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 78 毫秒
1.
2.
3.
4.
5.
We analyzed the temporal and spatial diversity of the microbiota in a low-usage and a high-usage hospital tap. We identified a tap-specific colonization pattern, with potential human pathogens being overrepresented in the low-usage tap. We propose that founder effects and local adaptation caused the tap-specific colonization patterns. Our conclusion is that tap-specific colonization represents a potential challenge for water safety.Humans are exposed to and consume large amounts of tap water in their everyday life, with the tap water microbiota representing a potent reservoir for pathogens (8). Despite the potential impact, our knowledge about the ecological diversification processes of the tap water microbiota is limited (4, 11).The aim of the present work was to determine the temporal and spatial distribution patterns of the planktonic tap water microbiota. We compared the summer and winter microbiota from two hospital taps supplied from the same water source. We analyzed 16S rRNA gene clone libraries by using a novel alignment-independent approach for operational taxonomic unit (OTU) designation (6), while established OTU diversity and richness estimators were used for the ecological interpretations.Tap water samples (1 liter) from a high-usage kitchen and a low-usage toilet cold-water tap in Akershus University Hospital, Lørenskog, Norway, were collected in January and July 2006. The total DNA was isolated and the 16S rRNA gene PCR amplified and sequenced. Based on the sequences, we estimated the species richness and diversity, we calculated the distances between the communities, and trees were constructed to reflect the relatedness of the microbiota in the samples analyzed. Details about these analytical approaches are given in the materials and methods section in the supplemental material.Our initial analysis of species composition was done using the RDPII hierarchical classifier. We found that the majority of pathogen-related bacteria in our data set belonged to the class Gammaproteobacteria. The genera encompassed Legionella, Pseudomonas, and Vibrio (Table (Table1).1). We found a significant overrepresentation of pathogen-related bacteria in the toilet tap (P = 0.04), while there were no significant differences between summer and winter samples. Legionella showed the highest relative abundance for the pathogen-related bacteria. With respect to the total diversity, we found that Proteobacteria dominated the tap water microbiota (representing 86% of the taxa) (see Table S1 in the supplemental material). There was, however, a large portion (56%) of the taxa that could not be assigned to the genus level using this classifier.

TABLE 1.

Cloned sequences related to human pathogensa
Sampling placeSampling timePathogenNCBI accession no.Identity (%)
ToiletSummerEscherichia coliEF41861499
ToiletSummerEscherichia sp.EF07430799
ToiletSummerLegionella sp.AY92415595
ToiletSummerLegionella sp.AY92415395
ToiletSummerLegionella sp.AY92415396
ToiletWinterLegionella sp.AY92406196
ToiletWinterLegionella sp.AY92415897
ToiletWinterLegionella sp.AY92415897
KitchenWinterLegionella sp.AY92399697
ToiletSummerPseudomonas fluorescensEF41307398
ToiletSummerPseudomonas fluorescensEF41307398
KitchenSummerPseudomonas fluorescensDQ20773199
ToiletWinterVibrio sp.DQ40838898
ToiletWinterVibrio sp.AB27476098
KitchenWinterVibrio sp.DQ40838898
KitchenWinterVibrio lentusAY29293699
KitchenWinterVibrio sp.AM18376597
ToiletWinterStenotrophomonas maltophiliaAY83773099
KitchenWinterStenotrophomonas maltophiliaDQ42487098
ToiletWinterStreptococcus suisAF28457898
ToiletWinterStreptococcus suisAF28457898
Open in a separate windowaThe relatedness between the cloned sequences and potential pathogens was determined by BLAST searches of the NCBI database, carried out using default settings.To obtain a better resolution of the uncharacterized microbiota, we analyzed the data using a clustering approach that is not dependent on a predefined bacterial group (see the materials and methods section in the supplemental material for details). These analyses showed that there were three relatively tightly clustered groups in our data set (Fig. (Fig.1A).1A). The largest group (n = 590) was only distantly related to characterized betaproteobacteria within the order Rhodocyclales. We also identified another large betaproteocaterial group (n = 320) related to Polynucleobacter. Finally, a tight group (n = 145) related to the alphaproteobacterium Sphingomonas was identified.Open in a separate windowFIG. 1.Tap water microbiota diversity, determined by use of a principal component analysis coordinate system. (A) Each bacterium is classified by coordinates, with the following color code: brown squares, kitchen summer; red diamonds, toilet summer; green triangles, kitchen winter; and green circles, toilet winter. (B and C) Each square represents a 1 × 1 (B) or 5 × 5 (C) OTU. PC1, first principal component; PC2, second principal component.The tap-specific distributions of the bacterial groups were investigated using density distribution analyses. A dominant population related to Polynucleobacter was identified for the toilet summer samples, while for the winter samples there was a dominance of the Rhodocyclales-related bacteria. The kitchen summer samples revealed a dominance of Sphingomonas. The corresponding winter samples did not reveal distinct high-density bacterial populations (see Table S2 in the supplemental material).Hierarchical clustering for the 1 × 1 OTU density distribution confirmed the relatively low overlap for the microbiota in the samples analyzed (Fig. (Fig.2).2). We found that the microbiota clustered according to tap and not season.Open in a separate windowFIG. 2.Hierarchical clustering for the density distribution of the tap water microbiota. The density of 1 × 1 OTUs was used as a pseudospecies for hierarchical clustering. The tree for the Cord distance matrix is presented, while the distances calculated using the three distance matrices Cord, Brad Curtis, and Sneath Sokal, respectively, are shown for each branch.We have described the species diversity and richness of the microbiota in Table S3 in the supplemental material. For the low taxonomic level, these analyses showed that the diversity and species richness were greater for the winter samples than for the summer samples. Comparing the two taps, the diversity and richness were greater in the kitchen tap than in the toilet tap. In particular, the winter sample from the kitchen showed great richness and diversity. The high taxonomic level, however, did not reveal the same clear differences as did the low level, and the distributions were more even. Rarefaction analyses for the low taxonomic level confirmed the richness and diversity estimates (see Fig. S1 in the supplemental material).Our final analyses sought to fit the species rank distributions to common rank abundance curves. Generally, the rank abundance curves were best fitted to log series or truncated log normal distributions (see Table S4 in the supplemental material). The log series distribution could be fit to all of the samples except the kitchen summer samples at the low taxonomic level, while the truncated log normal distribution could not be fit to the kitchen samples at the high taxonomic level. Interestingly, however, the kitchen winter sample was best fit to a geometric curve at both the high and the low taxonomic level.Diversifying, adaptive biofilm barriers have been documented for tap water bacteria (7), and it is known that planktonic bacteria can interact with biofilms in an adaptive manner (3). On the other hand, tap usage leads to water flowthrough and replacement of the global with the local water population by stochastic founder effects (1).Therefore, we propose that parts of the local diversity observed can be explained by local adaptation (10) and parts by founder effects (9).Most prokaryote diversity measures assume log normal or log series OTU dominance density distributions (5). The kitchen winter sample, however, showed deviations from these patterns by being correlated to geometric distributions (in addition to the log series and truncated log normal distributions for the high taxonomic level). This sample also showed a much greater species richness than the other samples. A possible explanation is that the species richness of the tap water microbiota can be linked to usage and that the kitchen tap is driven toward a founder microbiota by high usage.Since our work indicates an overrepresentation of Legionella in the low-usage tap, it would be of high interest to determine whether the processes for local Legionella colonization can be related to tap usage. Understanding the ecological forces affecting Legionella and other pathogens are of great importance for human health. At the Akerhus University Hospital, this was exemplified by a Pseudomonas aeruginosa outbreak in an intensive care unit, where the outbreak could be traced back to a single tap (2).  相似文献   

6.
7.
8.
9.
Bacteroidales species were detected in (tap) water samples from treatment plants with three different PCR assays. 16S rRNA gene sequence analysis indicated that the sequences had an environmental rather than fecal origin. We conclude that assays for Bacteroidales 16S rRNA genes are not specific enough to discern fecal contamination of drinking water in the Netherlands.Drinking water in many countries is routinely monitored for recent fecal contamination by testing for fecal indicator organisms Escherichia coli, thermotolerant coliforms, and/or intestinal enterococci to demonstrate microbial safety (13, 21, 42). Although these indicator organisms have been used for many decades, they have some limitations: the number of E. coli/coliform/enterococcus bacteria in feces is relatively low (18, 38), and they sometimes might be able to grow in the environment (10, 11, 14, 27). Consequently, scientists have been searching for alternative indicator organisms to determine fecal contamination of water. In 1967, bacteria belonging to the genus Bacteroides were suggested as alternative indicator organisms (26). Bacteroides spp. might have some advantages over the traditional indicator organisms. The numbers of Bacteroides spp. in the intestinal tract of humans and animals are 10 to 100 times higher than the numbers of E. coli or intestinal enterococci (1, 2, 12, 26). However, the use of Bacteroides spp. as indicator organisms was hampered by the complex cultivation conditions required (1, 2). The introduction of molecular methods made it possible to detect bacterial species that belong to the order Bacteroidales, an order that includes the genus Bacteroides, without cultivation. As a result, real-time PCR methods were developed for the quantitative detection of Bacteroidales in surface and recreation water and the potential of Bacteroidales species as an indication of fecal contamination of recreational waters was demonstrated (6, 12, 16, 19, 20, 29). Bacteroidales species might be useful indicator organisms for fecal contamination of drinking water as well. However, methods to detect fecal contamination in drinking water should be more sensitive, because people ingest more drinking water and the quality assessments and standards for fecal contamination are stricter than for bathing water. Studies exploring real-time PCR for the detection of Bacteroidales genes in drinking water have not been published to our knowledge. The objective of our study was, therefore, to determine if assays for the detection of Bacteroidales 16S rRNA genes can be used to detect fecal contamination in drinking water.Unchlorinated tap water samples were obtained in November 2007 and February 2010 from one or more locations in the distribution systems of nine different drinking water treatment plants (plants A to I; Table Table1)1) that produced unchlorinated drinking water from confined (plants B, C, E, F, and G) and unconfined (plants A, D, H, and I) groundwater. The treatment plants are located in the central part of the Netherlands within 90 km of each other. In addition, untreated groundwater from extraction wells and/or untreated raw groundwater (mixture of groundwater from different extraction wells) was sampled in March 2008 (Table (Table1).1). Water samples (100 ml) were filtered over a 25-mm polycarbonate filter (0.22-μm pore size, type GTTP; Millipore, Netherlands) and a DNA fragment was added as internal control to determine the recovery efficiency of DNA isolation and PCR analysis (2a, 40). DNA was isolated using a FastDNA spin kit for soil (Qbiogene, United States) according to the supplier''s protocol. Primer sets AllBac 296f and AllBac 412r, resulting in a PCR product of 108 bp, were used in combination with TaqMan probe AllBac375Bhqr to quantitatively determine the number of Bacteroidales 16S rRNA gene copies in the water samples using a real-time PCR instrument (20). The PCR cycle after which the fluorescence signal of the amplified DNA was detected (threshold cycle [CT]) was used to quantify the concentration of 16S rRNA gene copies. Quantification was based on comparison of the sample CT value with the CT values of a calibration curve graphed using known copy numbers of the Bacteroidales 16S rRNA gene, as previously described (12, 20). The correlation coefficient of the calibration curve was 0.99, and the efficiency of the PCR 95 to 105%. Finally, the Bacteroidales cell number was calculated by using the recovery rate of the internal standard and assuming five 16S rRNA gene copy numbers per cell (5). The detection limit of this gene assay was 50 Bacteroidales cells 100 ml−1 (corresponding to 10 16S rRNA gene copies per reaction mixture). Furthermore, the 16S rRNA genes that were obtained from several water samples from treatment plant C with the AllBac and TotBac (12) primer sets were sequenced, and the nearest relatives were obtained from the GenBank database using BLAST searches.

TABLE 1.

Numbers of Bacteroidales cells in extraction wells, raw groundwater, and unchlorinated tap water of nine different groundwater plants in the Netherlandsa
PlantSource of sampleNo. (100 ml−1) of Bacteroidales cells in:
200720082010
ATap water 1b5,948 ± 950
Tap water 22,682 ± 1,4591,254 ± 216
Tap water 34,362 ± 947439 ± 136
Raw water96 ± 15
BTap water 13,553 ± 9815,302 ± 2,952
Tap water 24,487 ± 3912,119 ± 1,367
Tap water 37,862 ± 4,5883,896 ± 3,003
Raw water3,209 ± 833
CTap water 1661 ± 75386 ± 199
Tap water 21,051 ± 626
Tap water 3831 ± 584
Tap water 41,254 ± 216
Extraction well 11,126 ± 262
Extraction well 22,666 ± 51
Extraction well 3<50
Raw water90 ± 44
DTap water1,103 ± 291,254 ± 216
Raw water48 ± 16
ETap water1,302 ± 2221,254 ± 216
Extraction well 1671 ± 97
FTap water1,317 ± 198
Raw water<50
GTap water 1675 ± 92439 ± 300
Tap water 2216 ± 65249 ± 98
Tap water 3154 ± 6322 ± 137
Raw water<50
HTap water7,073 ± 845
Raw water511 ± 254
ITap water1,577 ± 176
Raw water420 ± 66
Open in a separate windowaValues are the average results and standard deviations from replicate PCRs on the same water sample using the AllBac primer set (20). In November 2007, the distribution systems (tap water) of plants A, B, and G were sampled at three different locations, whereas for the other plants, one location in the distribution system was sampled. In March 2008, raw water of plants A to G was sampled, as well as one (plant E) or three (plant C) different extraction wells. Finally, in February 2010, the distribution systems of plants A, B, C, D, E, and G were sampled again.bMore than one tap water sample from a treatment plant means that samples were taken at different locations in the distribution system.The Bacteroidales 16S rRNA gene, quantified with the AllBac primer set, was detected in all tap water samples in November 2007 and February 2010. The number of cells varied between 154 and 7,862 Bacteroidales cells 100 ml−1, and the numbers in tap water of each plant were similar in 2007 and 2010 (Table (Table1).1). The Bacteroidales counts were high compared to the number of E. coli that are occasionally observed in fecally contaminated drinking water (17a) but low compared to numbers observed in surface water (4, 20, 22). Water from the extraction wells and raw water used for unchlorinated drinking water production were analyzed, and Bacteroidales species were detected in 10 out of 15 samples (Table (Table1).1). These results would imply that the extracted groundwater, raw water, and tap water were fecally contaminated. According to the Dutch drinking water decree (2b), both raw and tap water from the nine different treatment plants are regularly analyzed for fecal contamination by monitoring for E. coli, F-specific RNA phages, and somatic coliphages. For at least the last 10 years, these indicator organisms have not been detected in these waters.Additional qualitative PCR analyses using TotBac and BacUni primer sets (12, 19) targeting other parts of the Bacteroidales 16S rRNA gene were performed to confirm the presence of Bacteroidales species in the water samples of November 2007 and March 2008. Nine or 10 of the 11 samples that were positive with the AllBac primer set were also positive with the TotBac and BacUni primer sets (data not shown). The BacUni primer set has a higher detection limit (30 gene copies per PCR; 19), which could explain the difference from the results with the AllBac primer set. The TotBac primer set has the same detection limit as the AllBac primer set (12), but small differences in PCR efficiencies might have resulted in different results, since some water samples showed Bacteroidales 16S rRNA gene copy numbers around the detection limit (Table (Table1).1). Nevertheless, the additional PCR analyses demonstrated that the detection of Bacteroidales species in tap, raw, and extracted well water with the AllBac primer set was not an artifact. The primer sets used were developed in three different studies (12, 19, 20) but have been applied in a number of recent studies to detect fecal contamination of surface water (3, 4, 16, 22, 33, 34). The results from most of these studies showed that 16S rRNA genes of Bacteroidales were present in all surface water samples tested. Only Sinigalliano et al. (34) observed that 2 out of 4 water samples were negative with the TotBac primer set. However, the detection limit of the assay was not specified in that study.The nine different treatment plants tested in our study produce unchlorinated drinking water from groundwater, which is considered to be of high hygienic quality. In addition, the extraction wells are protected from fecal contamination by a protection zone where no activities related to human waste or animal manure are allowed. In the Netherlands, this protection zone is based on a 60-day residence time of the water. Previous studies have demonstrated that a residence time of 60 days is highly effective in removing fecal bacteria and viruses (30, 31, 39). Moreover, the Bacteroidales numbers in tap water in November 2007 were significantly higher than the numbers in raw groundwater in March 2008 (Mann-Whitney U test; P < 0.01). Because the recovery efficiency of the internal control was the same between raw water and tap water samples, this result demonstrates that Bacteroidales cell numbers increased during treatment and/or drinking water distribution. This result could suggest that the water was fecally contaminated during drinking water treatment and/or distribution. However, it is unlikely that the integrity of nine different treatment trains and/or supply systems was affected in the sampling period. The statutory monitoring did not show the presence of E. coli at these sites. Another hypothesis is that the increase of Bacteroidales cell numbers in tap water was caused by the growth of Bacteroidales species in (drinking) water systems. In summary, it is unexpected that the majority of the tap water, raw water, and extracted groundwater samples were fecally contaminated. These unexpected observations raise the question of whether the PCR methods detect only fecal Bacteroidales species and, thus, if the gene assays are suitable to discern fecal contamination in drinking water in the Netherlands.Sequence analyses of the Bacteroidales 16S rRNA genes were performed to determine the relatedness of sequences from the different sampling sites to sequences from the nearest relatives in the GenBank database. All sequences contained the primer regions, indicating that nonspecific amplification had not occurred in the PCRs. Because the PCR product from the AllBac primer set was small (108 bp), many 16S rRNA gene sequences (100 to 5,000) in the GenBank database were identical to the Bacteroidales 16S rRNA gene sequences obtained from groundwater and unchlorinated tap water samples from plant C. These identical 16S rRNA gene sequences were in general obtained from fecal sources, but some of them came from environmental rather than fecal sources (Table (Table2).2). The AllBac 16S rRNA gene sequences from tap water and groundwater had relative high similarities (96.3 to 100%) to sequences from bacterial species of the genera Bacteroides, Prevotella, and Tannerella (Table (Table2),2), which all belong to the order Bacteroidales.

TABLE 2.

Nearest relatives in GenBank to the Bacteroidales 16S rRNA gene sequences obtained from groundwater and unchlorinated tap water from plant C using different primer setsa
Primer set used, source of sample, and OTUsbGenBank sequence accession no.Source of sequence (GenBank sequence accession no.)SimilaritycNearest cultivated bacterium in GenBank (sequence accession no.)Similarity
AllBac
    Extraction well 1 (3/6)GQ169588Rhizosphere (EF605968)108/108Prevotella oralis (AY323522)105/108
    Extraction well 1 (3/6)GQ169589Water from watershed (DQ886209)108/108Tannerella forsythia(AB035460)107/108
    Extraction well 2 (1/6)GQ169590Phyllosphere Brazilian forest (DQ221468)108/108Tannerella forsythia(AB035460)106/108
    Extraction well 2 (5/6)GQ169591Bovine rumen (EU348207)108/108Tannerella forsythia(AB035460)106/108
    Extraction well 3 (1/6)GQ169592Phyllosphere Brazilian forest (DQ221468)108/108Prevotella oralis (AY323522)104/108
    Extraction well 3 (5/6)GQ169593Prevotella corporis (L16465)108/108Prevotella corporis (L16465)108/108
    Raw water (3/6)GQ169594Spitsbergen permafrost (EF034756)108/108Tannerella forsythia(AB035460)106/108
    Raw water (3/6)GQ169595Hindgut beetle larvae (FJ374179)108/108Tannerella forsythia(AB035460)107/108
    Tap water (6/6)GQ169596Prevotella timonensis (DQ518919)108/108Prevotella timonensis (DQ518919)108/108
    Prevotella buccalis (L16476)Prevotella buccalis (L16476)
    Prevotella ruminicola (AF218617)Prevotella ruminicola (AF218617)
    Bacteroides vulgatus (NC_009614)Bacteroides vulgatus (NC_009614)
TotBac
    Extraction well 1 (1/10)GQ169597Deep subsurface groundwater (AB237705)339/369Salinimicrobium terrae (EU135614)315/370
    Extraction well 1 (1/10)GQ169598Songhuajiang River sediment (DQ444125)363/377Paludibacter propionicigenes (AB078842)357/376
    Extraction well 1 (4/10)GQ169599Freshwater pond sediment (DQ676447)352/360Paludibacter propionicigenes (AB078842)313/372
    Extraction well 1 (4/10)GQ169600Pine River sediment (DQ833352)364/371Bacteroides oleiciplenus (AB490803)334/375
    Extraction well 2 (4/10)GQ169601Groundwater (AF273319)364/371Xanthobacillum maris (AB362815)338/375
    Extraction well 2 (6/10)GQ169602Human saliva (AB028385)381/382Prevotella intermedia (AY689226)380/382
    Extraction well 3 (1/10)GQ169603Pig manure (AY816766)354/377Bacteroides thetaiotaomicron (AE015928)311/380
    Extraction well 3 (3/10)GQ169604Pig manure (AY816867)371/376Butyricimonas virosa (AB443949)307/379
    Extraction well 3 (6/10)GQ169605Swedish lake (AY509350)343/362Parabacteroides distasonis (AB238927)320/374
    Raw water (10/10)GQ169606Prevotella timonensis (AF218617)378/379Prevotella timonensis (AF218617)378/379
    Tap water (1/10)GQ169607Deep subsurface groundwater (AB237705)338/369Salinimicrobium terrae (EU135614)312/370
    Tap water (2/10)GQ169608Yukon River, AK(FJ694652)367/372Psychroserpens burtonensis (U62913)312/375
    Tap water (7/10)GQ169609Deep subsurface groundwater (AB237705)341/369Salinimicrobium terrae (EU135614)315/370
Open in a separate windowaPrimer sets AllBac (20) and TotBac (12) were used in PCRs of samples, and GenBank was searched for relatives using BLAST.bOTUs are indicated by the values in parentheses (number of sequences belonging to the OTU/total number of sequences analyzed).cNumber of base pairs identical in both sequences/total number of base pairs in sequences.16S rRNA gene sequences obtained with the TotBac primer set were longer (∼370 bp) and did not show 100% similarity with the nearest relatives in the GenBank database (Table (Table2).2). Sequences from the GenBank database that showed the highest similarity (91.6% to 99.7%) with the 16S rRNA gene sequences from tap water and groundwater from plant C were in general isolated from environmental sources (Table (Table2).2). The 16S rRNA gene sequences from cultivated bacterial species that showed the highest similarity to the 16S rRNA gene sequences obtained in our study belonged to different genera (Table (Table2).2). Some of these genera (Salinimicrobium, Xanthobacillum, and Psychroserpens) did not belong to the order Bacteroidales. However, the 16S rRNA gene sequences from bacterial species of these genera showed low similarities with the sequences obtained in this study (83.2% to 90.1%) and six mismatches to the TotBac primers. Thus, it is unlikely that DNA from bacterial species belonging to Salinimicrobium, Xanthobacillum, and Psychroserpens was amplified in the gene assay. More importantly, the majority of the nearest environmental clone sequences retrieved from the GenBank database showed no or a single mismatch with the AllBac and TotBac primer and probe sequences. Thus, these primer sets are capable of amplifying 16S rRNA genes from bacteria that have been observed in ecosystems outside the intestinal tract of humans and animals.16S rRNA gene sequences related to Prevotella species were commonly observed in extracted groundwater, raw water, and tap water (Table (Table2).2). The isolation of Prevotella paludivivens from rice roots in a rice field soil (35) demonstrated the environmental nature of some Prevotella species. In addition, primer sequences developed for the detection of fecal Bacteroidales species (8, 12, 19, 20, 25, 29) showed no or a single mismatch with 16S rRNA gene sequences from P. paludivivens, Xylanibacterium oryzae, Paludibacter propionicigenes, Proteiniphilum acetatigenes, and Petrimonas sulfuriphila that are present in the GenBank database. These five Bacteroidales species have all been isolated from ecosystems other than the gastrointestinal tract. Consequently, primer sets for 16S rRNA genes of Bacteroidales species cannot always be used to discern fecal contamination in water.A number of 16S rRNA gene sequences observed in groundwater and tap water fell in the genus Bacteroides. The presence of Bacteroides 16S rRNA gene sequences in groundwater and tap water might also suggest that some Bacteroides species are capable of growth in the environment. However, until now, type strains of Bacteroides species growing outside the animal intestinal tract have not been published. Another possible explanation is that the observed 16S rRNA gene sequences originate from Bacteroides species that inhabit the anoxic intestinal tract of insects. Previous studies have shown that bacterial species belonging to the genus Bacteroides are common inhabitants of the hindguts of insects (15, 23, 24, 28, 32). Some of the 16S rRNA gene sequences obtained with the AllBac primer set in our study showed 100% similarity to 16S rRNA gene sequences from the hindgut of insects. Moreover, a number of 16S rRNA gene sequences isolated from the hindguts of insects (15, 23, 24, 32) showed no or a single mismatch with the TotBac and AllBac primer and probe sequences. In conclusion, these primer sets are capable of detecting Bacteroides species from the hindgut of insects as well. Water insects are normal inhabitants of groundwater and drinking water distribution systems (7, 41) and might be a source of Bacteroides species in water. Bacteroides species from insect feces do not indicate fecal pollution by warm-blooded animals, and insects do not normally shed human fecal pathogenic microorganisms. Bacteroides species from insect feces, therefore, can hamper Bacteroides gene assays developed for the detection of water fecally contaminated by warm-blooded animals. Additional cultivation techniques in combination with molecular tools are required to demonstrate the persistence or growth of Bacteroides bacteria in groundwater and drinking water or whether Bacteroides bacteria are present in water insects. However, these experiments were beyond the scope of our study.The three extraction wells of plant C are located close to each other and extract water from the same aquifer. Subsequently, extracted water from the three wells is mixed and enters the treatment plant as raw water. We hypothesize that if a fecal source in the vicinity of the extraction field of plant C contaminated the groundwater, water from the extraction wells and raw water should (partly) have the same Bacteroidales species. Although a relatively limited amount of clones was sequenced per sample (16), the diversity of Bacteroidales operational taxonomic units (OTU) within a sample was low (Table (Table2).2). In contrast, unique 16S rRNA gene sequences were observed between the different water types (e.g., extracted groundwater, raw water, and tap water) and sequence overlap between water types was low. These results demonstrate that the Bacteroidales 16S rRNA gene sequences at the sampling locations were not from the same fecal source and imply once again that Bacteroidales species were environmental rather than fecal.Finally, we hypothesized that if the Bacteroidales species observed in tap water were of nonfecal origin, human- and/or bovine-specific Bacteroidales strains should not be present in tap water. We tested for the presence of human- or bovine-specific Bacteroidales strains by using source-specific 16S rRNA gene assays (5) on tap water samples from February 2010. The results showed that human- and bovine-specific Bacteroidales 16S rRNA genes could not be detected in tap water, whereas a PCR product was always detected with the positive control. Again, these results indicate that the Bacteroidales species observed in tap water were of nonfecal origin.Overall, the results from our study indicate that gene assays for Bacteroidales detected environmental rather than fecal Bacteroidales species in groundwater and tap water from treatment plants in the Netherlands. First, Bacteroidales 16S rRNA gene sequences obtained from water samples taken at plant C showed (high) similarity to clone sequences that were isolated from environmental sources. The majority of these clone sequences and several Bacteroides clone sequences from the hindguts of insects showed no or a single mismatch with AllBac, TotBac, and BacUni primer and probe sequences. Second, the primer and probe sequences used for the gene assays have no or a single mismatch with 16S rRNA gene sequences of environmental Bacteroidales species P. paludivivens, X. oryzae, P. propionicigenes, P. acetatigenes, and/or P. sulfuriphila (9, 17, 35-37). Third, Bacteroidales 16S rRNA gene sequences from raw water and water from extraction wells were unique, and sequence overlap between water types was low. It is expected that in the case of fecal contamination of groundwater, different water types from the same groundwater area have similar Bacteroidales species. Fourth, the quantitative assays for Bacteroidales 16S rRNA genes commonly used to detect fecal contamination (3, 4, 12, 16, 19, 20, 22, 33, 34) detected Bacteroidales species in deep groundwater and tap water that have no history of fecal contamination. Fifth, Bacteroidales gene copy numbers were significantly higher in tap water than in raw groundwater, demonstrating an increase or growth of Bacteroidales species during the treatment and/or distribution of drinking water. Finally, human- and bovine-specific Bacteroidales strains were not detected in tap water. Consequently, (quantitative) assays for general Bacteroidales 16S rRNA genes are not suitable to discern fecal contamination in groundwater and unchlorinated drinking water in the Netherlands.Nucleotide sequence accession numbers.The 16S rRNA gene sequences obtained in this study were deposited in the GenBank database under accession numbers GQ169588 to GQ169609.  相似文献   

10.
Small-subunit ribosomal DNA (SSU rDNA) from 20 phenotypically distinct strains of 2,4-dichlorophenoxyacetic acid (2,4-D)-degrading bacteria was partially sequenced, yielding 18 unique strains belonging to members of the alpha, beta, and gamma subgroups of the class Proteobacteria. To understand the origin of 2,4-D degradation in this diverse collection, the first gene in the 2,4-D pathway, tfdA, was sequenced. The sequences fell into three unique classes found in various members of the beta and gamma subgroups of Proteobacteria. None of the α-Proteobacteria yielded tfdA PCR products. A comparison of the dendrogram of the tfdA genes with that of the SSU rDNA genes demonstrated incongruency in phylogenies, and hence 2,4-D degradation must have originated from gene transfer between species. Only those strains with tfdA sequences highly similar to the tfdA sequence of strain JMP134 (tfdA class I) transferred all the 2,4-D genes and conferred the 2,4-D degradation phenotype to a Burkholderia cepacia recipient.Bacteria capable of mineralizing 2,4-dichlorophenoxyacetic acid (2,4-D), a commonly used herbicide, are found in many different phylogenetic groups (2, 3, 7, 11, 22, 23). Evidence suggests that numerous variants of 2,4-D catabolic genes exist and that catabolic operons consist of a near-random mixing of these variants (7). Interspecies gene transfer is a well-documented phenomenon (13), and horizontal gene transfer of the 2,4-D-degrading plasmid pJP4 has been shown (3, 5). However, not all 2,4-D catabolic operons are found on plasmids (10, 11, 16, 20). The extent to which other 2,4-D genes have been exchanged in nature is unknown. The aim of this research was to assess the role of horizontal gene transfer in the evolution of 2,4-D-degrading strains. This article summarizes the results of two aspects of this work—the study of the transfer of the entire 2,4-D pathway by using standard mating experiments and a phylogenetic study of the tfdA gene. The tfdA gene codes for an α-ketoglutarate-dependent 2,4-D dioxygenase which converts 2,4-D into 2,4-dichlorophenol and glyoxylate (6). This 861-bp gene was first sequenced from Ralstonia eutropha JMP134 (19). Two more tfdA genes were cloned from chromosomal locations in Burkholderia strain RASC and Burkholderia strain TFD6 (16, 20). These proved to be identical to each other and 78.5% similar to the original. An alignment of the two variants allowed conserved areas to be identified and primers to be designed for the amplification of tfdA-like genes from other sources (24). Sequence analysis of putative tfdA fragments and the small-subunit ribosomal DNA (SSU rDNA) of the strains carrying them allowed us to construct phylogenies of the genes and their hosts and to look for congruency between them.

Mating experiments.

A collection of 2,4-D degraders containing 15 unique strains as determined by genomic fingerprinting (7) was used as a source of donors in a series of mating experiments (Table (Table1).1). Burkholderia cepacia D5, lacking the ability to grow on 2,4-D and not hybridizing to any tfd genes, was used as a recipient in mating experiments. Strain D5 contains neomycin phosphotransferase genes (nptII) carried on transposon Tn5 and is resistant to 50 μg each of kanamycin, carbenicillin, and bacitracin per ml. All of the 2,4-D strains used were sensitive to these antibiotics. Filter matings were performed with a donor-to-recipient ratio of 1:10. Colonies which grew on selective medium (500 ppm of 2,4-D in mineral salts agar [MMO] [23] including 50 μg of kanamycin, carbenicillin, and bacitracin per ml) were subjected to further tests. Their ability to catabolize 2,4-D was tested in liquid medium (same composition as that described above).

TABLE 1

2,4-D-degrading strains, geographic origins, and GenBank accession numbers
StrainGenBank accession no. (SSU rDNA)OriginMost similar to genus and/or speciesaTransferbtfdA typecGenBank accession no. (tfdA gene)Reference or source
JMP134AF049542AustraliaRalstonia eutropha+IM167303
EML1549AF049546OregonBurkholderia sp.+I2
TFD39AF049539SaskatchewanBurkholderia sp.+IU4319723
K712AF049543MichiganBurkholderia sp.+IU4327611
TFD9AF049537SaskatchewanAlcaligenes xylosoxidans+IU4327623
TFD41AF049541MichiganRalstonia eutropha+I23
TFD38AF049540MichiganRalstonia eutropha+NDc23
TFD23AF049536MichiganRhodoferax fermentans+IU4327623
RASCAF049544OregonBurkholderia sp.(+)IIU257172
TFD6AF049546MichiganBurkholderia sp.II23
TFD2AF049545MichiganBurkholderia sp.II23
TFD31AF049536SaskatchewanRhodoferax fermentansIII23
B6-9AF049538OntarioRhodoferax fermentansNDIIIU431969
I-18U22836OregonHalomonas sp.NDIIIU2249915
K1443AF049531MichiganSphingomonas sp.d11
2,4-D1AF049535MontanaSphingomonas sp.R. Sanford
B6-5AF049533OntarioSphingomonas sp.ND9
B6-10AF049534OntarioSphingomonas sp.ND9
EML146AF049532OregonSphingomonas sp.2
M1AF049530French PolynesiaRhodospeudomonas sp.NDR. Fulthorpe
Open in a separate windowaThe generus and/or species most similar to the strain is given based on similarities of SSU rDNA sequences. bSymbols: +, able to transfer 2,4-D degradation to B. cepacia D5; (+), able to transfer at very low frequency; −, no transfer detected. cND, not determined. d—, no amplificate was obtained. The disappearance of 2,4-D from the culture medium was monitored by high-performance liquid chromatography. Cells were removed by centrifugation, and the supernatant was filtered through 0.2-μm-pore-size filters. These samples were then analyzed on a Lichrosorb Rp-18 column (Anspec Co., Ann Arbor, Mich.) with 60% methanol–40% 0.1% H3PO4 as the eluant. 2,4-D was detected by measuring light absorption at 230 nm. The presence of tfd genes was detected by hybridizing colony blots with a DNA probe derived from the entire pJP4 plasmid. The identity of the colonies was confirmed by probing with the nptII gene of Tn5 (found in B. cepacia D5). Probes were labeled with random hexanucleotides incorporating [32P]dCTP (3,000 Ci/mmol; New England Nuclear, Boston, Mass.). Hybridizations were done under high-stringency conditions by using 50% formamide and Denhardt’s solution (18) at 42°C. Of the 15 unique strains tested, 9 transferred 2,4-D degradation abilities to D5. This transfer was confirmed by hybridization with pJP4 for eight of these strains. B. cepacia RASC could transfer degradative abilities, but neither it nor the transconjugant hybridized to the pJP4 probe. Work subsequent to this study has confirmed that the genes carried by RASC do not hybridize to those found on pJP4 under high-stringency conditions (7).

Phylogenetic analyses.

Total genomic DNA was isolated from 20 unique 2,4-D-degrading strains (including all 15 used for mating experiments) grown on 500 ppm of 2,4-D mineral salts medium amended with 50 ppm of yeast extract. SSU rDNA was amplified by using fD1 and rD1 as primers (25). Putative tfdA fragments were amplified by using primers TVU and TVL as previously described (24). PCR products were purified with a Gene Clean kit (Bio 101, La Jolla, Calif.). Sequencing was done with an Applied Biosystems model 373A automatic sequencer (Perkin-Elmer Cetus) by using fluorescently labeled dye termination at the Michigan State University Sequencing Facility. The sequencing primer used for SSU rDNA fragments was 519R (5′ GTA TTA CCG CGG CTG CTG G-3′). For tfdA fragments, the sequencing primers were the same as the amplification primers. GenBank accession numbers for these sequences are given in Table Table11.The SSU rDNA sequences were compared to sequences in GenBank by using the Basic Local Alignment Search Tool (BLAST) (1), and those strains with the highest maximal segment pair scores were retrieved from GenBank and included in the phylogenetic analysis. Sequences were aligned manually with the software SeqEd (Applied Biosystems) and with MacClade (14). Sites where nucleotides were not resolved for all sequences were deleted from the alignment, as were those nucleotides corresponding to the small loop in this region that is absent in the alpha subgroup of the class Proteobacteria. These deletions left 283 unambiguous sites for the construction of the SSU rDNA phylogenies. Phylogenetic trees were constructed by using the neighbor-joining analysis of pairwise Jukes-Cantor distances (4), and the topology was confirmed by using the maximum parsimony method PAUP (21). Desulfomonile tiedjei of the δ-Proteobacteria was used as an outgroup. Bootstrap analysis based on 100 replicates was used to place confidence estimates on the tree. Only bootstrap values of greater than 50 were used.

2,4-D degrader diversity.

The 2,4-D degraders in this study were distributed throughout the alpha, beta, and gamma subgroups of the Proteobacteria (Fig. (Fig.1).1). The lack of representation of gram-positive bacteria is likely a reflection of isolation methods, not of the lack of gram-positive 2,4-D degraders. The majority of these strains were members of the beta subgroup of Proteobacteria, five of which were most closely related to the genus Burkholderia, having at least 92% sequence similarity with each other. Three were closely related to Rhodoferax fermentans (close to the class Comamonadaceae), three were related to Ralstonia eutropha, and one was related to Alcaligenes xylosoxidans. TFD39 falls outside any clear cluster. One member of the γ-Proteobacteria, strain I-18, a haloalkaliphile, was found to be closely related to the salt-loving genus Halomonas (15). The remaining six strains all clustered in the alpha branch of Proteobacteria (Fig. (Fig.1).1). Of this subgroup, five were most closely related to the genus Sphingomonas. One member of the α-Proteobacteria, strain M1, which is the most oligotrophic and slow growing of all the strains used in this study, is 97% similar to Rhodopseudomonas palustris. The character of strain M1 correlates well with its phylogenetic placement near the slow-growing genus Bradyrhizobium. Open in a separate windowFIG. 1Neighbor-joining dendrogram (Jukes-Cantor distances) of SSU rDNA from 2,4-D-degrading bacteria (indicated in boldface type) and reference strains (indicated in italic type). Class I (•), class II (▴), and class III (■) types of tfdA genes are indicated. Bootstrap confidence limits (percentages) are indicated above each branch. Scale bar represents a Jukes-Cantor distance of 0.01.

tfdA gene fragments.

tfdA gene fragments were successfully amplified and sequenced from 10 strains of β-Proteobacteria and 1 strain of γ-Protobacteria. None of the strains from the α-Proteobacteria gave any amplificates with these primers. These 313 contiguous nucleotides were aligned with additional tfdA sequences from JMP134 and from strain RASC (Fig. (Fig.2).2). Three distinct classes of tfdA gene sequences with slight variations in each class were found. Class I included fragments from JMP134, TFD39, TFD23, K712, and TFD9 that differed from each other by 2 bp at the most. Class I tfdA genes are probably plasmid encoded. All strains with a class I tfdA gene examined so far contained broad-host-range, self-transmissible plasmids containing 2,4-D genes (2, 3, 11, 17). All of the strains with a class I tfdA gene were able to transfer the 2,4-D phenotype in the mating studies reported above. The class II tfdA sequences included identical fragments amplified from RASC, TFD6, and TFD2 which were 76% similar to those in class I. Class III included identical fragments from strains TFD31, B6-9, and I-18 which were 77% similar to class I genes and 80% similar to class II genes. Both class II and III tfdA genes differed from each other and from class I genes in the same nine sites corresponding to the third base pair of the codons. The tfdA phylogenetic tree is a simple one, with three distinct branches that are incongruent with the SSU rDNA-derived phylogeny (Fig. (Fig.3).3). Class I tfdA sequences were found in Burkholderia-like strains, in strains related to the Comamonas-Rhodoferax group, and in the Ralstonia-Acaligenes group, all in the β-Proteobacteria. Class II sequences are less widely distributed, found only in Burkholderia-like branches. However, even in this subgroup, this tfdA variant is found in strains that differ by 7% at the SSU rDNA level (RASC and TFD2). However, the class III sequences were most interesting, being found both in the Comamonas-Rhodoferax group and in a strain of the γ-Proteobacteria, I-18, strains that differ by 24% at the SSU rDNA level. Class III genes have since been found in a collection of randomly isolated non-2,4-D degraders, including gram-positive bacilli, as well as in various gram-negative bacteria, even though the gene is not expressed (10). Open in a separate windowFIG. 2Alignment of 313 nucleotides of internal fragments of tfdA genes from representative strains. Nucleotides identical to tfdA from pJP4 are represented by periods.Open in a separate windowFIG. 3Phylogenetic incongruency of tfdA genes and SSU rDNA from diverse 2,4-D-degrading bacteria. Dendrograms for tfdA and SSU rDNA are indicated. Shading indicates the type of tfdA sequence, either class I, II, or III. Note that branch lengths are not drawn to scale.An interesting result was the detection of two different tfdA gene variants in sibling strains. TFD23 and TFD31 are identical at the ribosomal gene level, but one harbors a class I gene and the other harbors a class III gene. Similarly, TFD6 and EML159 are rRNA siblings that carry a class II and class I gene, respectively.None of the α-Proteobacteria yielded a PCR product when amplified with the conserved tfdA primers. This finding complements our observation that none of these bacteria hybridized to the tfdA gene, even under conditions of low stringency, indicating that any tfdA-like genes in the α-Proteobacteria are likely to be more divergent from the ones sequenced here (7, 11). In addition, none of the Sphingomonas strains in the study hybridized with a whole pJP4 probe, and similarly, no Sphingomonas strains scored positive for transfer of 2,4-D-degrading ability to recipient B. cepacia D5. Together these results suggest a reduced gene flow between members of the α- and β- or γ-Proteobacteria or poor gene expression of β- or γ-derived genes by α-Proteobacteria. Although plasmid pJP4 is a broad-host-range plasmid and has been known to transfer to α-Proteobacteria such as Rhizobium and Agrobacterium species and to γ-Proteobacteria such as Pseudomonas putida, Pseudomonas fluorescens, and Pseudomonas aeruginosa, the 2,4-D pathway is not expressed in these strains of the α- or γ-Proteobacteria (3). Phylogenetically limited expression of plasmid-borne 3-chlorobenzoate-degradative genes has also been noted for the pseudomonads (8). Subsequent studies have found divergent but related sequences for the tfdB and tfdC genes in 2,4-D-degrading Sphingomonas strains (7, 12, 24).With the exceptions of the minor differences within the class I pJP4-like tfdA sequences, there were no intermediate tfdA sequences. The most likely explanation of this is that the rate of horizontal transfer of the tfd genes is high relative to the rate at which mutations can accumulate. Examination of sequences of tfdA genes from a greater variety of organisms may turn up more intermediate variation.  相似文献   

11.
Many plant species can be induced to flower by responding to stress factors. The short-day plants Pharbitis nil and Perilla frutescens var. crispa flower under long days in response to the stress of poor nutrition or low-intensity light. Grafting experiments using two varieties of P. nil revealed that a transmissible flowering stimulus is involved in stress-induced flowering. The P. nil and P. frutescens plants that were induced to flower by stress reached anthesis, fruited and produced seeds. These seeds germinated, and the progeny of the stressed plants developed normally. Phenylalanine ammonialyase inhibitors inhibited this stress-induced flowering, and the inhibition was overcome by salicylic acid (SA), suggesting that there is an involvement of SA in stress-induced flowering. PnFT2, a P. nil ortholog of the flowering gene FLOWERING LOCUS T (FT) of Arabidopsis thaliana, was expressed when the P. nil plants were induced to flower under poor-nutrition stress conditions, but expression of PnFT1, another ortholog of FT, was not induced, suggesting that PnFT2 is involved in stress-induced flowering.Key words: flowering, stress, phenylalanine ammonia-lyase, salicylic acid, FLOWERING LOCUS T, Pharbitis nil, Perilla frutescensFlowering in many plant species is regulated by environmental factors, such as night-length in photoperiodic flowering and temperature in vernalization. On the other hand, a short-day (SD) plant such as Pharbitis nil (synonym Ipomoea nil) can be induced to flower under long days (LD) when grown under poor-nutrition, low-temperature or high-intensity light conditions.19 The flowering induced by these conditions is accompanied by an increase in phenylalanine ammonia-lyase (PAL) activity.10 Taken together, these facts suggest that the flowering induced by these conditions might be regulated by a common mechanism. Poor nutrition, low temperature and high-intensity light can be regarded as stress factors, and PAL activity increases under these stress conditions.11 Accordingly, we assumed that such LD flowering in P. nil might be induced by stress. Non-photoperiodic flowering has also been sporadically reported in several plant species other than P. nil, and a review of these studies suggested that most of the factors responsible for flowering could be regarded as stress. Some examples of these factors are summarized in 1214

Table 1

Some cases of stress-induced flowering
Stress factorSpeciesFlowering responseReference
high-intensity lightPharbitis nilinduction5
low-intensity lightLemna paucicostatainduction29
Perilla frutescens var. crispainduction14
ultraviolet CArabidopsis thalianainduction23
droughtDouglas-firinduction30
tropical pasture Legumesinduction31
lemoninduction3235
Ipomoea batataspromotion36
poor nutritionPharbitis nilinduction3, 4, 13
Macroptilium atropurpureumpromotion37
Cyclamen persicumpromotion38
Ipomoea batataspromotion36
Arabidopsis thalianainduction39
poor nitrogenLemna paucicostatainduction40
poor oxygenPharbitis nilinduction41
low temperaturePharbitis nilinduction9, 12
high conc. GA4/7Douglas-firpromotion42
girdlingDouglas-firinduction43
root pruningCitrus sp.induction44
Pharbitis nilinduction45
mechanical stimulationAnanas comosusinduction46
suppression of root elongationPharbitis nilinduction7
Open in a separate window  相似文献   

12.
The number of species on Earth is highly uncertain. A recent study has suggested that there are less than 2 million prokaryotic species on Earth; this Formal Comment suggests instead that there are more likely hundreds of millions or billions of species, and that the majority of these are bacteria associated with insects and other animals.

The number of species on Earth is a fundamental number in science. Yet, estimates of global biodiversity have been highly uncertain. There are presently approximately 1.9 million described species [1]. Estimates of the actual number (both described and undescribed) have ranged from the low millions into the trillions [2,3]. Furthermore, described species richness [1] is dominated by animals (1.3 million; 68%), not bacteria (approximately 10,000 species; 0.5%). Larsen and colleagues [2] summarized evidence suggesting that the majority of species on Earth may be bacteria associated with insect hosts and that bacterial richness may push global biodiversity into the hundreds of millions of species or even low billions.Louca and colleagues [4] (LEA hereafter) have claimed instead that there are only 40,100 host-associated bacterial species among all animal species and 0.8 to 1.6 million prokaryotic species overall (see their “Author summary”). Strangely, they excluded bacterial species associated with animal hosts from their estimates of total prokaryotic diversity and justified this by claiming that the estimates of Larsen and colleagues [2] were “mathematically flawed.” Here, I examine their claims and present new estimates of global biodiversity.Remarkably, all projections by LEA for host-associated bacterial richness were based on an estimate from one ant genus (Cephalotes), an estimate that is demonstrably incorrect by orders of magnitude (S1 Text). Without examining the underlying data [5], LEA estimated only 40 bacterial species among all 130 ant species in this genus. Yet, simply counting the bacterial species among the 25 sampled ant species in that genus reveals 616 unique bacterial species, of which 539 appear to be unique to the genus and 369 each unique to a single ant species (using the standard 97% cutoff for 16S divergence and data from [5]). Thus, there were >500 bacterial species among 25 ant species, not 40 bacterial species among 130 ant species. This mistake was further exacerbated by inexplicably ignoring data from the other 2 insect genera analyzed by Larsen and colleagues [2], thus maximizing the impact of their incorrect estimate for this genus.Their overall estimate of bacterial richness was also strongly influenced by their questionable assumption that all animal genera can share bacterial species (i.e., reducing their estimate of 3 million host-associated bacterial species to only 40,100). They assumed “a conservative overlap of only 0.1% between any two randomly chosen genera” for the number of bacterial species shared between animal genera. No justification was given for this value of 0.1%, nor were any alternative values explored. Furthermore, they implicitly assumed that any bacterial species can be shared between any pair of animal genera, regardless of their phylogeny, habitat, or geographic range. So, for example, a bacterial species that is a gut endosymbiont of a terrestrial herbivorous insect species endemic to Madagascar could somehow be shared with a deep-sea worm in the northern Pacific Ocean. This is ridiculous: there must be a reason why bacterial species are shared among host species and genera (e.g., shared phylogeny, location, diet). For example, broad-scale studies show that sharing of bacteria among insect hosts is associated with both host phylogeny and diet [6].LEA stated “it is known that substantial overlap exists between the microbiota of different host genera and even of distantly related animal taxa.” However, they provided no numbers to justify this “substantial overlap.” In fact, none of the papers they cited as supporting this assumption actually do (S2 Text). For example, one study [7] found 5 bacterial species shared among 5 insect genera utilizing the same type of host plant (cycads). However, LEA do not mention that this study found 1,789 unique bacterial species among just these 5 insect species (or 177 after filtering). This seems inconsistent with their estimate of only 40,100 bacterial species across all animals. In summary, rather than estimating the overlap of bacterial species among host genera, LEA simply made a number up and combined this with unrealistic, unsupported assumptions about overlap. If LEA had considered Cephalotes (which all their estimates were based on), a survey of this genus and related genera [5] found 1,019 bacterial species, with only 77 of the 616 bacterial species in Cephalotes shared with other sampled genera, and the sharing of bacterial species among hosts strongly related to host phylogeny.Numerous surveys of bacterial diversity in insects strongly suggest that there are far more than 40,100 bacterial species among all animals (8] found roughly twice as many bacterial species as those of approximately 30 insect species [5,9], and the study of 218 insect species [6] found >3.5 times as many as the study of 62 insect species. The simple fact that a study found 9,301 bacterial species among only 218 sampled insect species strongly suggests that there are more than 40,100 bacteria among all animals.Table 1Surveys of bacterial diversity among insect species.LEA incorrectly estimated that a genus of 130 ant species (Cephalotes) hosts only 40 bacterial species and subsequently assumed that all animal genera have the same low number of bacterial species. These broad surveys of bacterial species among insects suggest that many insects (including Cephalotes) host much larger numbers of bacterial species.
Insect group sampledInsect species sampledUnique bacterial species foundReferences
Ants (Cephalotes and 3 related genera)291,019Sanders and colleagues [5]
Lycaenid butterflies311,156Whitaker and colleagues [9]
Native Hawaiian insects (beetles, flies, true bugs)131,094Poff and colleagues [10]
Various insect orders622,073Colman and colleagues [8]
21 insect orders2189,301Yun and colleagues [6]
Open in a separate windowGiven these problems with the estimate of LEA, what is the actual number of bacterial species on Earth? LEA were correct that Larsen and colleagues [2] only estimated the number of species-specific bacteria per insect host species, and those estimates could be wrong. I therefore recalculated those estimates based on more direct counts of species-specific bacteria from the original studies (S3 Text). In 2]. Specifically, Larsen and colleagues [2] projected 0.209 to 5.8 billion species on Earth, of which 66% to 91% are bacteria, whereas I project 0.183 to 4.2 billion, with 58% to 88% bacteria (2] and are explained below. For each scenario, the projected number of species for each group is shown, along with the percentage of the total number of species belonging to that group (note that plants are <0.5% and are rounded down to 0%). In addition to the 4 scenarios, 4 other assumptions were explored. The first 3 involve different estimated numbers of morphologically cryptic arthropod species per morphology-based insect species (from 6 to 2 to 0; for justification, see [2]). These impact the number of animal species, and all downstream estimates for other groups. The final, fourth set of analyses assumes 6 morphologically cryptic arthropod species and that mites host negligible numbers of nematode species. Scenario 1 assumes that all animal species have a full set of bacterial, protist, and fungal endosymbionts, even if they are parasites, but that microsporidian fungi and apicomplexan protists have little or no host-specific bacterial richness. Scenario 2 assumes that symbionts have limited numbers of symbionts themselves (i.e., nematodes have an average of only one host-specific bacterial species) and that microsporidians and apicomplexans have few or no bacterial species. Scenario 3 assumes that all animal species have a full set of symbiont species and that microsporidians and apicomplexans host (on average) as many bacterial species as animal species do. Scenario 4 is identical to Scenario 1, except that it assumes that mites have reduced species richness relative to other arthropods (0.25 mites∶1 other arthropod species). Note that there is an error in Table 3, Scenario 1 in Larsen and colleagues [2]: There should be 27.2 million animal species, not 20.4. The correct number is used here. Archaean species is considered to be limited overall [2], and so is not treated separately.
Scenario 1Scenario 2Scenario 3Scenario 4
Million species% of totalMillion species% of totalMillion species% of totalMillion species% of total
6 cryptic arthropod species
Animals163.29.4163.213.7163.23.9102.09.4
Plants0.300.300.300.30
Fungi165.69.6165.613.9165.63.9104.69.6
Protists163.29.4163.213.7163.23.9102.09.4
Bacteria1,240.371.6701.858.83,721.088.3775.271.5
Total1,732.71,194.14,213.31,084.1
2 cryptic arthropod species
Animals54.49.454.413.654.43.934.09.4
Plants0.300.300.300.30
Fungi56.89.856.814.256.84.036.410.0
Protists54.49.454.413.654.43.934.09.4
Bacteria413.471.4233.958.51,240.388.2258.471.1
Total579.4399.91,406.3363.1
0 cryptic arthropod species
Animals27.29.327.213.527.23.917.09.3
Plants0.300.300.300.30
Fungi29.610.229.614.729.64.219.410.6
Protists27.29.327.213.527.23.917.09.3
Bacteria206.771.0117.058.1620.288.0129.270.6
Total291.1201.3704.5182.9
Mites host limited nematode richness, 6 cryptic arthropod species
Animals122.49.4122.411.9122.43.991.89.4
Plants0.300.300.300.30
Fungi124.89.6124.812.1124.83.994.29.6
Protists122.49.4122.411.9122.43.991.89.4
Bacteria930.271.5661.064.12,790.788.3697.771.5
Total1,300.21,030.93,160.7975.8
Open in a separate windowIn summary, the conclusions of LEA are based on an initial estimate of bacterial richness for one genus that was clearly incorrect, combined with a made-up number (and unrealistic assumptions) to estimate overlap of bacterial species among host genera. Reanalyses here suggest that bacterial richness (and the diversity of life) is more likely in the hundreds of millions or billions.  相似文献   

13.
14.
The enzymes called lipoxygenases (LOXs) can dioxygenate unsaturated fatty acids, which leads to lipoperoxidation of biological membranes. This process causes synthesis of signaling molecules and also leads to changes in cellular metabolism. LOXs are known to be involved in apoptotic (programmed cell death) pathway, and biotic and abiotic stress responses in plants. Here, the members of LOX gene family in Arabidopsis and rice are identified. The Arabidopsis and rice genomes encode 6 and 14 LOX proteins, respectively, and interestingly, with more LOX genes in rice. The rice LOXs are validated based on protein alignment studies. This is the first report wherein LOXs are identified in rice which may allow better understanding the initiation, progression and effects of apoptosis, and responses to bitoic and abiotic stresses and signaling cascades in plants.Key words: apoptosis, biotic and abiotic stresses, genomics, jasmonic acid, lipidsLipoxygenases (linoleate:oxygen oxidoreductase, EC 1.13.11.-; LOXs) catalyze the conversion of polyunsaturated fatty acids (lipids) into conjugated hydroperoxides. This process is called hydroperoxidation of lipids. LOXs are monomeric, non-heme and non-sulfur, but iron-containing dioxygenases widely expressed in fungi, animal and plant cells, and are known to be absent in prokaryotes. However, a recent finding suggests the existence of LOX-related genomic sequences in bacteria but not in archaea.1 The inflammatory conditions in mammals like bronchial asthama, psoriasis and arthritis are a result of LOXs reactions.2 Further, several clinical conditions like HIV-1 infection,3 disease of kidneys due to the activation of 5-lipoxygenase,4,5 aging of the brain due to neuronal 5-lipoxygenase6 and atherosclerosis7 are mediated by LOXs. In plants, LOXs are involved in response to biotic and abiotic stresses.8 They are involved in germination9 and also in traumatin and jasmonic acid biochemical pathways.10,11 Studies on LOX in rice are conducted to develop novel strategies against insect pests12 in response to wounding and insect attack,13 and on rice bran extracts as functional foods and dietary supplements for control of inflammation and joint health.14 In Arabidopsis, LOXs are studied in response to natural and stress-induced senescence,15 transition to flowering,16 regulation of lateral root development and defense response.17The arachidonic, linoleic and linolenic acids can act as substrates for different LOX isozymes. A hydroperoxy group is added at carbons 5, 12 or 15, when arachidonic acid is the substrate, and so the LOXs are designated as 5-, 12- or 15-lipoxygenases. Sequences are available in the database for plant lipoxygenases (EC:1.13.11.12), mammalian arachidonate 5-lipoxygenase (EC:1.13.11.34), mammalian arachidonate 12-lipoxygenase (EC:1.13.11.31) and mammalian erythroid cell-specific 15-lipoxygenase (EC:1.13.11.33). The prototype member for LOX family, LOX-1 of Glycine max L. (soybean) is a 15-lipoxygenase. The LOX isoforms of soybean (LOX-1, LOX-2, LOX-3a and LOX-3b) are the most characterized of plant LOXs.18 In addition, five vegetative LOXs (VLX-A, -B, -C, -D, -E) are detected in soybean leaves.19 The 3-dimensional structure of soybean LOX-1 has been determined.20,21 LOX-1 was shown to be made of two domains, the N-terminal domain-I which forms a β-barrel of 146 residues, and a C-terminal domain-II of bundle of helices of 693 residues21 (Fig. 1). The iron atom was shown to be at the centre of domain-II bound by four coordinating ligands, of which three are histidine residues.22Open in a separate windowFigure 1Three-dimensional structure of soybean lipoxygenase L-1. The domain I (N-terminal) and domain II (C-terminal) are indicated. The catalytic iron atom is embedded in domain II (PDB ID-1YGE).21This article describes identification of LOX genes in Arabidopsis and rice. The Arabidopsis genome encodes for six LOX proteins23 (www.arabidopsis.org) (
LocusAnnotationNomenclatureA*B*C*
AT1G55020lipoxygenase 1 (LOX1)LOX185998044.45.2049
AT1G17420lipoxygenase 3 (LOX3)LOX3919103725.18.0117
AT1G67560lipoxygenase family proteinLOX4917104514.68.0035
AT1G72520lipoxygenase, putativeLOX6926104813.17.5213
AT3G22400lipoxygenase 5 (LOX5)LOX5886101058.86.6033
AT3G45140lipoxygenase 2 (LOX2)LOX2896102044.75.3177
Open in a separate window*A, amino acids; B, molecular weight; C, isoelectric point.Interestingly, the rice genome (rice.plantbiology.msu.edu) encodes for 14 LOX proteins as compared to six in Arabidopsis (and22). Of these, majority of them are composed of ∼790–950 aa with the exception for loci, LOC_Os06g04420 (126 aa), LOC_Os02g19790 (297 aa) and LOC_Os12g37320 (359 aa) (Fig. 2).Open in a separate windowFigure 2Protein alignment of rice LOXs and vegetative lipoxygenase, VLX-B,28 a soybean LOX (AA B67732). The 14 rice LOCs are indicated on left and sequence position on right. Gaps are included to improve alignment accuracy. Figure was generated using ClustalX program.

Table 2

Genes encoding lipoxygenases in rice
ChromosomeLocus IdPutative functionA*B*C*
2LOC_Os02g10120lipoxygenase, putative, expressed9271035856.0054
2LOC_Os02g19790lipoxygenase 4, putative29733031.910.4799
3LOC_Os03g08220lipoxygenase protein, putative, expressed9191019597.4252
3LOC_Os03g49260lipoxygenase, putative, expressed86897984.56.8832
3LOC_Os03g49380lipoxygenase, putative, expressed87898697.57.3416
3LOC_Os03g52860lipoxygenase, putative, expressed87197183.56.5956
4LOC_Os04g37430lipoxygenase protein, putative, expressed79889304.610.5125
5LOC_Os05g23880lipoxygenase, putative, expressed84895342.97.6352
6LOC_Os06g04420lipoxygenase 4, putative12614054.76.3516
8LOC_Os08g39840lipoxygenase, chloroplast precursor, putative, expressed9251028196.2564
8LOC_Os08g39850lipoxygenase, chloroplast precursor, putative, expressed9421044947.0056
11LOC_Os11g36719lipoxygenase, putative, expressed86998325.45.3574
12LOC_Os12g37260lipoxygenase 2.1, chloroplast precursor, putative, expressed9231046876.2242
12LOC_Os12g37320lipoxygenase 2.2, chloroplast precursor, putative, expressed35940772.78.5633
Open in a separate window*A, amino acids; B, molecular weight; C, isoelectric point.

Table 3

Percent homology of rice lipoxygenases against Arabidopsis
Loci (Os)Homolog (At)Identity/similarity (%)No. of aa compared
LOC_Os02g10120LOX260/76534
LOC_Os02g19790LOX554/65159
LOC_Os03g08220LOX366/79892
LOC_Os03g49260LOX556/73860
LOC_Os03g49380LOX560/75861
LOC_Os03g52860LOX156/72877
LOC_Os04g37430LOX361/75631
LOC_Os05g23880LOX549/66810
LOC_Os06g04420LOX549/62114
LOC_Os08g39840LOX249/67915
LOC_Os08g39850LOX253/70808
LOC_Os11g36719LOX552/67837
LOC_Os12g37260LOX253/67608
LOC_Os12g37320LOX248/60160
Open in a separate windowOs, Oryza sativa L.; At, Arabidopsis thaliana L.; aa, amino acids.In plants, programmed cell death (PCD) has been linked to different stages of development and senescence, germination and response to cold and salt stresses.24,25 To conclude, this study indicates that rice genome encodes for more LOX proteins as compared to Arabidopsis. The LOX members are not been thoroughly investigated in rice. The more advanced knowledge on LOXs function might spread light on the significant role of LOXs in PCD, biotic and abiotic stress responses in rice.  相似文献   

15.
Aluminum induced proteome changes in tomato cotyledons     
Suping Zhou  Roger Sauve  Theodore W Thannhauser 《Plant signaling & behavior》2009,4(8):769-772
Cotyledons of tomato seedlings that germinated in a 20 µM AlK(SO4)2 solution remained chlorotic while those germinated in an aluminum free medium were normal (green) in color. Previously, we have reported the effect of aluminum toxicity on root proteome in tomato seedlings (Zhou et al.1). Two dimensional DIGE protein analysis demonstrated that Al stress affected three major processes in the chlorotic cotyledons: antioxidant and detoxification metabolism (induced), glyoxylate and glycolytic processes (enhanced), and the photosynthetic and carbon fixation machinery (suppressed).Key words: aluminum, cotyledons, proteome, tomatoDifferent biochemical processes occur depending on the developmental stages of cotyledons. During early seed germination, before the greening of the cotyledons, glyoxysomes enzymes are very active. Fatty acids are converted to glucose via the gluconeogenesis pathway.2,3 In greening cotyledons, chloroplast proteins for photosynthesis and leaf peroxisomal enzymes in the glycolate pathway for photorespiration are metabolized.24 Enzymes involved in regulatory mechanisms such as protein kinases, protein phosphatases, and mitochondrial enzymes are highly expressed.3,5,6The chlorotic cotyledons are similar to other chlorotic counterparts in that both contains lower levels of chlorophyll, thus the photosynthetic activities are not as active. In order to understand the impact of Al on tomato cotyledon development, a comparative proteome analysis was performed using 2D-DIGE following the as previously described procedure.1 Some proteins accumulated differentially in Al-treated (chlorotic) and untreated cotyledons (Fig. 1). Mass spectrometry of tryptic digestion fragments of the proteins followed by database search has identified some of the differentially expressed proteins (Open in a separate windowFigure 1Image of protein spots generated by Samspot analysis of Al treated and untreated tomato cotyledons proteomes separated on 2D-DIGE.

Table 1

Proteins identified from tomato cotyledons of seeds germinating in Al-solution
Spot No.Fold (treated/ctr)ANOVA (p value)AnnotationSGN accession
12.340.00137412S seed storages protein (CRA1)SGN-U314355
22.130.003651unidentified
32.00.006353lipase class 3 familySGN-U312972
41.960.002351large subunit of RUBISCOSGN-U346314
51.952.66E-05arginine-tRNA ligaseSGN-U316216
61.950.003343unidentified
71.780.009219Monodehydroascorbate reductase (NADH)SGN-U315877
81.780.000343unidentified
91.754.67E-05unidentified
121.700.002093unidentified
131.680.004522unidentified
151.660.019437Glutamate dehydrogenase 1SGN-U312368
161.660.027183unidentified
171.622.01E-08Major latex protein-related, pathogenesis-relatedSGN-U312368
18−1.610.009019RUBisCo activaseSGN-U312543
191.610.003876Cupin family proteinSGN-U312537
201.600.000376unidentified
221.590.037216unidentified
0.003147unidentified
29−1.560.001267RUBisCo activaseSGN-U312543
351.520.001955unidentified
401.470.007025unidentified
411.470.009446unidentified
451.450.001134unidentified
59−1.405.91E-0512 S seed storage proteinSGN-U314355
611.391.96E-05MD-2-related lipid recognition domain containing proteinSGN-U312452
651.370.000608triosephosphate isomerase, cytosolicSGN-U312988
681.360.004225unidentified
811.320.001128unidentified
82−1.310.00140833 kDa precursor protein of oxygen-evolving complexSGN-U312530
871.300.002306unidentified
89−1.30.000765unidentified
921.290.000125superoxide dismutaseSGN-U314405
981.280.000246triosephosphate isomerase, cytosolicSGN-U312988
Open in a separate window  相似文献   

16.
Natural Infection of Burkholderia pseudomallei in an Imported Pigtail Macaque (Macaca nemestrina) and Management of the Exposed Colony     
Crystal H Johnson  Brianna L Skinner  Sharon M Dietz  David Blaney  Robyn M Engel  George W Lathrop  Alex R Hoffmaster  Jay E Gee  Mindy G Elrod  Nathaniel Powell  Henry Walke 《Comparative medicine》2013,63(6):528-535
Identification of the select agent Burkholderia pseudomallei in macaques imported into the United States is rare. A purpose-bred, 4.5-y-old pigtail macaque (Macaca nemestrina) imported from Southeast Asia was received from a commercial vendor at our facility in March 2012. After the initial acclimation period of 5 to 7 d, physical examination of the macaque revealed a subcutaneous abscess that surrounded the right stifle joint. The wound was treated and resolved over 3 mo. In August 2012, 2 mo after the stifle joint wound resolved, the macaque exhibited neurologic clinical signs. Postmortem microbiologic analysis revealed that the macaque was infected with B. pseudomallei. This case report describes the clinical evaluation of a B. pseudomallei-infected macaque, management and care of the potentially exposed colony of animals, and protocols established for the animal care staff that worked with the infected macaque and potentially exposed colony. This article also provides relevant information on addressing matters related to regulatory issues and risk management of potentially exposed animals and animal care staff.Abbreviations: CDC, Centers for Disease Control and Prevention; IHA, indirect hemagglutination assay; PEP, postexposure prophylacticBurkholderia pseudomallei, formerly known as Pseudomonas pseudomallei, is a gram-negative, aerobic, bipolar, motile, rod-shaped bacterium. B. pseudomallei infections (melioidosis) can be severe and even fatal in both humans and animals. This environmental saprophyte is endemic to Southeast Asia and northern Australia, but it has also been found in other tropical and subtropical areas of the world.7,22,32,42 The bacterium is usually found in soil and water in endemic areas and is transmitted to humans and animals primarily through percutaneous inoculation, ingestion, or inhalation of a contaminated source.8, 22,28,32,42 Human-to-human, animal-to-animal, and animal-to-human spread are rare.8,32 In December 2012, the National Select Agent Registry designated B. pseudomallei as a Tier 1 overlap select agent.39 Organisms classified as Tier 1 agents present the highest risk of deliberate misuse, with the most significant potential for mass casualties or devastating effects to the economy, critical infrastructure, or public confidence. Select agents with this status have the potential to pose a severe threat to human and animal health or safety or the ability to be used as a biologic weapon.39Melioidosis in humans can be challenging to diagnose and treat because the organism can remain latent for years and is resistant to many antibiotics.12,37,41 B. pseudomallei can survive in phagocytic cells, a phenomenon that may be associated with latent infections.19,38 The incubation period in naturally infected animals ranges from 1 d to many years, but symptoms typically appear 2 to 4 wk after exposure.13,17,35,38 Disease generally presents in 1 of 2 forms: localized infection or septicemia.22 Multiple methods are used to diagnose melioidosis, including immunofluorescence, serology, and PCR analysis, but isolation of the bacteria from blood, urine, sputum, throat swabs, abscesses, skin, or tissue lesions remains the ‘gold standard.’9,22,40,42 The prognosis varies based on presentation, time to diagnosis, initiation of appropriate antimicrobial treatment, and underlying comorbidities.7,28,42 Currently, there is no licensed vaccine to prevent melioidosis.There are several published reports of naturally occurring melioidosis in a variety of nonhuman primates (NHP; 2,10,13,17,25,30,31,35 The first reported case of melioidosis in monkeys was recorded in 1932, and the first published case in a macaque species was in 1966.30 In the United States, there have only been 7 documented cases of NHP with B. pseudomallei infection.2,13,17 All of these cases occurred prior to the classification of B. pseudomallei as a select agent. Clinical signs in NHP range from subclinical or subacute illness to acute septicemia, localized infection, and chronic infection. NHP with melioidosis can be asymptomatic or exhibit clinical signs such as anorexia, wasting, purulent drainage, subcutaneous abscesses, and other soft tissue lesions. Lymphadenitis, lameness, osteomyelitis, paralysis and other CNS signs have also been reported.2,7,10,22,28,32 In comparison, human''s clinical signs range from abscesses, skin ulceration, fever, headache, joint pain, and muscle tenderness to abdominal pain, anorexia, respiratory distress, seizures, and septicemia.7,9,21,22

Table 1.

Summary of reported cases of naturally occurring Burkholderia pseudomalleiinfections in nonhuman primates
CountryaImported fromDate reportedSpeciesReference
AustraliaBorneo1963Pongo sp.36
BruneiUnknown1982Orangutan (Pongo pygmaeus)33
France1976Hamlyn monkey (Cercopithecus hamlyni) Patas monkey (Erythrocebus patas)11
Great BritainPhilippines and Indonesia1992Cynomolgus monkey (Macaca fascicularis)10
38
MalaysiaUnknown1966Macaca spp.30
Unknown1968Spider monkey (Brachytelis arachnoides) Lar gibbon (Hylobates lar)20
Unknown1969Pig-tailed macaque (Macaca nemestrina)35
Unknown1984Banded leaf monkey (Presbytis melalophos)25
SingaporeUnknown1995Gorillas, gibbon, mandrill, chimpanzee43
ThailandUnknown2012Monkey19
United StatesThailand1970Stump-tailed macaque (Macaca arctoides)17
IndiaPig-tailed macaque (Macaca nemestrina)
AfricaRhesus macaque (Macaca mulatta) Chimpanzee (Pan troglodytes)
Unknown1971Chimpanzee (Pan troglodytes)3
Malaysia1981Pig-tailed macaque (Macaca nemestrina)2
Wild-caught, unknown1986Rhesus macaque (Macaca mulatta)13
Indonesia2013Pig-tailed macaque (Macaca nemestrina)Current article
Open in a separate windowaCountry reflects the location where the animal was housed at the time of diagosis.Here we describe a case of melioidosis diagnosed in a pigtail macaque (Macaca nemestrina) imported into the United States from Indonesia and the implications of the detection of a select agent identified in a laboratory research colony. We also discuss the management and care of the exposed colony, zoonotic concerns regarding the animal care staff that worked with the shipment of macaques, effects on research studies, and the procedures involved in reporting a select agent incident.  相似文献   

17.
Association of Campylobacter jejuni Cj0859c Gene (fspA) Variants with Different C. jejuni Multilocus Sequence Types     
C. P. A. de Haan  R. Kivist?  M. L. H?nninen 《Applied and environmental microbiology》2010,76(20):6942-6943
Cj0859c variants fspA1 and fspA2 from 669 human, poultry, and bovine Campylobacter jejuni strains were associated with certain hosts and multilocus sequence typing (MLST) types. Among the human and poultry strains, fspA1 was significantly (P < 0.001) more common than fspA2. FspA2 amino acid sequences were the most diverse and were often truncated.Campylobacter jejuni is the leading cause of bacterial gastroenteritis worldwide and responsible for more than 90% of Campylobacter infections (7). Case-control studies have identified consumption or handling of raw and undercooked poultry meat, drinking unpasteurized milk, and swimming in natural water sources as risk factors for acquiring domestic campylobacteriosis in Finland (7, 9). Multilocus sequence typing (MLST) has been employed to study the molecular epidemiology of Campylobacter (4) and can contribute to virulotyping when combined with known virulence factors (5). FspA proteins are small, acidic, flagellum-secreted nonflagellar proteins of C. jejuni that are encoded by Cj0859c, which is expressed by a σ28 promoter (8). Both FspA1 and FspA2 were shown to be immunogenic in mice and protected against disease after challenge with a homologous strain (1). However, FspA1 also protected against illness after challenge with a heterologous strain, whereas FspA2 failed to do the same at a significant level. Neither FspA1 nor FspA2 protected against colonization (1). On the other hand, FspA2 has been shown to induce apoptosis in INT407 cells, a feature not exhibited by FspA1 (8). Therefore, our aim was to study the distributions of fspA1 and fspA2 among MLST types of Finnish human, chicken, and bovine strains.In total, 367 human isolates, 183 chicken isolates, and 119 bovine isolates (n = 669) were included in the analyses (3). PCR primers for Cj0859c were used as described previously (8). Primer pgo6.13 (5′-TTGTTGCAGTTCCAGCATCGGT-3′) was designed to sequence fspA1. Fisher''s exact test or a chi-square test was used to assess the associations between sequence types (STs) and Cj0859c. The SignalP 3.0 server was used for prediction of signal peptides (2).The fspA1 and fspA2 variants were found in 62.6% and 37.4% of the strains, respectively. In 0.3% of the strains, neither isoform was found. Among the human and chicken strains, fspA1 was significantly more common, whereas fspA2 was significantly more frequent among the bovine isolates (Table (Table1).1). Among the MLST clonal complexes (CCs), fspA1 was associated with the ST-22, ST-45, ST-283, and ST-677 CCs and fspA2 was associated with the ST-21, ST-52, ST-61, ST-206, ST-692, and ST-1332 CCs and ST-58, ST-475, and ST-4001. Although strong CC associations of fspA1 and fspA2 were found, the ST-48 complex showed a heterogeneous distribution of fspA1 and fspA2. Most isolates carried fspA2, and ST-475 was associated with fspA2. On the contrary, ST-48 commonly carried fspA1 (Table (Table1).1). In our previous studies, ST-48 was found in human isolates only (6), while ST-475 was found in both human and bovine isolates (3, 6). The strict host associations and striking difference between fspA variants in human ST-48 isolates and human/bovine ST-475 isolates suggest that fspA could be important in host adaptation.

TABLE 1.

Percent distributions of fspA1 and fspA2 variants among 669 human, poultry, and bovine Campylobacter jejuni strains and their associations with hosts, STs, and CCs
Host or ST complex/ST (no. of isolates)% of strains witha:
P valueb
fspA1fspA2
Host
    All (669)64.335.4
    Human (367)69.530.0<0.001
    Poultry (183)79.220.8<0.001
    Bovine (119)25.274.8<0.0001
ST complex and STs
    ST-21 complex (151)2.697.4<0.0001
        ST-50 (76)NF100<0.0001
        ST-53 (19)NF100<0.0001
        ST-451 (9)NF100<0.0001
        ST-883 (11)NF100<0.0001
    ST-22 complex (22)100NF<0.0001
        ST-22 (11)100NF<0.01
        ST-1947 (9)100NF0.03
    ST-45 complex (268)99.30.7<0.0001
        ST-11 (7)100NFNA
        ST-45 (173)99.40.6<0.0001
        ST-137 (22)95.54.50.001
        ST-230 (14)100NF<0.0001
    ST-48 complex (18)44.455.6NA
        ST-48 (7)100NFNA
        ST-475 (8)NF100<0.001
    ST-52 complex (5)NF100<0.01
        ST-52 (4)NF1000.02
    ST-61 complex (21)NF100<0.0001
        ST-61 (11)NF100<0.0001
        ST-618 (3)NF1000.04
    ST-206 complex (5)NF100<0.01
    ST-283 complex (24)100NF<0.0001
        ST-267 (23)100NF<0.0001
    ST-677 complex (59)100NF<0.0001
        ST-677 (48)100NF<0.0001
        ST-794 (11)100NF<0.001
    ST-692 complex (3)NF1000.04
    ST-1034 complex (5)NF80NA
        ST-4001 (3)NF1000.04
    ST-1287 complex/ST-945 (8)100NFNA
    ST-1332 complex/ST-1332 (4)NF1000.02
    Unassigned STs
        ST-58 (6)NF100<0.01
        ST-586 (6)100NFNA
Open in a separate windowaIn 0.3% of the strains, neither isoform was found. NF, not found.bNA, not associated.A total of 28 isolates (representing 6 CCs and 13 STs) were sequenced for fspA1 and compared to reference strains NCTC 11168 and 81-176. All isolates in the ST-22 CC showed the same one-nucleotide (nt) difference with both NCTC 11168 and 81-176 strains, resulting in a Thr→Ala substitution in the predicted protein sequence (represented by isolate FB7437, GenBank accession number HQ104931; Fig. Fig.1).1). Eight other isolates in different CCs showed a 2-nt difference (isolate 1970, GenBank accession number HQ104932; Fig. Fig.1)1) compared to strains NCTC 11168 and 81-176, although this did not result in amino acid substitutions. All 28 isolates were predicted to encode a full-length FspA1 protein.Open in a separate windowFIG. 1.Comparison of FspA1 and FspA2 isoforms. FspA1 is represented by 81-176, FB7437, and 1970. FspA2 is represented by C. jejuni strains 76763 to 1960 (GenBank accession numbers HQ104933 to HQ104946). Scale bar represents amino acid divergence.In total, 62 isolates (representing 7 CCs and 35 STs) were subjected to fspA2 sequence analysis. Although a 100% sequence similarity between different STs was found for isolates in the ST-21, ST-45, ST-48, ST-61, and ST-206 CCs, fspA2 was generally more heterogeneous than fspA1 and we found 13 predicted FspA2 amino acid sequence variants in total (Fig. (Fig.1).1). In several isolates with uncommon and often unassigned (UA) STs, the proteins were truncated (Fig. (Fig.1),1), with most mutations being ST specific. For example, all ST-58 isolates showed a 13-bp deletion (isolate 3074_2; Fig. Fig.1),1), resulting in a premature stop codon. Also, all ST-1332 CC isolates were predicted to have a premature stop codon by the addition of a nucleotide between nt 112 and nt 113 (isolate 1960; Fig. Fig.1),1), a feature shared with two isolates typed as ST-4002 (UA). A T68A substitution in ST-1960 (isolate T-73494) also resulted in a premature stop codon. Interestingly, ST-1959 and ST-4003 (represented by isolate 4129) both lacked one triplet (nt 235 to 237), resulting in a shorter FspA2 protein. SignalP analysis showed the probability of a signal peptide between nt 22 and 23 (ACA-AA [between the underlined nucleotides]). An A24C substitution in two other strains, represented by isolate 76580, of ST-693 and ST-993 could possibly result in a truncated FspA2 protein as well.In conclusion, our results showed that FspA1 and FspA2 showed host and MLST associations. The immunogenic FspA1 seems to be conserved among C. jejuni strains, in contrast to the heterogeneous apoptosis-inducing FspA2, of which many isoforms were truncated. FspA proteins could serve as virulence factors for C. jejuni, although their roles herein are not clear at this time.  相似文献   

18.
Detection and Quantification of the Coral Pathogen Vibrio coralliilyticus by Real-Time PCR with TaqMan Fluorescent Probes     
F. Joseph Pollock  Pamela J. Morris  Bette L. Willis  David G. Bourne 《Applied and environmental microbiology》2010,76(15):5282-5286
A real-time quantitative PCR-based detection assay targeting the dnaJ gene (encoding heat shock protein 40) of the coral pathogen Vibrio coralliilyticus was developed. The assay is sensitive, detecting as little as 1 CFU per ml in seawater and 104 CFU per cm2 of coral tissue. Moreover, inhibition by DNA and cells derived from bacteria other than V. coralliilyticus was minimal. This assay represents a novel approach to coral disease diagnosis that will advance the field of coral disease research.Vibrio coralliilyticus has recently emerged as a coral pathogen of concern on reefs throughout the Indo-Pacific. It was first implicated as the etiological agent responsible for bleaching and tissue lysis of the coral Pocillopora damicornis on Zanzibar reefs (2). More recently, V. coralliilyticus has been identified as the causative agent of white syndrome (WS) outbreaks on several Pacific reefs (14). WS is a collective term describing coral diseases characterized by a spreading band of tissue loss exposing white skeleton on Indo-Pacific scleractinian corals (16). V. coralliilyticus is an emerging model pathogen for understanding the mechanisms linking bacterial infection and coral disease (13) and therefore provides an ideal model for the development of diagnostic assays to detect coral disease. Current coral disease diagnostic methods, which are based primarily upon field-based observations of macroscopic disease signs, often detect disease only at the latest stages of infection, when control measures are least effective. The development of diagnostic tools targeting pathogens underlying coral disease pathologies may provide early indications of infection, aid the identification of disease vectors and reservoirs, and assist managers in developing strategies to prevent the spread of coral disease outbreaks. In this paper, we describe the development and validation of a TaqMan-based real-time quantitative PCR (qPCR) assay that targets a segment of the V. coralliilyticus heat shock protein 40-encoding gene (dnaJ).Nucleotide sequences of the dnaJ gene were retrieved from relevant Vibrio species, including V. coralliilyticus (LMG 20984), using the National Center for Biotechnology Information''s (NCBI) Entrez Nucleotide Database search tool (http://www.ncbi.nlm.nih.gov/). Gene sequences of strains not available in public databases (V. coralliilyticus strains LMG 21348, LMG 21349, LMG 21350, LMG 10953, LMG 20538, LMG 23696, LMG 23691, LMG 23693, LMG 23692, and LMG 23694) were obtained through extraction of total DNA using a Promega Wizard Prep DNA Purification Kit (Promega, Sydney, Australia), PCR amplification, and sequencing using primers and thermal cycling parameters described by Nhung et al. (8). A 128-bp region (nucleotides 363 to 490) containing high concentrations of single nucleotide polymorphisms (SNPs), which were conserved within V. coralliilyticus strains but differed from non-V. coralliilyticus strains, was identified, and oligonucleotide primers Vc_dnaJ_F1 (5′-CGG TTC GYG GTG TTT CAA AA-3′) and Vc_dnaJ_R1 (5′-AAC CTG ACC ATG ACC GTG ACA-3′) and a TaqMan probe, Vc_dnaJ_TMP (5′-6-FAM-CAG TGG CGC GAA G-MGBNFQ-3′; 6-FAM is 6-carboxyfluorescein and MGBNFQ is molecular groove binding nonfluorescent quencher), were designed to target this region. The qPCR assay was optimized and validated using DNA extracted from V. coralliilyticus isolates, nontarget Vibrio species, and other bacterial species grown in marine broth (MB) (Table (Table1),1), under the following optimal conditions: 1× TaqMan buffer A, 0.5 U of AmpliTaq Gold DNA polymerase, 200 μM deoxynucleotide triphosphates (with 400 μM dUTP replacing deoxythymidine triphosphate), 0.2 U of AmpErase uracil N-glycosylase (UNG), 3 mM MgCl2, 0.6 μM each primer, 0.2 μM fluorophore-labeled TaqMan, 1 μl of template, and sterile MilliQ water for a total reaction volume to 20 μl. All assays were conducted on a RotoGene 300 (Corbett Research, Sydney, Australia) real-time analyzer with the following cycling parameters: 50°C for 120 s (UNG activation) and 95°C for 10 min (AmpliTaq Gold DNA polymerase activation), followed by 40 cycles of 95°C for 15 s (denaturation) and 60°C for 60 s (annealing/extension). During the annealing/extension phase of each thermal cycle, fluorescence was measured in the FAM channel (470-nm excitation and 510-nm detection).

TABLE 1.

Species, strain, and threshold cycle for all bacterial strains testeda
SpeciesStrainbOriginHost organismCT ± SEMcdnaJ gene sequence accession no.Reference
Vibrio coralliilyticusLMG 23696Nelly Bay, Magnetic Island, AustraliaMontipora aequituberculata12.43 ± 0.20HM21557014
LMG 23691Majuro Atoll, Republic of Marshall IslandsAcropora cytherea14.07 ± 1.33HM21557114
LMG 23693Nikko Bay, PalauPachyseris speciosa10.83 ± 2.76HM21557214
LMG 23692Nikko Bay, PalauPachyseris speciosa9.40 ± 0.36HM21557314
LMG 23694Nikko Bay, PalauPachyseris speciosad12.54 ± 0.24HM21557414
LMG 20984TIndian Ocean, Zanzibar, TanzaniaPocillopora damicornis12.80 ± 0.71HM2155752
LMG 21348Red Sea, Eilat, IsraelPocillopora damicornis13.81 ± 0.49HM2155763
LMG 21349Red Sea, Eilat,Pocillopora damicornis12.98 ± 0.94HM2155773
LMG 21350Red Sea, Eilat,Pocillopora damicornis11.49 ± 0.19HM2155783
LMG 10953Kent, United KingdomCrassostrea gigas (oyster) larvae10.53 ± 0.40HM2155793
LMG 20538Atlantic Ocean, Florianópolis, BrazilNodipecten nodosus (bivalve) larvae12.13 ± 0.50HM2155803
C1Caribbean Sea, La Parguera, Puerto RicoPseudopterogorgia americana14.53 ± 0.28HM21556815
C2Caribbean Sea, La Parguera, Puerto RicoPseudopterogorgia americanaNAHM21556915
Vibrio alginolyticusATCC 1774933.74 ± 0.33
Vibio brasiliensisDSM 1718437.84†
Vibrio calviensisDSM 1434727.06 ± 0.52
Vibrio campbelliiATCC 25920T39.10†
Enterovibrio campbelliiLMG 2136337.33 ± 2.41
Alliivibrio fischeriDSM 50731.36 ± 1.42
Vibrio fortisDSM 19133NA
Vibrio furnissiiDSM 19622NA
Vibrio harveyiDSM 19623NA
Vibrio natriegensATCC 1404828.56 ± 0.60
Vibrio neptuniusLMG 20536NA
Vibrio ordaliiATCC 3350925.56 ± 0.41
Vibrio parahaemolyticusATCC 17802NA
Vibrio proteolyticusATCC 1533830.00 ± 0.89††
Vibrio rotiferianusLMG 21460NA
Vibrio splendidusATCC 3312532.31 ± 0.82
Vibrio tubiashiiATCC 19109NA
Vibrio xuiiLMG 21346NA
Escherichia coliATCC 25922NA
Psychrobacter sp.AIMS 1618NA
Shewanella sp.AIMS C04125.34 ± 0.45
Open in a separate windowaOrigin, host organism, and dnaJ gene sequence accession numbers are shown for V. coralliilyticus strains.bStrain designations beginning with LMG were derived from the Belgian Coordinated Collections of Microorganisms, ATCC strains are from the American Type Culture Collection, DSM strains are from the Deutsche Sammlung von Mikroorganismen und Zellkulturen GmbH culture collection, AIMS strains are from the Australian Institute of Marine Science culture collection, and C1 and C2 were provided by Pamela Morris.c†, amplification in one of three reactions; ††, amplification in two of three reactions; NA, no amplification.dIsolated from seawater above coral.The qPCR assay specifically detected 12 out of 13 isolated V. coralliilyticus strains tested in this study (Table (Table1).1). The exception was one Caribbean strain (C2), which failed to give specific amplification despite repeated attempts. Positive detection of the target gene segment was determined by the increase in fluorescent signal beyond the fluorescence threshold value (normalized fluorescence, 0.010) at a specific cycle, referred to as the threshold cycle (CT). Specific detection was further confirmed by gel electrophoresis, which revealed a PCR product of the correct theoretical size (128 bp) (data not shown), and DNA sequencing, which confirmed the target amplified product to be a segment of the dnaJ gene. No amplification with the assay was detected for 13 other closely related Vibrio strains, including the closely related Vibrio neptunius and two non-Vibrio species (Table (Table1).1). A total of five other Vibrio strains and one non-Vibrio strain (Shewanella sp.) exhibited CT values less than the cutoff of 32 cycles. However, CT values for these strains (mean ± standard error of the mean [SEM], 27.96 ± 2.40) were all much higher than those for V. coralliilyticus strains (12.30 ± 1.52), and no amplicons were evident in post-qPCR gel electrophoresis (data not shown).The detection limit for purified V. coralliilyticus genomic DNA was 0.1 pg of DNA, determined by performing 10-fold serial dilutions (100 ng to 0.01 pg per reaction), followed by qPCR amplification. Similarly, qPCR assays of serial dilutions of V. coralliilyticus (LMG 23696) cells cultured overnight in MB (108 CFU ml−1 to extinction) were able to detect as few as 104 CFU (Fig. (Fig.1).1). Standard curves revealed a strong linear negative correlation between CT values and both DNA and cell concentrations of V. coralliilyticus over several orders of magnitude, with r2 values of 0.998 and 0.953 for DNA and cells, respectively (Fig. (Fig.11).Open in a separate windowFIG. 1.Standard curves delineating threshold (CT) values of fluorescence for indicators of pathogen presence: (A) concentration of V. coralliilyticus DNA and (B) number of V. coralliilyticus cells in pure culture. Error bars indicate standard error of the mean for three replicate qPCRs.Little interference of the qPCR assay was observed when purified V. coralliilyticus (LMG 23696) DNA (10 ng) was combined with 10-fold serial dilutions (0.01 to 100 ng per reaction) of non-V. coralliilyticus DNA (i.e., Vibrio campbellii [ATCC 25920T]). Over the entire range of nontarget DNA concentrations tested, the resulting CT values (mean ± SEM, 17.76 ± 0.53) were not significantly different from those of a control treatment containing 10 ng of V. coralliilyticus DNA and no nonspecific DNA (16.75 ± 0.18; analysis of variance [ANOVA], P = 0.51) (Table (Table2).2). Detection of V. coralliilyticus (LMG 23696) bacterial cells (104, 105, 106, 107, or 108 CFU per ml) in a background of non-V. coralliilyticus cells (i.e., V. campbellii [ATCC 25920T] at 0, 10, 104, or 107 CFU per ml) showed little reduction in assay sensitivity (see Fig. S1 in the supplemental material). For example, when V. coralliilyticus was seeded at 107 cells with similarly high concentrations of nontarget cells, little inhibition of the assay was observed.

TABLE 2.

Effect of nontarget bacterial DNA on the detection of 10 ng of purified V. coralliilyticus DNA
Amt of nontarget DNA (ng)CT (mean ± SEM)
10016.97 ± 0.33
1016.9 ± 0.08
116.74 ± 0.10
0.117 ± 0.09
0.0116.37 ± 0.43
0a16.75 ± 0.18
NTCb35.04 ± 0.02
Open in a separate windowaV. coralliilyticus (LMG 23696) DNA (10 ng) free of nontarget DNA and cells served as positive controls.bA qPCR mixture containing no bacterial DNA served as a no-template, or negative, control (NTC).The assay''s detection limit in seawater was tested by inoculating 10-fold serial dilutions of V. coralliilyticus (LMG 23696) cultures (grown overnight in MB medium, pelleted at 14,000 rpm for 10 min, and washed twice with sterile phosphate-buffered saline [PBS]) into 1 liter of seawater (equivalent final concentrations were 106 to 1 CFU ml−1). The entire volume of V. coralliilyticus-seeded seawater was filtered through a Sterivex-GP filter (Millipore), and DNA was extracted using the method described by Schauer et al. (11). The lowest detection limit for V. coralliilyticus cells seeded into seawater was 1 CFU ml−1 (Fig. (Fig.2),2), with no detection in a 1-liter volume of an unseeded seawater negative control. Standard curves revealed a strong correlation between CT values and the concentrations of V. coralliilyticus bacteria seeded into the seawater over several orders of magnitude (r2 of 0.968) (Fig. (Fig.22).Open in a separate windowFIG. 2.Standard curves showing CT values of the fluorescent signal versus the number of V. coralliilyticus cells per ml seawater (▿), and cells per cm2 of M. aequituberculata tissue, with (○) or without (·) enrichment. Each dot represents an independent experiment. Error bars indicate standard error of the mean for three replicate qPCR runs.The detection limit in seeded coral tissue homogenate was determined by seeding 10-fold dilutions (1010 to 103 CFU ml−1) of pelleted, PBS-washed and resuspended (in 10 ml of sterile PBS) V. coralliilyticus cells onto healthy fragments (∼10 cm2) of the coral Montipora aequituberculata collected from Nelly Bay (Magnetic Island, Australia). Corals were collected in March 2009 and maintained in holding tanks supplied with flowthrough ambient seawater. Resuspended cells were inoculated onto M. aequituberculata fragments, each contained in an individual 3.8-liter plastic bag, allowed to sit at room temperature for 30 min, and then air brushed with compressed air until only white skeleton remained. One-milliliter aliquots of the resulting slurry (PBS, bacteria, and coral tissue) was vortexed for 10 min at 14,000 rpm, and DNA was extracted using a PowerPlant DNA Isolation Kit (Mo Bio, Carlsbad, CA). The lowest detection limits for V. coralliilyticus cells seeded onto coral fragments was 104 CFU per cm2 of coral tissue (Fig. (Fig.2).2). Again, standard curves revealed a strong correlation between CT values and the concentrations of seeded bacteria over several orders of magnitude (r2 of 0.981) (Fig. (Fig.2).2). When a 1-ml aliquot of the slurry was also inoculated into 25 ml of MB and enriched for 6 h at 28°C (with shaking at 170 rpm), the detection limit increased by 1 order of magnitude, to 103 CFU of V. coralliilyticus per cm2 of coral tissue (Fig. (Fig.2).2). The slope of the standard curve reveals some inhibition, particularly at the highest V. coralliilyticus concentrations, which could result from lower replication rates in the cultures with the highest bacterial densities (i.e., 109 CFU). However, since this effect is most pronounced only at the highest bacterial concentrations, the detection limit is still valid. In all trials, unseeded coral fragments and enrichment cultures derived from uninoculated coral fragments served as negative controls.The current study describes the first assay developed to detect and quantify a coral pathogen using a real-time quantitative PCR (qPCR) approach. While previous studies have utilized antibodies or fluorescent in situ hybridization (FISH) to detect coral pathogens (1, 6), the combination of high sensitivity and specificity, low contamination risk, and ease and speed of performance (5) make qPCR technology an ideal choice for rapid pathogen detection in complex hosts, such as corals. The assay developed is highly sensitive for V. coralliilyticus, detecting as few as 1 CFU ml−1 of seawater and 104 CFU cm−2 of coral tissue (103 CFU cm−2 of coral tissue with a 6-h enrichment). These detection limits are likely to be within biologically relevant pathogen concentrations. For example, antibodies for specific detection of the coral bleaching pathogen Vibrio shiloi showed that bacterial densities reached 8.4 × 108 cells cm−3 1 month prior to maximum visual bleaching signs on the coral Oculina patagonica (6). Each seeded seawater and coral (enriched and nonenriched) dilution assay was performed in triplicate. The linearity of the resulting standard curves indicates consistent extraction efficiencies over V. coralliilyticus concentrations spanning 6 orders of magnitude (Fig. (Fig.2)2) and provides strong support for the robustness of the assay. In addition, the presence of competing, non-V. coralliilyticus bacterial cells and DNA had a minimal impact on the detection of V. coralliilyticus. This is an important consideration for accurate detection within the complex coral holobiont, where the target organism is present within a matrix of other microbial and host cells.V. coralliilyticus, like V. shiloi (10), is becoming a model pathogen for the study of coral disease. Recent research efforts have characterized the organism''s genome (W. R. Johnson et al., submitted for publication), proteome (N. E. Kimes et al., submitted for publication), resistome (15), and metabolome (4) and enhanced our understanding of the genetic (7, 9) and physiological (7, 13) basis of its virulence. Before effective management response plans can be formulated, however, continuing research on the genetic and cellular aspects of V. coralliilyticus must be complemented with knowledge of the epidemiology of this pathogen, including information on its distribution, incidence of infection, and rates of transmission throughout populations. The V. coralliilyticus-specific qPCR assay developed in this study will provide important insights into the dynamics of pathogen invasion and spread within populations (6) while also aiding in the identification of disease vectors and reservoirs (12). These capabilities will play an important role in advancing the field of coral disease research and effective management of coral reefs worldwide.   相似文献   

19.
Transcriptional and Functional Classification of the GOLVEN/ROOT GROWTH FACTOR/CLE-Like Signaling Peptides Reveals Their Role in Lateral Root and Hair Formation     
Ana Fernandez  Andrzej Drozdzecki  Kurt Hoogewijs  Anh Nguyen  Tom Beeckman  Annemieke Madder  Pierre Hilson 《Plant physiology》2013,161(2):954-970
  相似文献   

20.
Mouse Models of Osteoarthritis: A Summary of Models and Outcomes Assessment     
Sabine Drevet  Bertrand Favier  Emmanuel Brun  Gaëtan Gavazzi  Bernard Lardy 《Comparative medicine》2022,72(1):3
Osteoarthritis (OA) is a multidimensional health problem and a common chronic disease. It has a substantial impact on patient quality of life and is a common cause of pain and mobility issues in older adults. The functional limitations, lack of curative treatments, and cost to society all demonstrate the need for translational and clinical research. The use of OA models in mice is important for achieving a better understanding of the disease. Models with clinical relevance are needed to achieve 2 main goals: to assess the impact of the OA disease (pain and function) and to study the efficacy of potential treatments. However, few OA models include practical strategies for functional assessment of the mice. OA signs in mice incorporate complex interrelations between pain and dysfunction. The current review provides a comprehensive compilation of mouse models of OA and animal evaluations that include static and dynamic clinical assessment of the mice, merging evaluation of pain and function by using automatic and noninvasive techniques. These new techniques allow simultaneous recording of spontaneous activity from thousands of home cages and also monitor environment conditions. Technologies such as videography and computational approaches can also be used to improve pain assessment in rodents but these new tools must first be validated experimentally. An example of a new tool is the digital ventilated cage, which is an automated home-cage monitor that records spontaneous activity in the cages.

Osteoarthritis (OA) is a multidimensional health problem and a common chronic disease.36 Functional limitations, the absence of curative treatments, and the considerable cost to society result in a substantial impact on quality of life.76 Historically, OA has been described as whole joint and whole peri-articular diseases and as a systemic comorbidity.9,111 OA consists of a disruption of articular joint cartilage homeostasis leading to a catabolic pathway characterized by chondrocyte degeneration and destruction of the extracellular matrix (ECM). Low-grade chronic systemic inflammation is also actively involved in the process.42,92 In clinical practice, mechanical pain, often accompanied by a functional decline, is the main reason for consultations. Recommendations to patients provide guidance for OA management.22, 33,49,86 Evidence-based consensus has led to a variety of pharmacologic and nonpharmacologic modalities that are intended to guide health care providers in managing symptomatic patients. Animal-based research is of tremendous importance for the study of early diagnosis and treatment, which are crucial to prevent the disease progression and provide better care to patients.The purpose of animal-based OA research is 2-fold: to assess the impact of the OA disease (pain and function) and to study the efficacy of a potential treatment.18,67 OA model species include large animals such as the horse, goat, sheep, and dog, whose size and anatomy are expected to better reflect human joint conditions. However, small animals such as guinea pig, rabbit, mouse, and rat represent 77% of the species used.1,87 In recent years, mice have become the most commonly used model for studying OA. Mice have several advantageous characteristics: a short development and life span, easy and low-cost breeding and maintenance, easy handling, small joints that allow histologic analysis of the whole joint,32 and the availability of genetically modified lines.108 Standardized housing, genetically defined strains and SPF animals reduce the genetic and interindividual acquired variability. Mice are considered the best vertebrate model in terms of monitoring and controlling environmental conditions.7,14,15,87 Mouse skeletal maturation is reached at 10 wk, which theoretically constitutes the minimal age at which mice should be entered into an OA study.64,87,102 However, many studies violate this limit by testing mice at 8 wk of age.Available models for OA include the following (32,111 physical activity and exercise induced OA; noninvasive mechanical loading (repetitive mild loading and single-impact injury); and surgically induced (meniscectomy models or anterior cruciate ligament transection). The specific model used would be based on the goal of the study.7 For example, OA pathophysiology, OA progression, and OA therapies studies could use spontaneous, genetic, surgical, or noninvasive models. In addition, pain studies could use chemical models. Lastly, post-traumatic studies would use surgical or noninvasive models; the most frequently used method is currently destabilization of the medial meniscus,32 which involves transection of the medial meniscotibial ligament, thereby destabilizing the joint and causing instability-driven OA. An important caveat for mouse models is that the mouse and human knee differ in terms of joint size, joint biomechanics, and histologic characteristics (layers, cellularity),32,64 and joint differences could confound clinical translation.10 Table 1. Mouse models of osteoarthritis.
ModelsProsCons
SpontaneousWild type mice7,9,59,67,68,70,72,74,80,85,87,115,118,119,120- Model of aging phenotype
- The less invasive model
- Physiological relevance: mimics human pathogenesis
- No need for technical expertise
- No need for specific equipment
- Variability in incidence
- Large number of animals at baseline
- Long-term study: Time consuming (time of onset: 4 -15 mo)
- Expensive (husbandry)
Genetically modified mice2,7,25,40,50,52,67,72,79,80, 89,120- High incidence
- Earlier time of onset: 18 wk
- No need for specific equipment
- Combination with other models
- Time consuming for the strain development
- Expensive
Chemical- inducedMono-iodoacetate injection7,11,46,47,60,66,90,91,101,128- Model of pain-like phenotype
- To study mechanism of pain and antalgic drugs
- Short-term study: Rapid progression (2-7 wk)
- Reproducible
- Low cost
- Need for technical expertise
- Need for specific equipment
- Systemic injection is lethal
- Destructive effect: does not allow to study the early phase of pathogenesis
Papain injection66,67,120- Short-term study: rapid progression
- Low cost
- Need for technical expertise
- Need for specific equipment
- Does not mimic natural pathogenesis
Collagenase injection7,65,67,98- Short-term study: rapid progression (3 wk)
- Low cost
- Need for technical expertise
- Need for specific equipment
- Does not mimic natural pathogenesis
Non-invasiveHigh-fat diet (Alimentary induced obesity model)5,8,43,45,57,96,124Model of metabolic phenotype
No need for technical expertise
No need for specific equipment
Reproducible
Long-term study: Time consuming (8 wk–9 mo delay)
Expensive
Physical activity and exercise model45,73Model of post traumatic phenotype
No need for technical expertise
Long-term study: time consuming (18 mo delay)
Expensive
Disparity of results
Mechanical loading models Repetitive mild loading models Single-impact injury model7,16,23,24, 32,35,104,105,106Model of post traumatic phenotype
Allow to study OA development
Time of onset: 8-10 wk post injury
Noninvasive
Need for technical expertise
Need for specific equipment
Heterogeneity in protocol practices
Repetitive anesthesia required or ethical issues
SurgicalOvariectomy114Contested.
Meniscectomy model7,32,63,67,87 Model of post traumatic phenotype
High incidence
Short-term study: early time of onset (4 wk from surgery)
To study therapies
Need for technical expertise
Need for specific equipment
Surgical risks
Rapid progression compared to human
Anterior cruciate ligament transection (ACLT)7,39,40,61,48,67,70,87,126Model of posttraumatic phenotype
High incidence
Short-term study: early time of onset (3-10 wk from surgery)
Reproducible
To study therapies
Need for technical expertise
Need for specific equipment
Surgical risks
Rapid progression compared to human
Destabilization of medial meniscus (DMM)7,32,39,40Model of post traumatic phenotype
High incidence
Short-term study: early time of onset (4 wk from surgery)
To study therapies
The most frequently used method
Need for technical expertise
Need for specific equipment
Surgical risks
Rapid progression compared to human
Open in a separate windowSince all animal models have strengths and weaknesses, it is often best to plan using a number of models and techniques together to combine the results.In humans, the lack of correlation between OA imaging assessment and clinical signs highlights the need to consider the functional data and the quality of life to personalize OA management. Clinical outcomes are needed to achieve 2 main goals: to assess the impact of the OA in terms of pain and function and to study the efficacy of treatments.65 Recent reviews offer few practical approaches to mouse functional assessment and novel approaches to OA models in mice.7,32,67,75,79,83,87, 100,120 This review will focus on static and dynamic clinical assessment of OA using automatic and noninvasive emerging techniques (Test nameTechniquesKind of assessmentOutputSpecific equipment requiredStatic measurementVon Frey filament testingCalibrated nylon filaments of various thickness (and applied force) are pressed against the skin of the plantar surface of the paw in ascending order of forceStimulus- evoked pain-like behavior
Mechanical stimuli - Tactile allodynia
The most commonly used testLatency to paw withdrawal
and
Force exerted are recordedYesKnee extension testApply a knee extension on both the intact and affected knee
or
Passive extension range of the operated knee joint under anesthesiaStimulus-evoked pain-like behaviorNumber of vocalizations evoked in 5 extensionsNoneHotplateMouse placed on hotplate. A cutoff latency has been determined to avoid lesionsStimulus-evoked pain-like behavior
Heat stimuli- thermal sensitivityLatency of paw withdrawalYesRighting abilityMouse placed on its backNeuromuscular screeningLatency to regain its footingNoneCotton swab testBringing a cotton swab into contact with eyelashes, pinna, and whiskersStimulus-evoked pain-like behavior
Neuromuscular screeningWithdrawal or twitching responseNoneSpontaneous activitySpontaneous cage activityOne by one the cages must be laid out in a specific platformSpontaneous pain behavior
Nonstimulus evoked pain
ActivityVibrations evoked by animal movementsYesOpen field analysisExperiment is performed in a clear chamber and mice can freely exploreSpontaneous pain behavior
Nonstimulus evoked pain
Locomotor analysisPaw print assessment
Distance traveled, average walking speed, rest time, rearingYesGait analysisMouse is placed in a specific cage equipped with a fluorescent tube and a glass plate allowing an automated quantitative gait analysisNonstimulus evoked pain
Gait analysis
Indirect nociceptionIntensity of the paw contact area, velocity, stride frequency, length, symmetry, step widthYesDynamic weight bearing systemMouse placed is a specific cage. This method is a computerized capacitance meter (similar to gait analysis)Nonstimulus evoked pain
Weight-bearing deficits
Indirect nociceptionBody weight redistribution to a portion of the paw surfaceYesVoluntary wheel runningMouse placed is a specific cage with free access to stainless steel activity wheels. The wheel is connected to a computer that automatically record dataNonstimulus evoked pain
ActivityDistance traveled in the wheelYesBurrowing analysisMouse placed is a specific cage equipped with steel tubes (32 cm in length and 10 cm in diameter) and quartz sand in Plexiglas cages (600 · 340x200 mm)Nonstimulus evoked pain
ActivityAmount of sand burrowedYesDigital video recordingsMouse placed is a specific cage according to the toolNonstimulus evoked pain
Or
Evoked painScale of pain or specific outcomeYesDigital ventilated cage systemNondisrupting capacitive-based technique: records spontaneous activity 24/7, during both light and dark phases directly from the home cage rackSpontaneous pain behavior
Nonstimulus evoked pain
Activity-behaviorDistance walked, average speed, occupation front, occupation rear, activation density.
Animal locomotion index, animal tracking distance, animal tracking speed, animal running wheel distance and speed or rotationYesChallenged activityRotarod testGradual and continued acceleration of a rotating rod onto which mice are placedMotor coordination
Indirect nociceptionRotarod latency: riding time and speed with a maximum cut off.YesHind limb and fore grip strengthMouse placed over a base plate in front of a connected grasping toolMuscle strength of limbsPeak force, time resistanceYesWire hang analysisSuspension of the mouse on the wire and start the timeMuscle strength of limbs: muscle function and coordinationLatency to fall grippingNone
(self -constructed)
Open in a separate windowPain cannot be directly measured in rodents, so methods have been developed to quantify “pain-like” behaviors. The clinical assessment of mice should be tested both before and after the intervention (induced-OA ± administration of treatment) to take into account the habituation and establish a baseline to compare against.  相似文献   

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