disruption (82), and Peutz-Jeghers syndrome (48), although the latter association has been excluded recently (27). As a first step toward determining unique and redundant functions of Par-1 family members, mice disrupted for a second member of the family (Par-1a/MARK3/C-TAK1) were generated. We report that Par-1a−/− mice are viable and develop normally, and adult mice are hypermetabolic, have decreased white and brown adipose tissue mass, and unaltered glucose/insulin handling. However, when challenged by a high-fat diet (HFD), Par-1a−/− mice exhibit resistance to hepatic steatosis, resistance to glucose intolerance, and the delayed onset of obesity relative to that of control littermates. Strikingly, overnight starvation results in a complete depletion of glycogen and lipid stores along with an increase in autophagic vacuoles in the liver of Par-1a−/− but not Par-1b−/− mice. Correspondingly, Par-1a−/− mice develop hypoketotic hypoglycemia. These findings reveal unique metabolic functions of two Par-1 family members.  相似文献   

10.
11.
12.
Predator-prey relationships among prokaryotes have received little attention but are likely to be important determinants of the composition, structure, and dynamics of microbial communities. Many species of the soil-dwelling myxobacteria are predators of other microbes, but their predation range is poorly characterized. To better understand the predatory capabilities of myxobacteria in nature, we analyzed the predation performance of numerous Myxococcus isolates across 12 diverse species of bacteria. All predator isolates could utilize most potential prey species to effectively fuel colony expansion, although one species hindered predator swarming relative to a control treatment with no growth substrate. Predator strains varied significantly in their relative performance across prey types, but most variation in predatory performance was determined by prey type, with Gram-negative prey species supporting more Myxococcus growth than Gram-positive species. There was evidence for specialized predator performance in some predator-prey combinations. Such specialization may reduce resource competition among sympatric strains in natural habitats. The broad prey range of the Myxococcus genus coupled with its ubiquity in the soil suggests that myxobacteria are likely to have very important ecological and evolutionary effects on many species of soil prokaryotes.Predation plays a major role in shaping both the ecology and evolution of biological communities. The population and evolutionary dynamics of predators and their prey are often tightly coupled and can greatly influence the dynamics of other organisms as well (1). Predation has been invoked as a major cause of diversity in ecosystems (11, 12). For example, predators may mediate coexistence between superior and inferior competitors (2, 13), and differential trajectories of predator-prey coevolution can lead to divergence between separate populations (70).Predation has been investigated extensively in higher organisms but relatively little among prokaryotes. Predation between prokaryotes is one of the most ancient forms of predation (27), and it has been proposed that this process may have been the origin of eukaryotic cells (16). Prokaryotes are key players in primary biomass production (44) and global nutrient cycling (22), and predation of some prokaryotes by others is likely to significantly affect these processes. Most studies of predatory prokaryotes have focused on Bdellovibrionaceae species (e.g., see references 51, 55, and 67). These small deltaproteobacteria prey on other Gram-negative cells, using flagella to swim rapidly until they collide with a prey cell. After collision, the predator cells then enter the periplasmic space of the prey cell, consume the host cell from within, elongate, and divide into new cells that are released upon host cell lysis (41). Although often described as predatory, the Bdellovibrionaceae may also be considered to be parasitic, as they typically depend (apart from host-independent strains that have been observed [60]) on the infection and death of their host for their reproduction (47).In this study, we examined predation among the myxobacteria, which are also deltaproteobacteria but constitute a monophyletic clade divergent from the Bdellovibrionaceae (17). Myxobacteria are found in most terrestrial soils and in many aquatic environments as well (17, 53, 74). Many myxobacteria, including the model species Myxococcus xanthus, exhibit several complex social traits, including fruiting body formation and spore formation (14, 18, 34, 62, 71), cooperative swarming with two motility systems (64, 87), and group (or “wolf pack”) predation on both bacteria and fungi (4, 5, 8, 9, 15, 50). Using representatives of the genus Myxococcus, we tested for both intra- and interspecific variation in myxobacterial predatory performance across a broad range of prey types. Moreover, we examined whether prey vary substantially in the degree to which they support predatory growth by the myxobacteria and whether patterns of variation in predator performance are constant or variable across prey environments. The latter outcome may reflect adaptive specialization and help to maintain diversity in natural populations (57, 59).Although closely related to the Bdellovibrionaceae (both are deltaproteobacteria), myxobacteria employ a highly divergent mode of predation. Myxobacteria use gliding motility (64) to search the soil matrix for prey and produce a wide range of antibiotics and lytic compounds that kill and decompose prey cells and break down complex polymers, thereby releasing substrates for growth (66). Myxobacterial predation is cooperative both in its “searching” component (6, 31, 82; for details on cooperative swarming, see reference 64) and in its “handling” component (10, 29, 31, 32), in which secreted enzymes turn prey cells into consumable growth substrates (56, 83). There is evidence that M. xanthus employs chemotaxis-like genes in its attack on prey cells (5) and that predation is stimulated by close contact with prey cells (48).Recent studies have revealed great genetic and phenotypic diversity within natural populations of M. xanthus, on both global (79) and local (down to centimeter) scales (78). Phenotypic diversity includes variation in social compatibility (24, 81), the density and nutrient thresholds triggering development (33, 38), developmental timing (38), motility rates and patterns (80), and secondary metabolite production (40). Although natural populations are spatially structured and both genetic diversity and population differentiation decrease with spatial scale (79), substantial genetic diversity is present even among centimeter-scale isolates (78). No study has yet systematically investigated quantitative natural variation in myxobacterial predation phenotypes across a large number of predator genotypes.Given the previous discovery of large variation in all examined phenotypes, even among genetically extremely similar strains, we anticipated extensive predatory variation as well. Using a phylogenetically broad range of prey, we compared and contrasted the predatory performance of 16 natural M. xanthus isolates, sampled from global to local scales, as well as the commonly studied laboratory reference strain DK1622 and representatives of three additional Myxococcus species: M. flavescens (86), M. macrosporus (42), and M. virescens (63) (Table (Table1).1). In particular, we measured myxobacterial swarm expansion rates on prey lawns spread on buffered agar (31, 50) and on control plates with no nutrients or with prehydrolyzed growth substrate.

TABLE 1.

List of myxobacteria used, with geographical origin
Organism abbreviation used in textSpeciesStrainGeographic originReference(s)
A9Myxococcus xanthusA9Tübingen, Germany78
A23Myxococcus xanthusA23Tübingen, Germany78
A30Myxococcus xanthusA30Tübingen, Germany78
A41Myxococcus xanthusA41Tübingen, Germany78
A46Myxococcus xanthusA46Tübingen, Germany78
A47Myxococcus xanthusA47Tübingen, Germany78
A75Myxococcus xanthusA75Tübingen, Germany78
A85Myxococcus xanthusA85Tübingen, Germany78
TVMyxococcus xanthusTvärminneTvärminne, Finland79
PAKMyxococcus xanthusPaklenicaPaklenica, Croatia79
MADMyxococcus xanthusMadeira 1Madeira, Portugal79
WARMyxococcus xanthusWarwick 1Warwick, UK79
TORMyxococcus xanthusToronto 1Toronto, Ontario, Canada79
SUL2Myxococcus xanthusSulawesi 2Sulawesi, Indonesia79
KALMyxococcus xanthusKalalauKalalau, HI79
DAVMyxococcus xanthusDavis 1ADavis, CA79
GJV1Myxococcus xanthusGJV 1Unknown35, 72
MXFL1Myxococcus flavescensMx fl1Unknown65
MXV2Myxococcus virescensMx v2Unknown65
CCM8Myxococcus macrosporusCc m8Unknown65
Open in a separate window  相似文献   

13.
14.
15.
16.
Aerobic growth conditions significantly influenced anaerobic succinate production in two-stage fermentation by Escherichia coli AFP111 with knockouts in rpoS, pflAB, ldhA, and ptsG genes. At a low cell growth rate limited by glucose, enzymes involved in the reductive arm of the tricarboxylic acid cycle and the glyoxylate shunt showed elevated activities, providing AFP111 with intracellular redox balance and increased succinic acid yield and productivity.Succinic acid is valued as one of the key basic chemicals used in the preparation of biodegradable polymers or as raw material for chemicals of the C4 family (8, 19). The fermentative production of succinic acid from renewable resources is environmentally acceptable and sustainable (3). A breakthrough in genetically engineering Escherichia coli (6, 7, 11, 18) for succinate production was the isolation of strain AFP111 (1, 4), a mutant of NZN111 with a spontaneous ptsG mutation (pflAB ldhA double mutant). The process involves a two-stage fermentation, with aerobic cell growth followed by anaerobic conditions for succinate production (16, 21, 22). The aerobically induced enzymes can maintain their activity during the anaerobic phase and significantly affect succinate fermentation (22, 23). Using the best transition time based on the activities of the key enzymes and other physiological states, a two-stage fermentation using the recombinant AFP111 strain harboring pTrc99A-pyc achieved a final succinic acid concentration and productivity of 99.2 g·liter−1 and 1.3 g·liter−1·h−1, respectively (21).Aerobic cell growth is essential for the subsequent anaerobic fermentation. However, few studies have focused on the regulation of aerobic cell growth. As a regulation method, gluconeogenic carbon sources were used instead of glucose for the aerobic growth of Escherichia coli NZN111 and the activities of enzymes that are favorable for the anaerobic synthesis of succinate were enhanced (23, 24). Unfortunately, a gluconeogenic carbon source (e.g., sodium acetate) might increase the osmotic pressure of culture media, which would be detrimental to succinate production (23). As another regulation method, a glucose feeding strategy controlling the glucose concentration at about 0.5 g·liter−1 up to 1 g·liter−1 was reported to prevent excessive formation of acetic acid (16).In this study, we investigated different glucose feeding strategies for the aerobic growth phase of the two-phase process for succinate production by E. coli AFP111. Specifically, we compared several growth rates by using glucose limitation in addition to maximum growth under conditions of excess glucose.E. coli AFP111 [F+ λ rpoS396(Am) rph-1 ΔpflAB::Cam ldhA::Kan ptsG] (4, 16), which was a kind gift from D. P. Clark (Southern Illinois University), was the only strain used in this study. Luria-Bertani (LB) medium (60 ml) was used for inoculum culture in 1,000-ml flasks, and 3 liters of chemically defined medium (13, 14) was used for two-stage culture in a 7-liter fermentor. Two-stage fermentations were divided into three types, based on the glucose feeding strategy used during the aerobic stage. For type I culture, the glucose concentration was maintained at about 20 g·liter−1 during aerobic cell growth. Type II and III cultures comprised a batch process and subsequent glucose-limited fed-batch process (Fig. (Fig.1).1). The batch process initially contained 13 g/liter of glucose. The fed-batch process began when the dry cell weight (DCW) reached about 6 g/liter, with type II and type III cultures using a 600 g/liter glucose feed to achieve cell growth rates of 0.15 h−1 and 0.07 h−1, respectively (10). When the DCW reached 12 g·liter−1, the aerobically grown cells were directly transferred to anaerobic conditions (Fig. (Fig.1).1). For the anaerobic process, oxygen-free CO2 was sparged at 0.5 liter·min−1, the pH was controlled between 6.4 and 6.8 with intermittent supplementation of solid magnesium carbonate hydroxide, and the glucose concentration was maintained at about 20 g·liter−1 by supplying glucose in an 800-g·liter−1 solution.Open in a separate windowFIG. 1.Concentrations of glucose (circles), DCW (triangles), and succinic acid (squares) in the three types of two-stage fermentation by AFP111. μ, growth rate.The optical density at 600 nm was used to monitor cell growth, and this value was correlated to DCW. The concentration of glucose was assayed with an enzyme electrode analyzer, and organic acids were quantified by high-performance liquid chromatography (HPLC). The intracellular concentrations of NADH and NAD+ were assayed with a cycling method (12). The activities of isocitrate lyase (ICL) (20), pyruvate kinase (PYK) (17), phosphoenolpyruvate (PEP) carboxykinase (PCK) (20, 23), PEP carboxylase (PPC) (23), and malate dehydrogenase (MDH) (23) were measured spectrophotometrically at the end of the aerobic phase and 12 h after the onset of the anaerobic phase.All three types of fermentations were terminated when the succinate concentration increased less than 1 g·liter−1 in 5 h. Type III fermentation was terminated at a final succinic acid concentration of 101.2 g·liter−1 and an anaerobic-phase productivity of 1.89 g·liter−1·h−1 (Fig. (Fig.1).1). Trace amounts of by-products (such as acetate, ethanol, and pyruvate) accumulated and did not follow any trend in the anaerobic phase (data not shown).At the end of the aerobic culture phase, the specific enzyme activities of PCK, PYK, and ICL in type III culture were 2.9, 2.5, and 11.4 times higher, respectively, than the activities in type I culture (Table (Table1)1) . This phenomenon is consistent with published reports that suggest that the expression of enzymes involved in anaplerotic metabolism and the glyoxylate shunt (5, 15) is elevated in E. coli grown under glucose-limited conditions. These enzymes maintained their activities in the subsequent anaerobic phase (Table (Table1)1) and would be central to succinate production (22, 23). The elevated levels of PCK and PPC would provide the reductive branch of the tricarboxylic acid (TCA) cycle with oxaloacetate (OAA) at a higher rate (9), thereby supplying both malate and citrate (Table (Table11).

TABLE 1.

Activities of enzymes at the end of the aerobic culture phase and 12 h after the onset of the anaerobic phase
Fermentation typeaStagebMean sp act of enzyme ± SD (U/mg protein)c
PCKPPCMDHPYKICL
IAerobic0.82 ± 0.050.22 ± 0.0521.97 ± 0.151,175 ± 11.380.12 ± 0.00
Anaerobic0.55 ± 0.020.19 ± 0.0018.27 ± 1.05978 ± 12.330.09 ± 0.00
IIAerobic1.46 ± 0.100.23 ± 0.0425.69 ± 0.372,053 ± 3.650.73 ± 0.03
Anaerobic1.09 ± 0.010.20 ± 0.0135.55 ± 0.781,430 ± 13.780.41 ± 0.02
IIIAerobic2.38 ± 0.110.16 ± 0.0023.5 ± 0.132,955 ± 8.771.37 ± 0.00
Anaerobic1.75 ± 0.030.21 ± 0.0143.8 ± 0.622,501 ± 10.151.02 ± 0.01
Open in a separate windowaFermentation types were mentioned in culture conditions section.b“Aerobic” represents the data obtained at the end of aerobic culture; “Anaerobic” represents those obtained 12 h after transition to anaerobic fermentation.cThe standard deviations (SD) were calculated from triplicate samples of the same run.The reductive branch of the TCA cycle consumes 4 mol of electrons to form 2 mol of succinate based on 1 mol of glucose (1, 4). Therefore, the conversion of glucose to succinate through the reductive arm of the TCA cycle alone will lead to an intracellular imbalance of reducing equivalents (2, 18). Fortunately, the glyoxylate shunt (2, 18, 22) is available to provide 10 mol of electrons by converting 1 mol of glucose to 1 mol of succinate and 2 mol of CO2 (22). In the case of the ptsG mutant strain AFP111, when the molar flux at the PEP branch point flowing to OAA versus flowing to pyruvate reaches a ratio of 5:2, the intracellular redox balance is satisfied and the maximum theoretical mass yield of 1.12 g·g−1 succinic acid is achieved (22). Based on the elevated activities of PCK, PYK, and ICL (Table (Table1),1), both pathways leading to succinate were enhanced after glucose-limited growth. The succinic acid yields of 1.03 to 1.07 g·g−1 in the two glucose-limited processes approached the maximum theoretical yield for AFP111 (22), and these yields were about two times greater than the yield in the type I fermentation (Table (Table22).

TABLE 2.

Succinic acid production during anaerobic fermentation phasea
Fermentation typeMean ± SD
Succinic acid (g·liter−1)Yield (g·g−1)Productivity (g·liter−1·h−1)Specific productivity at 12 h (mg·g−1·h−1)NADH at 12 h mmol·(g DCW)−1NADH/NAD+ ratio at 12 h
I35.0 ± 0.740.43 ± 0.050.98 ± 0.04105 ± 150.88 ± 0.070.55 ± 0.08
II74.3 ± 3.241.03 ± 0.011.32 ± 0.05160 ± 81.95 ± 0.111.05 ± 0.10
III101.2 ± 1.041.07 ± 0.021.89 ± 0.07227 ± 111.97 ± 0.151.27 ± 0.13
Open in a separate windowaThe data were calculated only for the anaerobic stage. The standard deviations (SD) were calculated from two independent two-stage fermentations.In addition to differences in succinic acid yields, the glucose-limited and type I fermentations each resulted in significantly different specific succinic acid productivities (Table (Table2).2). A specific succinic acid productivity of 227 mg·g−1·h−1 was obtained at 12 h in type III fermentation. Because two pathways are needed for succinate production due to redox constraints, and enzyme activities in both pathways were elevated by glucose limitation, the results suggest that operating with glucose limitation provides the cells with greater metabolic flexibility to achieve a redox balance. Furthermore, the results suggest that one or more of these enzymes are limiting succinate formation under batch conditions (type I fermentation). Considering the NADH/NAD+ assays (Table (Table2),2), the results would support the hypothesis that succinate production was limited by insufficient NADH (2, 18).In summary, our study presented an efficient method of aerobic cell cultivation for two-stage succinate fermentation by engineered E. coli. Since the physiological state of aerobically grown cells was essential for their subsequent anaerobic succinate fermentation, some other environmental and physiology factors in the aerobic growth phase may also play an important role in improving succinate production.  相似文献   

17.
18.
19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号