14
views
0
recommends
+1 Recommend
0 collections
    0
    shares
      • Record: found
      • Abstract: found
      • Article: found
      Is Open Access

      Leptospira infection in rats: A literature review of global prevalence and distribution

      review-article

      Read this article at

      Bookmark
          There is no author summary for this article yet. Authors can add summaries to their articles on ScienceOpen to make them more accessible to a non-specialist audience.

          Abstract

          Background

          The role of rodents in Leptospira epidemiology and transmission is well known worldwide. Rats are known to carry different pathogenic serovars of Leptospira spp. capable of causing disease in humans and animals. Wild rats ( Rattus spp.), especially the Norway/brown rat ( Rattus norvegicus) and the black rat ( R. rattus), are the most important sources of Leptospira infection, as they are abundant in urban and peridomestic environments. In this study, we compiled and summarized available data in the literature on global prevalence of Leptospira exposure and infection in rats, as well as compared the global distribution of Leptospira spp. in rats with respect to prevalence, geographic location, method of detection, diversity of serogroups/serovars, and species of rat.

          Methods

          We conducted a thorough literature search using PubMed without restrictions on publication date as well as Google Scholar to manually search for other relevant articles. Abstracts were included if they described data pertaining to Leptospira spp. in rats ( Rattus spp.) from any geographic region around the world, including reviews. The data extracted from the articles selected included the author(s), year of publication, geographic location, method(s) of detection used, species of rat(s), sample size, prevalence of Leptospira spp. (overall and within each rat species), and information on species, serogroups, and/or serovars of Leptospira spp. detected.

          Findings

          A thorough search on PubMed retrieved 303 titles. After screening the articles for duplicates and inclusion/exclusion criteria, as well as manual inclusion of relevant articles, 145 articles were included in this review. Leptospira prevalence in rats varied considerably based on geographic location, with some reporting zero prevalence in countries such as Madagascar, Tanzania, and the Faroe Islands, and others reporting as high as >80% prevalence in studies done in Brazil, India, and the Philippines. The top five countries that were reported based on number of articles include India ( n = 13), Malaysia ( n = 9), Brazil ( n = 8), Thailand ( n = 7), and France ( n = 6). Methods of detecting or isolating Leptospira spp. also varied among studies. Studies among different Rattus species reported a higher Leptospira prevalence in R. norvegicus. The serovar Icterohaemorrhagiae was the most prevalent serovar reported in Rattus spp. worldwide. Additionally, this literature review provided evidence for Leptospira infection in laboratory rodent colonies within controlled environments, implicating the zoonotic potential to laboratory animal caretakers.

          Conclusions

          Reports on global distribution of Leptospira infection in rats varies widely, with considerably high prevalence reported in many countries. This literature review emphasizes the need for enhanced surveillance programs using standardized methods for assessing Leptospira exposure or infection in rats. This review also demonstrated several weaknesses to the current methods of reporting the prevalence of Leptospira spp. in rats worldwide. As such, this necessitates a call for standardized protocols for the testing and reporting of such studies, especially pertaining to the diagnostic methods used. A deeper understanding of the ecology and epidemiology of Leptospira spp. in rats in urban environments is warranted. It is also pertinent for rat control programs to be proposed in conjunction with increased efforts for public awareness and education regarding leptospirosis transmission and prevention.

          Author summary

          The role of rodents in the transmission of many diseases, including leptospirosis, is widely known. Rats abundant in urban and peridomestic environments are the most important reservoirs and sources of Leptospira infection in humans and animals. Leptospirosis is a significant but neglected disease of humans and animals that is increasing in incidence in regions affected by natural disasters. This paper summarizes the global prevalence and distribution of Leptospira infection in rats and will add to the literature that supports research, education, and public awareness regarding leptospirosis transmission and prevention.

          Related collections

          Most cited references155

          • Record: found
          • Abstract: found
          • Article: found
          Is Open Access

          Detection of Zoonotic Pathogens and Characterization of Novel Viruses Carried by Commensal Rattus norvegicus in New York City

          INTRODUCTION Zoonotic pathogens comprise a significant and increasing proportion of all new and emerging human infectious diseases (1, 2). Although zoonotic transmission is influenced by many factors, the frequency of contact between animal reservoirs and the human population appears to be a key element (3). Therefore, the risk of zoonotic transmission is increased by events that act to reduce the geographic or ecological separation between human and animal populations or increase the density and abundance of these populations where they coexist (2, 4). In this context, rapid and continuous urbanization constitutes a significant challenge to human health, as it creates irreversible changes to biodiversity that are driven by varied responses from animal species. In particular, species classified as urban exploiters and urban adapters may exist in unnaturally large and dense populations within urban environments and have above-average rates of contact with people (5 – 7). Of these, few species have been as successful at adapting to a peridomestic lifestyle as the Norway rat (Rattus norvegicus). In the urban environment, Norway rats closely cohabitate with humans—living inside buildings, feeding on refuse, and coming into contact with many aspects of the food supply (7 – 9). These characteristics, coupled with high levels of fecundity, growth rates, and population densities, suggest that urban Norway rats may be an important source of zoonotic pathogens (10, 11). Indeed, the Norway rat is a known reservoir of a range of human pathogens, including hantaviruses, Bartonella spp., and Leptospira interrogans; however, little is known about the microbial diversity present in urban rat populations or the risks they may pose to human health (12 – 16). As a first step toward understanding the zoonotic disease risk posed by rats in densely urban environments, we assessed the presence and prevalence of known and novel microbes in Norway rats in New York City (NYC). We took the unique approach of using both targeted molecular assays to detect known human pathogens and unbiased high-throughput sequencing (UHTS) to identify novel viruses related to agents of human disease. We quantified the tissue distribution of these novel viruses in the host using molecular methods and, in some cases, identified the site(s) of replication using strand-specific quantitative reverse transcription (RT)-PCR (ssqPCR). Unlike previous urban studies that have primarily relied on serological assays to assess the total prevalence of historic infection, our data provide a snapshot estimate of the current level of infection in the rat population, a parameter more closely related to the risk of zoonotic transmission (12, 17). Furthermore, as previous work has focused on rats found exclusively in outdoor locations, we concentrated our sampling within the built environment, where direct and indirect human-rodent contact is more likely to occur (18). RESULTS Sample collection. A total of 133 Norway rats were collected from five sites in NYC. Males (n = 72) were trapped slightly more often than females (n = 61), and juveniles were trapped more often than any other age category. Of the female rats, 43% were juveniles, 26% were subadults, and 31% were sexually mature adults, whereas 40% of the male rats were juveniles, 33% were subadults, and 26% were sexually mature. Targeted molecular analyses. Specific PCR-based assays were used to screen for the presence of 18 bacterial and 2 protozoan human pathogens (see Table S1 in the supplemental material). None of the samples tested was positive for Campylobacter coli, Listeria monocytogenes, Rickettsia spp., Toxoplasma gondii, Vibrio vulnificus, or Yersinia pestis, despite previous studies documenting most of these in multiple rodent species (15). All other bacterial and protozoan pathogens were detected in at least one animal (Table 1). Three bacterial pathogens were identified in more than 15% of animals: atypical enteropathogenic Escherichia coli (EPEC) was the most common (detected in 38% of rats), followed by Bartonella spp. (25% of rats) and Streptobacillus moniliformis (17% of rats) (Table 1). Phylogenetic analysis of a 327-nucleotide (nt) region of the gltA gene amplified from all infected animals revealed three distinct Bartonella species infecting NYC rats (Fig. S1). The most common of these was detected in 76% of Bartonella-positive animals and clustered within the Bartonella tribocorum group, which has previously been identified in multiple species of Rattus in Asia and North America. In addition, sequences 98% similar to those of Bartonella rochalimae were recovered from seven rats, and Bartonella elizabethae was identified in a single rat (Fig. S2). Infection with Bartonella was positively correlated with the age of the rat (P 90% similar at the nucleotide level to viruses known to infect Norway rats (e.g., Killham rat virus, rat astrovirus, and infectious diarrhea of infant rats [IDIR] agent [group B rotavirus]) and were not pursued further (Table 2). Viruses from an additional 13 families or genera that were 0.05). Only 10 rats were infected with more than two bacterial species, of which eight were female, and no rats were infected with more than four bacterial species (Table 4). In contrast, 53 rats were positive for more than two viral agents, and 13 of these carried more than five viruses (Table 4). As many as nine different viruses or four different bacterial species were identified in the same individual, with a maximum of 11 agents detected in a single rat. Patterns of coinfection between all agents were significantly nonrandom across the complete data set (C score, P = 0.0001); however, the only significantly positive association between any two bacteria occurred between Bartonella spp. and S. moniliformis (P = 0.005). Significantly positive associations were also observed between Bartonella spp. and multiple viruses, including NrKoV-1 and NrKoV-2 (P 180 g for females, >200 g for males) (65). The rats were necropsied, and the following tissues aseptically collected: brain, heart, kidney (83 rats only), liver, lung, inguinal lymph tissue, upper and lower intestine, salivary gland with associated lymph tissue, spleen, gonads (25 rats only), and urine or bladder (when <200 µl of urine was available). Oral and rectal swab samples were collected using sterile polyester swabs (Puritan Medical Products Company, Guilford, ME), and fecal pellets were collected when available. All samples were flash-frozen immediately following collection and stored at −80°C. All procedures described in this study were approved by the Institutional Animal Care and Use Committee at Columbia University (protocol number AC-AAAE6805). Targeted molecular analyses. DNA and RNA were extracted from each tissue and fecal sample using the AllPrep DNA/RNA minikit (Qiagen, Inc.) and from urine or serum using the QIAamp viral RNA minikit (Qiagen, Inc.). The extracted DNA was quantified and diluted to a working concentration of ≤400 ng/µl. Extracted RNA was quantified, and ≤5 µg used for cDNA synthesis with SuperScript III reverse transcriptase (Invitrogen) and random hexamers. Samples were tested by PCR for 10 bacterial, protozoan, and viral human pathogens previously associated with rodents using novel and previously published PCR assays, including Bartonella spp., L. interrogans, Rickettsia spp., S. moniliformis, Y. pestis, Cryptosporidium parvum, T. gondii, hepeviruses, hantaviruses (consensus assay), and SEOV (see Table S1 in the supplemental material). Each assay was performed using a subset of the sample types from each rat, selected to include known sites of replication or shedding (Table 1). Fecal samples were further analyzed for the presence of the following eight bacterial pathogens commonly associated with human gastrointestinal disease, using PCR-based assays: C. coli, Campylobacter jejuni, C. difficile, C. perfringens, L. monocytogenes, S. enterica, V. vulnificus, and Yersinia enterocolitica (see Table S1 in the supplemental material). PCR was also used to test for the presence of pathogenic E. coli, including enteroinvasive (EIEC, including Shigella), enterohemorrhagic (EHEC), enterotoxogenic (ETEC), enteroaggregative (EAEC), and enteropathogenic (EPEC) E. coli strains, using primers targeting virulence genes (Table S1) (66). In all cases, positive PCR products were confirmed by bidirectional dideoxy sequencing. Before Ro-SaV2 detection in intestinal samples was attempted, intestines were pretreated to remove fecal contamination by thorough washing with phosphate-buffered saline (PBS). To verify the absence of fecal material in the intestines, a PCR assay for cucumber green mottle mosaic virus (CGMMV) was performed on cDNA from paired fecal and intestinal samples from Ro-SaV2-infected animals. CGMMV was present in 10/11 Ro-SaV2-positive fecal samples and likely originated from the cucumber provided as a water source in the traps. However, all eight intestinal samples that were positive for Ro-SaV2 were negative for CGMVV, suggesting true intestinal infection by Ro-SaV2. UHTS. Serum samples and fecal pellets or rectal swab samples were also extracted, using a viral particle purification procedure, for UHTS. Briefly, each sample was successively passed through 0.45 µM and 0.22 µM sterile filters (Millipore) to remove bacterial and cellular debris and was treated with nucleases. Samples were lysed in NucliSENS buffer, extracted using the EasyMag platform (bioMérieux), and prepared for sequencing using the Ion Torrent Personal Genome Machine system, following the methods of Kapoor et al. (19). Sequencing was performed on pools of four to six samples, which were combined at the double-stranded DNA stage. Viral sequences were assembled using the Newbler or miraEST assemblers, and both contigs and unassembled reads were identified by similarity searches using BLASTn and BLASTx against the GenBank nonredundant nucleotide sequence database (67, 68). Viruses related to those known to cause disease in humans were selected for further study and verified by PCR on original (unpooled) sample material with primers derived from the UHTS sequence data. Confirmed positive results were followed by testing of the serum (n = 114) or fecal samples (n = 133) from remaining animals, and in some cases, subsequent screening of additional sample types from select positive animals (Table 3). One or more positive samples were chosen for further sequencing of phylogenetically relevant genes by overlapping PCR. The 5′ UTRs of NrPV, MPeV, and RPV were determined by rapid amplification of cDNA ends (RACE) using the SMARTer RACE cDNA amplification kit (Clontech). SEOV Baxter qPCR. Primers were designed to target a 121-nt region of the N gene (Baxter.qF, 5′ CATACCTCAGACGCACAC 3′; Baxter.qR, 5′ GGATCCATGTCATCACCG 3′; and Baxter. Probe, 5′-[6-FAM]CCTGGGGAAAGGAGGCAGTGGAT[TAMRA]-3′ [6-FAM, 6-carboxyfluorescein; TAMRA, 6-carboxytetramethylrhodamine]). For tissue samples, viral RNA copy numbers were normalized to the quantity of the reference gene encoding glyceraldehyde 3-phosphate dehydrogenase (GAPDH), whereas the viral RNA copy numbers in serum, oral, and rectal swab samples were reported per ml of serum or PBS wash, respectively (69). qPCR assays were run in duplicate on each sample, and the results were averaged. Samples with an average of ≤2 normalized copies were considered negative. ssqPCR. For ssqPCR, strand-specific synthetic standards were generated by transcribing positive- and negative-sense RNA in vitro from pCRII-TOPO dual promoter vectors (Life Technologies) containing 310 and 594 nt of the NS3 genes of NrHV-1 and NrHV-2, respectively. Positive- and negative-sense RNAs were synthesized from HindIII- or EcoRV-linearized plasmids by transcription from the T7 or SP6 RNA polymerase promoter. In vitro transcription was carried out for 2 h at 37°C using the RiboMax large-scale RNA production system (Promega) and 500 ng of linearized plasmid. Plasmid DNA was removed from the synthetic RNA transcripts by treatment with DNase I (Promega) for 30 min, followed by purification with the High Pure RNA purification kit (Roche). Purified RNA transcripts were analyzed on the Agilent 2100 Bioanalyzer, and RNA standards were prepared by serial dilution in human total RNA. cDNA from both strands was generated using strand-specific primers containing a tag sequence at the 5′ end (see Table S2 in the supplemental material) (70). The RNA was preheated at 70°C for 5 min with 10 pmol of specific primer and 1× reverse transcriptase buffer, followed by the addition of a preheated reaction mixture containing 1 mM MnCl2, 200 µM each deoxynucleoside triphosphate (dNTP), 40 U RNaseOUT, and 1 U Tth DNA polymerase (Promega). The reaction mixtures were incubated at 62°C for 2 min, followed by 65°C for 30 min. The cDNA was incubated with preheated 1× chelate buffer at 98°C for 30 min to inactivate the Tth reverse transcriptase before exonuclease I treatment to remove unincorporated RT primers (New England Biolabs). Reaction mixtures lacking RT primer were included to control for self-priming, the strand specificity of each primer was assessed by performing the RT step in the presence of the uncomplementary strand, and reaction mixtures lacking Tth DNA polymerase were included to control for plasmid DNA detection. ssqPCRs were performed using TaqMan universal master mix II with primers, probe, and 2 µl of cDNA under the following conditions: 50°C for 2 min, 95°C for 10 min, and then 40 cycles of 95°C for 15 s, 50°C for 20 s, and 72°C for 30 s (see Table S2 in the supplemental material). The specificity of the reaction was monitored by RT and amplification of serial dilutions of the uncomplementary strand. The sensitivities of the ssqPCR assays ranged from 0.35 × 103 to 3.5 × 103 RNA copies/reaction mixture volume, and nonstrand-specific amplification was not detected until 3.5 × 107 viral RNA copies of the uncomplementary strand per reaction mixture volume were present (Table S3). Phylogenetic and sequence analyses. Nucleotide or predicted amino acid sequences were aligned with representative members of the relevant family or genus using MUSCLE in Geneious version 7 (Biomatters Ltd.) and manually adjusted. Maximum-likelihood (ML) and Bayesian Markov chain Monte Carlo (MCMC) phylogenetic trees were constructed for each alignment using RAxML version 8.0 and MrBayes version 3.2, respectively (71, 72). ML trees were inferred using the rapid-search algorithm, either the general time-reversible (GTR) plus gamma model of nucleotide substitution or the Whelan and Goldman (WAG) plus gamma model of amino acid substitution, and 500 bootstrap replicates. MCMC trees were inferred using the substitution models described above, a minimum of 10 million generations with sampling every 10,000 generations and terminated when the standard deviation of split frequencies reached <0.01. Phylogenetic analysis of Bartonella was performed by trimming the gltA gene sequences to a 327-nt region (nt positions 801 to 1127) commonly used for taxonomic classification and constructing a neighbor-joining tree using the Hasegawa, Kishino, and Yano (HKY) plus gamma model of nucleotide substitution (13, 73). Phylogenetic analysis of the flaviviruses was performed by first constructing a tree that included representative viruses across the family using a highly conserved region of the NS5B protein (aa 462 to 802 of tick-borne encephalitis virus; GenBank accession number NP_775511.1), followed by complete NS3 and NS5B amino acid phylogenies constructed separately for the Pestivirus and Hepacivirus/Pegivirus genera. These were rooted based on the relative positions of each genus in the family-level tree. To estimate the temporal and geographic origin of SEOV Baxter in NYC, phylogeographic analysis of the N gene was performed using the MCMC method available in the BEAST package (version 1.8.0) (74). All available full-length or nearly full-length SEOV N gene sequences with published sampling times were downloaded from GenBank, aligned as described above, and randomly subsampled five times to include a maximum of 10 sequences per country per year. The codon-structured SDR06 model of nucleotide substitution was used along with a relaxed, uncorrelated lognormal molecular clock and a constant population size coalescent prior (best-fit model, data not shown). Two independent MCMC chains were run for 100 million generations each, and convergence of all relevant parameters was assessed using Tracer version 1.5. The runs were combined after removing a 10% burn-in, and the maximum-clade-credibility (MCC) tree, including ancestral location-state reconstructions, was summarized. Transmembrane domain prediction was performed using TMHHM 2.0, putative N-glycosylation sites were predicted with NetNGlyc 1.0, and the presence of N-terminal signal peptides was predicted using SignalP 4.1, all of which were accessed through the ExPASy web portal (http://www.expasy.org). RNA secondary structures were predicted by MFOLD and through homology searching and structural alignment with bases conserved in other pestiviruses for NrPV and with parechoviruses, hunniviruses, and rosavirus for RPV (75). RNA structures were initially drawn using PseudoViewer, followed by manual editing (76). Statistical analyses. Potential associations between the presence of a microbial agent and the age and sex of the rat were explored using chi-square tests performed with SPSS version 21 (IBM, Armonk, NY), with associations considered significant at a level of α = 0.05. Tests for the overall effect of the age category or sex on the number of viruses carried by an individual were conducted using the Kruskal-Wallis test for age and the Wilcoxon test for sex. Patterns of pathogen cooccurrence within a single host were explored using the Fortran software program PAIRS version 1.1, which utilizes a Bayesian approach to detect nonrandom associations between pairs of taxa (77). The C score statistic was employed as a measure of pathogen cooccurrence (78). Nucleotide sequence accession numbers. The GenBank accession numbers for the agents sequenced in this study are KJ950830 to KJ951004. SUPPLEMENTAL MATERIAL Figure S1 Neighbor-joining tree of a 327-nt region of the gltA gene of Bartonella. The sequences derived from this study are indicated by a circle, and those recovered from Rattus norvegicus rats in Los Angeles, China, and Southeast Asia are given by squares and triangles, respectively. Download Figure S1, PDF file, 0.3 MB Figure S2 Predicted RNA secondary structure of the 5′ UTR of NrPV. Completely conserved nucleotides across all pestiviruses are colored in pink, and the conserved structural domains are indicated. Numbers indicate nucleotide positions from the 5′ end. Download Figure S2, PDF file, 0.5 MB Figure S3 (A) Predicted RNA secondary structure of the 5′ UTR of RPV. Completely conserved nucleotides across all parechovirus and hunnivirus 5′ UTR sequences are colored in pink, and the conserved structural domains are indicated. Numbers indicate nucleotide positions from the 5′ end. (B) ML tree of the complete VP1 gene of RPV and feline, bat, and canine picornaviruses. Selected representatives of the Enterovirus and Sapelovirus genera are also shown. When the BSP and BPP values are both ≥70%, the nodal support value is given beneath associated nodes in the format BSP/BPP. Download Figure S3, PDF file, 0.8 MB Figure S4 Unrooted ML tree of the complete VP1 gene of Manhattan parechovirus (MPeV) (indicated by a red branch) and representative members of the Parechovirus genus. When the BSP and BPP values are both ≥70%, the nodal support is shown beneath the associated node in the format BSP/BPP. HPeV, human parechovirus. Download Figure S4, PDF file, 0.3 MB Table S1 Primers used for targeted molecular analysis in this study. Table S1, DOCX file, 0.1 MB. Table S2 Primer and probe sequences used in the strand-specific reverse transcription and quantitative PCR assays of NrHV-1 and NrHV-2. Table S2, DOCX file, 0.1 MB. Table S3 Sensitivities and specificities of the strand-specific qPCR assays for NrHV-1 and NrHV-2. Table S3, DOCX file, 0.04 MB.
            Bookmark
            • Record: found
            • Abstract: found
            • Article: not found

            Virulence of the zoonotic agent of leptospirosis: still terra incognita?

            Pathogenic leptospires are the bacterial agents of leptospirosis, which is an emerging zoonotic disease that affects both animals and humans worldwide. In this Review, the recent advances in our understanding of the epidemiology, taxonomy, genomics and the molecular basis of virulence in leptospires, and of how these properties contribute to the pathogenesis of leptospirosis, are discussed.
              Bookmark
              • Record: found
              • Abstract: found
              • Article: not found

              A survey of zoonotic pathogens carried by Norway rats in Baltimore, Maryland, USA.

              Norway rats (Rattus norvegicus) carry several zoonotic pathogens and because rats and humans live in close proximity in urban environments, there exists potential for transmission. To identify zoonotic agents carried by rats in Baltimore, Maryland, USA, we live-trapped 201 rats during 2005-2006 and screened them for a panel of viruses, bacteria, and parasites. Antibodies against Seoul virus (57.7%), hepatitis E virus (HEV, 73.5%), Leptospira interrogans (65.3%), Bartonella elizabethae (34.1%), and Rickettsia typhi (7.0%) were detected in Norway rats. Endoparasites, including Calodium hepatica (87.9%) and Hymenolepis sp. (34.4%), and ectoparasites (13.9%, primarily Laelaps echidninus) also were present. The risk of human exposure to these pathogens is a significant public health concern. Because these pathogens cause non-specific and often self-limiting symptoms in humans, infection in human populations is probably underdiagnosed.
                Bookmark

                Author and article information

                Contributors
                Role: Editor
                Journal
                PLoS Negl Trop Dis
                PLoS Negl Trop Dis
                plos
                plosntds
                PLoS Neglected Tropical Diseases
                Public Library of Science (San Francisco, CA USA )
                1935-2727
                1935-2735
                9 August 2019
                August 2019
                : 13
                : 8
                : e0007499
                Affiliations
                [001]Ross University School of Veterinary Medicine, Basseterre, St. Kitts, West Indies
                Mahidol University, THAILAND
                Author notes

                The authors have declared that no competing interests exist.

                [¤]

                Current address: College of Veterinary Medicine, University of Florida, Gainesville, Florida, United States of America

                Author information
                http://orcid.org/0000-0002-6329-2333
                http://orcid.org/0000-0001-6600-9102
                Article
                PNTD-D-19-00277
                10.1371/journal.pntd.0007499
                6688788
                31398190
                b111d731-12ba-4a42-b8c8-4a71049e3571
                © 2019 Boey et al

                This is an open access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

                History
                Page count
                Figures: 1, Tables: 11, Pages: 24
                Funding
                The author(s) received no specific funding for this work.
                Categories
                Review
                Biology and Life Sciences
                Organisms
                Bacteria
                Leptospira
                Biology and Life Sciences
                Microbiology
                Medical Microbiology
                Microbial Pathogens
                Bacterial Pathogens
                Leptospira
                Medicine and Health Sciences
                Pathology and Laboratory Medicine
                Pathogens
                Microbial Pathogens
                Bacterial Pathogens
                Leptospira
                Research and Analysis Methods
                Animal Studies
                Experimental Organism Systems
                Model Organisms
                Rats
                Research and Analysis Methods
                Model Organisms
                Rats
                Research and Analysis Methods
                Animal Studies
                Experimental Organism Systems
                Animal Models
                Rats
                Biology and Life Sciences
                Organisms
                Eukaryota
                Animals
                Vertebrates
                Amniotes
                Mammals
                Rodents
                Rats
                Biology and Life Sciences
                Molecular Biology
                Molecular Biology Techniques
                Artificial Gene Amplification and Extension
                Polymerase Chain Reaction
                Research and Analysis Methods
                Molecular Biology Techniques
                Artificial Gene Amplification and Extension
                Polymerase Chain Reaction
                Medicine and Health Sciences
                Infectious Diseases
                Bacterial Diseases
                Leptospirosis
                Medicine and Health Sciences
                Tropical Diseases
                Neglected Tropical Diseases
                Leptospirosis
                Medicine and Health Sciences
                Infectious Diseases
                Zoonoses
                Leptospirosis
                Biology and Life Sciences
                Anatomy
                Renal System
                Kidneys
                Medicine and Health Sciences
                Anatomy
                Renal System
                Kidneys
                Biology and Life Sciences
                Organisms
                Eukaryota
                Animals
                Vertebrates
                Amniotes
                Mammals
                Rodents
                Ecology and Environmental Sciences
                Terrestrial Environments
                Urban Environments
                Biology and Life Sciences
                Population Biology
                Population Dynamics
                Geographic Distribution

                Infectious disease & Microbiology
                Infectious disease & Microbiology

                Comments

                Comment on this article