13.01.2014 Views

University of Veterinary Medicine Hannover Institute of Parasitology ...

University of Veterinary Medicine Hannover Institute of Parasitology ...

University of Veterinary Medicine Hannover Institute of Parasitology ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong><br />

<strong>Institute</strong> <strong>of</strong> <strong>Parasitology</strong><br />

Fish Disease Research Unit<br />

Epidemiological Study on Yersinia ruckeri Isolates from Rainbow Trout<br />

(Oncorhynchus mykiss, Walbaum) in North West Germany<br />

THESIS<br />

Submitted in partial fulfillment <strong>of</strong> the requirements for the degree<br />

DOCTOR OF PHILOSOPHY (PhD)<br />

Awarded by the <strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong><br />

by<br />

Yidan Huang<br />

Chengdu, P.R. China<br />

<strong>Hannover</strong>, Germany [2013]


Supervisor:<br />

Pr<strong>of</strong>. Dr. Dieter Steinhagen<br />

Fish Disease Research Unit<br />

<strong>Institute</strong> <strong>of</strong> <strong>Parasitology</strong><br />

<strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong>, Foundation, Germany<br />

Supervision Group:<br />

Pr<strong>of</strong>. Dr. Lothar Kreienbrock<br />

<strong>Institute</strong> for Biometry, Epidemiology and Information Processing<br />

<strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong>, Foundation, Germany<br />

Pr<strong>of</strong>. Dr. Martin Runge<br />

Lower Saxony State Office for Consumer Protection and Food safety<br />

(LAVES)<br />

Food and <strong>Veterinary</strong> <strong>Institute</strong> Braunschweig/<strong>Hannover</strong>, Germany<br />

External Referee<br />

Pr<strong>of</strong>. Dr. Claus-Peter Czerny<br />

Department <strong>of</strong> Animal Sciences, Microbiology and Animal Hygiene<br />

Georg-August-<strong>University</strong> <strong>of</strong> Göttingen, Germany<br />

Date <strong>of</strong> final exam: 31.10.2013<br />

Sponsorship: Yidan Huang held the scholarship <strong>of</strong>fered by China Scholarship Council.


To my Family


Publications during the PhD program:<br />

Yidan Huang, Martin Ryll, Charles Walker, Arne Jung, Martin Runge, Dieter<br />

Steinhagen.<br />

Fatty acid composition <strong>of</strong> Yersinia ruckeri isolates from aquaculture ponds in North West<br />

Germany (Berliner Münchener Tierärztliche Wochenschrift 127, Heft 1/2(2014), 10-15).<br />

Yidan Huang, Martin Runge, Geovana Brenner Michael, Stefan Schwarz, Arne Jung,<br />

Dieter Steinhagen.<br />

Biochemical and Molecular Heterogeneity among Isolates <strong>of</strong> Yersinia ruckeri from Rainbow<br />

Trout (Oncorhynchus mykiss, Walbaum) in North West Germany (BMC <strong>Veterinary</strong> Research,<br />

2013, 9:215).<br />

Yidan Huang, Geovana Brenner Michael, Roswitha Becker, Heike Kaspar, Joachim<br />

Mankertz, Stefan Schwarz, Martin Runge, Dieter Steinhagen<br />

Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Yersinia ruckeri from<br />

fish (accepted by <strong>Veterinary</strong> Microbiology)<br />

Posters and oral presentation in national or international conferences:<br />

Yidan Huang , Stefan Schwarz , Silke Braune , Martin Runge, Dieter Steinhagen<br />

Poster: Biochemical and Molecular Differences <strong>of</strong> Yersinia ruckeri isolates from Rainbow<br />

trout (Oncorhynchus mykiss, Walbaum) in Northwestern Germany (The 3 rd<br />

poster<br />

prize).Tagung der DVG-Fachgruppe "Bakteriologie und Mykologie", Leipzig, Germany,<br />

2012<br />

Yidan Huang, Martin Runge, Stefan Schwarz, Silke Braune, Dieter Steinhagen<br />

Poster: Epidemiological Study on Yerinia ruckeri Isolates from Trout Hatcheries in<br />

Northwestern Germany. XIV. Gemeinschaftstagung der Deutschen, Österreichischen und<br />

Schweizer Sektionen der EAFP, Bautzen, Germany, 2012.<br />

Yidan Huang, Geovana Brenner Michael, Stefan Schwarz, Silke Braune, Martin Ryll,<br />

Martin Runge, Dieter Steinhagen


Oral presentation: Investigation on Genetic Diversity <strong>of</strong> Yersinia ruckeri Isolates from Trout<br />

Hatcheries in Northwestern Germany. 16 th Intern. Symposium <strong>of</strong> the World Association <strong>of</strong><br />

<strong>Veterinary</strong> Laboratory Diagnosticians, Berlin, Germany, 2013<br />

Yidan Huang, Geovana Brenner Michael, Heike Kaspar, Stefan Schwarz, Martin<br />

Runge, Dieter Steinhagen.<br />

Poster: Analysis <strong>of</strong> Resistance Pheno- and Genotypes <strong>of</strong> Yersinia ruckeri Isolated from<br />

Diseased Fish. 5 th<br />

Symposium on Antimicrobial Resistance in Animals and the<br />

Environment ARAE, Gent, Belgium, 2013.


Tables <strong>of</strong> Content<br />

CHAPTER 1. INTRODUCTION............................................................................................................. 1<br />

CHAPTER 2.<br />

BIOCHEMICAL AND MOLECULAR HETEROGENEITY AMONG ISOLATES OF<br />

YERSINIA RUCKERI FROM RAINBOW TROUT (ONCORHYNCHUS MYKISS, WALBAUM) IN<br />

NORTH WEST GERMANY ............................................................................................................................ 35<br />

CHAPTER 3.<br />

FIELD STUDY ON YERSINIA RUCKERI ISOLATES FROM TROUT<br />

HATCHERIES IN NORTH WEST GERMANY .............................................................................................. 47<br />

CHAPTER 4.<br />

FATTY ACID COMPOSITION OF YERSINIA RUCKERI ISOLATES FROM<br />

AQUACULTURE PONDS IN NORTH WEST GERMANY ............................................................................ 65<br />

CHAPTER 5.<br />

IN VITRO CYTOTOXICITY AND MULTIPLEX PCR DETECTION OF<br />

VIRULENCE FACTORS OF YERSINIA RUCKERI ISOLATED FROM RAINBOW TROUT IN NORTH<br />

WEST GERMANY………………………………………… ........ ………………………………………………81<br />

CHAPTER 6.<br />

PHENO- AND GENOTYPIC ANALYSIS OF ANTIMICROBIAL RESISTANCE<br />

PROPERTIES OF YERSINIA RUCKERI FROM FISH ................................................................................... 109<br />

CHAPTER 7. GENERAL DISCUSSION ............................................................................................ 135<br />

CHAPTER 8. SUMMARY .................................................................................................................. 149<br />

CHAPTER 9. ZUSAMMENFASSUNG .............................................................................................. 153<br />

ACKNOWLEDGEMENT............................................................................................................................... 157


Chapter 1. Introduction<br />

1


Chapter 1 Introduction<br />

1 Fisheries and Aquaculture<br />

Fisheries and aquaculture contribute greatly to the world‟s food supply, with about 148<br />

million tonnes <strong>of</strong> fish in 2010 and 154 million in 2011 (FAO, 2012). Fishery products are<br />

among the most traded food commodities worldwide, and the trading <strong>of</strong> those products<br />

reached a new peak in 2011. Different aquatic species are farmed worldwide, using freshwater,<br />

brackish water and marine water. Some species are successful internationally introduced,<br />

including tilapias, Chinese carps, Atlantic salmon, Pangasius catfishes, largemouth black bass,<br />

turbot, piarapatinga, pacu, and rainbow trout.<br />

1.1 Rainbow trout aquaculture<br />

Rainbow trout (Oncorhynchus mykiss, Walbaum) belongs to salmonid species and has been<br />

introduced for food and sport to different countries. The first rainbow trout hatchery was<br />

established in the US in the 1870s. Since the 1950s, as the requirement grew rapidly, rainbow<br />

trout production is on the third place for the whole diadromous fish production (Fig.1-1 ).<br />

Particularly in Europe, the production <strong>of</strong> rainbow trout has increased more and more to supply<br />

domestic markets in the countries such as Italy, France, Germany, Denmark, Spain and so on<br />

(FAO, 2009).<br />

Fig.1-1 Production <strong>of</strong> major diadromous fish species in 2010 (taken from FAO, 2012)<br />

Rainbow trout is one <strong>of</strong> the traditional aquaculture species in Germany. They are farmed in<br />

earthen ponds, raceways and other modern indoor and outdoor facilities(ROSENTHAL and<br />

HILGE 2000). Since 1880, when rainbow trout was first introduced to Germany from North<br />

2


Chapter 1 Introduction<br />

America, it has become the most important farmed species in Germany. The development <strong>of</strong><br />

artificial food, flow-through-systems, artificial oxygen enrichment <strong>of</strong> production water and<br />

effective disease control have boosted the production. Therefore, the production figures for<br />

this species have increased annually. At present, some small-scale producers still operate<br />

earthen ponds, but the most <strong>of</strong> trout are reared in flow through units made <strong>of</strong> concrete or<br />

plastic at a higher density level.<br />

3


Chapter 1 Introduction<br />

2. Common Salmonid bacterial diseases<br />

No matter among the cultured or wild population, salmonids can be infected by different<br />

bacteria, including enteric redmouth (Yersinia ruckeri), bacteria gill disease (Flavobacterium<br />

branchiophilum), furunculosis (Aeromonas salmonicida), vibriosis (Vibrio anguillarum and<br />

Vibrio salmonicida), bacterial kidney disease (Renibacterium salmoninarum), bacterial<br />

cold-water disease (F. psychrophilum), etc.(FARALDO-GOMEZ and SANSON 2003). In the<br />

following chapter, enteric redmouth disease will be discussed mainly.<br />

2.1 Enteric Redmouth Disease (ERM)<br />

2.1.1 Epidemiological facts<br />

Enteric redmouth disease (ERM), reported in Idaho, USA, as early as 1950s(RUCKER 1966),<br />

is one <strong>of</strong> the important diseases <strong>of</strong> aquatic animals. Subsequently, in the late 1970s to the<br />

early 1980s, it was first introduced to Europe from the USA (HORNE and BARNES 1999).<br />

Recently, it is proposed that Yersinia ruckeri, the aetiological agent <strong>of</strong> ERM, was introduced<br />

separately into the UK and mainland Europe and subsequent cross-border transfer <strong>of</strong> UK and<br />

European isolates has been limited(WHEELER et al. 2009).<br />

Since ERM was first reported, knowledge <strong>of</strong> the host and geographic ranges has increased. So<br />

far, there are many reports about ERM, especially in European countries, ranging from<br />

different hosts (Table1-1). And now, it is one <strong>of</strong> the most important infectious diseases in<br />

rainbow trout (Oncorhynchus mykiss) aquaculture in Europe.<br />

Table1-1 Worldwide reports <strong>of</strong> Yersinia ruckeri outbreaks<br />

Country Species Years <strong>of</strong> Publication References<br />

Chile Carp 1987 (ENRIQUEZ and<br />

ZAMORA 1987)<br />

Canada Burbot 1987 (DWILOW et al.<br />

1987)<br />

Germany Rainbow trout 1994 (FUHRMANN and<br />

BOEHM 1983;<br />

KLEIN et al. 1994)<br />

U.K. Otter, Arctic char 1996 (CLLINS et al.<br />

4


Chapter 1 Introduction<br />

1996)<br />

Rainbow trout 2003 (FRERICHS et al.<br />

1985; D A AUSTIN<br />

et al. 2003)<br />

Peru Rainbow trout 2004,2011 (BRAVO and<br />

KOJAGURA 2004;<br />

BASTARDO et al.<br />

2011b)<br />

Spain Rainbow trout 2006,1986 (CRUZ et al. 1986;<br />

FOUZ et al. 2006)<br />

Turkey Black sea salmon 2006 (SAVAS et al.<br />

2006)<br />

Rainbow trout 2004 (KARATAS et al.<br />

2004)<br />

Czechoslovakia Rainbow trout 1990 (VLADIK and<br />

PROUZA 1990)<br />

France Sturgeon 1987 (VUILLAUME et<br />

al. 1987)<br />

Italy Rainbow trout 1985 (GIORGETTI et al.<br />

1985)<br />

Finland<br />

Perch, brown trout<br />

rainbow trout and<br />

white fish<br />

1992 (VALTONEN et al.<br />

1992)<br />

USA Channel catfish 1999 (DANLEY et al.<br />

1999)<br />

P.R. China<br />

Channel catfish<br />

Amur Sturgeon<br />

2009,2013 (K Y WANG et al.<br />

2009; LI et al. 2013)<br />

Switzerland Rainbow trout 1986 (MEIER 1986)<br />

Denmark Norwegian salmon 1986 (SPARBOE et al.<br />

1986)<br />

Turkey Rainbow trout 2011 (ONUK et al. 2011)<br />

5


Chapter 1 Introduction<br />

2.1.2 Clinical and Pathological facts<br />

Clinical signs <strong>of</strong> infected rainbow trout may include dark coloration, hemorrhages in the<br />

mouth, the presence <strong>of</strong> pale s<strong>of</strong>t liver and kidney, muscle degradation, swollen abdomen and<br />

gastro-enteritis. Clinical infection can also be characterized by a yellow discharge from the<br />

vent. Y. ruckeri has also been isolated from freshwater invertebrates (DULIN et al. 1976), and<br />

therefore, it probably persists in the environment <strong>of</strong> ponds which were previously stocked<br />

with infected fish populations(BERGH 2008). On the histopathological level, several changes<br />

were observed in infected fish, including bacteremia with inflammation, glomerulonephritis<br />

and necrotic foci in kidney, necrosis in liver, telangiectasis in spleen, hemorrhages and<br />

hyperemia in the intestinal mucosa, myocardial degeneration, atrophy and edema in the heart,<br />

atrophy <strong>of</strong> pancreatic tissues and melanophore hyperplasia in the skin (MAHJOOR and<br />

AKHLAGHI 2012).<br />

In Europe, the bacterium is endemic in many trout farms and can cause severe losses. The<br />

disease most commonly affects younger rainbow trout at temperatures above 10 °C.<br />

Outbreaks are <strong>of</strong>ten related to adverse situations or stressed carrier fish which initiate the<br />

infection. Protection is provided by commercially available vaccines, which reduce losses to a<br />

large extent(ROBERTS 2001; BERGH 2008).<br />

6


Chapter 1 Introduction<br />

3 General background <strong>of</strong> Yersinia ruckeri<br />

3.1 Taxonomic position<br />

In Bergey‟s Manual <strong>of</strong> Determinative Bacteriology, (8th ed) (BUCHANAN and GIBBONS<br />

1974), the genus Yersinia is placed in the family Enterobacteriaceae, in which Y. ruckeri has<br />

been proposed, but this has not been accepted generally. Ewing et al (EWING et al. 1978)<br />

found that Y. ruckeri strains were about 30% related to species <strong>of</strong> both Serratia and Yersinia,<br />

and the G+C contents <strong>of</strong> this bacteria were approximately 47.5 to 48.5%, which were similar<br />

to those <strong>of</strong> Yersinia but markedly different from those <strong>of</strong> Serratia. Austin et al (B AUSTIN et<br />

al. 1982) discovered that a number <strong>of</strong> phenotypic traits <strong>of</strong> Y. ruckeri are not consistent with<br />

the inclusion <strong>of</strong> this organism in the genus Yersinia. A similar result could be seen in the<br />

research <strong>of</strong> Green and Austin (GREEN and AUSTIN 1983), emphasizing that comparatively<br />

few phenotypic traits distinguished Y. ruckeri from Salmonella arizonae. According to the<br />

results <strong>of</strong> multilocus sequence typing (MLST) and 16S rRNA analysis, Y. ruckeri was the<br />

most distant species within the genus Yersinia (KOTETISHVILI et al. 2005).<br />

Basing on the genomic characterization, Y. ruckeri has the smallest genome (3.7 Mb),<br />

although it shares the same core set <strong>of</strong> approximately 2,500 genes with the other members <strong>of</strong><br />

the genus, whose genomes range in size from 4.3 to 4.8 Mb (CHEN et al. 2010). Romalde et<br />

al(ROMALDE et al. 1991b) mentioned that the chromosome size <strong>of</strong> Y. ruckeri was about<br />

4460 to 4770 kbp, so that the different chromosome size with the other Yersinia species<br />

indicated again that the controversial taxonomic position <strong>of</strong> Y. ruckeri should be reconsidered.<br />

And unlike most other Yersinia species, Y. ruckeri was found missing the mtnKADCBEU<br />

gene cluster (CHEN et al. 2010). Méndez et al (M NDEZ et al. 2009) found a component <strong>of</strong><br />

type IV secretion systems (T4SS), which is absent from human pathogenic Yersinia, is<br />

associated with the virulence <strong>of</strong> Y. ruckeri, which proved again, that Y. ruckeri is a very<br />

homogeneous species that is quite different from the other members <strong>of</strong> the genus Yersinia.<br />

Nevertheless, in Bergey‟s Manual <strong>of</strong> Systematic Bacteriology in 2005 (BOTTONE et al.<br />

2005), Y. ruckeri is still included in the family Enterobacteriaceae, the genus Yersinia.<br />

Therefore, the taxonomic standing <strong>of</strong> Y. ruckeri may need to be reevaluated, perhaps as a new<br />

genus within the Enterobacteriaceae.<br />

7


Chapter 1 Introduction<br />

3.2 Isolation, Identification and Biochemical Properties<br />

3.2.1 Morphology and Biochemical Properties<br />

As the etiological agent <strong>of</strong> ERM, Y. ruckeri is a gram-negative, rod-shaped, 1.0×2.0-3.0 μm in<br />

size, oxidase-negative, peritrichous, fermentative bacterium with facultative-aerobic<br />

metabolism (ROSS et al. 1966; SCHÄPERCLAUS 1991; STRÖM-BESTOR et al. 2010). A<br />

positive result can be observed in the methyl red test and voges-proskauer reaction; gelatin is<br />

degraded, but not aesculin, chitin, DNA, elastin, pectin, tributyrin or urea; it can grow in 0-3%<br />

(w/v) sodium chloride, and utilize sodium citrate; Y. ruckeri can also use maltose, fructose,<br />

trehalose, mannitol and glucose to produce acid, but cannot use inositol, lactose, raffinose,<br />

salicin or sucrose; it could produce catalase, β-galactosidase, but not H 2 S, indole, oxidase<br />

phenylalanine deaminase or phosphatase (B AUSTIN and AUSTIN 1999). The motility <strong>of</strong> Y.<br />

ruckeri depends upon the temperature <strong>of</strong> incubation and the biotypes. Cultures in the log<br />

phase <strong>of</strong> growth are motile between 18-27 °C, although the optimum temperature for the<br />

growth <strong>of</strong> Y. ruckeri is 22 - 25 °C (EWING et al. 1978). Biotype 2 strains are non-motile<br />

because <strong>of</strong> the lacking <strong>of</strong> flagella and were isolated first by Davis and Frerichs (DAVIES and<br />

FRERICHS 1989) in the 1980s. More and more attention was paid to this kind <strong>of</strong> strains since<br />

it can cause disease to vaccinated fish and caused outbreaks in 2003 (D A AUSTIN et al.<br />

2003). It was later proved that biotype 2 strains had come out, because the population<br />

structure in the natural environment has changed, instead <strong>of</strong> a direct vaccination-induced<br />

mutation (JOHN TINSLEY 2010).<br />

3.2.2 Isolation and Identification<br />

Y.ruckeri grows well on trypticase soy agar (TSA) and brain heart infusion agar (BHI agar),<br />

forming colonies <strong>of</strong> 1-2 mm in diameter, smooth, round, raised, white cream in color,<br />

translucent and which exhibit a butyrous type <strong>of</strong> growth with entire edges(FURONES et al.<br />

1993b). On deoxycholate-citrate-mannitol agar, Y. ruckeri forms typical red to magenta<br />

colonies (HUNTER et al. 1980). A selective medium developed by Waltman and Shotts<br />

(WALTMAN and SHOTTS 1984) containing Tween 80, sucrose, and bromothymol blue was<br />

used for the isolation <strong>of</strong> Y. ruckeri colonies, which are green with a zone <strong>of</strong> Tween 80<br />

hydrolysis. However, one problem <strong>of</strong> this selective medium is that it is limited to isolate<br />

biotype 2 (BT2) strains, which are Tween80-negative and non-motile(DAVIES and<br />

FRERICHS 1989). Rodger (B AUSTIN et al. 1982) described a ribose ornithine deoxycholate<br />

8


Chapter 1 Introduction<br />

citrate (ROD) medium, which is both selective and differential for Y. ruckeri according to<br />

yellow deposits surrounding the colony. However, Furones et al (FURONES 1990;<br />

FURONES et al. 1993a) found that not all Y. ruckeri strains, especially HSF- strains, produce<br />

the characteristic yellow deposit on ROD medium. Basing on the ROD medium, Furones et al<br />

(FURONES et al. 1993a) simplified it by supplying TSA with 1% (w/v) SDS, to differentiate<br />

isolates <strong>of</strong> Y. ruckeri, and discovered that Y. ruckeri produces a creamy deposit around the<br />

colonies on this medium. Furthermore, they added 100μg/ml <strong>of</strong> coomassie brilliant blue into<br />

TSA-SDS medium and made it simpler and more economic to screen for virulent serotype I<br />

(HSF+) strains <strong>of</strong> Y. ruckeri (FURONES et al. 1993a).<br />

Bacterial sensitivity to bacteriophages provides the possibility <strong>of</strong> using phage typing as an<br />

epidemiological tool and as a tool for rapid diagnosis. Therefore, Stevenson and Airdrie<br />

(STEVENSON and AIRDRIE 1984b) screened eight bacteriophages against Y. ruckeri, which<br />

have a potential value for fish health-monitoring programs. Noga et al (E J NOGA et al. 1988)<br />

found that the sensitivity and specificity <strong>of</strong> kidney biopsy were high and that there were no<br />

significant differences between results from both kidney biopsy and necropsy, suggesting a<br />

non-lethal way to diagnose Y. ruckeri infection.<br />

Due to its relative cheap price and simplicity for operation, the API-20E system is used in<br />

aquaculture for rapid diagnosis <strong>of</strong> fish diseases. This test was initially designed for the<br />

identification <strong>of</strong> members <strong>of</strong> the family Enterobacteriaceae. Y. ruckeri can be correctly<br />

identified by pr<strong>of</strong>iles 1104100 and 5104100 (SANTOS et al. 1993). Isolates in Chile shared<br />

the pr<strong>of</strong>ile numbers 5107100, 5307100, 5107500 and 5105500(BASTARDO et al. 2011a).<br />

The pr<strong>of</strong>ile number 51051010 cannot discriminate between Y. ruckeri and Hafnia alvei, and<br />

can misidentify isolates <strong>of</strong> Y. ruckeri as H. alvei (SANTOS et al. 1993; DANLEY et al. 1999).<br />

However, this problem can be solved by performing supplementary test to determine<br />

xylose-fermenting ability: Y. ruckeri is negative, whereas H. alvei is positive for xylose<br />

fermentation (BULLER 2004).<br />

Based on 16S ribosomal DNA (rDNA) polymorphism, Waren et al (WARSEN et al. 2004)<br />

developed a DNA microarray suitable for simultaneous detection and discrimination between<br />

multiple bacterial species, including Y. ruckeri, and this methodology permitted 100%<br />

specificity for 15 fish pathogens, which means it is suitable for detection and surveillance for<br />

commercially important fish pathogens. Seker et al (SEKER et al. 2012) established a<br />

9


Chapter 1 Introduction<br />

species-specific PCR for rapid identification <strong>of</strong> Y. ruckeri. Yugueros et al (YUGUEROS et al.<br />

2001) found that only the Y. ruckeri strains presented identical RFLPs (two fragments <strong>of</strong> 900-<br />

and 265-bp), which can be used as a tool <strong>of</strong> identification and differentiation <strong>of</strong> at least three<br />

Yersinia spp., as well as E. coli and S. enteritidis (YUGUEROS et al. 2001). Wortberg et al<br />

characterized Y. ruckeri by Fourier transform infrared spectroscopy (FT-IR) and found,<br />

although it was not successful to identify the motility <strong>of</strong> Y. ruckeri, this method was still a<br />

powerful, economic, fast and reliable tool for the identification <strong>of</strong> Y. ruckeri (WORTBERG et<br />

al. 2012).<br />

Souza et al (SOUZA et al. 2010)compared four molecular typing methodologies,<br />

Enterobacterial Repetitive Intergenic Consensus PCR (ERIC-PCR), Pulsed-Field Gel<br />

Electrophoresis (PFGE), 16S rRNA gene sequencing and Multilocus Sequence Analysis<br />

(MLSA), for identifying species among Yersinia isolates, and found ERIC-PCR, 16S rRNA<br />

gene sequencing and MLSA seem to be valuable techniques for use in taxonomic and<br />

identification studies <strong>of</strong> the genus Yersinia, whereas PFGE does not. Recently, a novel assay<br />

was developed to identify and differentiate Y. ruckeri BT2 strains, basing on the mutant<br />

allele-specific changes in restriction enzyme cleavage sites(WELCH 2011). Bastardo et al.<br />

(BASTARDO et al. 2012)used MLST and identified 30 different STs, separated into two<br />

clonal complexes, indicating the genotypic diversity present in Y. ruckeri. In the same study, it<br />

was suggested that initial steps <strong>of</strong> Y. ruckeri clonal diversification may have occurred by<br />

recombination(BASTARDO et al. 2012).<br />

3.3 Serological Studies and Distribution<br />

Y. ruckeri can be distinguished by serotypes, biotypes and outer membrane protein<br />

(OMP)-types. Different types are associated with different virulence. In the early 1980s, using<br />

the whole-cell antigens conjugated with unabsorbed and cross-absorbed whole-cell antisera,<br />

six serotypes (I-VI) were distinguished by Stevenson and Airdrie (STEVENSON and<br />

AIRDRIE 1984a) and Daly et al (DALY et al. 1986).However, Davies (DAVIES 1990)<br />

suggested a classification, which did not correspond with that suggested by the authors<br />

mentioned above. In Davies‟ classification, Y. ruckeri isolates are divided by heat-stable<br />

O-antigens into five serotypes, designated serotypes O1, O2, O5, O6 and O7, with different<br />

outer-membrane protein types (OMP-types) separately (Table1-2)(DAVIES 1990, 1991a).<br />

From the serotype O1, OMP-type 3 isolates are virulent, so-called „Hagerman‟<br />

10


Chapter 1 Introduction<br />

isolates(DAVIES 1991a). All five O-serotypes were present in north America, whereas only<br />

serotype O1 isolates were identified in Australia and South Africa(DAVIES 1990). In<br />

European countries, serotype O1 isolates accounted for 91% <strong>of</strong> the European isolates; while<br />

serotype O2 isolates were common and widely distributed in France, Norway, Great Britain<br />

and West Germany; Serotype O5 and O7 isolates were obtained from Great Britain, Poland<br />

and Denmark respectively, whereas serotype O6 isolates occurred in Finland and West<br />

Germany(DAVIES 1990; PĘKALA et al. 2010). However, so far, there is no uniform<br />

classification for O-serotypes. One <strong>of</strong> the most accepted to date is proposed by Romalde et<br />

al(ROMALDE et al. 1993), who distinguished four different O-serotypes: Serotype O1 can be<br />

divided into two subgroups O1a (serovar I) and O1b (serovar III) and serotype O2 (serovar II)<br />

into three subgroups O2a, O2b and O2c. The other two are serotype O3 (serovar V) and<br />

serotype O4 (serovar VI) (ROMALDE et al. 1993).<br />

Besides, Y. ruckeri can be defined as two biotypes: biotype 1 is motile and positive for<br />

phospholipase activity and Tween 80 hydrolysis; while biotype 2 is non-motile and negative<br />

for the two tests above(DAVIES and FRERICHS 1989).<br />

Table1-2 Serotypes, OMP-types and Biotypes <strong>of</strong> Yersinia ruckeri according to Davies<br />

(DAVIES 1991a)<br />

Serotypes OMP-types Biotypes<br />

O1 1,2,3,4 1,2<br />

O2 1,2 1,2<br />

O5 1,2 1<br />

O6 1,2 1<br />

O7 1,5 1<br />

3.4 Pathogenesis<br />

Despite the importance and extensive knowledge <strong>of</strong> Yersinia species in pathogenesis <strong>of</strong><br />

mammals, the precise mechanisms <strong>of</strong> virulence are practically unknown. Recent reports about<br />

the pathogenesis <strong>of</strong> Y. ruckeri are mainly concentrating on host factors, mechanisms <strong>of</strong><br />

11


Chapter 1 Introduction<br />

invasion and research on possible bacterial virulent factors.<br />

3.4.1 Host factor<br />

Y. ruckeri can affect salmonids and other fish in both freshwater and seawater, persists in the<br />

fish farm which suffers from an outbreak <strong>of</strong> ERM and has a high ability to adhere on<br />

materials commonly found in fish farm tanks (L COQUET et al. 2002a; L COQUET et al.<br />

2002b). Rainbow trout are especially susceptible, but steelhead, lake, cutthroat, brown and<br />

brook trout, and coho, sockeye, Chinook and Atlantic salmon are also affected(EDWARD J<br />

NOGA 2010). In the epidemiological investigation performed by Good et al (GOOD et al.<br />

2001) from 1981 to 1997, Brook trout (S. fontinalis), Lake trout (S. namaycush), Rainbow<br />

trout (O. mykiss) and Splake (S. fontinalis×S. namatcysh) were found positive for Y. ruckeri.<br />

Although any age salmonid is susceptible, the early-production lots (< 6 months) were<br />

significantly more likely to be positive for Y. ruckeri(GOOD et al. 2001).<br />

3.4.2 Route <strong>of</strong> infection<br />

Y. ruckeri first adheres to gill mucus and thereafter invades the branchial vasculature leading<br />

rapidly to septicemia and colonization <strong>of</strong> the internal organs(TOBBACK et al. 2009).<br />

However, some bacteria were noted in the intestinal crypts, indicating that these bacteria<br />

evaded the first line <strong>of</strong> host defenses in the gut mucosa (TOBBACK et al. 2009), which need<br />

further studies to prove the role <strong>of</strong> skin and gut as portals <strong>of</strong> entry. Only virulent strains are<br />

probably able to defeat the host‟s immune mechanisms(TOBBACK et al. 2009), which<br />

indicate that immune evasion is a major virulence property <strong>of</strong> Y. ruckeri. Survival and growth<br />

in macrophages is the first step for Y. ruckeri to invade the host and trout macrophages<br />

provide a safe niche for Y. ruckeri(RYCKAERT et al. 2010). Y. ruckeri lacks urease,<br />

methionine salvage genes, and B12-related metabolism and there may be an alternative mode<br />

<strong>of</strong> infection for Y. ruckeri(CHEN et al. 2010). Some other findings suggested that serum<br />

resistance plays a role in the pathogenesis <strong>of</strong> Y. ruckeri infections, which is probably<br />

important in the extracellular survival <strong>of</strong> the pathogen in the host (DAVIES 1991b;<br />

TOBBACK et al. 2010)<br />

12


Chapter 1 Introduction<br />

3.4.3 Virulence factors<br />

Previously, genes associated with invasion in two other members <strong>of</strong> Yersinia, inv from Y.<br />

pseudotuberculosis (ISBERG and FALKOW 1985)and ail from Y. enterocolitica(MILLER<br />

and FALKOW 1988), were described. However, there is no evidence for inv or ail<br />

homologues in Y. ruckeri (KAWULA et al. 1996), which proved that Y. ruckeri use a cell<br />

invasion pathway that is distinct from that used by the other pathogenic Yersinia. Due to this<br />

difference, since then, searching for the main virulence factors <strong>of</strong> Y. ruckeri became one<br />

major part <strong>of</strong> pathogenicity researches.<br />

3.4.3.1 Extracellular Products<br />

The extracellular products <strong>of</strong> Y. ruckeri, which are strongly toxic for fish, include lipase,<br />

protease, and cytotoxic and hemolytic activities, and reproduce some characteristic symptoms,<br />

such as hemorrhage in the mouth and intestine(ROMALDE and TORANZO 1993). The<br />

extracellular serralysin metalloprotease Yrp1 was identified in some Y. ruckeri isolates,<br />

leading to the classification <strong>of</strong> two different groups: Azo + (presence <strong>of</strong> the Yrp1 proteolytic<br />

activity) and Azo - ( absence <strong>of</strong> the Yrp1 proteolytic activity)(SECADES and GUIJARRO<br />

1999). Subsequently, it was found that the yrp1 operon in an Azo - strain was blocked at the<br />

transcriptional level rather than its absence in this group(FERNANDEZ et al. 2003). The<br />

Yrp1 protein contains a ZnMc superfamily conserved domain which is associated to<br />

virulence(K WANG et al. 2012). Fernández et al (FERNANDEZ et al. 2002) showed that the<br />

gene encoding Yrp1 is part <strong>of</strong> an operon containing a type I ABC transporter involved in<br />

protein secretion, encoded by three genes (yrpD, yrpE, and yrpF), together with gene inh, that<br />

encodes a protease inhibitor. As Yrp1 protease can hydrolyze laminin, this degradation may<br />

be the cause <strong>of</strong> membrane alterations leading to erosion and pores in capillary vessels, which<br />

results in the hemorrhages in mouth and intestines(FERNANDEZ et al. 2003). A trout model<br />

was used to study the participation <strong>of</strong> the protease in pathogenesis and it was shown that<br />

inactivation <strong>of</strong> either yrp1 or yrpE lead to a significant increase in the 50% lethal<br />

dose(FERNANDEZ et al. 2002), indicating the protease being involved in virulence.<br />

Moreover, Fernandez et al (FERNANDEZ et al. 2007) detected other genes relative to<br />

hemolysin and found that the gene YhlB precedes another ORF (Yhl A) encoding a<br />

Serratia-type hemolysin, and in the LD50 experiment, the value <strong>of</strong> the mutant strains<br />

150RyhlA was approximately 100-fold higher than that <strong>of</strong> the parental strain, which<br />

suggested that Yhl A is another virulence factor <strong>of</strong> Y. ruckeri (FERNANDEZ et al. 2007).<br />

13


Chapter 1 Introduction<br />

3.4.3.2 Flagella and Type three secretion system (T3SS)<br />

Genetic studies on bacterial virulence factors demonstrated that pathogens posses specific<br />

pathogenic genes. These clusters <strong>of</strong> genes can be transferred even horizontally, which leads to<br />

pathogens with distant relationships carrying similar virulence genes. This kind <strong>of</strong> view has<br />

become particularly obvious for a set <strong>of</strong> approximately 20 genes which together encode a<br />

pathogenic mechanism called type III secretion system (T3SS) (HUECK 1998). Type three<br />

secretion systems have been found in various Gram-negative bacteria to transport proteins<br />

from the cytoplasm to the external environment(G P SALMOND and REEVES 1993), and<br />

have been proved to be essential for bacterial pathogenesis(CORNELIS 2002; TOBE et al.<br />

2006; VILCHES et al. 2009). A T3SS is mainly composed <strong>of</strong> four different parts:<br />

transcriptional regulators, chaperones, the components <strong>of</strong> the secretion apparatus, and the<br />

components <strong>of</strong> an extracellular filamentous organelle, ranging from a short needle complex to<br />

flagellar hooks and filaments (HUECK 1998; KNUTTON et al. 1998; FRANCIS et al. 2002;<br />

PAGE and PARSOT 2002). Pallen et al(PALLEN et al. 2005) suggested to divide the T3SS<br />

into two different groups: flagellar T3SS and Non-flagellar T3SS, the former associated with<br />

flagellum biosynthesis and the latter mediate the interactions between bacteria and eukaryotic<br />

cells. As the following Fig.1-2 shows, flagella are connected, or related to the type three<br />

injectisome, which is a complex nanomachine ensuring Gram-negative bacteria to deliver<br />

effector proteins into host cells (CORNELIS 2006). Non-flagellar T3SS (NF-T3SS), also<br />

called as injectisome, attracted attention for the first time, because it plays a important role in<br />

the pathogenesis <strong>of</strong> many diseases(G P SALMOND and REEVES 1993; HUECK 1998).<br />

NF-T3SS is homogenous to the flagellar T3SS, for examples, FliFHIKNPQR and FlhAB<br />

(flagellar secrection apparatus) in flagellar T3SS are homologous to the components in the<br />

non-flagellar T3SS, but more structural data is still needed (DESVAUX et al. 2006;<br />

ERHARDT et al. 2010). YscJLNPRSTUV in the Ysc-Yop system <strong>of</strong> Yersinia species share<br />

homology with flagellar components(DESVAUX et al. 2006). The internal part <strong>of</strong> the Ysc<br />

injectisome, found in Yersinia species contains some proteins which have counterparts in the<br />

basal body <strong>of</strong> the flagellum, suggesting that these two organelles share the same evolutionary<br />

origin(CORNELIS 2002).<br />

14


Chapter 1 Introduction<br />

Fig.1-2 Type III secretion systems. (a) A translocation-associated type III secretion system<br />

(T3aSS) from the Ysc–Yop system found in Yersinia species. (b) The bacterial flagellum<br />

(T3bSS). Export and assembly hook and filament rely on a dedicated flagellar type III<br />

secretion system that is associated with the basal body (DESVAUX et al. 2006).<br />

The tool used by bacteria to move in their environment is the flagellum, which is a rigid,<br />

helical filament (BERG and ANDERSON 1973), composed <strong>of</strong> about 25 different proteins<br />

(CHEVANCE and HUGHES 2008). The ancestor <strong>of</strong> all T3SS was recognized to be the<br />

flagellar T3SS, which is present in both Gram-positive and Gram-negative bacteria, and it<br />

was suggested that the type 3 secretion needed for pathogenesis evolved from<br />

flagellar-specific T3SS(HUECK 1998; MACNAB 2004).<br />

Milton et al (MILTON et al. 1996) found that a flagellin gene flaA is needed for crossing the<br />

fish integument and may play a role in virulence after invasion <strong>of</strong> the host. The aflagellate<br />

mutant <strong>of</strong> Y. enterocolitica decreased its ability to invade cultured epithelial cells and to<br />

colonize the porcine intestinal tissue in vitro, and could survive within cultured human<br />

macrophages over 3 h (MCNALLY et al. 2007), which proved again the importance <strong>of</strong><br />

flagella to motile bacteria. Flagella contribute to the virulence <strong>of</strong> Y. ruckeri, but at best only in<br />

a marginal way and at the early stage when the pathogen in the environment first contacts the<br />

host and begins to invade(KIM 2000). However, the Y. ruckeri strains isolated from outbreaks<br />

in Southern England are belonging to biotype 2 and lacking both flagellar motility and<br />

15


Chapter 1 Introduction<br />

secreted lipase activity(D A AUSTIN et al. 2003). Recently, Evenhuis et al (EVENHUIS et al.<br />

2009) constructed a mutation flhA::Tn5, lacking <strong>of</strong> both motility and secreted lipase activity,<br />

and proved that flagellar secretion is unnecessary for virulence in this organism, including<br />

early steps in the process <strong>of</strong> infection. Moreover, it was found that all serogroups and biotypes<br />

were virulent to rainbow trout, suggesting flagellin may not be required for rainbow trout<br />

proinflammatory innate immune response(JOHN TINSLEY 2010).<br />

Three pathogenic Yersinia spp., Y. pestis, Y. enterocolitica and Y. pseudotuberculosis, exhibit<br />

the ability to resist the host‟s primary immune defense, most notably by inhibiting their own<br />

uptake by pr<strong>of</strong>essional phagocytes(BURROWS and BACON 1956). The antiphagocytic<br />

effect is mediated by the Yersinia type III secretion system and specifically requires a protein<br />

called YopH(ANDERSSON et al. 1996). Mills et al (MILLS et al. 1997) showed that Y.<br />

enterocolitica induces apoptosis in macrophages, which requires type III secretion and<br />

depends on YopP, which appears to be a novel effector. Y. enterocolitica maintains three<br />

different pathways for type three protein secretions, Ysa T3SS, Ysc T3SS, and flagellar<br />

T3SS(M B YOUNG and YOUNG 2002). As the proteins secreted by Ysc T3SS referred to as<br />

Yops (Yersinia outer proteins), those extracellular proteins secreted by Ysa T3SS are called<br />

Ysps (Yersinia secreted proteins). It was found that the flagellar defect didn‟t affect the<br />

production <strong>of</strong> Ysps or Yops(M B YOUNG and YOUNG 2002). Phospholipase YplA, which<br />

has been implicated in Y. enterocolitica virulence, can be a substrate for the Ysc, Ysa, and<br />

flagellar T3SSs(M B YOUNG and YOUNG 2002), indicating that the sharing <strong>of</strong> substrates by<br />

different T3SSs <strong>of</strong> Y. enterocolitica may be important during the course <strong>of</strong> an infection.<br />

Recently, an Ysa (Yersinia secretion apparatus) -like T3SS, different from other human<br />

pathogenic Yersinia species, was found to be present in Y. ruckeri (GUNASENA et al. 2003).<br />

Additionally, more research is needed to reveal the presence and the function <strong>of</strong> the Ysa T3SS<br />

in Y. ruckeri.<br />

3.4.3.3 Ruckerbactin<br />

Iron is necessary for bacterial growth, because it is essential for enzymatic<br />

reactions(CASTIGNETTI and SMARRELLI-JR. 1986). Iron uptake is essential for successful<br />

colonization and invasion by many microbial pathogens and therefore they developed high<br />

affinity iron transport systems to compete for iron with the host (FARALDO-GOMEZ and<br />

SANSON 2003). It is believed that the production <strong>of</strong> hemolysin is helpful to provide iron<br />

16


Chapter 1 Introduction<br />

from red blood cells(MARTINEZ et al. 1990). Counterparts <strong>of</strong> the loci fct, entC, fepDGC,<br />

fepB, entS, and exbB, all involved in iron piracy, were identified in Y. ruckeri, confirming that<br />

Y. ruckeri produces a catechol siderophore (ruckerbactin)(ROMALDE et al. 1991a;<br />

FERNANDEZ et al. 2004). It is proved that this iron uptake system ruckerbactin is involved<br />

in the virulence <strong>of</strong> Y. ruckeri(FERNANDEZ et al. 2004).<br />

3.4.3.4 Type Four Secretion System (T4SS)<br />

Type Four secretion systems, built from core components <strong>of</strong> conjugation machines and<br />

transport proteins or protein-DNA complexes, have been demonstrated to be involved in<br />

virulence in other microorganisms (G P C SALMOND 1994; ZINK et al. 2002). Two loci,<br />

iviXII and iviXIII, were detected; the former shows homology with the lipoproteins TraI and<br />

DotC, involved in secretion systems <strong>of</strong> this type; the latter is associated with the tight<br />

adherence properties (FERNANDEZ et al. 2004). These proteins have been proposed to<br />

compose a new subfamily <strong>of</strong> type IV secretion systems (BHATTACHARJEE et al. 2001).<br />

Recently, the tra operon was initially identified as an ivi gene, and a traI mutant strain, which<br />

showed a significantly lower recovery in vivo competition experiments, was approximately 10<br />

times more attenuated than the parental strain(MENDEZ et al. 2009). It is proved that the tra<br />

operon, as a component <strong>of</strong> a T4SS, contributes to the virulence <strong>of</strong> Y. ruckeri(MENDEZ et al.<br />

2009).<br />

3.4.3.5 ZnuABC Operon and BarA-UvrY two-component system<br />

As a tool for investigating bacterial virulence genes, signature-tagged mutagenesis is used to<br />

indentify mutants that can survive in vitro but not in the host fish. Dahiya and Stevenson<br />

(DAHIYA and STEVENSON 2010b) applied this method on the study <strong>of</strong> virulence factors <strong>of</strong><br />

Y. ruckeri and found different factors including the genes with sequence homologies to genes<br />

for ZnuA, a periplasmic zinc-binding protein <strong>of</strong> the ZnuABC transporter, and the UvrY<br />

response regulator <strong>of</strong> the BarA-UvrY two-component system. Furthermore, they studied the<br />

ZnuABC operon and found the ΔznuABC mutant was unable to compete with the parental<br />

strain and survived poorly in rainbow trout kidney in a competitive challenge by<br />

immersion(DAHIYA and STEVENSON 2010c), indicating that the ZnuABC transporter<br />

plays a role in establishing and maintaining a rainbow trout infection by Y. ruckeri. The<br />

function <strong>of</strong> the BarA-UvrY two-component system was also investigated by the same<br />

authors(DAHIYA and STEVENSON 2010a), which showed that the BarA–UvrY TCS<br />

contributes to the pathogenesis <strong>of</strong> Y. ruckeri in its natural host rainbow trout, possibly by<br />

17


Chapter 1 Introduction<br />

regulating the invasion <strong>of</strong> epithelial cells and the sensitivity to oxidative stress induced by<br />

immune cells.<br />

3.4.3.6 Other factors<br />

Y. ruckeri has a preference for different cell lines. It can adhere to and invade fish derived<br />

cultured cells, such as rainbow trout kidney (RTPK) or rainbow trout gonad (RTG)cell, and<br />

has strongly preference for fathead minnow epithelial (FHM), but not human derived Hep-2<br />

cells(KAWULA et al. 1996). Romalde and Toranzo(ROMALDE and TORANZO 1993)<br />

proved the ability <strong>of</strong> Y. ruckeri to adhere to and effectively invade fish cell lines cultured in<br />

vitro. However, the virulence <strong>of</strong> Y. ruckeri for rainbow trout does not seem to correlate with in<br />

vitro invasiveness <strong>of</strong> cell lines(ROMALDE and TORANZO 1993). It is obvious that<br />

colchicines and cytochalasin-D could be used to reduce the ability <strong>of</strong> Y. ruckeri to invade<br />

cultured cell lines (FERNANDEZ et al. 2002). It is believed that microtubules and<br />

micr<strong>of</strong>ilaments play a role in in vitro invasion, but there was no clear relation with the<br />

virulence found, due to cell type and strain-dependent internalization mechanisms triggered<br />

by Y. ruckeri(TOBBACK et al. 2010).<br />

3.5 Control<br />

As aquaculture is getting increasingly expanding and intensified, infectious diseases are<br />

always a major issue and can cause great losses and problems. Although probiotics, essential<br />

oils and phage therapy were brought up as alternatives methods (IMBEAULT et al. 2006;<br />

YEH et al. 2009; SUN et al. 2010), the main treatment <strong>of</strong> bacterial diseases are still relying on<br />

antimicrobial compounds and vaccines. Antibiotics are usually used either prophylactically or<br />

therapeutically. In aquaculture, broad-spectrum antibiotics are generally used. Licensed<br />

agents vary from country to country. The antibiotics licensed to treat food animals in<br />

European Union were according to VO(EU) 37/2010. Table1-3 showed an overview <strong>of</strong> the<br />

most important antibiotics used in the aquaculture industry.<br />

18


Chapter 1 Introduction<br />

Table1-3<br />

2005)<br />

Examples <strong>of</strong> antibiotics used in aquaculture(LALUMERA et al. 2004; SERRANO<br />

Antibiotics Classification Route <strong>of</strong> administration<br />

Amoxicillin Beta-lactams Oral<br />

Florfenicol<br />

Florfenicol(Fluorinated Oral<br />

derivative <strong>of</strong> thiamphenicol)<br />

Erythromycin Macrolides Oral/Bath/Injection<br />

Streptomycin Spectinomycin Bath<br />

Neomycin Aminoglycosides Bath<br />

Oxolinic acid Quinolones Oral<br />

Enr<strong>of</strong>loxacine<br />

Oral/Bath<br />

Flumequine Flumequine Oral/Bath/Injection<br />

Oxytetracycline Tetracycline Oral/Bath/Injection<br />

Tetracycline<br />

Oral/Bath/Injection<br />

Doxycycline<br />

Oral<br />

Chlortetracycline Chlortetracycline Oral/Bath/Injection<br />

Sulfachlorpyridazine Sulphonamides Oral<br />

Sulfadimethoxine<br />

Sulfamethazine<br />

Sulfanilamide<br />

Sulfaquinoxaline<br />

Sulfathiazole<br />

However, as the usage <strong>of</strong> antibiotics spreads in aquaculture, increasing problems are arising,<br />

such as drug residue and antibiotic resistance. Guichard and Licek (GUICHARD and LICEK<br />

2006) detected the agents which were licensed by various countries contained in fish products,<br />

including oxytetracycline, first-generation quinolones, potentiated sulphonamides, florfenicol,<br />

and amoxicillin. A growing number <strong>of</strong> antibiotic resistant bacteria were recently found in<br />

aquaculture (SON et al. 1997; SCHMIDT et al. 2000; AKINBOWALE et al. 2006; KADLEC<br />

et al. 2011).<br />

3.5.1 Antimicrobial susceptibility <strong>of</strong> Y. ruckeri<br />

The use <strong>of</strong> antimicrobial compounds in aquaculture remains emotive, because <strong>of</strong> the<br />

possibility <strong>of</strong> residues in fish tissues and environment, as well as sustained increasing <strong>of</strong><br />

19


Chapter 1 Introduction<br />

bacterial resistance. General methods for using antimicrobial compounds comprise the oral<br />

route, bath, dip, injection, bioencapsuation and flush. In rainbow trout aquaculture, several<br />

antimicrobial compounds are applied to deal with ERM, for example, chloramphenicol,<br />

flumequine, oxolinic acid, oxytetracycline, potentiated sulphonamide and tiamulin (B<br />

AUSTIN and AUSTIN 2007).<br />

Compared with Flavobacterium psychrophilum, Y. ruckeri still remained higher sensitivity to<br />

antimicrobial substances(SCHMIDT et al. 2000). In 1983, Bosse et al. (BOSSE and POST<br />

1983) showed that Tribrissen (trimethoprim and sulphadiazine) and Tiamulin could be used to<br />

control ERM effectively, at a dosis <strong>of</strong> 1.0mg <strong>of</strong> Tribrissen/kg fish/day and 5.0 mg <strong>of</strong><br />

tiamulin/kg fish/day, respectively. However, in 2006, it was reported that some Y. ruckeri<br />

isolates were resistant to trimethoprim and sulphamethaxazole (KIRKAN et al. 2006). Tinsley<br />

(JOHN TINSLEY 2010) used different serotypes and biotypes <strong>of</strong> Y. ruckeri to test for<br />

antibiotic sensitivity and found all isolates were resistant to sulphatriad and cephalothin, but<br />

sensitive to tetracycline and cotrimoxazole. A 36-MDa plasmid found in some Y. ruckeri<br />

isolates carried the resistant determinants to tetracycline and sulfonamide(DE-GRANDIS and<br />

STEVENSON 1985). It was also shown that all serovar Ⅱ, Ⅲ and Ⅴ isolates exhibited a<br />

high resistance to polymyxin B(DE-GRANDIS and STEVENSON 1985). Shah et al(SHAH et<br />

al. 2012) found that quinolone resistant Y. ruckeri strains from Norway revealed a single bp<br />

mutation by replacing serine by arginine at position 83 in the GyrA protein. Streptomycin and<br />

tetracycline resistant genes, strAB and tetDCAR were discovered on plasmid pYR1 <strong>of</strong> Y.<br />

ruckeri (NCBI Reference Sequence: NC_009139.1).<br />

3.5.2 Vaccination<br />

Y. ruckeri is an opportunistic pathogen commonly present in water and antibiotic treatments<br />

would be problematic as they induce antibiotic resistance. Consequently, the best way to deal<br />

with the infection is vaccination. Formalin killed cells (FKC) and O-Antigen (Ag-O) <strong>of</strong> Y.<br />

ruckeri can significantly increase the immune indicators <strong>of</strong> rainbow trout, and in particular,<br />

directly activate phagocyte activity, such as phagocytosis(ISPIR et al. 2009). So far, a certain<br />

progress has been achieved with the control <strong>of</strong> infections caused by Y. ruckeri. Particularly,<br />

ERM may be controlled effectively by inactivated whole cell vaccines(ROSS et al. 1966),<br />

which are available commercially and are considered to be successful(TEBBIT et al. 1981).<br />

20


Chapter 1 Introduction<br />

Field trials with an existing commercially available ERM vaccine were reported from<br />

Germany, prior to registration and licensing <strong>of</strong> the vaccine(SCHLOTFELDT et al. 1986).<br />

Vaccines for ERM have been used successfully for nearly three decades and consist <strong>of</strong><br />

immersion-applied, killed whole-cell preparations <strong>of</strong> motile serovar 1 Y. ruckeri<br />

strains(STEVENSON 1997). Raida und Buchmann(RAIDA and BUCHMANN 2008) found<br />

that temperature played an significant role in affecting the results <strong>of</strong> vaccination, namely, no<br />

protective effects <strong>of</strong> vaccination were detected in rainbow trout reared at 5 and 25 ℃ after<br />

bath with a bacterin <strong>of</strong> Y. ruckeri O1.<br />

Recently, ERM outbreaks have been reported in vaccinated fish at trout farms in the United<br />

Kingdom(D A AUSTIN et al. 2003), Spain(FOUZ et al. 2006), and the United States(ARIAS<br />

et al. 2007). The Y. ruckeri strains isolated from the outbreaks in Southern England are<br />

atypical serovar 1 isolates lacking both flagellar motility and secreted lipase activity(D A<br />

AUSTIN et al. 2003). These Y. ruckeri biotype 2 (BT2) variants are believed to have a<br />

reduced sensitivity to immersion vaccination (D A AUSTIN et al. 2003). It was hypothesized<br />

that vaccination exerted a selective pressure that enabled the emergence <strong>of</strong> non-motile strains<br />

that are resistant to the commercial vaccines. Therefore, there is a high risk that these<br />

non-motile vaccine-resistant strains spread and originate severe outbreaks <strong>of</strong> disease in trout<br />

farms. From this reason, new approaches based on subunit or DNA vaccines could be used as<br />

an additional way to eliminate or minimize these outbreaks. Nonetheless, DNA vaccines<br />

could not be used together with the original vaccines, because Lorenzen et al (LORENZEN et<br />

al. 2009) described that intraperitoneal injection <strong>of</strong> a commercial vaccine against ERM made<br />

from formalin-killed bacteria immediately followed by a DNA-vaccine against viral<br />

hemorrhagic septicemia virus (VHSV), reduced the protective effect <strong>of</strong> the DNA-vaccine<br />

against challenge with VHSV. Fernández et al (FERNANDEZ et al. 2003) found the relative<br />

percent survival <strong>of</strong> the Yrp1 toxoid-treated fish was 79%, confirming the role <strong>of</strong> the Yrp1<br />

toxoid as a subunit immunogen. It is hopefully that, in the near future, with further research <strong>of</strong><br />

the pathogenesis <strong>of</strong> different Y. ruckeri strains, several new generations <strong>of</strong> vaccines, such as<br />

genetic engineered vaccines, would be developed to control ERM.<br />

4 The Aim <strong>of</strong> the Research<br />

Recently, since vaccine resistant strains <strong>of</strong> Y. ruckeri were isolated on some incidents from<br />

trout hatcheries in North West Germany, an epidemiological survey on the occurrence and<br />

21


Chapter 1 Introduction<br />

phenotypic/molecular characterization <strong>of</strong> Y. ruckeri strains from trout hatcheries in North<br />

West Germany was planned.<br />

The aims <strong>of</strong> the study are:<br />

To provide data on the distribution <strong>of</strong> Y. ruckeri strains in rainbow trout hatcheries and<br />

correlate this to outbreaks <strong>of</strong> ERM in the farms.<br />

To study the diversity <strong>of</strong> both biochemical and molecular characteristics <strong>of</strong> Y. ruckeri isolates<br />

from North West Germany.<br />

To find out the relationship between the fatty acid composition <strong>of</strong> Y. ruckeri isolates and the<br />

epidemiological characters<br />

To detect in vitro cytotoxicity <strong>of</strong> both motile and non-motile strains <strong>of</strong> Y. ruckeri under<br />

different conditions, e.g. temperature and time.<br />

To study virulence factor genes and their expression in collected Y. ruckeri isolates which<br />

would provide a basic virulent background <strong>of</strong> these North West Germany isolates.<br />

To test the antimicrobial susceptibility <strong>of</strong> collected isolates and <strong>of</strong>fer more data on<br />

antimicrobial resistant Y. ruckeri isolates.<br />

It would allow the development <strong>of</strong> a concept for strategic, preventive health monitoring in<br />

fish farms on the basis <strong>of</strong> risk analysis. It would be also helpful with identifying and<br />

characterizing pathogenic Y. ruckeri strains and examine the distribution <strong>of</strong> these strains in the<br />

field as a basis for preventive disease monitoring plans.<br />

22


Chapter 1 Introduction<br />

References<br />

AKINBOWALE, O. L., H. PENG and M. D. BARTON (2006):<br />

Antimicrobial resistance in bacteria isolated from aquaculture sources in Australia.<br />

J Appl Microbiol 100, 1103-1113<br />

ANDERSSON, K., N. CARBALLEIRA, K. E. MAGNUSSON, C. PERSSON, O. STENDAHL, H.<br />

WOLF-WATZ and M. F LLMANN (1996):<br />

YopH <strong>of</strong> Yersinia pseudotuberculosis interrupts early phosphotyrosine signalling associated with phagocytosis.<br />

Mol. Microbiol. 20, 1057-1069<br />

ARIAS, C. R., O. OLIVARES-FUSTER, K. HAYDEN, C. A. SHOEMAKER, J. M. GRIZZLE and P. H.<br />

KLESIUS (2007):<br />

First report <strong>of</strong> Yersinia ruckeri biotype 2 in the USA.<br />

J Aquat Anim Health 19, 35-40<br />

AUSTIN, B. and D. A. AUSTIN (1999):<br />

Bacterial Fish Pathogens: Disease <strong>of</strong> Farmed and Wild Fish.<br />

Praxis Publishing Ltd., Chichester, UK<br />

AUSTIN, B. and D. A. AUSTIN (2007):<br />

Bacterial Fish Pathogens: Diseases <strong>of</strong> Farmed and Wild Fish Praxis Publishing, Chichester, UK<br />

AUSTIN, B., M. GREEN and C. J. RODGERS (1982):<br />

Morphological diversity among strains <strong>of</strong> Yersinia ruckeri.<br />

Aquaculture 27, 73-78<br />

AUSTIN, D. A., P. A. W. ROBERTSON and B. AUSTIN (2003):<br />

Recovery <strong>of</strong> a new biogroup <strong>of</strong> Yersinia ruckeri from diseased rainbow trout(Oncorhynchus mykiss, Walbaum).<br />

Syst Appl Microbiol 26, 127-131<br />

BASTARDO, A., H. BOHLE, C. RAVELO, A. E. TORANZO and J. L. ROMALDE (2011a):<br />

Serological and molecular heterogenetity among Yersinia ruckeri strains isolated from farmed Atlantic salmon<br />

Salmo salar in Chile.<br />

Dis Aqua Organ 93, 207-214<br />

BASTARDO, A., C. RAVELO and J. L. ROMALDE (2012):<br />

Multilocus sequence typing reveals high genetic diversity and epidemic population structure for the fish<br />

pathogen Yersinia ruckeri.<br />

Environ microbiol 14, 1888-1897<br />

BASTARDO, A., V. SIERRALTA, J. LEON, C. RAVELO and J. L. ROMALDE (2011b):<br />

Phenotypical and genetic characterization <strong>of</strong> Yersinia ruckeri strains isolated from recent outbreaks in farmed<br />

rainbow trout Oncorhynchus mykiss (Walbaum) in Peru.<br />

Aquaculture 317, 229-232<br />

BERG, H. C. and R. A. ANDERSON (1973):<br />

Bacteria weim by rotating their flagellar filaments.<br />

Nature 245, 380-382<br />

BERGH, O. (2008):<br />

Bacterial disease <strong>of</strong> fish.<br />

Science Publishers, Enfield, NJ<br />

BHATTACHARJEE, M. K., S. C. KACHLANY, D. H. FINE and D. H. FIGURSKI (2001):<br />

Nonspecific adherence and fibril biogenesis by Actinobacillus actinomycetemcomitans:TadA protein is an<br />

ATPase.<br />

J. Bacteriol. 183, 5927-5936<br />

23


Chapter 1 Introduction<br />

BOSSE, M. P. and G. POST (1983):<br />

Tribrissen and tiamulin for control <strong>of</strong> enteric redmouth disease.<br />

J Fish Dis 6, 27-32<br />

BOTTONE, E. J., H. BERCOVIER and H. H. MOLLARET (2005):<br />

Genus. XLI. Yersinia.<br />

In: Bergey‟s Manual <strong>of</strong> Systematic Bacteriology<br />

Springer-Verlag, New York, S. 838-848<br />

BRAVO, S. and V. KOJAGURA (2004):<br />

First Isolation <strong>of</strong> Yersinia ruckeri from rainbow trout (Oncorhynchus mykiss) in Peru.<br />

Bull Eur Assn Fish P 24, 104-108<br />

BUCHANAN, R. E. and N. E. GIBBONS (1974):<br />

Bergey‟s manual <strong>of</strong> determinative bacteriology.<br />

The Williams & Wilkins Co., Baltimore<br />

BULLER, N. B. (2004):<br />

Interpretation <strong>of</strong> Biochemical Identification Tests and Sets.<br />

In: Bacteria from Fish and Other Aquatic Animals: A Practical Identification Manual<br />

CABI Publishing, London, UK, S. 117-136<br />

BURROWS, T. and G. A. BACON (1956):<br />

The basis <strong>of</strong> virulence in Pasteurella pestis: an antigen determinig virulence.<br />

Br J Exp Pathol. 37, 481-493<br />

CASTIGNETTI, D. and J. SMARRELLI-JR. (1986):<br />

Siderophores, the iron nutrition <strong>of</strong> plants, and nitrate reductase.<br />

FEMS Microbiol Lett 209, 147-151<br />

CHEN, P. E., C. COOK, A. C. STEWART, N. NAGARAJAN, D. D. SOMMER, M. POP, B. THOMASON, M.<br />

P. K. THOMASON, S. LENTZ, N. NOLAN, A. SULAKVELIDZE, A. MATECZUN, L. DU, M. E. ZWICK<br />

and T. D. READ (2010):<br />

Genomic characterization <strong>of</strong> the Yersinia genus.<br />

Genome Biology 11, R1<br />

CHEVANCE, F. F. V. and K. T. HUGHES (2008):<br />

Coordinating assembly <strong>of</strong> a bacterial macromolecular machine.<br />

Nat Rev Microbiol 6, 455-465<br />

CLLINS, R. O., G. FOSTER and H. M. ROSS (1996):<br />

Isolation <strong>of</strong> Yersinia ruckeri from an otter and salmonid fish from adjacent freshwater catchments.<br />

Vet Rec 139, 169<br />

COQUET, L., P. COSETTE, G.-A. JUNTER, E. BEUCHER, J.-M. SAITER, T. JOUENNE and J.-M. S. D<br />

(2002a):<br />

Adhesion <strong>of</strong> Yersinia ruckeri to fish farm materials:influence <strong>of</strong> cell and material surface properties.<br />

Colloids and surfaces B 26, 373-378<br />

COQUET, L., P. COSETTE, L. QUILLET, F. PETIT, G. A. JUNTER and T. JOUENNE (2002b):<br />

Occurrence and phenotypic charaterization <strong>of</strong> Yersinia ruckeri strains with bi<strong>of</strong>ilm-forming capacity in a<br />

rainbow trout Farm.<br />

Appl Environ Microbiol 68, 470-475<br />

CORNELIS, G. R. (2002):<br />

Yersinia type III secretion: send in the effectors.<br />

J Cell Biol 158, 401-408<br />

24


Chapter 1 Introduction<br />

CORNELIS, G. R. (2006):<br />

The type III secretion injectisome.<br />

Nat Rev Microbiol 4, 811-825<br />

CRUZ, J. A. D. L., A. RODRIGUEZ, C. TEJEDOR, E. D. LUCAS and L. R. OROZCO (1986):<br />

Isolation and identification <strong>of</strong> Yersinia ruckeri, causal agent <strong>of</strong> enteric redmouth disease (ERM), for the first time<br />

in Spain.<br />

Bull Eur Assn Fish P 6, 43-44<br />

DAHIYA, I. and R. M. W. STEVENSON (2010a):<br />

The UvrY response regulator <strong>of</strong> the BarA–UvrY two-component system contributes to Yersinia ruckeri infection<br />

<strong>of</strong> rainbow trout(Oncorhynchus mykiss).<br />

Arch. Microbiol. 192, 541-547<br />

DAHIYA, I. and R. M. W. STEVENSON (2010b):<br />

Yersinia ruckeri genes that attenuate survival in rainbow trout(Oncorhynchus mykiss) are identified using<br />

signature-tagged mutants.<br />

Vet Microbiol 144, 399-404<br />

DAHIYA, I. and R. M. W. STEVENSON (2010c):<br />

The ZnuABC operon is important for Yersinia ruckeri infections <strong>of</strong> rainbow trout, Oncorhynchus mykiss<br />

(Walbaum).<br />

J Fish Dis 33, 331-340<br />

DALY, J. G., B. LINDVIK and R. M. W. STEVENSON (1986):<br />

Serological heterogeneity <strong>of</strong> recent isolates <strong>of</strong> Yersinia ruckeri from Ontario and British Columbia.<br />

Dis Aqua Organ 1, 151-153<br />

DANLEY, M. L., A. E. GOODWIN and H. S. KILLIAN (1999):<br />

Epizootics in farm-raised channel catfish, Ictalurus punctatus(Rafinesque), caused by the enteric redmouth<br />

bacterium Yersinia ruckeri.<br />

J Fish Dis 22, 451-456<br />

DAVIES, R. L. (1990):<br />

O-Serotyping <strong>of</strong> Yersinia ruckeri with Special Emphasis on European Isolates.<br />

Vet Microbiol 22, 299-307<br />

DAVIES, R. L. (1991a):<br />

Clonal analysis <strong>of</strong> Yersinia ruckeri based on biotypes, serotypes and outer membrane protein-types.<br />

J Fish Dis 14, 221-228<br />

DAVIES, R. L. (1991b):<br />

Virulence and serum-resistance in different clonal groups and serotypes <strong>of</strong> Yersinia ruckeri.<br />

Vet Microbiol 29, 289-297<br />

DAVIES, R. L. and G. N. FRERICHS (1989):<br />

Morphological and biochemical differences among isolates <strong>of</strong> Yersinia ruckeri obtained from wide geographical<br />

areas.<br />

J Fish Dis 12, 357-365<br />

DE-GRANDIS, S. A. and R. M. W. STEVENSON (1985):<br />

Antimicrobial susceptibility patterns and R plasmid-mediated resistance <strong>of</strong> the fish pathogen Yersinia ruckeri.<br />

Antimicrob Agents Ch 27, 938-942<br />

DESVAUX, M., M. HEBRAUD, I. R. HENDERSON and M. J. PALLEN (2006):<br />

Type III secretion: what's in a name?<br />

Trends in Microbiology 14, 157-160<br />

DULIN, M. P., T. HUDDLESTON, R. E. LARSON and G. W. KLONTZ (1976):<br />

Enteric redmouth disease.<br />

25


Chapter 1 Introduction<br />

<strong>University</strong> <strong>of</strong> Idaho, College <strong>of</strong> Forestry, Wild and Range Science Bulletin No 8 Moscow, ID 8:15pp<br />

DWILOW, A. G., B. W. SOUTER and K. KNIGHT (1987):<br />

Isolation <strong>of</strong> Yersinia ruckeri from burbot, Lota Lota (L), from the mackenzie river, Canada.<br />

Journal <strong>of</strong> Fish Disease 10, 315-317<br />

ENRIQUEZ, R. and J. ZAMORA (1987):<br />

Isolation <strong>of</strong> Yersinia ruckeri from Carps (Cyprinus carpio) in Valdiva.<br />

Archivos Se Medicina Veterinaria 19, 33-36<br />

ERHARDT, M., K. NAMBA and K. T. HUGHES (2010):<br />

Bacterial nanomachines: The flagellum and type III injectisome.<br />

Cold Spring Harb Perspect Biol 2, a000299<br />

EVENHUIS, J. P., S. E. LAPATRA, D. W. VERNER-JEFFREYS, I. DALSGAARD and T. J. WELCH (2009):<br />

Identification <strong>of</strong> Flagellar Motility Genes in Yersinia ruckeri by Transposon Mutagenesis.<br />

Antimicrob Agents Ch 75, 6630-6633<br />

EWING, W. H., A. J. ROSS, D. J. BRENNER and G. R. FANNING (1978):<br />

Yersinia ruckeri sp. nov., the Redmouth (RM) Bacterium.<br />

International Journal <strong>of</strong> Systematic Bacteriology 28, 37-44<br />

FARALDO-GOMEZ, J. and S. P. SANSON (2003):<br />

Acquisition <strong>of</strong> siderophores in Gram-negative bacteria.<br />

National Review 4, 105-116<br />

FERN NDEZ, L., J. R. L PEZ, P. SECADES, A. MEN NDEZ, I. M RQUEZ and J. A. GUIJARRO (2003):<br />

In vitro and in vivo studies <strong>of</strong> the Yrp1 protease from Yersinia ruckeri and its role in protective immunity against<br />

enteric redmouth disease <strong>of</strong> salmonids.<br />

Appl. Environ. Microbiol. 69, 7328-7335<br />

FERN NDEZ, L., I. M RQUEZ and J. A. GUIJARRO (2004):<br />

Identification <strong>of</strong> Specific In Vivo-Induced (ivi) Genes in Yersinia ruckeri and Analysis <strong>of</strong> Ruckerbactin, a<br />

Catecholate Siderophore Iron Acquisition System.<br />

Appl Environ Microbiol 70, 5199-5207<br />

FERN NDEZ, L., M. PRIETO and J. A. GUIJARRO (2007):<br />

The iron- and temperature-regulated haemolysin YhlA is a virulence factor <strong>of</strong> Yersina ruckeri.<br />

Microbiology 153, 483-489<br />

FERN NDEZ, L., P. SECADES, J. R. LOPEZ, I. MARQUEZ and J. A. GUIJARRO (2002):<br />

Isolation and analysis <strong>of</strong> a protease gene with an ABC transport system in the fish pathogen Yersinia ruckeri:<br />

insertional mutagenesis and involvement in virulence.<br />

Microbiology 148, 2233-2243<br />

FOUZ, B., C. ZARZA and C. AMARO (2006):<br />

First description <strong>of</strong> non-motile Yersinia ruckeri serovar I strains causing disease in rainbow trout, Oncorhynchus<br />

mykiss (Walbaum), cultured in Spain.<br />

J Fish Dis 29, 339-346<br />

FRANCIS, M. S., H. WOLF-WATZ and A. FORSBERG (2002):<br />

Regulation <strong>of</strong> type III secretion systems.<br />

Current Opinnion in Microbiology 5, 166-172<br />

FRERICHS, G. N., J. A. STEWART and R. O. COLLINS (1985):<br />

Atypical infection <strong>of</strong> rainbow trout, Salmo gairdneri Richardson, with Yersinia ruckeri.<br />

J Fish Dis 8, 383-387<br />

FUHRMANN, H. and K. H. BOEHM (1983):<br />

An outbreak <strong>of</strong> enteric redmouth disease in West Germany.<br />

J Fish Dis 6, 309-311<br />

26


Chapter 1 Introduction<br />

FURONES, M. D. (1990)<br />

Factors affecting the pathogenicity <strong>of</strong> Yersinia ruckeri.<br />

Plymouth, UK,<br />

FURONES, M. D., M. L. GILPIN and C. B. MUNN (1993a):<br />

Culture media for the differentiation <strong>of</strong> isolates <strong>of</strong> Yersinia ruckeri, based on detection <strong>of</strong> a virulence factor.<br />

J Appl Microbiol 74, 360-366<br />

FURONES, M. D., C. J. RODGERS and C. B. MUNN (1993b):<br />

Yersina ruckeri, the causal agent <strong>of</strong> enteric redmouth disease(ERM) in fish.<br />

Annu Rev Fish Dis 105-125<br />

GOOD, C. M., M. A. THORBURN and R. M. W. STEVENSON (2001):<br />

Host factors associated with the detection <strong>of</strong> Aeromonas salmonicida and Yersinia ruckeri in Ontario, Canada<br />

government fish hatcheries.<br />

Prev Vet Med 49, 165-173<br />

GREEN, M. and B. AUSTIN (1983):<br />

The identification <strong>of</strong> Yersinia ruckeri and its relationship to other representatives <strong>of</strong> the enterobacteriaceae.<br />

Aquaculture 34, 185-192<br />

GUICHARD, B. and E. LICEK (2006):<br />

A comparative study <strong>of</strong> antibiotics registered for use in farmed fish in European countries.<br />

Poster presented at the First OIE Global Conference on Aquatic Animal Health 10 October, Bergen, Norway<br />

GUNASENA, D., J. KOMROWER and S. MACINTYRE (2003):<br />

The Fish Pathogen Yersinia ruckeri Possesses a TTS System.<br />

In: The Genus Yersinia: Entering the Functional Genomic Era<br />

Kluwer Academic/Plenum Publishers, New York, USA, S. 105-107<br />

HORNE, M. T. and A. C. BARNES (1999):<br />

Enteric redmouth disease (Yersinia ruckeri).<br />

In: Fish diseases and disorders<br />

CABI Publishing, Wallingford, S. 445-477<br />

HUECK, C. J. (1998):<br />

Type III Protein Secretion Systems in Bacterial Pathogens <strong>of</strong> Animals and Plants.<br />

Microbiology and Molecular Biology Reviews 62, 379-433<br />

HUNTER, V. A., M. D. KNITTEL and J. L. FRYER (1980):<br />

Stress-induced transmission <strong>of</strong> Yersinia ruckeri infection from carriers to recipient steelhead trout Salmo<br />

gairdneri Richardson.<br />

J Fish Dis 3, 467-472<br />

IMBEAULT, S., S. PARENT, M. LAGACE, C. F. UHLAND and J.-F. BLAIS (2006):<br />

Using bacteriophages to prevent furunculosis caused by Aeromonas salmonicida in farmed brook trout.<br />

J Aqua Anim Health 18, 203-214<br />

ISBERG, R. R. and S. FALKOW (1985):<br />

A single genetic locus encoded by Yersinia psuedotuberculosis permits invasion <strong>of</strong> cultured animal cells by<br />

Escherichia coli K-12.<br />

Nature 317, 262-264<br />

ISPIR, U., H. B. GOKHAN, M. OZCAN, M. DORUCU and N. SAGLAM (2009):<br />

Immune Response <strong>of</strong> Rainbow Trout (Oncorhynchus mykiss) to Selected Antigens <strong>of</strong> Yersinia ruckeri.<br />

Acta Vet. Brno 78, 145-150<br />

27


Chapter 1 Introduction<br />

KADLEC, K., E. V. CZAPIEWSKI, H. KASPAR, J. WALLMANN, G. B. MICHAEL, U. STEINACKER and S.<br />

SCHWARZ (2011):<br />

Molecular basis <strong>of</strong> sulfonamide and trimethoprim resistance in fish-pathogenic Aeromonas isolates.<br />

Appl Environ Microbiol 77, 7147-7150<br />

KARATAS, S., A. CANDAN and D. DEMIRCAN (2004):<br />

Enteric Red Mouth Disease in Cultured Rainbow trout (Oncorhynchus Mykiss) on the black sea coast <strong>of</strong> Turkey.<br />

Isr J Aquacult-Bamid 56, 226-231<br />

KAWULA, T. H., M. J. LELIVELT and P. E. ORNDORFF (1996):<br />

Using a new inbred fish model and cultured fish tissue cells to study Aeromonas hydrophila and Yersinia ruckeri<br />

pathogenesis.<br />

Microb Pathogenesis 20, 119-125<br />

KIM, W. (2000)<br />

Investigation <strong>of</strong> the role <strong>of</strong> flagella in the virulence <strong>of</strong> Yersinia ruckeri.<br />

Canada,<br />

KIRKAN, S., E. Ö. G KSOY, O. KAYA and S. TEKBIYIK (2006):<br />

In-vitro antimicrobial susceptibility <strong>of</strong> pathogenic bacteria in rainbow trout (Oncorhynchus mykiss, Walbaum).<br />

Turkish Journal <strong>of</strong> <strong>Veterinary</strong> and Animal Sciences 30, 337-341<br />

KLEIN, B. U., D. W. KLEINGELD and K. H. BOHM (1994):<br />

First isolation <strong>of</strong> a non-motile/tween 80 negative Yersinia ruckeri strain in Germany.<br />

Bull Eur Assn Fish P 14, 165-166<br />

KNUTTON, S., I. ROSENSHINE, M. J. PALLEN, I. NISAN and B. C. NEVES (1998):<br />

A novel EspA-associated surface organelle <strong>of</strong> enteropatogenic Escherichia coli involved in protein translocation<br />

into epithelial cells.<br />

The EMBO Journal 17, 2166-2176<br />

KOTETISHVILI, M., A. KREGER, G. WAUTERS, J. G. JR. MORRIS, A. SULAKVELIDZE and O. C. STINE<br />

(2005):<br />

Multilocus Sequence Typing for Studying Genetic Relationships among Yersinia Species.<br />

American Society for Microbiology 43, 2674-2684<br />

LALUMERA, G. M., D. CALAMARI, P. GALLI, S. CASTIGLIONI, G. CROSA and R. FANELLI (2004):<br />

Preliminary investigation on the enviromental occurrence and effects <strong>of</strong> antibiotics used in aquaculture in Italy.<br />

Chemosphere 54, 661-668<br />

LI, S., D. WANG, H. LIU and T. LU (2013):<br />

Isolation <strong>of</strong> Yersinia ruckeri strain H01 from farm-raised Amur Sturgeon Acipenser schrencki in China.<br />

J Aqua Anim Health 25, 9-14<br />

M NDEZ, J., L. FERN NDEZ, A. MEN NDEZ, P. REIMUNDO, D. P REZ-PASCUAL, R. NAVAIS and J. A.<br />

GUIJARRO (2009):<br />

A Chromosomally Located traHIJKCLMN Operon Encoding a Putative Type IV Secretion System Is Involved<br />

in the Virulence <strong>of</strong> Yersinia ruckeri.<br />

Appl Environ Microbiol 75, 937-945<br />

MACNAB, R. M. (2004):<br />

Type III flagellar protein export and flagellar assembly.<br />

Biochemica et Biophysica Acta 1694, 207-217<br />

MAHJOOR, A. A. and M. AKHLAGHI (2012):<br />

A pathological study <strong>of</strong> rainbow trout organs naturally infected with enteric redmouth disease.<br />

Asian Journal <strong>of</strong> Animal Sciences 6, 147-153<br />

MARTINEZ, J. L., A. DELGADO-IRIBARREN and F. BAQUERO (1990):<br />

Mechanisms <strong>of</strong> iron acquisition and bacteril virulence.<br />

FEMS Microbiology Reviews 75,<br />

28


Chapter 1 Introduction<br />

MCNALLY, A., R. M. L. RAGIONE, A. BEST, G. MANNING and D. G. NEWELL (2007):<br />

An aflagellate mutant Yersinia enterocolitica biotype 1A strain displays altered invasion <strong>of</strong> epithelial cells,<br />

persistence in macrophages, and cyokine secretion pr<strong>of</strong>iles in vitro.<br />

J Med Microbiol 153, 1339-1349<br />

MEIER, W. (1986):<br />

Enteric redmouth disease: outbreak in rainbow trout in Switzerland.<br />

Dis Aqua Organ 2, 81-82<br />

MILLER, V. L. and S. FALKOW (1988):<br />

Evidence for two genetic loci from Yersinia enterocolitica that can promote invasion <strong>of</strong> epithelial cells.<br />

Infect Immun 56, 1242-1248<br />

MILLS, S. D., A. BOLAND, M. P. SORY, P. VAN DER SMISSEN, C. KERBOURCH, B. B. FINLAY and G.<br />

R. CORNELIS (1997):<br />

Yersinia enterocolitica induces apoptosis in macrophages by a process requiring functional type III secretion and<br />

translocation mechanisms and involving YopP, presumably acting as an effector protein.<br />

Proc Natl Acad Sci USA 94, 12638-12643<br />

MILTON, D. L., R. O'TOOLE, P. HORSTEDT and H. WOLF-WATZ (1996):<br />

Flagellin A is essential for the virulence <strong>of</strong> Vibrio anguillarum.<br />

J Bacteriol 178, 1310-1319<br />

NOGA, E. J. (2010):<br />

Fish Disease: Diagnosis and Treatment<br />

Wiley-Blackwell, Iowa<br />

NOGA, E. J., J. F. LEVINE, K. TOWNSEND, R. A. BULLIS, C. P. CARLSON and W. T. CORBETT (1988):<br />

Kidney biopsy: a nonlethal method for diagnosing Yersinia ruckeri infection (enteric redmouth disease) in<br />

rainbow trout (Salmo gairdneri).<br />

Am J Vet Res 49, 363-365<br />

ONUK, E. E., A. CIFTCI, A. FINDIK, G. CIFTCI, S. ALTUN, F. BALTA, S. OZER and A. Y. COBAN (2011):<br />

Phenotypic and molecular characterization <strong>of</strong> Yersinia ruckeri isolates from Rainbow trout (Oncorhynchus<br />

mykiss, Walbaum,1792) in Turkey.<br />

Berliner und Münchener Tierärztliche Wochenschrift 124, 320-328<br />

PAGE, A. L. and C. PARSOT (2002):<br />

Chaperones <strong>of</strong> the type III secretion pathway: jacks <strong>of</strong> all trades.<br />

Mol Microbiol 46, 1-11<br />

PALLEN, M. J., S. A. BEATSON and C. M. BAILEY (2005):<br />

Bioinformatics, genomics and evolution <strong>of</strong> non-flagellar type-III secretion systems: a Darwinian perpective<br />

FEMS Microbiology Reviews 29, 201-229<br />

PĘKALA, A., A. KOZIŃSKA and J. ANTYCHOWICZ (2010):<br />

Serological Variation Among Polish Isolates <strong>of</strong> Yersinia ruckeri.<br />

Bull Vet Inst Pulawy 54, 305-308<br />

RAIDA, M. K. and K. BUCHMANN (2008):<br />

Bath vaccination <strong>of</strong> rainbow trout (Oncorhynchus mykiss Walbaum) against Yersinia ruckeri: Effects <strong>of</strong><br />

temperature on protection and gene expression.<br />

Vaccine 26, 1050-1062<br />

ROBERTS, R. J. (2001):<br />

Fish Pathology.<br />

WB Sauners, London<br />

ROMALDE, J. L., R. F. CONCHAS and A. E. TORANZO (1991a):<br />

Evidence that Yersinia ruckeri possesses a high affinity iron uptake system.<br />

29


Chapter 1 Introduction<br />

FEMS Microbiol Lett 80, 121-126<br />

ROMALDE, J. L., I. ITEMAN and E. CARNIEL (1991b):<br />

Use <strong>of</strong> pulsed field gel electrophoresis to size the chromosome <strong>of</strong> the bacterial fish pathogen Yersinia ruckeri<br />

FEMS Microbiol Lett 84, 217-226<br />

ROMALDE, J. L., B. MARGARINOS, J. L. BARJA and A. E. TORANZO (1993):<br />

Antigenic and molecular characterization <strong>of</strong> Yersinia ruckeri Proposal for a new intraspecies classification.<br />

Syst Appl Microbiol 16, 411-419<br />

ROMALDE, J. L. and A. E. TORANZO (1993):<br />

Pathological activities <strong>of</strong> Yersinia ruckeri, the enteric redmouth (ERM) bacterium.<br />

FEMS Microbiol Lett 112, 291-300<br />

ROSENTHAL, H. and V. HILGE (2000):<br />

Aquaculture production and environmental regulations in Germany.<br />

J appl Ichthyol 16, 163-166<br />

ROSS, A. J., R. R. RUCKER and W. H. EWING (1966):<br />

Description <strong>of</strong> a bacterium associated with redmouth disease <strong>of</strong> rainbow trout (Salmo Gairdneri).<br />

Can J Microbiol 12, 763-770<br />

RUCKER, R. R. (1966):<br />

Redmouth disease <strong>of</strong> rainbow trout (Salmo gairdner).<br />

Bull <strong>of</strong>f int Epizoot. 65, 825-830<br />

RYCKAERT, J., P. BOSSIER, K. D. HERDE, A. DIEZ-FRAILE, P. SORGELOOS, F. HAESEBROUCK and F.<br />

PASMANS (2010):<br />

Persistence <strong>of</strong> Yersinia ruckeri in trout macrophages.<br />

Fish Shellfish Immun 29, 648-655<br />

SALMOND, G. P. and P. J. REEVES (1993):<br />

Membrane Traffic Wardens and Protein Secretion in Gram-negative Bacteria.<br />

Trends in Biochemical Sicences 18, 7-12<br />

SALMOND, G. P. C. (1994):<br />

Secretion <strong>of</strong> extracellular virulence factors by plant pathogenic bacteria.<br />

Annu Rev Phytopathol 32, 181-200<br />

SANTOS, Y., J. L. ROMALDE, I. BANDIN, B. MAGARIFIOS, S. NUIIEZ, J. L. BARJA and A. A. E.<br />

TORANZO (1993):<br />

Usefulness <strong>of</strong> the API-20E system for the identification <strong>of</strong> bacterial fish pathogens.<br />

Aquaculture 116, 111-120<br />

SAVAS, H., I. ALTINOK, E. CAKMAK and S. FIRIDIN (2006):<br />

Isolation <strong>of</strong> Renibacterium salmoninarum from cultured Black Sea salmon (Salmo trutta labrax): first report in<br />

Turkey.<br />

Bull Eur Assn Fish P 26, 238-246<br />

SCH PERCLAUS, W. (1991):<br />

Special Part A. Diseases caused by pathogens.<br />

In: Fish Diseases<br />

Amerind Publishing Co. Pvt. Ltd, New Delhi, S. 531<br />

SCHLOTFELDT, H. J., K. H. B HM, F. PFORTM LLER and K. PFORTM LLER (1986):<br />

Enteric redmouth disease/ERM <strong>of</strong> trout and other farmed fish in northwest Germany.<br />

Fischkrankheiten 37, 66-70<br />

SCHMIDT, A. S., M. S. BRUUN, I. DALSGAARD, K. PEDERSEN and J. L. LARSEN (2000):<br />

30


Chapter 1 Introduction<br />

Occurence <strong>of</strong> antimicrobial resistance in fish-pathogenic and enviromental bacteria associated with four Danish<br />

rainbow trout farms.<br />

Appl Environ Microbiol 66, 4908-4915<br />

SECADES, P. and J. A. GUIJARRO (1999):<br />

Purification and characterization <strong>of</strong> an extracellular protease from the fish pathogen Yersinia ruckeri and effect<br />

<strong>of</strong> culture conditions on production.<br />

Appl Environ Microbiol 65, 3969-3975<br />

SEKER, E., M. KARAHAN, U. ISPIR, B. CATINKAYA, N. SAGLAM and M. SARIEYYUPOGLU (2012):<br />

Investigation <strong>of</strong> Yersinia ruckeri infection in rainbow trout (Oncorhynchus mykiss Walbaum 1792) farms by<br />

polymerase chain reaction (PCR) and bacteriological culture.<br />

Kafkas Universitesi Veteriner Fakultesi Dergisi 18, 913-916<br />

SERRANO, P. H. (2005):<br />

Responsible use <strong>of</strong> antibiotics in aquaculture.<br />

FAO Fisheries Technical Paper 469, 97<br />

SHAH, S. Q. A., S. KARATAS, H. NILSEN and T. M. STEINUM (2012):<br />

Characterization and expression <strong>of</strong> the gyrA gene from quinolone resistant Yersinia ruckeri strains isolated from<br />

Atlantic salmon (Slamo salar L.) in Norway.<br />

Aquaculture 350-353, 37-41<br />

SON, R., G. RUSUL, A. M. SAHILAH, A. ZAINURI, A. R. RAHA and I. SALMAH (1997):<br />

Antibiotic resistance and plasmid pr<strong>of</strong>ile <strong>of</strong> Aeromonas hydrophila isolates from cultured fish, Telapia (Telapia<br />

mossambica).<br />

Lett Appl Microbiol 24, 479-482<br />

SOUZA, R. A., A. PITONDO-SILVA, D. P. FALC O and J. P. FALC O (2010):<br />

Evaluation <strong>of</strong> four molecular typing methodologies as tools for determining taxonomy relations and for<br />

identifying species among Yersinia isolates.<br />

J Microbiol Meth 82, 141-150<br />

SPARBOE, O., C. KOREN, T. HAASTEIN, T. POPPE and H. STENWIG (1986):<br />

The first isolation <strong>of</strong> Yersinia ruckeri from farmed Norwegian salmon.<br />

Bull Eur Assn Fish P 6, 41-42<br />

STEVENSON, R. M. W. (1997):<br />

Immuniztion with bacterial antigens: yersiniosis.<br />

Dev Biol Stand 90, 117-124<br />

STEVENSON, R. M. W. and D. E. AIRDRIE (1984a):<br />

Serological variation among Yersinia ruckeri strains.<br />

J Fish Dis 7, 247-254<br />

STEVENSON, R. M. W. and D. W. AIRDRIE (1984b):<br />

Isolation <strong>of</strong> Yersinia ruckeri bacteriophages.<br />

Appl Environ Microbiol 47, 1201-1205<br />

STR M-BESTOR, M., N. MUSTAM KI, S. HEINIKAINEN, V. HIRVEL -KOSKI, D. VERNER-JEFFREYS<br />

and T. WIKLUND (2010):<br />

Introduction <strong>of</strong> Yersinia ruckeri biotype 2 into Finnish fish farms.<br />

Aquaculture 308, 1-5<br />

SUN, Y. Z., H. L. YANG, R. L. MA and W. Y. LIN (2010):<br />

Probiotic applications <strong>of</strong> two dominant gut Bacillus strains with antagonistic activity improved the growth<br />

performance nd immune responses <strong>of</strong> grouper Epinephelus coioides.<br />

Fish Shellfish Immun 29, 803-809<br />

TINSLEY, J. (2010)<br />

31


Chapter 1 Introduction<br />

Studies on the pathogenicity <strong>of</strong> Yersinia ruckeri biotype 2 to rainbow trout (Oncorhynchus mykiss, Walbaum)<br />

Edinburgh, UK School <strong>of</strong> Life Sciences<br />

TOBBACK, E., A. DECOSTERE, K. HERMANS, J. RYCKAERT, L. DUCHATEAU, F. HAESEBROUCK<br />

and K. CHIERS (2009):<br />

Route <strong>of</strong> entry and tissue distribution <strong>of</strong> Yersinia ruckeri in experimentally infected rainbow trout Oncorhynchus<br />

mykiss.<br />

Dis Aqua Organ 84, 219-228<br />

TOBBACK, E., A. DECOSTERE, K. HERMANS, W. VAN DEN BROECK, F. HAESEBROUCK and K.<br />

CHIERS (2010):<br />

In vitro markers for virulence in Yersinia ruckeri.<br />

J Fish Dis 33, 197-209<br />

TOBE, T., S. A. BEASON, H. TANIGUCHI, H. ABE, C. M. BAILEY, A. FIVIAN, R. YOUNIS, S.<br />

MATTHEWS, O. MARCHES, G. FRANKEL, T. HAYASHI and M. J. PALLEN (2006):<br />

An extensive repertoire <strong>of</strong> type III secretion effectors in Escherichia coli O157 and the role <strong>of</strong> lambdoid phages<br />

in their dissemination.<br />

Proceedings <strong>of</strong> the National Academy <strong>of</strong> Sciences <strong>of</strong> the United States <strong>of</strong> America 103, 14941-14946<br />

VALTONEN, E. T., P. RINTAMAKI and M. KOSKIVAARA (1992):<br />

Occurrence and pathogenicity <strong>of</strong> Yersinia ruckeri at fish farms in northern and central Finland.<br />

J Fish Dis 15, 163-171<br />

VILCHES, S., N. JIMENEZ, J. M. TOM S and S. MERINO (2009):<br />

Aeromonas hydrophila AH-3 Type III Secretion System Expression and Regulatory Network.<br />

Appl Environ Microbiol 75, 6382-6392<br />

VLADIK, P. and A. PROUZA (1990):<br />

The Finding <strong>of</strong> Yersinia ruckeri in rainbow trout in Czechoslovakia.<br />

Veterinarni Medicina 35, 437-442<br />

VUILLAUME, A., R. BRUN, P. CHENE, E. SOCHON and R. LESEL (1987):<br />

First isolation <strong>of</strong> Yersinia ruckeri from sturgeon, Acipenser baeri Brandt, in south west <strong>of</strong> France.<br />

Bull Eur Assn Fish P 7, 18-19<br />

WALTMAN, W. D. and E. B. SHOTTS (1984):<br />

A medium for the isolation and differentiation <strong>of</strong> Yersinia ruckeri.<br />

Can J Fish Aquat Sci 41, 804-806<br />

WANG, K., H. LIAN, D. CHEN, J. HUANG, J. WANG and Q. MOU (2012):<br />

Cloning, identification and molecular characteristics analysis <strong>of</strong> p1 gene <strong>of</strong> Yersinia ruckeri isolated from<br />

channel catfish (Ictalurus punctatus).<br />

Asian Journal <strong>of</strong> Animal and <strong>Veterinary</strong> Advances 7, 1067-1078<br />

WANG, K. Y., F. L. FAN, X. L. HUANG, Y. GENG, D. F. CHEN and J. L. HUANG (2009):<br />

Occurrence and treatment <strong>of</strong> a channel catfish(Ictalurus punctatus(Rafinesque)) disease, caused by Yersinia<br />

ruckeri.<br />

Scientific Fish Farming 5, 50-51<br />

WARSEN, A. E., M. J. KRUG, S. LAFRENTZ, D. R. STANEK, F. J. LOGE and D. R. CALL (2004):<br />

Simultaneous siscrimination between 15 fish pathogens by using 16S ribosomal DNA PCR and DNA<br />

microarrays.<br />

Appl Environ Microbiol 70, 4216-4221<br />

WELCH, T. J. (2011):<br />

Rapid genotyping assays for the identification and differentiation <strong>of</strong> Yersinia ruckeri biotype 2 strains.<br />

Lett Appl Microbiol 53, 383-385<br />

WHEELER, R. W., R. L. DAVIES, I. DALSGAARD, J. GARCIA, T. J. WELCH, S. WAGLEY, K. S.<br />

BATEMAN and D. W. VERNER-JEFFREYS (2009):<br />

32


Chapter 1 Introduction<br />

Yersinia ruckeri biotype 2 isolates from mainland Europe and the UK represent different clonal groups.<br />

Dis Aqua Organ 84, 25-33<br />

WORTBERG, F., E. NARDY, M. CONTZEN and J. RAU (2012):<br />

Identification <strong>of</strong> Yersinia ruckeri from diseased salmonid fish by fourier transform infrared spectroscopy.<br />

J Fish Dis 35, 1-10<br />

YEH, R. Y., Y. L. SHIU, S. C. SHEI, S. C. CHENG, S. Y. HUANG, J. C. LIN and C. H. LIU (2009):<br />

Evaluation <strong>of</strong> the antibacterial activity <strong>of</strong> leaf and twig extracts <strong>of</strong> stout camphor tree, Cinnamomun kanehirae,<br />

and the effects on immunity and disease resistance <strong>of</strong> white shrimp, Litopenaeus vannamei.<br />

Fish Shellfish Immun 27, 26-32<br />

YOUNG, M. B. and G. M. YOUNG (2002):<br />

YplA Is Exported by the Ysc, Ysa, and Flagellar Type III Secretion Systems <strong>of</strong> Yersinia enterocolitica.<br />

J Bacteriol 184, 1324-1334<br />

YUGUEROS, J., A. TEMPRANO, J. M. LUENGO and G. NAHARRO (2001):<br />

Molecular cloning <strong>of</strong> Yersinia ruckeri aroA gene: a useful taxonomic tool.<br />

J Fish Dis 24, 383-390<br />

ZINK, S. D., L. PEDERSEN, N. P. CIANCIOTTO and Y. A. KWAIK (2002):<br />

The Dot/Icm type IV secretion system <strong>of</strong> Legionella pneumophila is essential for the induction <strong>of</strong> apoptosis in<br />

human macrophages.<br />

Infect Immun 70, 1657-1663<br />

33


Chapter 2.<br />

Biochemical and Molecular Heterogeneity among<br />

Isolates <strong>of</strong> Yersinia ruckeri from Rainbow Trout<br />

(Oncorhynchus mykiss, Walbaum) in North West<br />

Germany<br />

Yidan Huang, Martin Runge, Geovana Brenner Michael,<br />

Stefan Schwarz, Arne Jung, Dieter Steinhagen<br />

BMC <strong>Veterinary</strong> Research, 2013, 9:215<br />

35


The extent <strong>of</strong> Yidan Huang‟s contribution to the article is evaluated according to the<br />

following scale:<br />

A. has contributed to collaboration (0-33%).<br />

B. has contributed significantly (34-66%).<br />

C. has essentially performed this study independently (67-100%).<br />

1. Design <strong>of</strong> the project including design <strong>of</strong> individual experiments: B<br />

2. Performing <strong>of</strong> the experimental part <strong>of</strong> the study: C<br />

3. Analysis <strong>of</strong> the experiments: B<br />

4. Presentation and discussion <strong>of</strong> the study in article form: C<br />

36


Chapter 2 Biochemical and Molecular Heterogeneity <strong>of</strong> Y. ruckeri<br />

37


Chapter 2 Biochemical and Molecular Heterogeneity <strong>of</strong> Y. ruckeri<br />

38


Chapter 2 Biochemical and Molecular Heterogeneity <strong>of</strong> Y. ruckeri<br />

39


Chapter 2 Biochemical and Molecular Heterogeneity <strong>of</strong> Y. ruckeri<br />

40


Chapter 2 Biochemical and Molecular Heterogeneity <strong>of</strong> Y. ruckeri<br />

41


Chapter 2 Biochemical and Molecular Heterogeneity <strong>of</strong> Y. ruckeri<br />

42


Chapter 2 Biochemical and Molecular Heterogeneity <strong>of</strong> Y. ruckeri<br />

43


Chapter 2 Biochemical and Molecular Heterogeneity <strong>of</strong> Y. ruckeri<br />

44


Chapter 2 Biochemical and Molecular Heterogeneity <strong>of</strong> Y. ruckeri<br />

45


Chapter 3.<br />

Field study on Yersinia ruckeri isolates from trout<br />

hatcheries in North West Germany<br />

Yidan Huang, Arne Jung, Geovana Brenner Micheal,<br />

Martin Runge, Stefan Schwarz, Dieter Steinhagen<br />

47


The extent <strong>of</strong> Yidan Huang‟s contribution to the article is evaluated according to the<br />

following scale:<br />

A. has contributed to collaboration (0-33%).<br />

B. has contributed significantly (34-66%).<br />

C. has essentially performed this study independently (67-100%).<br />

1. Design <strong>of</strong> the project including design <strong>of</strong> individual experiments: B<br />

2. Performing <strong>of</strong> the experimental part <strong>of</strong> the study: B<br />

3. Analysis <strong>of</strong> the experiments: C<br />

4. Presentation and discussion <strong>of</strong> the study in article form: C<br />

48


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

Field study on Yersinia ruckeri isolates from trout hatcheries in North<br />

West Germany<br />

Yidan Huang 1 , Arne Jung 2 , Geovana Brenner Michael 3 , Martin Runge 4 , Stefan Schwarz 3 , Dieter<br />

Steinhagen 1<br />

1 Fish Disease Research Unit, <strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong>, Germany<br />

2 Clinic for Poultry, <strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong>, Foundation, <strong>Hannover</strong>, Germany<br />

3 <strong>Institute</strong> <strong>of</strong> Farm Animal Genetics, Friedrich-Loeffler-Institut (FLI), Neustadt-Mariensee, Germany<br />

4 Lower Saxony State Office for Consumer Protection and Food Safety (LAVES), Food and <strong>Veterinary</strong> <strong>Institute</strong> Braunschweig/<strong>Hannover</strong>, Germany<br />

ABSTRACT: Enteric Redmouth Disease (ERM), caused by Yersinia ruckeri, is one <strong>of</strong> the most<br />

important infectious diseases in rainbow trout (Oncorhynchus mykiss) aquaculture in Europe. Our<br />

aim was to analyse the distribution <strong>of</strong> Y. ruckeri in North West Germany, as well as the variation <strong>of</strong><br />

their biochemical and molecular characteristics as a basis for strain differentiation. During different<br />

seasons from June 2011 until June 2012, samples were taken from different rainbow trout hatcheries.<br />

12 rainbow trout hatcheries were visited during the sampling period. Y. ruckeri was isolated<br />

frequently from rainbow trouts <strong>of</strong> about 21-30cm length. 48 isolates were collected during this field<br />

study in North West Germany and 33 strains were obtained from LAVES, <strong>Hannover</strong> and LHL<br />

Hessen strain collection. 92% <strong>of</strong> the isolates from the field study were non-motile. Moreover, during<br />

the sampling period in winter and early spring, all the isolates recovered from the field were<br />

non-motile. In several trout farms, more than two isolates with different genetic features were present.<br />

Isolates from one designated typing group were obtained in field study during every sampling<br />

campaign throughout the year, while motile isolates from another typing group were present only<br />

during summer and autumn samplings. Our results will provide data on identifying and<br />

characterizing pathogenic Y. ruckeri strains and examining the distribution <strong>of</strong> these strains in the<br />

field as a basis for preventive disease monitoring plans.<br />

KEY WORDS: Epidemiology, Sampling, Distribution, Variation, Enteric redmouth disease<br />

INTRODUCTION<br />

Enteric Redmouth Disease (ERM) was reported in Idaho, USA, as early as 1950s (Rucker<br />

1966). Subsequently, in the late 1970s to the early 1980s, it was first introduced to Europe from<br />

the USA (Horne & Barnes 1999). So far, most <strong>of</strong> European countries have reported ERM cases<br />

49


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

in aquaculture (Lesel & Lesel 1983, Valtonen et al. 1992, Oraić et al. 2002), ranging from<br />

different hosts (Valtonen et al. 1992, Danley et al. 1999, Noga 2000), but rainbow trout<br />

(Oncorhynchus mykiss) is especially susceptible. Today ERM is one <strong>of</strong> the important diseases<br />

<strong>of</strong> aquatic animals. Since Yersinia ruckeri is the causative pathogen <strong>of</strong> ERM, many biochemical<br />

and molecular methods are applied to detect and identify this bacteria (Gibello et al. 1999,<br />

Altinok et al. 2001, Saleh et al. 2008, Glenn et al. 2011), so that the protection and treatment <strong>of</strong><br />

ERM can be applied as quickly as possible.<br />

In order to control ERM, commercial vaccines made from inactivated whole cells <strong>of</strong> Y.<br />

ruckeri were developed (Ross et al. 1966), and the vaccines are considered to be successful<br />

(Tebbit et al. 1981). However, the usage <strong>of</strong> vaccines and the treatment with antibiotics may<br />

cause a selective pressure on Y. ruckeri and it could be the reason for recent outbreaks caused by<br />

non-motile strains <strong>of</strong> Y. ruckeri, which seems to be not sensitive to several vaccines made for<br />

motile Y. ruckeri (Austin et al. 2003, Fouz et al. 2006). In 1994, Klein et al. (Klein et al. 1994)<br />

reported the first isolation <strong>of</strong> a non-motile strain <strong>of</strong> Y. ruckeri in Germany. Furthermore, the<br />

outbreak <strong>of</strong> ERM was not any more constrained by lower temperature. Outbreaks were<br />

observed even in winter when the temperature is lower than 10°C. Therefore, an<br />

epidemiological study <strong>of</strong> Y. ruckeri in North West Germany was performed, in order to provide<br />

data on the diversity <strong>of</strong> Y. ruckeri in this region.<br />

MATERIALS AND METHODS<br />

Sample collection and bacteria isolation<br />

Rainbow trouts (Oncorhynchus mykiss) were sampled from keeping units <strong>of</strong> 12 different trout<br />

farms in the German federal state North Rhine-Westphalia (NRW) during the four seasons <strong>of</strong><br />

2011-2012. Farms ranged from concrete tanks to flow-through units. The freshwater was supplied<br />

from different rivers, including the River Spring, River Rur, River Lambach and others, with water<br />

temperature varying from 4℃ in winter to 18℃ in summer during sampling time. From the ponds,<br />

fish with clinical signs indicating a bacterial infection were selected. In the farms MU and MS,<br />

where diseased fish were not present, fish were randomly sampled from ponds. An overview <strong>of</strong><br />

50


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

visited farms, sampled keeping units and collected trout is given in Table 1. Y. ruckeri isolates were<br />

obtained from liver and kidney <strong>of</strong> sampled trout.<br />

Other isolates used in this study were obtained from the strain collection <strong>of</strong> LAVES, Food and<br />

<strong>Veterinary</strong> <strong>Institute</strong> Braunschweig/<strong>Hannover</strong>, and the fish disease service <strong>of</strong> Hessen at Giessen.<br />

Samples from LAVES, <strong>Hannover</strong>, originated mainly from the federal state <strong>of</strong> Lower Saxony and<br />

were collected in 2004 to 2009 from fish with clinical signs <strong>of</strong> disease. The isolates from Hessen<br />

were isolated from 2010 to 2011 from diseased rainbow trout without detailed data about the original<br />

geographic sites. The isolates were kept at -80℃.<br />

Identification <strong>of</strong> Isolates<br />

All isolates were examined for the Gram-staining and cytochrome oxidase reaction, and were<br />

biochemically characterized by API 20E tests (Biomerieux, France). Five isolates were identified by<br />

both the API 20E system and 16s rDNA sequencing.<br />

Typing <strong>of</strong> Isolates<br />

For typing, API 20E pr<strong>of</strong>iles <strong>of</strong> the isolates were recorded, four repetitive sequence-based PCRs<br />

were performed, including BOX-A1R-based repetitive extragenic palindromic-PCR (BOX-PCR),<br />

(GTG) 5 -PCR, enterobacterial repetitive intergenic consensus (ERIC-PCR) and repetitive extragenic<br />

palindromic (REP-PCR) and pulsed-field electrophoresis was performed as described previously<br />

(Huang et al. 2013b).<br />

RESULTS<br />

Bacteria isolates<br />

In total 81 Y. ruckeri isolates were obtained, including 48 isolates from the field study performed<br />

in trout hatcheries <strong>of</strong> NRW, 33 isolates from LAVES, Food and <strong>Veterinary</strong> <strong>Institute</strong><br />

Braunschweig/<strong>Hannover</strong> and the fish disease service <strong>of</strong> Hessen at Giessen. One non-motile strain<br />

isolated also in NRW in 2008 was <strong>of</strong>fered by Dr. Gould from MSD Animal Heath and a reference<br />

strain DSM 18506 was included. The isolates were identified as Y. ruckeri by API 20E. Five isolates<br />

were identified by 16S rDNA sequencing, additionally. In Fig. 1, the geographic origins <strong>of</strong> isolates<br />

were marked in red; the locations <strong>of</strong> the different sample collection sites in NRW were labelled with<br />

the acronyms for the different fish farms (Fig.1). No detailed information was received about the<br />

geographic origins <strong>of</strong> isolates from Hessen.<br />

51


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

Sample collection in North Rhine-Westphalia<br />

During the different seasons from June 2011 until June 2012, 12 rainbow trout hatcheries were<br />

visited and rainbow trout were sampled from different 91 keeping units including ponds and<br />

raceways. Y. ruckeri could be isolated from 39 keeping units from 9 trout hatcheries. Rainbow trout<br />

collected from 3 hatcheries were not infected by Y. ruckeri (see table 2). In infected ponds, the<br />

prevalence <strong>of</strong> the infection ranged from 100% (12 infected from 12 individuals examined) to 12% (3<br />

infected from 25 examined individuals). During the whole sampling period two farms (SR and P)<br />

had high positive rates <strong>of</strong> Y. ruckeri infections. Y. ruckeri could be isolated from 46 out <strong>of</strong> 76 trouts<br />

(60.5%), and 30 out <strong>of</strong> 46 trouts examined (65.2 %), respectively (Table 3), while three other farms<br />

(FR, MU and MS) were not harbouring infected rainbow trouts. However, all farms except FR, MU<br />

and MS had experienced ERM outbreaks before.<br />

During the different seasons <strong>of</strong> the sampling period, the overall prevalence <strong>of</strong> Y. ruckeri isolation<br />

ranged from 61.9 % (60 out <strong>of</strong> 97 trout examined) in June 2011 and 17.3 % (14 out <strong>of</strong> 81 trout<br />

examined) in July 2012. In ponds sampled in February, an overall prevalence <strong>of</strong> Y .ruckeri <strong>of</strong> 25.0 %<br />

(15 out <strong>of</strong> 60 individuals sampled) was recorded, while in September and April the bacterium was<br />

detected in 31.9 % (15 out <strong>of</strong> 47) and 31.6 % (12 out <strong>of</strong> 38) <strong>of</strong> the rainbow trout sampled (Table 3).<br />

The outbreak peak <strong>of</strong> ERM was mainly in summer and early autumn (Table 3). In winter and early<br />

spring, the water temperatures were below 10 °C , however, positive isolates could also be<br />

recovered from some fish farms, and the isolates were proved non-motile in further tests.<br />

Y. ruckeri was detected in rainbow trout from all size groups examined, from smaller than 5 cm up<br />

to 36 cm body length. However, most infections were recorded from trout in the size range <strong>of</strong> 21-25<br />

cm and 26-30 cm body length (Fig. 2).<br />

Distribution <strong>of</strong> non-motile strains<br />

About 60% <strong>of</strong> the isolates collected from Lower Saxony and Hessen were non-motile; while over<br />

90% <strong>of</strong> the isolates collected during the field study performed in trout hatcheries in NRW were<br />

non-motile (Fig. 3). During the field study, non-motile strains were dominant during the whole<br />

sampling year, but were especially isolated in the cold seasons, in February and April, when water<br />

temperature was below 10 °C (Fig.4).<br />

52


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

Variation <strong>of</strong> Y. ruckeri<br />

According to the results <strong>of</strong> API 20E, PFGE and the repetitive sequence-based PCR assays, the<br />

isolates could be divided into 27 different typing groups (Huang et al. 2013b) (Table 4). Most <strong>of</strong> the<br />

isolates belonged to tp2. In 5 fish farms, isolates <strong>of</strong> at least two different typing groups were present;<br />

in farm SR isolates from six different typing groups were present, in farm B isolates from five, and in<br />

farm KP isolates from four typing groups (Table 4). Farms where isolate tp 2 and tp8 were present<br />

had a selling and purchasing relationship with each other. However, no correlations were found<br />

among the farms where tp20 was present. In 4 farms, Y. ruckeri from a single typing group were<br />

isolated (Table 4). In different months, isolates from three or more different typing groups were<br />

recovered (Table 4). In the same fish farm, isolates from two or more typing groups could be<br />

recovered at the same time; and isolates from typing group 20 were obtained in samples collected at<br />

every sampling time (Table 4). Additionally, eight typing groups were found only in Lower Saxony,<br />

two typing groups were found only in Hessen, and 10 typing groups only in NRW. Two typing<br />

groups (Tp2 and Tp6) were distributed in all federal states investigated.<br />

DISCUSSIONS<br />

Since Y. ruckeri isolates was reported for the first time in the USA (Rucker 1966) and then<br />

introduced to Europe (Horne & Barnes 1999), it has become an important infectious disease in trout<br />

hatcheries (Busch 1978, Fuhrmann & Boehm 1983a, Roberts 1983). In infected rainbow trout<br />

populations, up to 25 % <strong>of</strong> the fish might carry the bacterium (Busch 1978). The bacterium was also<br />

recorded from bi<strong>of</strong>ilms <strong>of</strong> tanks in trout farms (Coquet et al. 2002) which could serve as a source for<br />

disease outbreaks. In the current study, rainbow trout from ponds in 12 trout hatcheries located in<br />

NRW, North West Germany, were examined for the presence <strong>of</strong> Y. ruckeri. From these nine farms<br />

had repeated outbreaks with ERM over several years, while 3 farms had no history with this<br />

infection. In the current study, Y. ruckeri was present in rainbow trout from all farms, which had<br />

previous ERM outbreaks, but not in the farms which previously were free <strong>of</strong> ERM. Overall, 82 Y.<br />

ruckeri isolates were recorded from the farms in the region <strong>of</strong> North West Germany, which according<br />

to biochemical and molecular identification methods could be divided into 26 different typing groups<br />

(Huang et al. 2013b). In the majority <strong>of</strong> the farms, which were repeatedly sampled, bacteria from<br />

53


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

different typing groups were found, and on several occasions, bacteria <strong>of</strong> different typing groups<br />

were simultaneously present in rainbow trout from the same farm.<br />

In previous studies, mainly motile strains <strong>of</strong> Y. ruckeri were found associated with clinical<br />

infections <strong>of</strong> ERM in rainbow trout (Meier 1986), but non-motile isolates were increasingly<br />

recognized as causative agents for disease outbreaks (Giorgetti et al. 1985, Fouz et al. 2006). In the<br />

current study, isolates from most trout farms were non-motile, which might indicate that after the<br />

first detection <strong>of</strong> a non-motile isolate in Germany in 1994 (Klein et al. 1994), non-motile bacteria<br />

were distributed over several farms in North West Germany. In particular bacteria from the typing<br />

groups 2 (which were found most frequently) and 8 were isolated from 5 farms. These farms had a<br />

trading relationship between each other, which could facilitate the distribution <strong>of</strong> particular types <strong>of</strong><br />

the pathogens. In addition, bacteria from typing group 20 were found in several farms, which<br />

however had no obvious trading relation. It is accepted that Y. ruckeri outbreaks in farms can<br />

originate from latently infected rainbow trout (Rodgers 1992), or from environmental samples<br />

(Romalde et al. 1994, Coquet et al. 2002) but its transmission was also related to wild fish or aquatic<br />

invertebrates as putative vectors (McDanniel 1971, Fuhrmann & Boehm 1983b, Willumsen 1989). A<br />

transmission by birds could also be possible (Bangert et al. 1988). Thus the distribution <strong>of</strong> Y. ruckeri<br />

by such possible vectors has to be considered, but in the current study, those farms, which had no<br />

previous history with Y. ruckeri infection, did not experience an ERM outbreak throughout the<br />

observation period. This might emphasise the importance <strong>of</strong> latent infections with Y. ruckeri in<br />

farmed carrier fish or survival in bi<strong>of</strong>ilms <strong>of</strong> the farm for the spreading <strong>of</strong> the disease. This view<br />

could be supported by the findings <strong>of</strong> the present study. In farms with repeated outbreaks <strong>of</strong> ERM, Y.<br />

ruckeri from the same or from genetically similar typing groups (Huang et al. 2013b) were isolated.<br />

Outbreaks <strong>of</strong> ERM were usually associated with challenging environmental conditions, such as<br />

poor water quality, excessive stocking densities, and high water temperature (Horne & Barnes 1999).<br />

In this study, outbreaks were mainly observed in summer and early autumn, at a water temperature<br />

between 10 °C and 18 °C. In particular outbreaks caused by the motile bacteria from typing group 16<br />

were observed during this period <strong>of</strong> time, supporting the hypothesis that motile strains are more<br />

active during warmer seasons (Huang et al. 2013a). In the farms, rainbow trout ranging from 5 cm to<br />

54


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

36 cm body length were found to be infected, but ongrowing rainbow trout at a body length <strong>of</strong> 21cm<br />

to 30cm were mostly affected. Fish under the length <strong>of</strong> 5cm could also be infected. Kawula et al used<br />

Y. ruckeri to infect platy fish and found that this small fry died without clinical signs (Kawula et al.<br />

1996). From June 2011 to July 2011, the rate <strong>of</strong> Y. ruckeri positive rainbow trout decreased from<br />

61.2% to 17.3%, as a result from the usage <strong>of</strong> vaccines and antibiotic treatments. However, these<br />

measures could not provide 100% <strong>of</strong> protection, and that may be why the positive rate increased<br />

back to 31.9% in September 2011. In some ponds, outbreaks <strong>of</strong> ERM, were even observed during<br />

February and April, when the water temperature was below 10°C. These outbreaks were associated<br />

with non-motile Y. ruckeri strains, which previously were found to be more active at lower<br />

temperature (Huang et al. 2013a). Wheeler et al. (Wheeler et al. 2009) discovered that Y. ruckeri was<br />

introduced separately into the UK and Europe from the USA and these isolates from the UK and<br />

mainland Europe may represent different clonal groups. In this study, each federal state in North<br />

West Germany had its own array <strong>of</strong> typing groups <strong>of</strong> Y. ruckeri, however, two typing groups could<br />

be found in farms from all three states. In a previous study (Huang et al. 2013b), 35.4% <strong>of</strong> 82 isolates<br />

were associated with these two groups. This result suggested the bacteria might share a common<br />

origin but subsequently developed independently in different geographic regions.<br />

CONCLUSIONS<br />

Y. ruckeri could be isolated from rainbow trout all around the year. Non-motile strains were<br />

dominant in North West Germany, especially in winter and early spring when water temperature was<br />

below 10 °C. Our results will provide data on identifying and characterizing pathogenic Y. ruckeri<br />

strains and examining the distribution <strong>of</strong> these strains in the field as a basis for preventive disease<br />

monitoring plans.<br />

ACKNOWLEDGEMENTS<br />

We thank Dr. S. Braune (LAVES, <strong>Hannover</strong>), Dr. C. Gould (MSD Animal Health) and Dr. A Nilz<br />

(LHL Hessen) for the generous provision <strong>of</strong> isolates that used in this study. Dr. W. Schäfer and D.<br />

Mock (LANUV, Albaum) gave support during the field sampling. Dr. G. Brenner Michael, R.<br />

Becker (FLI, Mariensee) and S. Baumann (LAVES) provided valuable technical assistance. This work<br />

was supported by Landesamt für Natur, Umwelt und Verbraucherschutz Nordrhein-Westfalen (LANUV).<br />

55


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

REFERENCES<br />

Altinok I, Grizzle JM, Liu Z (2001) Detection <strong>of</strong> Yersinia ruckeri in rainbow trout blood by use <strong>of</strong> the<br />

polymerase chain reaction. Dis Aquat Organ 44:29-34<br />

Austin DA, Robertson PAW, Austin B (2003) Recovery <strong>of</strong> a New Biogroup <strong>of</strong> Yersinia ruckeri from<br />

Diseased Rainbow Trout(Oncorhynchus mykiss, Walbaum). Systematic and Applied Microbiology<br />

Bangert 26:127-131 RL, Ward AC, Stauber EH, Cho BR, Widders PR (1988) A survey <strong>of</strong> the aerobic bacteria in the feces<br />

<strong>of</strong> captive raptors. Avian Dis 32:53-62<br />

Busch RA (1978) Enteric redmouth disease(Hagerman strain). Marine Fisheries Review 40:42-51<br />

Coquet L, Cosette P, Quillet L, Petit F, Junter GA, Jouenne T (2002) Occurrence and phenotypic<br />

charaterization <strong>of</strong> Yersinia ruckeri strains with bi<strong>of</strong>ilm-forming capacity in a rainbow trout Farm.<br />

Appl Environ Microbiol 68:470-475<br />

Danley ML, Goodwin AE, Killian HS (1999) Epizootics in farm-raised channel catfish, Ictalurus<br />

punctatus(Rafinesque), caused by the enteric redmouth bacterium Yersinia ruckeri. Journal <strong>of</strong> Fish<br />

Diseases 22:451-456<br />

Fouz B, Zarza C, Amaro C (2006) First description <strong>of</strong> non-motile Yerinia ruckeri serovar I strains causing<br />

disease in rainbow trout, Oncorhynchus mykiss (Walbaum), cultured in Spain. Journal <strong>of</strong> Fish Disease<br />

29:339-346<br />

Fuhrmann H, Boehm KH (1983a) An outbreak <strong>of</strong> enteric redmouth disease in West Germany. J Fish Dis<br />

6:309-311<br />

Fuhrmann H, Boehm KH (1983b) An outbreak <strong>of</strong> enteric redmouth disease in West Germany. Journal <strong>of</strong> Fish<br />

Diseases 6:309-311<br />

Gibello A, Blanco MM, Moreno MA, Cutuli MT, Domenech A, Moniguez L, Fernandez-Garayzabal JF (1999)<br />

Development <strong>of</strong> a PCR Assay for Detection <strong>of</strong> Yersinia ruckeri in Tissues <strong>of</strong> Inoculated and Naturally<br />

Infected Trout. Applied and Enviromental Microbiology 65:346-350<br />

Giorgetti G, Ceschia G, Bovo G (1985) First isolation <strong>of</strong> Yersinia ruckeri in farmed rainbow trout in Italy. In:<br />

Ellis AE (ed) Fish and shellfish pathology Proceedings <strong>of</strong> the European Association <strong>of</strong> Fish<br />

Pathologists, Plymouth<br />

Glenn RA, Taylor PW, Hanson KC (2011) The use <strong>of</strong> a real-time PCR primer/probe set to observe infectivity<br />

<strong>of</strong> Yersinia ruckeri in Chinook salmon, Oncorhynchus tshawytscha (Walbaum), and steelhead trout,<br />

Oncorhynchus mykiss (Walbaum). Journal <strong>of</strong> Fish Disease 34:783-791<br />

Horne MT, Barnes AC (1999) Enteric redmouth disease (Yersinia ruckeri). In: Woo PTK, Bruno DW (eds)<br />

Fish diseases and disorders, Book 3. CABI Publishing, Wallingford<br />

Huang Y, Adamek M, Walker C, Runge M, Steinhagen D (2013a) In vitro cytotoxicity and multiplex PCR<br />

detection <strong>of</strong> virulence factors <strong>of</strong> Yersinia ruckeri isolated from rainbow trout in North West Germany.<br />

Berliner Münchener Tierärztliche Wochenschrift (submitted for publicaiton, Chapter 5)<br />

Huang Y, Runge M, Michael GB, Schwarz S, Jung A, Steinhagen D (2013b) Biochemical and Molecular<br />

Heterogeneity among Isolates <strong>of</strong> Yersinia ruckeri from Rainbow Trout (Oncorhynchus mykiss,<br />

Walbaum) in North West Germany. BMC <strong>Veterinary</strong> Research 9:215<br />

Kawula TH, Lelivelt MJ, Orndorff PE (1996) Using a new inbred fish model and cultured fish tissue cells to<br />

study Aeromonas hydrophila and Yersinia ruckeri pathogenesis. Microbial Pathogenesis 20:119-125<br />

Klein BU, Kleingeld DW, Bohm KH (1994) First isolation <strong>of</strong> a non-motile/tween 80 negative Yersinia ruckeri<br />

strain in Germany. Bulletin <strong>of</strong> the European Association <strong>of</strong> Fish Pathologists 14:165-166<br />

56


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

Lesel R, Lesel M (1983) Outbreak <strong>of</strong> enteric redmouth disease in rainbow trout, Salmo gairdneri Richardson,<br />

in France. Journal <strong>of</strong> Fish Diseases 6:385-387<br />

McDanniel DW (1971) Hagerman redmouth. American Fishes US Trout News 15:14-28<br />

Meier W (1986) Enteric redmouth disease: outbreak in rainbow trout in Switzerland. Diseases <strong>of</strong> Aquatic<br />

Organisms 2:81-82<br />

Noga EJ (2000) Enteric Redmouth Disease(ERM, Redmouth, Yersiniosis, Blood Spot, Yersinia Ruckeri<br />

Infection), Vol. Iowa State <strong>University</strong> Press, Ames<br />

Oraić D, Zrnčić S, Šoštarić B, Bažulić D, Lipej Z (2002) Occurrence <strong>of</strong> enteric redmouth disease in rainbow<br />

trout(Oncorhynchus Mykiss) on Farms in Croatia. Acta Veterinaria Hungarica 50:283-291<br />

Roberts MS (1983) A report <strong>of</strong> an epizootic in hatchery-reared rainbow trout, Salmo gairdneri Richardson, at<br />

an English trout farm, caused by Yersinia ruckeri. Journal <strong>of</strong> Fish Diseases 6:551-552<br />

Rodgers CJ (1992) Development <strong>of</strong> a selective-differential medium for the isolation <strong>of</strong> Yersinia ruckeri and its<br />

application in epidemiological studies. Journal <strong>of</strong> Fish Disease 15:243-254<br />

Romalde JL, Magarinos B, Pazos F, Silva A, Toranzo AE (1994) Incidence <strong>of</strong> Yersinia ruckeri in two farms in<br />

Galicia (NW Spain) during a one-year period. Journal <strong>of</strong> Fish Disease 17:533-539<br />

Ross AJ, Rucker RR, Ewing WH (1966) Description <strong>of</strong> a bacterium associated with redmouth disease <strong>of</strong><br />

rainbow trout (Salmo Gairdneri). Can J Microbiol 12:763-770<br />

Rucker RR (1966) Redmouth disease <strong>of</strong> rainbow trout (Salmo gairdneri). Bull <strong>of</strong>f int Epizoot 65:825-830<br />

Saleh M, Soliman H, El-Matbouli M (2008) Loop-mediated isothermal amplification as an emerging<br />

technology for detection <strong>of</strong> Yersinia ruckeri the causative agent <strong>of</strong> enteric red mouth disease in fish.<br />

BMC <strong>Veterinary</strong> Research 4<br />

Tebbit GL, Erickson JD, Water RBvd (1981) Development and use <strong>of</strong> Yersinia ruckeri bacterins to control<br />

enteric redmouth disease. In: International symposium on fish biologies: serodiagnostics and vaccines,<br />

Book 49. Dev Biol Stand<br />

Valtonen ET, Rintamaki P, Koskivaara M (1992) Occurrence and pathogenicity <strong>of</strong> Yersinia ruckeri at fish<br />

farms in northern and central Finland. J Fish Dis 15:163-171<br />

Wheeler RW, Davies RL, Dalsgaard I, Garcia J, Welch TJ, Wagley S, Bateman KS, Verner-Jeffreys DW<br />

(2009) Yersinia ruckeri biotype 2 isolates from mainland Europe and the UK represent different<br />

clonal groups. Diseases <strong>of</strong> Aquatic Organisms 84:25-33<br />

Willumsen B (1989) Birds and wild fish as potential vectors <strong>of</strong> Yersinia ruckeri. Journal <strong>of</strong> Fish Disease<br />

12:275-277<br />

57


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

Fig.1 Geographic origins <strong>of</strong> isolates<br />

Table 1 Overview fish species as host for Yersinia ruckeri in fish farms in North West Germany and year <strong>of</strong> isolation<br />

(82 isolates without the reference strain)<br />

Fish Species<br />

Sampling year<br />

2004 2005 2006 2007 2008 2009 2010 2011 2012 unknown Total<br />

Rainbow trout 3 6 7 1 2 7 33 16 75<br />

Brown trout 1 1<br />

Koi 1 1<br />

Pike 1 1<br />

Trout 1 1 2<br />

Not specified 1 1 2<br />

58


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

Table 2 Overview <strong>of</strong> rainbow trout farms, rearing units and rainbow trout populations sampled in North Rhine-Westphalia, Germany<br />

Given is the body length <strong>of</strong> fish (cm).<br />

Fish Farms Unit June<br />

2011<br />

Unit July 2011 Unit September<br />

2011<br />

Unit February<br />

2012<br />

Unit April<br />

2012<br />

Unit<br />

June<br />

2012<br />

SR H 28-30 + T 6 18-22 ON 11-13 T 23 22-24 T 23 12-15<br />

T 7 25-28 T 3 25-30 T10 26-30 T 4 23-25 T 6 21-23<br />

T 11 18-20 FT 25-28 T 2 22-24 T 8 10-12<br />

T 21 18-20 T14 18-22<br />

K1 28-30<br />

L 1 25-28 T 4 12-16 T 2 10-12 20-22<br />

T 10 20 T 3 13<br />

T 13 22-26 T 12 19-22<br />

FR 22-25 T 5 15-18 23-25<br />

T 13 18-20<br />

KP T 1 28 GT 21-26 T 1 22-24 L 3 30-32 B 2-4<br />

T 6 24-26 ST 32-35 T 3 23-25 K 25-27 T 3 6-8<br />

L 3 28-30 L 3 26-28 2-3 T 2 2-3<br />

L 2 25-27 L 1 25<br />

5-9 6-8<br />

H 30<br />

F 28-30 31-35<br />

P T 1 15-20 T 25-30 T 15-18 PZ2 28-30 T 12-15 TZ3 10-16<br />

T 2 22-24 T 2 28 T W 29-30 25 T L 29-31<br />

T 3 28-30 24-30 B3 34-36<br />

59


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

B 20-22 18-22 18-26 16-20 12-28 16-22<br />

7-9<br />

MU 1 45 B6 24-26 T 30-36<br />

2 20 H 31-33<br />

3 31-32<br />

4 6-7<br />

MS D3 26 B 10-12<br />

A3 7-8 TA1 11-13<br />

H 5-6 B3 12-14<br />

N N4 28-30<br />

N11 23-25<br />

T 23 19-21<br />

A +<br />

S 8-10 22-24<br />

Total 1 13/16 7/21 6/17 3/10 3/9 7/18<br />

+ No detailed information about the fish size <strong>of</strong> positive samples.<br />

1 Positive population/total populations sampled<br />

Size range in Bold indicated the positive samples found in the size group.<br />

Table 3 Rates <strong>of</strong> Yersinia ruckeri positive rainbow trout in different farms from North Rhine-Westphalia, Germany during the sampling time<br />

Fish Farms<br />

06.2011<br />

12-15°C<br />

07.2011<br />

15-18°C<br />

09.2011<br />

10-15°C<br />

02.2012<br />

4-5°C<br />

04.2012<br />

6-8°C<br />

06.2012<br />

8-12°C<br />

Total<br />

60


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

SR 81.0%(17/21) × 1 33.3%(3/9) 50.0%(7/14) 56.3%(9/16) 62.5%(10/16) 60.5%(46/76)<br />

L 57.1%(8/14) 22.2%(2/9) 33.3%(2/6) 0(0/3) 37.5%(12/32)<br />

FR 0%(0/17) 0%(0/10) 0%(0/10) 0%(0/37)<br />

KP 60.9%(14/23) 12.0%(3/25) 0%(0/6) 17.7%(3/17) 0%(0/21) 1 21.7%(20/92)<br />

F 0%(0/3) 60.0%(3/5) 37.5%(3/8)<br />

P 100%(12/12) 100%(4/4) 55.6%(5/9) 0%(0/4) 0%(0/5) 75%(9/12) 65.2% (30/46)<br />

B 81.8%(9/11) 62.5%(5/8) 37.5%(3/8) 41.7%(5/12) 33.3%(3/9) 0%(0/11) 42.4%(25/59)<br />

MU 0%(0/8) 0%(0/8) 0%(0/5) 0%(0/21)<br />

MS 0%(0/17) 0%(0/32) 0%(0/49)<br />

N 33.3%(2/6) 33.3%(2/6)<br />

A × 1<br />

S 0%(0/3) × 1 0%(0/3)<br />

Total 61.9%(60/97) 17.3%(14/81) 31.9%(15/47) 25.0%(15/60) 31.6%(12/38) 21.0%(22/102) 31.9%(137/429)<br />

1 Positive isolates were obtained before the sampling period.<br />

Table 4 Different genetic groups <strong>of</strong> Yersinia ruckeri present in rainbow trout farms in North Rhine-Westphalia, Germany<br />

Fish Farms 06.2011 07.2011 09.2011 02.2012 04.2012 06.2012 Total<br />

SR tp2, tp19,tp20 tp3 1 ,tp20 1 tp2 tp6 tp2 tp2,tp8 tp2,3,6,8,19,20<br />

L tp16 tp16 tp16 - tp16<br />

FR - - - -<br />

KP tp2 tp2 - tp7,tp8 - tp26 1 tp2,7,8,26<br />

F - - tp20 tp20<br />

P tp6,tp15 tp2 tp2,tp22 - - tp2 tp2,6,15,22<br />

61


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

B tp17 tp4 tp20 tp20 tp20 - tp4,17,20<br />

MU - - - -<br />

MS - - -<br />

N tp21 tp21<br />

A tp2 1 ,tp20 1 tp2,20<br />

S - tp8 1 tp8<br />

Total<br />

tp2,6,15,16,17, tp1,2,3,4,16,20 tp2,5,16,20,21,22 tp6,7,8,20 tp2,8,20 tp2,8,20,26<br />

19,20<br />

1 Positive isolates were obtained before the sampling period.<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

No. <strong>of</strong> pos. cases<br />

No. <strong>of</strong> neg. cases<br />

Fig. 2 Infection <strong>of</strong> rainbow trout with Yersinia ruckeri: Distribution <strong>of</strong> positive and negative cases in rainbow trout <strong>of</strong> different sizes<br />

62


Chapter 3 Field study on Y. ruckeri in North West Germany<br />

100%<br />

80%<br />

60%<br />

40%<br />

20%<br />

LS & H (Before<br />

2012)<br />

NRW (2011-<br />

2012)<br />

0%<br />

LS & H (Before<br />

2012)<br />

NRW (2011-<br />

2012)<br />

Fig.3 Distribution <strong>of</strong> non-motile strains in different sample collections. LS: Lower Saxony, H:<br />

Hessen, NRW: North Rhein-Westphalia<br />

Fig.4 Distribution <strong>of</strong> non-motile strains in field study<br />

63


Chapter 4.<br />

Fatty acid composition <strong>of</strong> Yersinia ruckeri isolates from<br />

aquaculture ponds in North West Germany<br />

Yidan Huang, Martin Ryll, Charles Walker, Arne Jung,<br />

Martin Runge, Dieter Steinhagen<br />

Berliner Münchener Tierärztliche Wochenschrift 2014,127 (1/2):10-15<br />

65


The extent <strong>of</strong> Yidan Huang‟s contribution to the article is evaluated according to the<br />

following scale:<br />

A. has contributed to collaboration (0-33%).<br />

B. has contributed significantly (34-66%).<br />

C. has essentially performed this study independently (67-100%).<br />

1. Design <strong>of</strong> the project including design <strong>of</strong> individual experiments: B<br />

2. Performing <strong>of</strong> the experimental part <strong>of</strong> the study: B<br />

3. Analysis <strong>of</strong> the experiments: B<br />

4. Presentation and discussion <strong>of</strong> the study in article form: C<br />

66


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

Short communication<br />

Fish Disease Research Unit, <strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong>, Germany 1<br />

Clinic for Poultry, <strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong>, Germany 2<br />

School <strong>of</strong> Life Science, Keele <strong>University</strong>, Staffordshire, UK 3<br />

Lower Saxony State Office for Consumer Protection and Food Safety (LAVES), Food and<br />

<strong>Veterinary</strong> <strong>Institute</strong> Braunschweig/<strong>Hannover</strong>, Germany 4<br />

Fatty acid composition <strong>of</strong> Yersinia ruckeri isolates from aquaculture ponds in North<br />

West Germany<br />

Fettsäure-Pr<strong>of</strong>ile von Yersinia ruckeri-Isolaten aus Fischteichen in Nord-Westdeutschland<br />

Yidan Huang 1 , Martin Ryll 2 , Charles Walker 1,3 , Arne Jung 2 , Martin Runge 4 , Dieter<br />

Steinhagen 1<br />

Abstract<br />

Enteric Redmouth Disease (ERM), caused by Yersinia (Y.) ruckeri is one <strong>of</strong> the most<br />

important diseases in salmonid aquaculture. Outbreaks <strong>of</strong> ERM were controlled by vaccines<br />

directed against motile strains <strong>of</strong> the bacterium, until recently non-motile vaccine-resistant<br />

strains evolved and caused severe outbreaks. Non-motile isolates were found widespread in<br />

aquaculture populations in North West Germany. In the present study, 82 Y. ruckeri isolates<br />

were isolated from trout hatcheries in North Rhine Westfalia, Lower Saxony and Hessen and<br />

only 20 % <strong>of</strong> them were motile. In order to further characterise the Y. ruckeri isolates from<br />

fish aquaculture populations in North West Germany, the fatty acid compositions <strong>of</strong> 82 Y.<br />

ruckeri field isolates from this area and <strong>of</strong> the Y. ruckeri reference strain DSM18506 were<br />

analysed by gas chromatography. All Y. ruckeri isolates exhibited 15 major fatty acids,<br />

including 12:0, 13:0, 13.957 (equivalent chain length, ECL unknown), 14:0, 14.502 (ECL<br />

unknown), 15:0, 16:1ω5c, 16:0, 17:1ω8c, 17:0 CYCLO, 17:0, 16:1 2OH, 18:1ω9c, 18:1ω7c<br />

and 18:0. From a dendrogram, all isolates were close to one another, clustering together;<br />

while slight differences were detected among the isolates and the reference strain DSM18506.<br />

Compared to their epidemiological and biochemical characteristics, there was no relationship<br />

found between the fatty acid pr<strong>of</strong>iles, API 20E pr<strong>of</strong>iles, motility and geographic distribution.<br />

67


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

Our results show that the fatty acid composition <strong>of</strong> Y. ruckeri isolates from North West<br />

Germany is highly homogenous.<br />

Keywords: Enteritic red mouth disease, rainbow trout, homogeneity, motility, identification<br />

Zusammenfassung<br />

Die von Yersinia ruckeri verursachte Rotmaulseuche (engl. „enteritic red mouth disease“,<br />

ERM) ist eine der wichtigsten Erkrankungen in der Aquakultur von Salmoniden. Ausbrüche<br />

von ERM wurden durch den Einsatz von Vakzinen, die gegen motile Stämme des<br />

Bakteriums gerichtet sind, sicher bekämpft, bis sich vakzin-resistente, unbewegliche Stämme<br />

entwickelten, die dann schwere Krankheitsausbrüche verursachten. Nicht bewegliche Isolate<br />

von Yersinia ruckeri wurden in einigen Forellenpopulationen aus Teichwirtschaften in<br />

Nord-Westdeutschland nachgewiesen. In der vorliegenden Studie wurden Forellen aus<br />

Teichwirtschaften in Nordrhein-Westfalen, Niedersachsen und Hessen auf Yersinia ruckeri<br />

untersucht. Lediglich 20 % der isolierten Y. ruckeri waren beweglich. Um die in der<br />

Feldstudie erhaltenen Isolate weitergehend zu charakterisieren, wurden Fettsäurepr<strong>of</strong>ile von<br />

den 82 Y. ruckeri Isolaten aus Forellenbeständen in Nord-Westdeutschland und vom<br />

Referenzstamm DSM 18506 mittels Gaschromatographie bestimmt. Alle Isolate von Y.<br />

ruckeri wiesen 15 charakteristische Fettsäuren auf, darunter 12:0, 13:0, 13,957 („equivalent<br />

chain lenght“, ECL unbekannt), 14:0, 14,502 (ECL ungekannt), 15:0, 16:1ω5c, 16:0, 17:1ω8c,<br />

17:0 CYCLO, 17:0, 16:1 2OH, 18:1ω9c, 18:1ω7c und 18:0. Der Vergleich der<br />

Fettsäurepr<strong>of</strong>ile in einem Dendrogram zeigte die Ähnlichkeit der Isolate. Die Isolate aus<br />

Nord-Westdeutschland bildeten hier einen homogenen Ast. Davon abgesetzt zeigte sich der<br />

Referenzstamm DSM 18506. Wurden epidemiologische und biochemische Charakteristika<br />

der Isolate mit den Fettsäurepr<strong>of</strong>ilen verglichen, ließ sich keine Beziehung zwischen den<br />

Fettsäuremustern, den API 20E Pr<strong>of</strong>ilen, Beweglichkeit und geographischer Herkunft<br />

erkennen. Unsere Ergebnisse zeigen eine hohe Homogenität im Fettsäuremuster von Yersinia<br />

ruckeri-Isolaten aus Nord-Westdeutschland.<br />

Schlüsselwörter: Rotmaulseuche; Regenbogenforellen, Homogenität, Beweglichkeit,<br />

Identifizierung<br />

Introduction, Material and Methods<br />

68


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

Enteric redmouth Disease (ERM) was reported in Idaho, USA, as early as in the 1950s<br />

(Rucker, 1966). Subsequently, in the late 1970s to the early 1980s, the disease caused by the<br />

enterobacterium Yersinia (Y.) ruckeri was first introduced to Europe from the USA (Horne<br />

and Barnes, 1999). In Rainbow trout the disease is associated with a swollen abdomen, small<br />

haemorrhages in the mouth corners, the gum, palate, tongue (red mouth) and the paired fins.<br />

The gill operculum can turn red and haemorrhages can be seen in the eye (Schlotfeldt et al.,<br />

1995). The disease causes significant economic losses in freshwater aquaculture worldwide<br />

and is one <strong>of</strong> the most important diseases <strong>of</strong> aquatic animals (Austin and Austin, 2007). In<br />

addition to salmonids, Y. ruckeri infects different fish species, including catfish, carp, pike<br />

and even freshwater crayfish (Coquet et al., 2002; Karatas et al., 2004; Lesel and Lesel, 1983;<br />

Oraić et al., 2002; Ross et al., 1966; Valtonen et al., 1992; Wang et al., 2009).<br />

Initially, Y. ruckeri was considered a phenotypically homogenous species (Arias et al., 2007),<br />

but later on in an extensive study on Y. ruckeri isolates from different geographical origins,<br />

approximately 20 % <strong>of</strong> the isolates were found non-motile and negative for Tween hydrolysis<br />

(Davies and Frerichs, 1989). These isolates were assigned as Y. ruckeri biotype 2 strains<br />

(Davies and Frerichs, 1989), and in subsequent years were found to be responsible for<br />

diseases outbreaks in rainbow trout (Oncorhynchus (O.) mykiss), despite them having been<br />

vaccinated with a commercial vaccine against the typical motile biotype 1 strains. After the<br />

first detection <strong>of</strong> these strains in rainbow trout aquaculture in England (Austin et al., 2003),<br />

biotype 2 strains caused ERM outbreaks in Spain (Fouz et al., 2006), the US (Arias et al.,<br />

2007) as well as in Germany.<br />

The most common way to diagnose a Y. ruckeri infection is identified by biochemical tests,<br />

such as API 20E, and, further considering <strong>of</strong> outer membrane protein and lipopolysaccharide<br />

pr<strong>of</strong>iles (Tinsley et al., 2011). In addition to a biochemical identification <strong>of</strong> the bacterium,<br />

further diagnostic methods were developed, including the identification <strong>of</strong> Y. ruckeri<br />

bacteriophages (Stevenson and Airdrie, 1984), DNA microarray technology (Warsen et al.,<br />

2004), species-specific PCR (Seker et al., 2012), or Fourier transform infrared spectroscopy<br />

(FT-IR) (Wortberg et al., 2012).<br />

The determination <strong>of</strong> the whole cell fatty acid composition by gas chromatography (GC) has<br />

been described as a simple way to identify and classify bacteria as early as 1963 (Abel et al.,<br />

1963). Canonica and Pisano (1985) suggested the use <strong>of</strong> gas-liquid chromatographic analysis<br />

for classifying motile Aeromonas strains <strong>of</strong> clinical origin from nonclinical strains according<br />

to their fatty acid methyl ester (FAME) pr<strong>of</strong>iles (Canonica and Pisano, 1985). Leclercq et al.<br />

(2000) found the composition <strong>of</strong> Y. pestis fatty acid was highly uniform (Leclercq et al., 2000),<br />

69


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

and subsequently it was possible to differentiate Y. pestis from various outbreaks <strong>of</strong> plague<br />

from Y. pseudotuberculosis according to characteristic major cellular fatty acid (CFA) pr<strong>of</strong>iles<br />

(Tan et al., 2010). In addition to API 20E, the analysis <strong>of</strong> the composition <strong>of</strong> cellular fatty<br />

acids was also proposed as one <strong>of</strong> the phenotypic characteristics for the identification <strong>of</strong> Y.<br />

ruckeri (Arias et al., 2007).<br />

The current study was undertaken in order to analyse the heterogeneity <strong>of</strong> Y. ruckeri present in<br />

fish populations from aquaculture ponds in North West Germany. In this study, 82 Y. ruckeri<br />

field isolates including both motile and non-motile strains from populations with different<br />

epidemiological characteristics and a reference strain were analysed. In addition to<br />

conventional biochemical testing, cellular fatty acids pr<strong>of</strong>iles were assessed in order to<br />

explore whether differences in biotype or epidemiological characteristics would be correlated<br />

with altered fatty acid composition as an additional phenotypic trait <strong>of</strong> Y. ruckeri.<br />

From the 82 isolates <strong>of</strong> Y. ruckeri used, 48 isolates were collected from Rainbow trout (O.<br />

mykiss) in a field survey, performed during 2011-2012 in trout hatcheries in the German<br />

federal state North-Rhine-Westphalia. Additional 34 isolates were obtained from the strain<br />

collections <strong>of</strong> LAVES Niedersachsen, Food and <strong>Veterinary</strong> <strong>Institute</strong> Braunschweig/<strong>Hannover</strong>,<br />

the fish disease service at Landesbetrieb Hessisches Landeslabor (LHL) in Giessen and <strong>of</strong><br />

MSD Animal Heath, who kindly provided the non-motile isolate G1S1, which was isolated<br />

from a trout hatchery in North-Rhine-Westphalia. These isolates were obtained from different<br />

fishes suffering from disease, mainly from rainbow trout during 2004 and 2009 (Tab. 1). The<br />

reference strain DSM 18506 was <strong>of</strong>fered by the Clinic for Poultry, <strong>University</strong> <strong>of</strong> <strong>Veterinary</strong><br />

<strong>Medicine</strong>, <strong>Hannover</strong>. All strains were previously identified as Y. ruckeri by biochemical<br />

testing using the API 20E test at 25°C for 48h (BioMerieux, France). Six isolates with an<br />

uncertain identification on the basis <strong>of</strong> their API pr<strong>of</strong>iles were confirmed further by using 16S<br />

rDNA gene sequencing. In biochemical identification, some variation was observed among<br />

the isolates, which resulted in 8 different numeric pr<strong>of</strong>iles, with the API 20E pr<strong>of</strong>iles 5107100<br />

(27 isolates), 5306100 (23 isolates) and 5307100 (20 isolates) were most abundant. Motility<br />

was checked by phase-contrast microscopy (1000x). To confirm the motility, bacteria were<br />

stab inoculated with a needle into the bottoms <strong>of</strong> API M medium (bioMérieux).and incubated<br />

at 25°C for 24 h. From the 82 field isolates tested, 16 isolates were motile and able to<br />

hydrolyse Tween 20/80 and thus were assigned as typical biotype I isolates. The general<br />

characteristics <strong>of</strong> the isolates used are shown in Table 1. Additionally, three Y. enterocolitica<br />

reference strains (DSM 11502, DSM 11503, DSM 11504), three Y. enterocolitica field<br />

70


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

isolates, the Y. pseudotuberculosis reference strain DSM 8992 and the Y. ruckeri reference<br />

strain DSM 18506 were included in the analysis.<br />

Bacteria were inoculated on Columbia blood agar using the quadrant streak pattern and grown<br />

for 24 hours at 25 °C (Y. ruckeri) or at 35 °C (Y. enterocolitica and Y. pseudotuberculosis).<br />

Bacteria in quadrant 3 (approx. 40 mg) were harvested. Cellular fatty acids were extracted and<br />

transformed into fatty acid methyl esters (FAMEs) according to the procedures outlined in the<br />

Sherlock microbial identification systems (MIS) operating manual (version 4.0, Microbial<br />

IDentification Inc, Delaware, USA). After the extraction, FAMEs were injected into a HP<br />

5890A gas chromatograph equipped with an automatic injector (injector HP 7673), sample<br />

controller, a gas chromatograph capillary column (Ultra 2, crosslinked 5% phenyl methyl<br />

silicone, 25 m x 0.2mm x 0.11µm film thickness) and a flame ionizations detector (FID).<br />

Fatty acids with up to 20 carbon atoms (C-9 to C-20) were measured in a hydrogen phase. A<br />

calibration mix (Hewlett Packhard) was included in every run as an internal reference. For<br />

data analysis, we used the CLIN40 method (Sherlock MIS Version 4.0, Microbial<br />

IDentification Inc., USA). A library validation report was generated by the MIDI procedure<br />

using the Sherlock MIS operating system (Version 4.0, Microbial IDentification Inc, USA). A<br />

dendrogram for possible biotype tracking was generated by the same system.<br />

Results and Discussion<br />

A phenotypic characterisation <strong>of</strong> Y. ruckeri strains isolated from fish in North West Germany<br />

was carried out by comparing different FAME pr<strong>of</strong>iles. The following 15 fatty acids were<br />

found as most abundant (>3% <strong>of</strong> the total amount) in all isolates: 12:0 (dodecanoic acid), 13:0<br />

(tridecanoic acid), 13.957 (ECL unknown), 14:0 (tetradecanic acid), 14.502 (ECL unknown),<br />

15:0 (pentadecanoic acid), 16:1ω5c, 16:0 (hexadecanoic acid), 17:1ω8c, 17:0 CYCLO<br />

(methylenhexadecanoic acid), 17:0 (heptadecanoic acid), 16:1 2OH, 18:1ω9c, 18:1ω7c<br />

(octadecenoic acids) and 18:0 (octadecanoic acid). Table 2 displays the relative amounts <strong>of</strong><br />

these fatty acids detected in the isolates <strong>of</strong> Y. ruckeri. Between the isolates, only slight<br />

qualitative and quantitative differences were observed. There were no differences between<br />

motile and non-motile isolates. In a dendrogram displaying the relatedness among the isolates<br />

studied (Fig. 1), most Y. ruckeri isolates formed a large cluster which included motile and<br />

non-motile strains from various locations within the studied geographic area, and which, in<br />

addition to isolates from rainbow trout (RT) also comprised isolates from brown trout and koi<br />

carp with different API 20E pr<strong>of</strong>iles. Within this group, the Euclidian distance between Y.<br />

ruckeri isolates was not larger than 3.75. The Euclidean distance is <strong>of</strong>ten used to compare<br />

71


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

pr<strong>of</strong>iles <strong>of</strong> respondents across variables. A distance <strong>of</strong> 1 or less then suggests the bacterial<br />

clusters were not well separated (Barshick et al., 1999) and a distance <strong>of</strong> 10 or less indicates<br />

all the isolates belong to the same species (Hallaksela and Niemostö, 1998). In the<br />

dendrogram a second cluster comprising motile and non-motile isolates from RT could be<br />

seen. However, the Euclidian distance to the first group was less than 5. One Y. ruckeri isolate<br />

from pike was separated by a Euclidian distance <strong>of</strong> about 5, and the reference strain, obtained<br />

from the Leibniz-<strong>Institute</strong> DSMZ-German Collection <strong>of</strong> Microorganisms and Cell Cultures,<br />

separated with a Euclidian distance <strong>of</strong> approx. 7.5 (Fig. 1). The clear separation <strong>of</strong><br />

DSM18506 from other Y. ruckeri strains in the dendrogram was reflected in slight differences<br />

in the pattern <strong>of</strong> cellular fatty acids (Tab. 2). In Figure 1, it is shown that Y. ruckeri isolates<br />

were well separated from other Yersinia species, with an Euclidian distance <strong>of</strong> 15 to Y.<br />

pseudotuberculosis and <strong>of</strong> 35 to Y. enterocolitica, respectively. This confirmed a high<br />

phylogenetic difference among Yersinia species.<br />

When our results were compared to previous reports on FAME composition <strong>of</strong> Y. ruckeri<br />

isolates from the USA (Arias et al., 2007), the fatty acids 12:0, 14:0, 15:0, 16:0, 17:0 CYCLO,<br />

17:0 and 18:1ω7c were found as main components in both studies.<br />

Jantzen and Lassen (1980) observed that strains from Y. enterocolitica, Y. pseudotuberculosis<br />

and Y. pestis all exhibited a similar fatty acid composition <strong>of</strong> 3-OH-14:0, 16:1, 16:0,17:0<br />

CYCLO and 18:1, and 16:0 was the most prominent fatty acid present in Y. pestis and Y.<br />

pseudotuberculosis (Tan et al., 2010). In the present study and in the study performed by<br />

Arias et al. (2007), 16:0 and 17:0 CYCLO were present in Y. ruckeri as well, with 16:0 as the<br />

most abundant fatty acid, but in contrast to Arias et al. (2007), in our isolates the level <strong>of</strong> 17:0<br />

CYCLO was relatively low compared to other Yersinia species (Jantzen and Lassen, 1980).<br />

While Arias et al. (2007) in their isolates found a clear distinction by FAME between<br />

non-motile Y. ruckeri and biotype I strains, in our analysis, motile and non-motile isolates<br />

clustered together. While the survey performed by Arias et al. (2007) mainly concentrated on<br />

an analysis <strong>of</strong> non-motile, biotype 2 strains, our material included a large array <strong>of</strong> field<br />

isolates from both biotypes. FAME pr<strong>of</strong>iles from other Yersinia species, in particular Y. pestis,<br />

were also found as highly conserved under standard preparation and analysis methods, even<br />

though samples from a wide geographical and temporal range were analyzed (Leclercq et al.,<br />

2000). In our study, the Euclidian distance between the main clusters <strong>of</strong> Y. ruckeri from the<br />

dendrogram was as low as 5 and the distance <strong>of</strong> these Y. ruckeri isolates to the Y. ruckeri<br />

reference strain was 7.5. The DSM reference strain and the Y. ruckeri isolate from pike were<br />

different from the majority <strong>of</strong> the isolates from RT, which in our study clustered together<br />

72


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

(Fig. 1). However, the Euclidian distance <strong>of</strong> Y. ruckeri isolates to the main clusters is lower<br />

than 10, indicating a high homogeneity <strong>of</strong> Y. ruckeri isolates from North West Germany based<br />

on FAME pr<strong>of</strong>iles, which not reflected differences in motility, host species, geographic<br />

distribution and API 20E pr<strong>of</strong>iles. Whether this phenotypic homogeneity is found in genetic<br />

traits as well will be analysed in further molecular studies.<br />

Acknowledgments<br />

We acknowledge the generous support provided during sample collection in aquaculture<br />

ponds by Dr. W. Schäfer and D. Mock from the Fish Health Service, North Rhine Westphalia.<br />

Dr. S. Braune (LAVES Niedersachsen), Dr. A. Nilz (LHL, Giessen) and Dr. C. Gould (MSD<br />

Animal Health) provided Yersinia ruckeri samples from their collection <strong>of</strong> fish pathogenic<br />

bacteria. We thank Pr<strong>of</strong>. Atanassova (<strong>Institute</strong> <strong>of</strong> Food Quality) for providing Y.<br />

enterocolitica and Y. pseudotuberculosis isolates. This study was financially supported by the<br />

Landesamt für Natur, Umwelt und Verbraucherschutz Nordrhein-Westfalen (LANUV). Y. H.<br />

received a scholarship from the China Scholarship Council.<br />

Conflicts <strong>of</strong> interest<br />

None <strong>of</strong> the authors had a financial or personal relationship with other people or organisations<br />

that can inappropriately influence or bias the content <strong>of</strong> this communication.<br />

References<br />

Abel K, Schmertzing HD, Peterson JI (1963): Classification <strong>of</strong> microorganisms by analysis<br />

<strong>of</strong> chemical composition. I . Feasibility <strong>of</strong> utilizing gas chromatography. J Bacteriol 85:<br />

1039–1044.<br />

Arias CR, Olivares-Fuster O, Hayden K, Shoemaker CA, Grizzle JM, Klesius PH (2007):<br />

First report <strong>of</strong> Yersinia ruckeri biotype 2 in the USA. J Aquat Anim Health 19: 35–40.<br />

Austin B, Austin DA (2007): Bacterial fish pathogens: Diseases <strong>of</strong> farmed and wild fish<br />

Praxis Publishing, Chichester, UK.<br />

Austin DA, Robertson PAW, Austin B (2003): Recovery <strong>of</strong> a new biogroup <strong>of</strong> Yersinia<br />

ruckeri from diseased rainbow trout(Oncorhynchus mykiss, Walbaum). Syst Appl<br />

Microbiol 26: 127–131.<br />

Barshick SA, Wolf DA, Vass AA (1999): Differentiation <strong>of</strong> microorganisms based on<br />

Pyrolysis-Ion trap mass spectrometry using chemical ionization. Anal Chem 71: 633–641.<br />

73


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

Canonica FP, Pisano MA (1985): Identification <strong>of</strong> hydroxy fatty acids in Aeromonas<br />

hydrophila, Aeromonas sobria and Aeromonas caviae. J Clin Microbiol 22: 1061–1062.<br />

Coquet L, Cosette P, Quillet L, Petit F, Junter GA, Jouenne T (2002): Occurrence and<br />

phenotypic charaterization <strong>of</strong> Yersinia ruckeri strains with bi<strong>of</strong>ilm-forming capacity in a<br />

rainbow trout Farm. Appl Environ Microbiol 68: 470–475.<br />

Davies RL, Frerichs GN (1989): Morphological and biochemical differences among isolates<br />

<strong>of</strong> Yersinia ruckeri obtained from wide geographical areas. J Fish Dis 12: 357–365.<br />

Fouz B, Zarza C, Amaro C (2006): First description <strong>of</strong> non-motile Yersinia ruckeri serovar I<br />

strains causing disease in rainbow trout, Oncorhynchus mykiss (Walbaum), cultured in<br />

Spain. J Fish Dis 29: 339–346.<br />

Hallaksela A-M, NiemostöP (1998): Stem discoloration <strong>of</strong> planted silverbirth. Scan J Forest<br />

Res 13: 169–176.<br />

Horne MT, Barnes AC (1999): Enteric redmouth disease (Yersinia ruckeri). In: Woo PTK,<br />

Bruno DW (eds.), Fish diseases and disorders. CABI Publishing, Wallingford, p. 445–477.<br />

Jantzen E, Lassen J (1980): Characterization <strong>of</strong> Yersinis species by analysis <strong>of</strong> whole-cell<br />

fatty acid. Int J Syst Bacteriol 30: 421–428.<br />

Karatas S, Candan A, Demircan D (2004): Enteric red mouth disease in cultured rainbow<br />

trout (Oncorhynchus mykiss) on the black sea coast <strong>of</strong> Turkey. Isr J Aquacult-Bamidgeh 56:<br />

226–231.<br />

Leclercq A, Guiyoule A, Lioui ME, Carniel E, Decallonne J (2000): High homogeneity <strong>of</strong><br />

the Yersinia pestis fatty acid composition. J Clin Microbiol 38: 1545–1551.<br />

Lesel R, Lesel M (1983): Outbreak <strong>of</strong> enteric redmouth disease in rainbow trout, Salmo<br />

gairdneri Richardson, in France. J Fish Dis 6: 385–387.<br />

Oraić D, Zrnčić S, Šoštarić B, Bažulić D, Lipej Z (2002): Occurrence <strong>of</strong> enteric redmouth<br />

disease in rainbow trout (Oncorhynchus mykiss) on farms in Croatia. Acta Vet Hung 50:<br />

283–291.<br />

Ross AJ, Rucker RR, Ewing WH (1966): Description <strong>of</strong> a bacterium associated with<br />

redmouth disease <strong>of</strong> rainbow trout (Salmo gairdneri). Can J Microbiol 12: 763–770.<br />

Rucker RR (1966): Redmouth disease <strong>of</strong> rainbow trout (Salmo gairdneri). Bull Off Int<br />

Epizoot 65: 825–830.<br />

Schlotfeldt H-J, Alderman DJ, Baudin-Laurencin F, Bernoth EM, Bruno DW (1995):<br />

What should I do? : a practical guide for the fresh water fish farmer. Bull Eur Ass Fish<br />

Pathol 15: suppl.<br />

74


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

Seker E, Karahan M, Ispir U, Catinkaya B, Saglam N, Sarieyyupoglu M (2012):<br />

Investigation <strong>of</strong> Yersinia ruckeri infection in rainbow trout (Oncorhynchus mykiss<br />

Walbaum 1792) farms by polymerase chain reaction (PCR) and bacteriological culture.<br />

Kafkas Univ Vet Fak Derg 18: 913–916.<br />

Stevenson RMW, Airdrie DW (1984): Isolation <strong>of</strong> Yersinia ruckeri bacteriophages. Appl<br />

Environ Microbiol 47: 1201–1205.<br />

Tan Y, Wu M, Liu H, Dong X, Guo Z, Song Z, Li Y, Cui Y, Song Y, Du Z, Yang R<br />

(2010): Cellular fatty acids as chemical markers for differentiation <strong>of</strong> Yersinia pestis and<br />

Yersinia Pseudotuberculosis. Lett Appl Microbiol 50: 104–111.<br />

Tinsley JW, Austin DA, Lyndon AR, Austin B (2011) Novel non-motile phenotypes <strong>of</strong><br />

Yersinia ruckeri suggest expansion <strong>of</strong> the current clonal complex theory. J Fish Dis 34:<br />

311–317.<br />

Valtonen ET, Rintamaki P, Koskivaara M (1992): Occurrence and pathogenicity <strong>of</strong><br />

Yersinia ruckeri at fish farms in northern and central Finland. J Fish Dis 15: 163–171.<br />

Wang KY, Fan FL, Huang XL, Geng Y, Chen DF, Huang JL (2009): Occurrence and<br />

treatment <strong>of</strong> a channel catfish(Ictalurus punctatus (Rafinesque)) disease, caused by<br />

Yersinia ruckeri. Scientific Fish Farming 5: 50–51.<br />

Warsen AE, Krug MJ, LaFrentz S, Stanek DR, Loge FJ, Call DR (2004): Simultaneous<br />

siscrimination between 15 fish pathogens by using 16S ribosomal DNA PCR and DNA<br />

microarrays. Appl Environ Microbiol 70: 4216–4221.<br />

Wortberg F, Nardy E, Contzen M, Rau J (2012): Identification <strong>of</strong> Yersinia ruckeri from<br />

diseased salmonid fish by fourier transform infrared spectroscopy. J Fish Dis 35: 1–10.<br />

Address for correspondence<br />

Pr<strong>of</strong>. Dr. Dieter Steinhagen<br />

Fish Disease Research Unit<br />

<strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong><br />

Bünteweg 17,<br />

30559 <strong>Hannover</strong><br />

Germany<br />

dieter.steinhagen@tiho-hannover.de<br />

TABLE 1: Characteristics <strong>of</strong> Yersinia ruckeri isolates from Northern Germany used for<br />

cellular fatty acid analysis<br />

75


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

Isolate No. Host Isolation Motility API20E Isolate Host Isolation Motility API20E<br />

year<br />

Pr<strong>of</strong>ile No.<br />

year<br />

Pr<strong>of</strong>ile<br />

DSM18506 RT 1961 + g L-1 RT 2011 + c<br />

*<br />

194-1 RT 2004 - a K1 RT 2011 - a<br />

204-1 Trout 2004 + b lay3 RT 2011 - b<br />

205-1 RT 2004 - d P2 RT 2011 - b<br />

275-1 RT 2004 - a P1 RT 2011 - a<br />

70-1 Koi 2005 + a LT13-1 RT 2011 + c<br />

72-2 RT 2005 + b P1-1 RT 2011 - b<br />

101-1 RT 2005 - a KPST3-1 RT 2011 - b<br />

111-1 - 2005 + e KPL1-1 RT 2011 - b<br />

115-1 RT 2005 - a B1-1 RT 2011 - d<br />

178-1 B 2005 + e B1-2 RT 2011 - a<br />

182-1 RT 2005 - b SRF1-1 RT 2011 - c<br />

285-1 Pike 2005 + f SRHK1-4 RT 2011 - c<br />

287-1 RT 2005 + b N1 RT 2011 - a<br />

339-1 RT 2005 - a AT7-2 RT 2011 - b<br />

317-1 Trout 2006 - a LT12-1 RT 2011 + c<br />

01-1 RT 2007 - b P2-1 RT 2011 - c<br />

16-1 RT 2007 + b AT7-1 RT 2011 - b<br />

20-1 RT 2007 + a B3-1-1 RT 2011 - c<br />

157-1 RT 2007 + g N2 RT 2011 - a<br />

337-1 RT 2007 - b AVTA1 RT 2011 - b<br />

365-1 RT 2007 - b SFFT2 RT 2011 - b<br />

494-1 RT 2007 - a B5-2 RT 2011 - h<br />

201-1 RT 2008 + g ST1 RT 2011 - c<br />

109-2 RT 2009 - h PT2-2 RT 2011 - b<br />

87-1 RT 2009 - a KPK1 RT 2012 - c<br />

1024 RT 2010 - b KPK2 RT 2012 - b<br />

3279 RT 2010 - c SRON1 RT 2012 - a<br />

2846 RT 2010 - a B4-1 RT 2012 - c<br />

2946 RT 2010 - a G1S1 RT 2008 - c<br />

1521 RT 2011 + e SRT2-2 RT 2012 - c<br />

3484 RT 2010 - b B1-5 RT 2012 - c<br />

3554 RT 2010 - a STR3 RT 2012 - c<br />

2948 RT 2010 - a SRT13-1 RT 2012 - c<br />

21-1 RT 2011 - a SRT14-1 RT 2012 - c<br />

11-1 RT 2011 - b KP6-1 RT 2012 + e<br />

kp6 RT 2011 - b SRT23-1 RT 2012 - b<br />

kp1 RT 2011 - b PB3-1 RT 2012 - b<br />

B1 RT 2011 - d PZ3-2 RT 2012 - b<br />

P3 RT 2011 - a F1-1 RT 2012 - c<br />

HK1 RT 2011 - c SRT6-1 RT 2012 - c<br />

HK2 RT 2011 - a<br />

76


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

RT: Rainbow trout. B: Brown trout. API pr<strong>of</strong>iles: a 5306100, b 5107100, c 5307100, d 5106100, e 5307500, f 5305700, g 5304100, h 5104100.<br />

TABLE 2: Major components <strong>of</strong> the cellular fatty acid pr<strong>of</strong>ile <strong>of</strong> Yersinia ruckeri isolates from<br />

different fish species from North West Germany detected by the CLIN 40 method<br />

Feature Mean SD DSM18506<br />

12:0 4.66 0.19 4.75<br />

13:0 0.19 0.10 0.99<br />

ECL Unknown 13.957 0.69 0.03 0.76<br />

14:0 1.36 0.09 1.38<br />

ECL Unknown 14.502 0.70 0.11 0.58<br />

15:0 1.61 0.32 4.14<br />

16:1ω5c 0.27 0.05 0.30<br />

16:0 25.42 0.95 20.86<br />

16:1 2OH 0.06 0.08 -<br />

17:1ω8c 0.23 0.12 0.55<br />

17:0 CYCLO 5.88 1.34 2.89<br />

17:0 0.75 0.11 1.14<br />

18:1ω9c 1.53 0.23 1.36<br />

18:1ω7c 12.46 0.71 10.84<br />

18:0 0.89 0.14 0.72<br />

77


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

78


Chapter 4 Fatty Acid composition <strong>of</strong> Y. ruckeri in North West Germany<br />

FIGURE 1: Dendrogram <strong>of</strong> Yersina ruckeri isolates<br />

* Motile isolates; B: Isolates from Brown trout; K: Isolates from Koi carp; T: Isolates from trout;P:<br />

Isolates from Pike; N: Isolates from non specific fish species; Other isolates were from rainbow trout.<br />

79


Chapter 5.<br />

In vitro cytotoxicity and multiplex PCR detection <strong>of</strong><br />

virulence factors <strong>of</strong> Yersinia ruckeri isolated from<br />

rainbow trout in North West Germany<br />

Yidan Huang, Mikolaj Adamek, Charles Walker, Martin Runge,<br />

Dieter Steinhagen<br />

81


The extent <strong>of</strong> Yidan Huang‟s contribution to the article is evaluated according to the<br />

following scale:<br />

A. has contributed to collaboration (0-33%).<br />

B. has contributed significantly (34-66%).<br />

C. has essentially performed this study independently (67-100%).<br />

1. Design <strong>of</strong> the project including design <strong>of</strong> individual experiments: B<br />

2. Performing <strong>of</strong> the experimental part <strong>of</strong> the study: B<br />

3. Analysis <strong>of</strong> the experiments: C<br />

4. Presentation and discussion <strong>of</strong> the study in article form: C<br />

82


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

Fish Disease Research Unit, <strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong>, Germany 1<br />

School <strong>of</strong> Life Science, Keele <strong>University</strong>, Staffordshire, UK 2<br />

Lower Saxony State Office for Consumer Protection and Food Safety (LAVES), Food and<br />

<strong>Veterinary</strong> <strong>Institute</strong> Braunschweig/<strong>Hannover</strong>, Germany 3<br />

In vitro cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Yersinia ruckeri<br />

isolated from rainbow trout in North West Germany<br />

Untersuchungen zur in-vitro-Zytotoxizität und zum Nachweis von Virulenzfaktoren mittels<br />

multiplex- PCR bei Yersinia ruckeri-Isolaten aus Teichwirtschaften in Nord-Westdeutschland<br />

Yidan Huang 1 , Mikolaj Adamek 1 , Charles Walker 1, 2 , Martin Runge 3 , Dieter Steinhagen 1§<br />

§Corresponding author<br />

Abstract The aim <strong>of</strong> this study was to investigate differences in presence and expression <strong>of</strong> virulence factors<br />

between biotype 1 and 2 strains <strong>of</strong> 82 Yersinia ruckeri isolates, collected from North West Germany during<br />

period <strong>of</strong> 2004-2012, and to analyze the cytotoxicity <strong>of</strong> these strains to different fish cell lines. The common<br />

virulence factor genes, such as yhlA and yhlB encoding for hemolysin YhlA, rucC and rupG encoding for<br />

ruckerbactin, yrp1 and yrpDEF for ABC exporter protein system, and two flagellar genes, including flgA for<br />

flagellar secretion chaperones and flhA for flagellar secretion apparatus, were found present in both biotype 1<br />

and 2 isolates <strong>of</strong> Y. ruckeri collected from North West Germany using multiplex PCR. mRNA expression <strong>of</strong><br />

these genes was compared between the two biotypes <strong>of</strong> Y. ruckeri. There was no significant diversity (p>0.05)<br />

in the expression <strong>of</strong> these genes between biotype 1 and 2 strains. 27 Y. ruckeri isolates from different typing<br />

groups were analysed in cytotoxicity tests to common carp brain (CCB), epithelioma papulosum cyprini<br />

fathead minnow epithelial cell (FHM) and rainbow trout gonad-2 (RTG-2), respectively. In vitro cytotoxicity<br />

the isolates to CCB, EPC and FHM was higher than that to RTG-2 (p


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

however, after 48h, there was no significant difference (p>0.05) <strong>of</strong> cytotoxicity between those two biotypes.<br />

Our results suggest that biotype 2 strains from North West Germany are homogenous with biotype 1 strains on<br />

the basis <strong>of</strong> genetic virulence factor genes. At lower temperature non-motile Y. ruckeri isolates were found<br />

active than motile strains, which could explain why in winter non-motile strains were found more responsible<br />

for ERM outbreaks than motile strains.<br />

Key Words: Enteric red mouth disease, Biotype, cytotoxicity, mRNA expression, Flagellar genes,<br />

Zusammenfassung<br />

In der vorliegenden Studie wurden 82 Yersinia ruckeri Isolate, die 2004 bis 2012 von Fischen aus<br />

Teichwirtschaften in Nord-Westdeutschland isoliert wurden, auf Vorkommen und Expression von<br />

Virulenzfaktoren sowie auf Zytotoxizität gegenüber unterschiedlichen Fischzellen untersucht. Dabei sollten<br />

allem Unterschiede zwischen Isolaten aus den Biotypen 1 und 2 analysiert werden. Untersucht wurden Gene<br />

beschriebene Virulenzfaktoren: yhlA und yhlB, die das Hämolysin YhlA codieren, rucC und rupC, die das<br />

Ruckerbactin codieren, yrp1 und yrpDEF, die Elemente des ABC-Protein-Exkretionssystem codieren, sowie<br />

zwei Geißel-assoziierte Gene, flgA, das ein Sekretions-Chaperon codiert sowie flhA, das ein Protein des<br />

Geißel-assoziierten Sekretionsapparates codiert. Alle Gene ließen sich in allen untersuchten Isolaten aus<br />

Nord-Westdeutschland mittels multiplex PCR nachweisen. Zudem wurde die mRNA-Expression dieser Gene<br />

bei Isolaten aus beiden Biotypen von Y. ruckeri verglichen. Alle analysierten Gene waren bei den untersuchten<br />

Isolaten nicht unterschiedlich exprimiert. Des weiteren wurden 27 Y. ruckeri Isolate aus unterschiedlichen<br />

Typisierungsgruppen in vitro auf ihre Zytotoxizität auf die Fischzelllinen „common carp brain“ (CCB),<br />

Epithelioma papulosum cyprini (EPC), „fathead minnow epithelial cells“ (FHM), und „rainbow trout<br />

gonad-2“ (RTG-2) geprüft. Die Mehrzahl der Y. ruckeri Isolate wies eine geringe in vitro Zytotoxizität auf.<br />

Zytotoxizität war gegenüber den CCB, EPC und FHM Zellen höher als gegenüber RTG-2 Zellen (p< 0,05).<br />

15°C war nach 24 h Inkubation die maximale Zytotoxizität nicht motiler Y. ruckeri Isolate gegenüber EPC<br />

FHM Zellen höher als bei motilen Isolaten (p< 0,05). Nach 48 h Inkubationszeit war kein Unterschied<br />

Isolaten aus beiden Biotypen erkennbar. Unsere Ergebnisse lassen vermuten, dass bei Y. ruckeri Isolaten in<br />

Nord-Westdeutschland das Vorkommen von Virulenzfaktoren in beiden Biotypen einheitlich ist. Bei<br />

84


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

Temperaturen scheinen nicht motile Y. ruckeri Isolate eine höhere Zytotoxizität aufzuweisen, was erklären<br />

könnte, warum im Winter nicht motile Y. ruckeri Isolate häufiger für Ausbrücke der Rotmaulseuche<br />

verantwortlich waren, als motile Isolate.<br />

Schlüsselwörter: Rotmaulseuche, Biotyp, Zytotoxizität, mRNA-Expression, Geißel-assoziierte Gene<br />

Introduction<br />

Yersinia (Y.) ruckeri, the causative agent <strong>of</strong> enteric redmouth disease (ERM), is considered to be one <strong>of</strong> the<br />

important pathogens in trout aquaculture. Since this bacterium was first isolated from rainbow trout in the<br />

in 1950 (Rucker 1966), it has kept causing disease outbreaks in trout populations in various geographic<br />

although generally the disease can be well controlled by means <strong>of</strong> vaccination. Y. ruckeri comprises 2 biotypes<br />

(BT) and several serotypes (Romalde et al. 1993). Strains positive for motility and lipase activity are grouped<br />

into biotype 1, while biotype 2 strains are non-motile and exhibit no lipase activity. Biotype 2 strains were<br />

responsible <strong>of</strong> disease outbreaks also in rainbow trout populations, which had been vaccinated with a<br />

commercial vaccine against the typical biotype 1 strains (Davies and Frerichs 1989).<br />

The bacterium can infect salmonids and other fishes from both freshwater and marine environments; it persists<br />

in fish farms which suffer from ERM outbreaks and has a high bi<strong>of</strong>ilm- forming capability to adhere to<br />

commonly found in fish ponds or fish tanks (Coquet et al. 2002a, Coquet et al. 2002b). During the infection<br />

process, Y. ruckeri adheres first to mucus (Tobback et al. 2010a), subsequently to epithelial cells <strong>of</strong> the gills<br />

to the villi <strong>of</strong> intestinal enterocytes (Tobback et al. 2010b). The intestine, together with the gills are considered<br />

as important site <strong>of</strong> entry (Méndez and Guijarro 2013, Tobback et al. 2009), with a subsequent systemic<br />

dissemination <strong>of</strong> the pathogen (Méndez and Guijarro 2013). In infection experiments, using the immersion <strong>of</strong><br />

juvenile rainbow trout into a suspension <strong>of</strong> Y. ruckeri, virulent strains induced mortality while avirulent strains<br />

did not cause mortality or persistent infection (Tobback et al. 2009, Tobback et al. 2010b). Both virulent and<br />

avirulent Y. ruckeri strains were able to cross the gill or intestinal epithelium(Tobback et al. 2010b).<br />

So far, differences in virulence could be identified by the induction <strong>of</strong> mortality in experimental<br />

infections(Davies 1991, Tobback et al. 2009). Up to now, only a few mechanisms related to the pathogenicity<br />

Y. ruckeri have been analysed (Fernández et al. 2007, Méndez and Guijarro 2013). Type III secretion systems<br />

85


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

(T3SS) have been proved to be essential for the pathogenesis <strong>of</strong> different bacterial organisms (Cornelis 2002,<br />

Tobe et al. 2006, Vilches et al. 2009). recently, a Ysa (Yersinia secretion apparatus) -like T3SS, which is<br />

different to the T3SS from human pathogenic Yersinia species, was found in Y. ruckeri (Gunasena et al. 2003).<br />

However, there is no detailed information about the structure <strong>of</strong> this Ysa-like T3SS. FlhA is an essential<br />

component <strong>of</strong> the flagellar secretion apparatus, and FlgA is one <strong>of</strong> the flagellar secretion chaperones<br />

and Hughes 2008). Since biotype 2 strains <strong>of</strong> Y. ruckeri are lacking both flagellar motility and secreted lipase<br />

activity(Austin et al. 2003), the question whether the presence or expression <strong>of</strong> flagellar genes in Y. ruckeri in<br />

North West Germany needs further consideration.<br />

Virulence factors <strong>of</strong> Yersinia species include genes encoding extracellular toxins and iron acquisition.<br />

and Toranzao (1993) reported that the injection <strong>of</strong> extracellular products <strong>of</strong> Y. ruckeri induced haemorrhages<br />

and necrotic areas at the injection site. The extracellular products had a cytotoxic and proteolytic activity. The<br />

responsible Yrp1 protease was found to be encoded by the gene yrp1, and requires three genes (yrpD, yrpE,<br />

yrpF) to be secreted (Fernández et al. 2002). Yrp1 can hydrolyze laminin, thus may cause membrane<br />

leading to erosion and pores in capillary vessels, which results in the hemorrhages in mouth and intestines<br />

(Fernández et al. 2003). Another virulence factor <strong>of</strong> Y. ruckeri, a Serratia-type hemolysin, is encoded by the<br />

open-reading frame Yhl A (Fernández, Prieto and Guijarro 2007).<br />

For successful colonization and invasion iron acquisition is essential for many microbial pathogens and<br />

therefore they developed high affinity iron transport systems(Faraldo-Gomez and Sanson 2003). In Yersinia<br />

ruckeri, the genes rucC and rupG were identified as putative counterparts <strong>of</strong> Escherichia coli entC and fepG.<br />

These genes are all involved in iron piracy, confirming that Y. ruckeri produces a catechol siderophore<br />

(ruckerbactin) (Fernández et al. 2004, Romalde et al. 1991a).<br />

So far, a limited number <strong>of</strong> motile and non-motile Y. ruckeri-strains were analysed for differences in the<br />

expression <strong>of</strong> genes encoding those virulence factors. In the present study, these genes were detected in the 82<br />

Y. ruckeri isolates collected from North West Germany and a reference strain by a multiplex PCR and for the<br />

first time their expression differences were checked between biotype 1 and 2 isolates. In addition, the<br />

cytotoxicity <strong>of</strong> selected the isolates to fish cells was analysed in an in vitro assay.<br />

Materials and Methods<br />

86


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

Bacteria<br />

A total <strong>of</strong> 83 Y. ruckeri isolates (including one reference strain DSM18506) were included in this study. From<br />

these 48 isolates were collected during a field study in rainbow trout populations in the German federal state<br />

North Rhine-Westphalia (NRW) with a history <strong>of</strong> ERM outbreaks, which took place during the four seasons<br />

<strong>of</strong> 2011-2012. Furthermore, 33 isolates from clinical cases <strong>of</strong> ERM were obtained from LAVES<br />

Niedersachsen, Food and <strong>Veterinary</strong> <strong>Institute</strong> Braunschweig/<strong>Hannover</strong>, Germany (Lower Saxony) and<br />

Landesbetrieb Hessisches Landeslabor (LHL) Giessen (Hessen), one non-motile strain previously isolated in<br />

NRW in 2008 was <strong>of</strong>fered by Dr. Gould from MSD Animal Heath and the reference strain DSM 18506, was<br />

included in the analysis. All isolates were cultured on Tryptone soy agar (TSA) and incubated at 23±2 °C for<br />

24 h. All the isolates were identified as Y. ruckeri and differentiated according to motility and lipase activity in<br />

motile biotype 1 and non-motile biotype 2 isolates as well as by API 20 E pr<strong>of</strong>iles, repetitive sequence-based<br />

PCR and pulsed field electrophoresis. With these methods, 27 different typing groups could be recognized,<br />

including the reference strain DSZM 18506 as typing group (tp) 27.<br />

Analysis <strong>of</strong> flagellar genes and virulence factors<br />

All 82 isolates and the reference strain DSM 18506 were used for the detection <strong>of</strong> genes encoding virulence<br />

flagellar secretion system and factors using a multiplex PCR. Subsequently, 6 isolates from biotype 1 and 2<br />

respectively, including the reference strain DSM 18506 were chosen according to their epidemiological<br />

characteristics and used for RNA isolation and detection <strong>of</strong> gene expression.<br />

DNA isolation<br />

Bacterial cells were lysed and DNA was extracted following the manufacturer‟s protocol <strong>of</strong> the innuPREP<br />

Bacteria DNA Kit (Analytikjena, Germany). Yield and purity <strong>of</strong> isolated DNA was analysed by means <strong>of</strong> a<br />

NanoDrop spectrophotometer (Delaware, USA). The DNA was stored at -20°C until further use.<br />

RNA isolation<br />

12 Y. ruckeri isolates (6 isolates <strong>of</strong> biotype 1 and 6 isolates <strong>of</strong> biotype 2) were chosen according to their<br />

motility and geographic origins for RNA isolation. Total RNA was isolated using the TRI Reagent (Sigma,<br />

USA) according to the manufacturer‟s instructions. The total RNA pellet was air-dried and suspended in<br />

100μl sterile double distilled water. RNA concentration and purity were determined by a NanoDrop<br />

spectrophotometer (Delaware, USA).<br />

87


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

cDNA synthesis<br />

The RNA mix (600 ng) <strong>of</strong> each Y. ruckeri isolate, respectively, 1 μl <strong>of</strong> 10x buffer with MgCl 2 , 2 units <strong>of</strong><br />

DNase Ⅰ, and 0.25 μl <strong>of</strong> ribolock RNase inhibitor (20 u) (Thermo Scientific, Germany) were incubated at 37 ℃<br />

for 30 min. After incubation, the DNase Ⅰ was inactivated by adding 1 μl <strong>of</strong> 25 mM EDTA and incubating at<br />

65 ℃ for 10 min. Subsequently, cDNA was synthetized using Maxima Reverse Transcriptase (Fermentas,<br />

Germany) according to the manufacturer‟s instruction. 5 μl <strong>of</strong> the RNA samples after DNAse treatment were<br />

mixed with 0.25 μl random hexamer (100 pmol), 0.25 μl <strong>of</strong> Oligo dT (100 pmol), 1 μl <strong>of</strong> dNTPs (10 mM each)<br />

and nuclease-free water (until a total volume <strong>of</strong> 15 μl), and incubated at 65 ℃ for 5 min. Subsequently, the<br />

samples were mixed with 4 μl <strong>of</strong> 5x RT buffer, 0.5 μl ribolock (20 u), 0.5 μl <strong>of</strong> Maxima Reverse Transcriptase<br />

(100 u). The mixture was incubated for 10 minutes at 25 ℃, followed by 25 minutes at 50 ℃. Finally, the<br />

reaction was ended by heating at 85 ℃ for 5 minutes. Then the cDNA samples were diluted 1:20 and kept at<br />

-20 ℃ for further use. All non reverse transcribed samples were checked by PCR in order to confirm no<br />

contamination <strong>of</strong> genomic DNA.<br />

Primer design<br />

Genes encoding for hemolysin YhlA (yhlA and yhlB, NCBI Accession: AY576533.3), ruckerbactin (rucC and<br />

rupG, NCBI Accession: AY576531.1), ABC exporter protein system (yrp1 and yrpD, E, F, NCBI Accession:<br />

JQ890543.1), the two flagellar factors flagellar secretion chaperones FlgA (flgA,NCBI Accession:<br />

NZ_ACCC01000021.1) and the essential component <strong>of</strong> the flagellar secretion apparatus FlhA (flhA, NCBI<br />

Accession: NZ_ACCC01000021.1), were chosen to be analysed by multiplex PCR. Primers were designed<br />

according to related Y. ruckeri nucleotide sequences obtained from the NCBI data base.<br />

Primers were designed using the s<strong>of</strong>tware Primer 3, available online at http://frodo.wi.mit.edu/. The melting<br />

temperature <strong>of</strong> the primers was calculated using the s<strong>of</strong>tware <strong>of</strong> Integrated DNA technologies OligoAnalyzer<br />

3.1 available at http://eu.idtdna.com/. Primers with detailed information used in this study were listed in<br />

Table 1.<br />

Multiplex PCR<br />

88


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

Multiplex PCRs were performed in 25 μl volumes containing 0.5 μl <strong>of</strong> cDNA template, 0.2 mM<br />

concentrations <strong>of</strong> deoxynucleoside triphosphates, 5 μl <strong>of</strong> 5× reaction buffer A (PeQlab, Germany), the primer<br />

mixture containing 0.5 μl <strong>of</strong> each primer (10 μM) in the same group and 5 U <strong>of</strong> KAPA2G Robust DNA<br />

Polymerase (PeQlab, Germany). The reaction was performed in a Biometra T3000 thermocycler (Analytik<br />

Jena, Germany) with an initial denaturation cycle at 95°C for 3 min, followed by 30 cycles <strong>of</strong> amplification<br />

(denaturation at 95°C for 15 s annealing at 60°C for 30 s, and extension at 68°C for 1 min), and a final<br />

7-min-elongation at 72°C. The PCR products were detected by electrophoresis by using 1% agarose gels pre<br />

stained with GelRed (Biotium,USA). Images were captured under ultraviolet light by a digital camera system<br />

(INTAS Science Imaging Instruments GmbH, Göttingen, Germany).<br />

Detection <strong>of</strong> gene expressions<br />

For a transcription analysis <strong>of</strong> the genes encoding virulence factors, PCRs were performed for each gene<br />

separately in 25 μl volumes containing 5 μl <strong>of</strong> cDNA template, 0.2 mM concentrations <strong>of</strong> deoxynucleoside<br />

triphosphates, 5 μl <strong>of</strong> 5× reaction buffer A (PeQlab, Germany), 0.5 μl <strong>of</strong> each forward and reverse primer<br />

(10μM) and 5 U <strong>of</strong> KAPA2G Robust DNA Polymerase (PeQlab, Germany).<br />

The reaction conditions were<br />

the same as for the multiplex PCR described above, except the amplification cycles were 40 instead <strong>of</strong> 30. The<br />

expression <strong>of</strong> the 16S rRNA gene was used as reference gene. PCR products were detected by electrophoresis<br />

in 1% agarose gels pre stained with GelRed (Biotium, USA). Images were captured under ultraviolet light by<br />

a digital camera system (INTAS Science Imaging Instruments GmbH, Göttingen, Germany).<br />

Semi-quantitative analysis <strong>of</strong> PCR products<br />

The amount <strong>of</strong> PCR products was determined by densitometric analysis using a computerized image analysis<br />

system, CP ATLAS 2.0 Cross-platform Advanced Thin Layer and Gel analysis s<strong>of</strong>tware. Grading the ratio<br />

between the densitometric results <strong>of</strong> the target genes and the 16S rRNA gene generated semiquantitative PCR<br />

results.<br />

In vitro- testing <strong>of</strong> isolates for cytotoxicity<br />

For an in vitro testing <strong>of</strong> isolates for cytotoxicity, 4 different fish cell lines were used: Common carp brain<br />

(CCB), epithelioma papulosum cyprini (EPC), fathead minnow epithelial cell (FHM) are derived from<br />

cyprinid fishes and rainbow trout gonad-2 (RTG-2) cells were <strong>of</strong> salmonid origin. 27 Y. ruckeri isolates were<br />

randomly selected from each <strong>of</strong> the different typing groups. From each isolate, 10 5 , 10 6 and10 7 bacterial cells<br />

89


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

per well were inoculated into the wells <strong>of</strong> a microtiter plate with a confluent monolayer <strong>of</strong> the cells. Thereafter,<br />

the plates were incubated at 15 °C for 6h, 24h and 48h as well as at 25 °C (20 °C for RTG-2 cells) for 2h, 6h,<br />

and 24h. Then the cell free supernatant was collected from the wells and used for the detection <strong>of</strong> cytotoxicity<br />

by means <strong>of</strong> a cytotoxicity detection kit (TaKaRa, Japan) on the basis <strong>of</strong> Lactate dehydrogenase (LDH)<br />

release from destructed cell. The measurement was done according to the manufacturer‟s instructions.<br />

Triplicate wells inoculated with bacteria and without a cell monolayer and triplicate wells with a cell<br />

monolayer treated with Triton X 100 served as low and high controls respectively. For wells inoculated with<br />

bacteria, the percentage <strong>of</strong> cytotoxicity was calculated according to the following equation:<br />

The ability <strong>of</strong> Y. ruckeri isolates to cause cytotoxicity was then grouped into four categories: cytotoxicity<br />

under 25%, 25%-50%, 50%-75 % and above 75%.<br />

Statistical analysis<br />

All data are presented as means±SD. Statistical significant differences between groups were determined by<br />

two way-ANOVA followed by multiple t-tests using Graphpad prism 6.0 s<strong>of</strong>tware. Differences at p< 0.05<br />

were considered significant.<br />

Results<br />

Multiplex PCR<br />

A multiplex PCR was used to test Y. ruckeri isolates for the presence <strong>of</strong> the genes encoding virulence factors.<br />

All targeted genes were found to be present in all 83 Y. ruckeri isolates from fish populations in North West<br />

Germany. PCR products for all the genes under study could be detected in isolates in biotype 1 and 2, as well<br />

as in the reference strain from the DSMZ collection.<br />

Gne Transcription<br />

When the transcription <strong>of</strong> the flagellar and virulence factor genes was analysed, gene transcripts could be<br />

detected in motile and non-motile Y. ruckeri isolates. A semi-quantitative analysis <strong>of</strong> the level <strong>of</strong> gene<br />

transcripts <strong>of</strong> target genes in relation to the transcription <strong>of</strong> the 16S rRNA indicated a higher expression <strong>of</strong><br />

most <strong>of</strong> the genes in biotype 1 (motile) isolates compared to the non-motile biotype 2 isolates (Fig 2). Slight<br />

differences were observed in the RNA expression between biotype 1 and biotype 2 isolates in all tested genes.<br />

These differences however, were not significant (p> 0.05).<br />

90


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

The flagellar genes flhA and flgA were expressed in isolates from both motile and non-motile biotypes (Fig. 2).<br />

In biotype 1 isolates, flhA expression (component <strong>of</strong> the flagellar secretion apparatus) was significantly higher<br />

than flgA (flagellar secretion chaperone) expression (P0.05, Fig. 2).<br />

There was no significant difference (p>0.05) in the expression <strong>of</strong> rucC and rupG genes, coding together for<br />

ruckerbactin, and <strong>of</strong> yhlA and yhlB genes, essential for the hemolysin YhlA (P>0.05) between isolates from<br />

biotype 1 or biotype 2. In all isolates, genes from the ABC reporter system (yrp1, yrpD, yrpE, yrpF) were<br />

differently expressed. In both biotype 1 and biotype 2isolates the expression <strong>of</strong> yrp1 was significantly (p


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

(Tab. 3). These differences in cytotoxicity were significantly different in experiments performed at 15 °C, but<br />

at 25 °C. Overall, cytotoxicity <strong>of</strong> Y. ruckeri isolates was significantly higher (p


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

species share homology with the flagellar T3SS such as FliFHIKNPQR and FlhAB (Desvaux et al. 2006).<br />

According the flagellar cascade theory in enterobacteria, the expression <strong>of</strong> the operon flhDC, a class 1 gene,<br />

ensures the expression <strong>of</strong> all subsequent genes, and the expression <strong>of</strong> class 2 genes results in the synthesis <strong>of</strong><br />

the basal body-hook structure and the activator <strong>of</strong> class 3 genes (Gillen and Hughes 1991, Helmann 1991).<br />

The flhA gene, which was analysed here, is a flagellar class 1 gene and involved in type three export <strong>of</strong><br />

flagellar components(Minamino and Macnab 2000). The flhA mutant <strong>of</strong> Bacillus thuringiensis had the ablity<br />

<strong>of</strong> synthesizing flagellin but was not able to export it (Ghelardi et al. 2002). In Y. ruckeri flagella contributed<br />

to the virulence during the early stage <strong>of</strong> infection, when the pathogen in the environment first contacts the<br />

host and begins to invade(Kim 2000). Recently, Evenhuis et al (Evenhuis et al. 2009) constructed a flhA::Tn5<br />

mutant <strong>of</strong> Y. ruckeri, which was lacking both motility and secreted lipase activity. The mutation loss <strong>of</strong><br />

flagellar secretion was considered as a possible mechanism for the emergence <strong>of</strong> non-motile biotype 2 Y.<br />

ruckeri isolates, but it did not affect the virulence <strong>of</strong> the isolates in rainbow trout (Evenhuis et al. 2009). In our<br />

study, the flagellar genes flgA and flhA were detected in both motile and non-motile isolates <strong>of</strong> Y. ruckeri, and<br />

there were no differences in the expression <strong>of</strong> these genes between biotype1 and biotype 2 strains. The biotype<br />

2 isolates from our study were lacking a flagellum as confirmed by a flagellar silver-staining (unpublished<br />

observations), which confirms the findings <strong>of</strong> Evenhuis et al (Evenhuis et al. 2009) that a lack <strong>of</strong> flagellar<br />

motility in these biotype 2 strains, is not related to a complete absence <strong>of</strong> flgA and flhA genes. Welch et al<br />

(Welch et al. 2011) found mutations present in flhA and flhB genes from European strains <strong>of</strong> Y. ruckeri,<br />

therefore, it is believed that these mutations may lead to a functional suppression <strong>of</strong> these genes <strong>of</strong> the isolates<br />

from north west Germany as well.<br />

Extracellular products <strong>of</strong> Y. ruckeri are highly toxic to fish and were found to induce clinical signs <strong>of</strong> the<br />

disease like darkening <strong>of</strong> the skin and haemorrhages in a rainbow trout injection model (Romalde and Toranzo<br />

1993). An important component in these extracelluar products is the extracellular protease yrp1, which is<br />

secreted by a type 1 Gram-negative bacterial ABC exporter protein system, comprised <strong>of</strong> the genes yrpD, E, F<br />

and a protease inhibitor (Fernández et al. 2002). Site-directed insertion mutations <strong>of</strong> yrp1 and yrpE resulted in<br />

a loss <strong>of</strong> protease activity and decreased virulence <strong>of</strong> bacterial strains to rainbow trout (Fernández et al. 2002).<br />

In the present study, the genes encoding the protease and the protein exporter system were found in the<br />

genome <strong>of</strong> all the isolates tested. In these isolates, the expression <strong>of</strong> yrp1 and yrp F were relatively higher than<br />

93


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

the expression <strong>of</strong> the other two genes, suggesting that the expression <strong>of</strong> yrpF is highly required for the<br />

secretion <strong>of</strong> Yrp1 protease. Furthermore, a haemolysin YhlA and the activation <strong>of</strong> a protein (YhlB) associated<br />

with the secretion <strong>of</strong> this haemolysin were identified as additional virulence factors <strong>of</strong> Y. ruckeri (Fernández et<br />

al. 2007). The genes encoding these proteins were present in all isolates examined in the present study and the<br />

level <strong>of</strong> expression was similar between biotype 1 and biotype 2 isolates. The expression <strong>of</strong> YhlB was induced<br />

under iron-restricted conditions (Fernández et al. 2007), which could result in a higher iron release form<br />

erythrocytes by haemolysis. An important iron uptake system <strong>of</strong> Y. ruckeri is ruckerbactin which also was<br />

identified as virulence factor in the host (Fernández et al. 2004). The expression <strong>of</strong> rucC and rucG genes,<br />

encoding for ruckerbactin showed only slight differences in biotype 1 and biotype 2 strains, suggesting<br />

biotype 2 mutated from biotype 1 without changing the virulence factors at the genetic level.<br />

The presence <strong>of</strong> all genes encoding for different mechanisms involved in the virulence <strong>of</strong> Y. ruckeri in all<br />

isolates from North West Germany indicate a genetic homogeneity <strong>of</strong> this species in this geographic area,<br />

even though several different typing groups could be identified by pulsed field electrophoresis. This confirms<br />

the data from previous studies based on genetic fingerprinting (Romalde et al. 1993), multi-locus sequencing<br />

(Kotetishvili et al. 2005) and gene expression (Fernández et al. 2003, Fernández et al. 2007, Fernández et al.<br />

2002). Furthermore, the virulence associated genes were not differently transcribed between isolates from the<br />

two biotypes.<br />

In several fish pathogenic bacteria, including aeromonads, virulent bacterial strains also displayed an<br />

increased cytotoxicity towards fish cells in vitro (Burr et al. 2002, Burr et al. 2003, Wahli et al. 2005). In the<br />

present study, the cytotoxicity <strong>of</strong> Y. ruckeri was tested towards four different fish cell lines derived from<br />

cyprinids (CCB, EPC and FHM) and from rainbow trout (RTG-2) tissues. Most <strong>of</strong> the tested isolates (88.89%)<br />

exhibited some cytotoxicity to at least two <strong>of</strong> these fish cell lines, EPC and FHM cells were found more<br />

sensitive to Y. ruckeri isolates than CCB and RTG-2 cells. This confirms findings from previous studies in<br />

which Y. ruckeri adhered to and invaded several cultured cells derived from salmonid fishes, such as Chinook<br />

salmon embryo (CHSE-214), RTG, Rainbow trout kidney (RTPK), or Rainbow trout liver (R1) cells<br />

(Tobback 2009, Tobe et al. 2006), but exhibited a strong host cell preference to FHM cells (Kawula et al.<br />

1996) as observed in the present study. Y. ruckeri did not invade human derived Hep-2 cells(Kawula et al.<br />

1996). Several Y. ruckeri isolates were able to adhere to and to invade into CHSE 24 or R1 cells, but after 6 h<br />

94


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

<strong>of</strong> incubation showed a lower intracellular survival, suggesting Y. ruckeri may not be able to multiply or<br />

survive inside <strong>of</strong> these cultured cells (Tobback 2009). In our study, an obvious cytotoxicity could not be<br />

observed after an incubation time <strong>of</strong> just 6 h even when the cells were cultivated at 25℃. Increase cytotoxicity<br />

was seen after 24 h, indicating that intracellular survival was not crucial to Y. ruckeri isolates and that the<br />

bacterium could cause cytotoxicity to cultured cells even extracellular (Tobback 2009). Romalde and Toranzo<br />

(Romalde and Toranzo 1993) proved the ability <strong>of</strong> Y. ruckeri to adhere to and effectively invade fish cell lines<br />

cultured in vitro. However, the virulence <strong>of</strong> Y. ruckeri isolates to rainbow trout was not observed to correlate<br />

with their in vitro invasiveness in cell lines (Tobback et al. 2010a). In these studies (Kawula et al. 1996,<br />

Romalde et al. 1991b, Tobback et al. 2010a) a relatively high percentage <strong>of</strong> Y. ruckeri invaded and survived in<br />

FHM cells, which corresponds to findings in our study, where all isolates showed elevated cytotoxicity to<br />

FHM cells. Yersinia ruckeri strains are generally considered as poor producers <strong>of</strong> cytotoxins (Romalde and<br />

Toranzo 1993). This corresponds to observation in the present study where a large proportion <strong>of</strong> the isolates<br />

had a low cytotoxicity, in particular to CCB and RTG-2 cells.<br />

After incubation for 24 h at 15 ℃, non-motile strains caused a higher cytotoxicity than motile strains. This<br />

difference between strains from the two biotypes was not seen at 25 °C, suggesting that at low temperature<br />

non-motile strains were more active than motiles. It may explain why non-motile strains were more <strong>of</strong>ten<br />

found responsible for ERM outbreak in winter than motile strains.<br />

Conclusions<br />

Common virulence factor genes <strong>of</strong> Y. ruckeri, such as yhlA and yhlB encoding for the hemolysin YhlA, rucC<br />

and rupG encoding for ruckerbactin, yrp1 and yrpDEF for an ABC exporter protein system, and two flagellar<br />

genes, including flgA for flagellar secretion chaperones and flhA for flagellar secretion apparatus, were found<br />

present in all biotype 1 and 2 isolates <strong>of</strong> Y. ruckeri collected from North West Germany during the period <strong>of</strong><br />

2004-2012. There was also no significant diversity (p>0.05) in the expression <strong>of</strong> these genes between biotype<br />

1 and 2 strains. Our results suggest that in biotype 2 strains, lack <strong>of</strong> flagellar motility, is not because <strong>of</strong> the<br />

absence <strong>of</strong> flgA and flhA genes, but probably results from the functional suppression <strong>of</strong> these genes. These<br />

data indicates that biotype 2 strains are homogenous with biotype 1 strains on the basis <strong>of</strong> genetic virulence<br />

factor genes expression. Y. ruckeri isolates could cause higher cytotoxicity to EPC and FHM cell lines than<br />

CCB and RTG-2. Non-motile strains were more active than motile strains when the temperature was low.<br />

95


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

Acknowledgement<br />

We acknowledge the generous support provided during sample collection in aquaculture ponds by Dr. W.<br />

Schäfer and D. Mock from the Fish Health Service, North Rhine Westphalia. Dr. S. Braune (LAVES<br />

Niedersachsen) and Dr. A. Nilz (LHL, Giessen) provided Yersinia ruckeri samples from their collection <strong>of</strong><br />

fish pathogenic bacteria. This study was financially supported by the Landesamt für Natur, Umwelt und<br />

Verbraucherschutz Nordrhein-Westfalen (LANUV).<br />

References<br />

Austin DA, Robertson PAW, and Austin B (2003): Recovery <strong>of</strong> a new biogroup <strong>of</strong> Yersinia ruckeri from diseased<br />

rainbow trout(Oncorhynchus mykiss, Walbaum). Syst Appl Microbiol 26: 127-131.<br />

Burr SE, Stuber K, Wahli T, and Frey J (2002): Evidence for a type III secretion system in Aeromonas salmonicida<br />

subsp. salmonicida. J Bacteriol 184: 5966-5970.<br />

Burr SE, Wahli T, Segner H, Pugovkin D, and Frey J (2003): Association <strong>of</strong> type III secretion genes with virulence <strong>of</strong><br />

Aeromonas salmonicida subsp. salmonicida. Dis Aquat Org 57: 167-171.<br />

Chevance FF, and Hughes KT (2008): Coordinating assembly <strong>of</strong> a bacterial macromolecular machine. Nat Rev<br />

Microbiol 6: 455-465.<br />

Coquet L, Cosette P, Junter G-A, Beucher E, Saiter J-M, Jouenne T, and D J-MS (2002a): Adhesion <strong>of</strong> Yersinia<br />

ruckeri to fish farm materials:influence <strong>of</strong> cell and material surface properties. Colloids Surf B 26: 373-378.<br />

Coquet L, Cosette P, Quillet L, Petit F, Junter GA, and Jouenne T (2002b): Occurrence and phenotypic<br />

charaterization <strong>of</strong> Yersinia ruckeri strains with bi<strong>of</strong>ilm-forming capacity in a rainbow trout Farm. Appl Environ<br />

Microbiol 68: 470-475.<br />

Cornelis GR (2002): Yersinia type III secretion: send in the effectors. J Cell Biol 158: 401-408.<br />

Davies RL (1991): Virulence and serum-resistance in different clonal groups and serotypes <strong>of</strong> Yersinia ruckeri. Vet<br />

Microbiol 29: 289-297.<br />

Davies RL, and Frerichs GN (1989): Morphological and biochemical differences among isolates <strong>of</strong> Yersinia ruckeri<br />

obtained from wide geographical areas. J Fish Dis 12: 357-365.<br />

Desvaux M, Hebraud M, Henderson IR, and Pallen MJ (2006): Type III secretion: what's in a name? Trends<br />

Microbiol 14: 157-160.<br />

96


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

Erhardt M, Namba K, and Hughes KT (2010): Bacterial nanomachines: The flagellum and type III injectisome. Cold<br />

Spring Harb Perspect Biol 2: a000299.<br />

Evenhuis JP, Lapatra SE, Verner-Jeffreys DW, Dalsgaard I, and Welch TJ (2009): Identification <strong>of</strong> Flagellar<br />

Motility Genes in Yersinia ruckeri by Transposon Mutagenesis. Antimicrob Agents Chemother 75: 6630-6633.<br />

Faraldo-Gomez J, and Sanson SP (2003): Acquisition <strong>of</strong> siderophores in Gram-negative bacteria. National Review 4:<br />

105-116.<br />

Fernández L, López JR, Secades P, Menéndez A, Márquez I, and Guijarro JA (2003): In vitro and in vivo studies <strong>of</strong><br />

the Yrp1 protease from Yersinia ruckeri and its role in protective immunity against enteric redmouth disease <strong>of</strong> salmonids.<br />

Appl Environ Microbiol 69: 7328-7335.<br />

Fernández L, Márquez I, and Guijarro JA (2004): Identification <strong>of</strong> Specific In Vivo-Induced (ivi) Genes in Yersinia<br />

ruckeri and Analysis <strong>of</strong> Ruckerbactin, a Catecholate Siderophore Iron Acquisition System. Appl Environ Microbiol 70:<br />

5199-5207.<br />

Fernández L, Prieto M, and Guijarro JA (2007): The iron- and temperature-regulated haemolysin YhlA is a virulence<br />

factor <strong>of</strong> Yersina ruckeri. Microbiology 153: 483-489.<br />

Fernández L, Secades P, Lopez JR, Marquez I, and Guijarro JA (2002): Isolation and analysis <strong>of</strong> a protease gene with<br />

an ABC transport system in the fish pathogen Yersinia ruckeri: insertional mutagenesis and involvement in virulence.<br />

Microbiology 148: 2233-2243.<br />

Ghelardi E, Celandroni F, Salvetti S, Beecher DJ, Gominet M, Lereclus D, Wong ACL, and Senesi S (2002):<br />

Requirment <strong>of</strong> flhA for swarming differentiation, flagellin export, and secretion <strong>of</strong> virulence-associated proteins in<br />

Bacillus thuringiensis. J Bacteriol 184: 6424-6433.<br />

Gibello A, Blanco MM, Moreno MA, Cutuli MT, Domenech A, Moniguez L, and Fernandez-Garayzabal JF (1999):<br />

Development <strong>of</strong> a PCR Assay for Detection <strong>of</strong> Yersinia ruckeri in Tissues <strong>of</strong> Inoculated and Naturally Infected Trout.<br />

Appl Environ Microbiol 65: 346-350.<br />

Gillen KL, and Hughes KT (1991): Negative regulatory loci coupling flagellin synthesis to flagellar assembly in<br />

Salmonella typhimurium. J Bacteriol 173: 2301-2310.<br />

Gunasena D, Komrower J, and Macintyre S (2003): The Fish Pathogen Yersinia ruckeri Possesses a TTS System. In:<br />

Skurnik M, Benoechea JA, and Granfors K (eds.), The Genus Yersinia: Entering the Functional Genomic Era. Kluwer<br />

Academic/Plenum Publishers, New York, USA, p. 105-107.<br />

97


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

Helmann JD (1991): Alternative sigma factors and the regulation <strong>of</strong> flagellar gene expression. Mol Microbiol 5:<br />

2875-2882.<br />

Kawula TH, Lelivelt MJ, and Orndorff PE (1996): Using a new inbred fish model and cultured fish tissue cells to<br />

study Aeromonas hydrophila and Yersinia ruckeri pathogenesis. Microb Pathog 20: 119-125.<br />

Kim W (2000): Investigation <strong>of</strong> the role <strong>of</strong> flagella in the virulence <strong>of</strong> Yersinia ruckeri. Canada, The <strong>University</strong> <strong>of</strong><br />

Guelph.<br />

Kotetishvili M, Kreger A, Wauters G, Jr. Morris JG, Sulakvelidze A, and Stine OC (2005): Multilocus Sequence<br />

Typing for Studying Genetic Relationships among Yersinia Species. J Clin Microbiol 43: 2674-2684.<br />

Méndez J, and Guijarro JA (2013): In vivo monitoring <strong>of</strong> Yersinia ruckeri in fish tissues: progression and virulence<br />

gene expression. Environ Microbiol Rep 5: 179-185.<br />

Minamino T, and Macnab RM (2000): Interaction among components <strong>of</strong> the Samonella flagellar export apparatus and<br />

its substrates. Mol Microbiol 35: 1052-1064.<br />

Romalde JL, Conchas RF, and Toranzo AE (1991a): Evidence that Yersinia ruckeri possesses a high affinity iron<br />

uptake system. FEMS Microbiol Lett 80: 121-126.<br />

Romalde JL, Iteman I, and Carniel E (1991b): Use <strong>of</strong> pulsed field gel electrophoresis to size the chromosome <strong>of</strong> the<br />

bacterial fish pathogen Yersinia ruckeri FEMS Microbiology Letters 84: 217-226.<br />

Romalde JL, Margarinos B, Barja JL, and Toranzo AE (1993): Antigenic and molecular characterization <strong>of</strong> Yersinia<br />

ruckeri Proposal for a new intraspecies classification. System Appl Microbiol 16: 411-419.<br />

Romalde JL, and Toranzo AE (1993): Pathological activities <strong>of</strong> Yersinia ruckeri, the enteric redmouth (ERM)<br />

bacterium. FEMS Microbiol Lett 112: 291-300.<br />

Rucker RR (1966): Redmouth disease <strong>of</strong> rainbow trout (Salmo gairdneri). Bull <strong>of</strong>f int Epizoot 65: 825-830.<br />

Salmond GP, and Reeves PJ (1993): Membrane Traffic Wardens and Protein Secretion in Gram-negative Bacteria.<br />

Trends Bioch Sci 18: 7-12.<br />

Tobback E (2009): Early pathogenesis <strong>of</strong> Yersinia ruckeri infections in rainbow trout (Oncorhynchus mykiss, Walbaum)<br />

Ghent, Ghent <strong>University</strong>.<br />

Tobback E, Decostere A, Hermans K, Ryckaert J, Duchateau L, Haesebrouck F, and Chiers K (2009): Route <strong>of</strong><br />

entry and tissue distribution <strong>of</strong> Yersinia ruckeri in experimentally infected rainbow trout Oncorhynchus mykiss. Dis<br />

Aquat Org 84: 219-228.<br />

98


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

Tobback E, Decostere A, Hermans K, Van den Broeck W, Haesebrouck F, and Chiers K (2010a): In vitro markers<br />

for virulence in Yersinia ruckeri. J Fish Dis 33: 197-209.<br />

Tobback E, Hermans K, Decostere A, Van-den-Broeck W, Haesebrouck F, and Chiers K (2010b): Interactions <strong>of</strong><br />

virulent and avirulent Yersinia ruckeri strains with isolated gill arches and intestinal explants <strong>of</strong> rainbow trout<br />

Oncorhynchus mykiss. Dis Aquat Org 90: 175-179.<br />

Tobe T, Beason SA, Taniguchi H, Abe H, Bailey CM, Fivian A, Younis R, Matthews S, Marches O, Frankel G,<br />

Hayashi T, and Pallen MJ (2006): An extensive repertoire <strong>of</strong> type III secretion effectors in Escherichia coli O157 and<br />

the role <strong>of</strong> lambdoid phages in their dissemination. Proc Nat Acad Sci U S A 103: 14941-14946.<br />

Vilches S, Jimenez N, Tomás JM, and Merino S (2009): Aeromonas hydrophila AH-3 Type III Secretion System<br />

Expression and Regulatory Network. Appl Environ Microbiol 75: 6382-6392.<br />

Wahli T, Burr SE, Pugovkin D, Mueller O, and Frey J (2005): Aeromonas sobia, a causative agent <strong>of</strong> disease in<br />

farmed perch, Perca fluviatilis L. J Fish Dis 28.<br />

Welch TJ, Verner-Jeffreys DW, Dalsgaard I, Wiklund T, Evenhuis JP, Cabrera JAG, Hinshaw JM, Drennan JD,<br />

and LaPatra SE (2011): Independent emergence <strong>of</strong> Yersinia ruckeri biotype 2 in th United States and Europe. Appl<br />

Environ Microbiol 77: 3493-3499.<br />

Address <strong>of</strong> correspondence<br />

Pr<strong>of</strong>. Dr. Dieter Steinhagen<br />

Fish Disease Research Unit<br />

<strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong><br />

Buenteweg 17,<br />

30559 <strong>Hannover</strong><br />

Germany<br />

dieter.steinhagen@tiho-hannover.de<br />

99


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

TABLE 1: PCR primers for the analysis <strong>of</strong> Yersinia ruckeri virulence factor targets<br />

PCR<br />

Target<br />

Function<br />

Primer<br />

Sequence(5'→3')<br />

Ampliconsize<br />

References<br />

Name<br />

Gene<br />

Name<br />

(bp)<br />

16S<br />

16S<br />

Identification YER8 GCGAGGAGGAAGGGTTAAGTG 575 (Gibello et<br />

rRNA<br />

rRNA<br />

YER10<br />

GAAGGCACCAAGGCATCTCTG<br />

al. 1999)<br />

Multiplex<br />

flgA<br />

flagellar<br />

flgA fw GTGCCGCTGACAATCTGG 217 This study<br />

PCR<br />

secretion<br />

Group 1<br />

chaperones<br />

flgA rv<br />

CCAAGGGAACTCTGGCTTTG<br />

rucC ruckerbactin rucCfw CGAAAGGCTCCAACTGACTG 414 This study<br />

rucCrv<br />

CAGAAGGCGGTGTTTTGCTC<br />

rupG ruckerbactin rupGfw AGTGCCGCTTCCAGTGATG 510 This study<br />

rupGrv<br />

GCCTCGCAGTGTATGGGTATG<br />

yhlB hemolysin yhlBfw CGCCTGACCAATGAACTGAC 744 This study<br />

yhlBrv<br />

GAGTGTGGGGCTGCTGATAC<br />

flhA<br />

flagellar<br />

flhA fw TATGCCGGGTAAACAGATGG 809 This study<br />

secretion<br />

apparatus<br />

flhA rv<br />

TGCCAATTTCAACCCCTTTC<br />

yhlA hemolysin yhlA CGGCTAACACCCACACAATC 990 This study<br />

100


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

fw<br />

yhlArv<br />

CGTTGGCTGCATCTTGTTTC<br />

Multiplex<br />

yrpE<br />

ABC<br />

yrpEfw CAGTTGGCCGAGTTACAGTCC 209 This study<br />

PCR<br />

exporter<br />

yrpErv<br />

ACCCGTGCATCAACAAACAG<br />

Group 1<br />

yrpF protein yrpFfw CGGTCTTGTCTTTGCGTCAG 364 This study<br />

system<br />

yrpFrv<br />

CGGCCAATAATGTTTTCCAG<br />

yrp1 yrp1fw ATTACCCGCACCAACCAAAC 609 This study<br />

yrp1rv<br />

GGCCCCGTACAAATGCTG<br />

yrpD yrpDfw TTTGATGCGCCTTGGTTTC 799 This study<br />

yrpDrv<br />

TTCATATCGGCACCATCCAG<br />

TABLE 2: Mean and maximum cytotoxicity <strong>of</strong> Yersinia ruckeri (10 7 cells/ well) to different fish cell lines (%)<br />

Temperature<br />

(℃)<br />

Time<br />

(h)<br />

CCB EPC FHM RTG-2<br />

mean Max. mean Max. mean Max. mean Max.<br />

15 6 0.1 1.16 0.41 6.67 0.43 3.47 0.23 2.42<br />

24 7.21 41.84 3.83 20.28 6.45 69.52 4.84 48.93<br />

48 49.26 64.89 50.23 82.39 55.41 88.3 43.85 87.07<br />

25/20 2 0.02 0.32 0.67 9.49 0.45 5.04 0.58 6.09<br />

6 1.25 8.06 4.23 16.55 1.97 9.15 1.6 11.45<br />

101


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

24 31.63 57.76 56.9 87.41 58.3 95.23 13.48 52.67<br />

TABLE 3: Yersinia ruckeri isolates from North West Germany: Comparison <strong>of</strong> molecular typing and in vitro cytotoxicity<br />

Typing<br />

Group<br />

(tp) No.<br />

Incubation time and temperature<br />

24 h at 25 ℃/20 ℃* 48 h at 15 ℃<br />

CCB EPC FHM RTG2 CCB EPC FHM RTG2<br />

tp1 + ++ ++ - + + +++ ++<br />

tp2 - - - - ++ - +++ ++<br />

tp3 - ++ ++ - + +++ +++ -<br />

tp4 - - +++ - ++ ++ +++ ++<br />

tp5 + ++ ++ - ++ ++ +++ ++<br />

tp6 + ++ ++ + ++ ++ ++ ++<br />

tp7 + ++ +++ + + ++ +++ -<br />

tp8 + ++ - - - ++ + -<br />

tp9 - +++ +++ - + ++ +++ -<br />

tp10 - + +++ - ++ ++ +++ ++<br />

tp11 ++ ++ +++ - ++ ++ ++ +++<br />

102


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

tp12 - +++ + - ++ - + ++<br />

tp13 ++ ++ +++ - ++ ++ ++ +++<br />

tp14 - - - - - - - ++<br />

tp15 + ++ +++ - ++ ++ ++ ++<br />

tp16 + +++ - - ++ + - +<br />

tp17 + - - - - - - -<br />

tp18 + + +++ ++ + ++ +++ -<br />

tp19 - - - - ++ + + -<br />

tp20 ++ ++ +++ - ++ ++ ++ ++<br />

tp21 - - - - - - - -<br />

tp22 + ++ +++ - ++ +++ +++ -<br />

tp23 + +++ +++ - ++ ++ +++ ++<br />

tp24 + +++ +++ - ++ + + ++<br />

tp25 - +++ - - ++ - - -<br />

tp26 + +++ +++ - ++ ++ ++ ++<br />

tp27 ++ +++ +++ - ++ ++ ++ ++<br />

Note: RTG-2 cells were tested at 20 ℃, instead <strong>of</strong> 25 ℃. „-‟: cytotoxicity was lower than 25%; „+‟: cytotoxicity was between 25% and 50%; „++‟:<br />

cytotoxicity was between 50% and 75%; „+++‟: cytotoxicity was above 75%.<br />

103


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

TABLE 4: Mean and maximum cytotoxicity <strong>of</strong> Yersinia ruckeri isolates at 15℃ to different cell lines (%)<br />

Motility<br />

Incubation time<br />

24h<br />

48h<br />

CCB EPC FHM RTG2 CCB EPC FHM RTG2<br />

mean max mean max mean max mean max mean max mean max mean max mean max<br />

Motile 4.14 11.04 3.36 11.98 0.64 3.78 1.01 9.08 62.04 71.08 43.05 70.38 44.97 87.43 48.05 66.8<br />

Non-motile 8.75 41.84 4.07 20.28 9.35 69.52 6.76 48.93 42.7 64.89 53.81 77.72 60.63 88.3 41.74 94.26<br />

104


Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

M 1 2 3 4 5 6 7 8 9 10<br />

yhlA<br />

flhA<br />

yhlB<br />

rupG<br />

rucC<br />

yrpD<br />

yrp1<br />

yrpF<br />

flgA<br />

yrpE<br />

FIGURE 1:<br />

Detection <strong>of</strong> virulence factor genes in Yersinia ruckeri isolated from fish in aquaculture<br />

populations from North West Germany by multiplex PCR. It showed the results from 4 Y. ruckeri isolates<br />

including the reference strain DSM 18506 as an example. DSM 18506 (lanes 1, 6) and isolate 111-1 (lanes<br />

4, 9) were motile strains, while G1S1 (lanes 2, 7) and KP6-2 (lanes 3, 8) were non-motile strains. In PCR 1<br />

(lanes 1-4), six bands present for six different genes were detected and in PCR 2 (lanes 5-9) four bands with<br />

products <strong>of</strong> different sizes present different genes were detected. M: Maker, Lane 1-5 Multiplex PCR group<br />

1: Lane 1 DSM 18506, Lane 2 G1S1, Lane 3 KP6-2, Lane 4 111-1, Lane 5 negative control; Lane 6-10<br />

Multiplex PCR group 2: Lane 6 DSM 18506, Lane 7 G1S1, Lane 8 KP6-2, Lane 9 111-1, Lane 10 negative<br />

control (without DNA template).<br />

105


Number <strong>of</strong> isolates<br />

Number <strong>of</strong> isolates<br />

Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

FIGURE 2: Relative expression <strong>of</strong> different virulence factor genes in Yersinia ruckeri isolates from biotype 1<br />

and biotype 2.<br />

24h 25℃ 48h 15℃<br />

C<br />

30<br />

30<br />

C<br />

B<br />

25<br />

20<br />

15<br />

10E5<br />

10E6<br />

10E7<br />

25<br />

20<br />

15<br />

10E5<br />

10E6<br />

10E7<br />

10<br />

10<br />

5<br />

5<br />

0<br />

Below 25% 25%-50% 50%-75% Above 75%<br />

Cytotoxicity<br />

0<br />

Below 25% 25%-50% 50%-75% Above 75%<br />

Cytotoxicity<br />

106


Number <strong>of</strong> isolates<br />

Number <strong>of</strong> isolates<br />

Number <strong>of</strong> isolates<br />

Number <strong>of</strong> isolates<br />

Number <strong>of</strong> isolates<br />

Number <strong>of</strong> isolates<br />

Chapter 5 Cytotoxicity and multiplex PCR detection <strong>of</strong> virulence factors <strong>of</strong> Y. ruckeri<br />

E<br />

25<br />

25<br />

P<br />

C<br />

20<br />

15<br />

10<br />

10E5<br />

10E6<br />

10E7<br />

20<br />

15<br />

10<br />

10E5<br />

10E6<br />

10E7<br />

5<br />

5<br />

0<br />

Below 25% 25%-50% 50%-75% Above 75%<br />

Cytotoxicity<br />

0<br />

Below 25% 25%-50% 50%-75% Above 75%<br />

Cytotoxicity<br />

F<br />

H<br />

M<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Below 25% 25%-50% 50%-75% Above 75%<br />

Cytotoxicity<br />

10E5<br />

10E6<br />

10E7<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Below 25% 25%-50% 50%-75% Above 75%<br />

Cytotoxicity<br />

10E5<br />

10E6<br />

10E7<br />

R<br />

30<br />

25<br />

T<br />

G-<br />

2<br />

25<br />

20<br />

15<br />

10<br />

5<br />

10E5<br />

10E6<br />

10E7<br />

20<br />

15<br />

10<br />

5<br />

10E5<br />

10E6<br />

10E7<br />

0<br />

Below 25% 25%-50% 50%-75% Above 75%<br />

Cytotoxicity<br />

0<br />

Below 25% 25%-50% 50%-75% Above 75%<br />

Cytotoxicity<br />

FIGURE 3: Cytotoxicity <strong>of</strong> Yersinia ruckeri isolates to different fish cell lines after incubation for 24h<br />

at 25℃ (20℃ for RTG-2 cells) and at 48 h at 15 ℃. 10E5, 10E6 and 10E7: bacterial concentration <strong>of</strong><br />

10 5 cells/ well, 10 6 cells/well and 10 7 cells/well.<br />

107


108


Chapter 6.<br />

Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance<br />

properties <strong>of</strong> Yersinia ruckeri from fish<br />

Yidan Huang, Geovana Brenner Michael, Roswitha Becker,<br />

Heike Kaspar, Joachim Mankertz, Stefan Schwarz,<br />

Martin Runge, Dieter Steinhagen<br />

Accepted by <strong>Veterinary</strong> Microbiology<br />

109


The extent <strong>of</strong> Yidan Huang‟s contribution to the article is evaluated according to the following<br />

scale:<br />

A. has contributed to collaboration (0-33%).<br />

B. has contributed significantly (34-66%).<br />

C. has essentially performed this study independently (67-100%).<br />

1. Design <strong>of</strong> the project including design <strong>of</strong> individual experiments: B<br />

2. Performing <strong>of</strong> the experimental part <strong>of</strong> the study: B<br />

3. Analysis <strong>of</strong> the experiments: B<br />

4. Presentation and discussion <strong>of</strong> the study in article form: B<br />

110


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties<br />

<strong>of</strong> Yersinia ruckeri from fish<br />

Yidan Huang a , Geovana Brenner Michael b , Roswitha Becker b , Heike Kaspar c ,<br />

Joachim Mankertz c , Stefan Schwarz b , Martin Runge d , Dieter Steinhagen a,§<br />

a<br />

Fish Disease Research Unit, <strong>University</strong> <strong>of</strong> <strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong>, Foundation, Germany<br />

b<br />

<strong>Institute</strong> <strong>of</strong> Farm Animal Genetics, Friedrich-Loeffler-Institut (FLI), Neustadt-Mariensee,<br />

Germany<br />

c<br />

Federal Office <strong>of</strong> Consumer protection and Food safety (BVL), Berlin, Germany<br />

d<br />

Lower Saxony State Office for Consumer Protection and Food Safety (LAVES), Food and<br />

<strong>Veterinary</strong> <strong>Institute</strong> Braunschweig/<strong>Hannover</strong>, Germany<br />

Short title: Antimicrobial resistance <strong>of</strong> Yersinia ruckeri<br />

Keywords: minimal inhibitory concentration, antimicrobial susceptibility testing, target site<br />

mutation, plasmid, gene cassette<br />

111


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

§ Corresponding author: Tel: +49 511 953-8560. Fax: +49 511 953-8587<br />

E-mail address: dieter.steinhagen@tiho-hannover.de<br />

Abstract<br />

Enteric red-mouth disease, caused by Yersinia ruckeri, is an important disease in rainbow trout<br />

aquaculture. Antimicrobial agents are frequently used in aquaculture, thereby causing a selective pressure<br />

on bacteria from aquatic organisms under which they may develop resistance to antimicrobial agents. In<br />

this study, the distribution <strong>of</strong> minimal inhibitory concentrations (MICs) <strong>of</strong> antimicrobial agents for 83<br />

clinical and non-clinical epidemiologically unrelated Y. ruckeri isolates from North West Germany was<br />

determined. Antimicrobial susceptibility was conducted by broth microdilution at 22 ± 2 °C for 24, 28 and<br />

48 h. Incubation for 24 h at 22 ± 2 °C appeared to be suitable for susceptibility testing <strong>of</strong> Y. ruckeri. In<br />

contrast to other antimicrobial agents tested, enr<strong>of</strong>loxacin and nalidixic acid showed a bimodal<br />

distribution <strong>of</strong> MICs, with one subpopulation showing lower MICs for enr<strong>of</strong>loxacin (0.008-0.015 µg/mL)<br />

and nalidixic acid (0.25-0.5 µg/mL) and another subpopulation exhibiting elevated MICs <strong>of</strong> 0.06-0.25 and<br />

8-64 µg/mL, respectively. Isolates showing elevated MICs revealed single amino acid substitutions in the<br />

quinolone resistance-determining region (QRDR) <strong>of</strong> the GyrA protein at positions 83 (Ser83-Arg or -Ile)<br />

or 87 (Asn87-Tyr), which raised the MIC values 8- to 32-fold for enr<strong>of</strong>loxacin or 32- to 128-fold for<br />

nalidixic acid. An isolate showing elevated MICs for sulfonamides and trimethoprim harbored a ~8.9 kb<br />

plasmid, which carried the genes sul2, strB and a dfrA14 gene cassette integrated into the strA gene. These<br />

observations showed that Y. ruckeri isolates were able to develop mutations that reduce their<br />

susceptibility to (fluoro)quinolones and to acquire plasmid-borne resistance genes.<br />

112


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

1. Introduction<br />

Yersinia ruckeri is a Gram-negative rod-shaped bacterium which belongs to the family<br />

Enterobacteriaceae and is the causative agent <strong>of</strong> enteric red mouth disease (ERM). Several fish species<br />

can be infected by Y. ruckeri, especially rainbow trout (Oncorhynchus mykiss) (Noga, 2010). Because <strong>of</strong><br />

the economic losses, ERM is one <strong>of</strong> the most important diseases in trout aquaculture in Europe. To<br />

harmonize antimicrobial susceptibility testing <strong>of</strong> bacteria from aquatic animals and to make the obtained<br />

results more comparable, two documents have been published by the Clinical and Laboratory Standards<br />

<strong>Institute</strong> (CLSI): VET03-A (CLSI, 2006a) for susceptibility testing by agar disk diffusion and VET04-A<br />

for susceptibility testing by broth dilution methods (CLSI, 2006b). In addition, a third CLSI document,<br />

VET03/VET04-S1 (CLSI, 2010) that contains the first clinical breakpoints and also epidemiological<br />

cut-<strong>of</strong>f values applicable to Aeromonas salmonicida has been published. While clinical breakpoints are<br />

important to predict the clinical outcome <strong>of</strong> an antimicrobial treatment (Bywater et al., 2006; Schwarz et<br />

al., 2010), epidemiological cut-<strong>of</strong>f values only differentiate between a wildtype and a non-wildtype<br />

population, the latter <strong>of</strong> which usually has acquired resistance mechanisms. In aquaculture,<br />

agents, such as quinolones, sulfonamides and trimethoprim, were used to treat fish diseases (Angulo,<br />

2000; Rigos and Troisi, 2005), thereby furthering the development <strong>of</strong> antimicrobial resistance among<br />

bacteria <strong>of</strong> aquatic organisms. As little information about the antimicrobial susceptibility <strong>of</strong> Y. ruckeri is<br />

currently available (reviewed in Rodgers, 2001), we investigated Y. ruckeri isolates obtained from<br />

different fish species in North West Germany for their minimal inhibitory concentrations (MICs) for<br />

different antimicrobial agents. Furthermore, selected Y. ruckeri isolates, that showed distinctly higher<br />

113


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

MIC values for specific antimicrobial agents than the remaining isolates, were investigated for the<br />

presence <strong>of</strong> resistance-mediating mutations or resistance genes.<br />

2. Materials and Methods<br />

2.1 Bacterial cultures<br />

Eighty-two Y. ruckeri isolates collected from different fish species, including rainbow trout<br />

(Oncorhynchus mykiss), brown trout (Salmo trutta), koi (Cyprinus carpio) and pike (Esox lucius) by<br />

field and monitoring studies in the period from 2004 to 2012 in North West Germany in the federal<br />

states Lower Saxony (n=25), Hessen (n=8) and North Rhine-Westphalia (n=49) were investigated.<br />

Among them, 33 clinical isolates were obtained from diseased rainbow trout suffering from ERM.<br />

These isolates were provided by the Lower Saxony State Office for Consumer Protection and Food<br />

Safety (LAVES), the Food and <strong>Veterinary</strong> <strong>Institute</strong> Braunschweig/Hanover and the fish disease service<br />

at the Landesbetrieb Hessisches Landeslabor. The remaining 49 isolates were obtained from rainbow<br />

trout during different ERM outbreaks in the federal state North Rhine-Westphalia. In addition, the type<br />

strain Y. ruckeri DSM 18506 (ATCC ® 29473) provided by the Clinic for Poultry at the <strong>University</strong> <strong>of</strong><br />

<strong>Veterinary</strong> <strong>Medicine</strong> <strong>Hannover</strong>, Foundation was included in this study. In a previous study, all these<br />

isolates were identified as Y. ruckeri and were investigated for their phenotypic and genomic<br />

relationships (Huang et al., 2013). The Y. ruckeri isolates were maintained on tryptone soy agar (TSA,<br />

Sigma-Aldrich, Taufenkirchen, Germany) and incubated at 25 °C.<br />

2.2 Antimicrobial susceptibility testing<br />

114


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Susceptibility tests were performed by broth microdilution according to the CLSI guideline<br />

VET04-A (CLSI, 2006b) and recommendations <strong>of</strong> Miller and colleagues (Miller et al., 2003; 2005).<br />

Custom-made microtiter plates (MCS Diagnostics, Swalmen, The Netherlands) served to test a total <strong>of</strong><br />

24 antimicrobial agents or combinations. The antimicrobial agents tested and the test ranges<br />

corresponded to those previously described by Feßler et al. (2012). E. coli ATCC ® 25922 served as a<br />

quality control strain. All plates were incubated at 22 ± 2 °C and the results were read at 24, 28 and 48h<br />

(CLSI, 2006b; Miller et al., 2005).<br />

2.3 Molecular analysis <strong>of</strong> antimicrobial resistance <strong>of</strong> selected Y. ruckeri isolates<br />

Based on previously described epidemiological characteristics, motility and pulsed-field gel<br />

electrophoresis (PFGE) patterns (Huang et al., 2013), ten representative isolates that showed elevated<br />

MICs for enr<strong>of</strong>loxacin and nalidixic acid, and three isolates with low MICs (control isolates) were<br />

investigated for mutations in the QRDR <strong>of</strong> the target gene gyrA. All selected isolates originated from<br />

different ERM outbreaks. An internal part <strong>of</strong> the gyrA gene was amplified by PCR (Shah et al., 2012)<br />

and the amplicon sequenced (Eur<strong>of</strong>ins MWG Operon, Germany).<br />

The unique isolate 1521, that showed elevated MICs for the sulfonamide sulfamethoxazole and<br />

the combination sulfamethoxazole/trimethoprim, was investigated by PCR assays for the presence <strong>of</strong><br />

sulfonamide resistance genes sul1, sul2 and sul3 (Kehrenberg and Schwarz, 2001, Grape et al., 2003),<br />

various trimethoprim resistance genes (dfrA1, dfrA5, dfrA7 and dfrA14-A17) (Kadlec et al., 2005),<br />

streptomycin resistance genes strA and strB (Kehrenberg and Schwarz, 2001; Gebreyes and Althier,<br />

115


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

2002) and for the linkage <strong>of</strong> sul2 and strA (Kehrenberg and Schwarz, 2001). In addition, this isolate was<br />

investigated for the presence <strong>of</strong> class 1 and 2 integrons (Kadlec et al., 2005).<br />

2.4 Plasmid transformation, analysis and sequencing<br />

The plasmid <strong>of</strong> Y. ruckeri isolate 1521 was isolated by anion-exchange chromatography on<br />

NucleoBond Xtra Midi columns (Macherey-Nagel, Düren, Germany). Plasmids <strong>of</strong> Escherichia coli<br />

V517 served as size standards (Macrina et al., 1978). An additional size estimation <strong>of</strong> the plasmid was<br />

done by summing up the sizes <strong>of</strong> fragments obtained with the restriction enzymes DraI, EcoRI, EcoRV,<br />

HindIII or PstI.<br />

The plasmid <strong>of</strong> isolate 1521 was transformed into E. coli JM107 by the CaCl 2 method as<br />

described elsewhere (Kehrenberg and Schwarz, 2001). Luria-Bertani (LB) agar plates supplemented<br />

with trimethoprim (30 µg/mL) were used to select for transformants. Transformants were checked for<br />

their plasmid content and their antimicrobial susceptibility pr<strong>of</strong>ile. Moreover, the aformentioned PCR<br />

assays were applied for the detection <strong>of</strong> sulfonamide, streptomycin and trimethoprim resistance genes.<br />

For the characterization <strong>of</strong> the resistance region, a 5,021 bp segment <strong>of</strong> the transformed plasmid (named<br />

pYR1521) was sequenced by primer-walking starting in the dfrA14 gene. The segment was deposited at<br />

the EMBL database under accession no. HG423538. Comparative analysis <strong>of</strong> nucleotide and amino acid<br />

sequences was performed using BLAST at the National Center for Biotechnology Information (NCBI)<br />

site (www.ncbi.nlm.nih.gov/BLAST/).<br />

116


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

3. Results<br />

3.1 Distribution <strong>of</strong> MIC values among Y. ruckeri isolates<br />

According to the recommendations given in the CLSI guideline VET04-A (CLSI, 2006b), the<br />

microtitre plates should be incubated at 22 2 °C and the results should be read after either 24-28 h or<br />

after 44-48 h. As far as CLSI-approved QC ranges for the antimicrobial agents tested were available, the<br />

MICs obtained for the QC strain E. coli ATCC ® 25922 were within these QC ranges (Supplemental<br />

table S1). A trend towards higher MICs after prolonged incubation for 48 h was observed for most <strong>of</strong><br />

the antimicrobial agents tested (Table 1). Since only a slight increase – if at al – was seen in the MIC<br />

values after 28 h in comparison to those read after 24 h, the MIC data for the 28 h incubation are not<br />

shown in Table 1. Table 1 also shows the calculated MIC 50 and MIC 90 after 24 h and 48 h. The lowest<br />

MICs were found for cefquinome, cefotaxime and enr<strong>of</strong>loxacin. For most <strong>of</strong> the antimicrobial agents<br />

tested, a unimodal distribution <strong>of</strong> MICs was observed. In contrast, for enr<strong>of</strong>loxacin and nalidixic acid a<br />

bimodal MIC distribution was observed with one subpopulation (n=11) showing low enr<strong>of</strong>loxacin and<br />

nalidixic acid MICs <strong>of</strong> 0.008-0.015 µg/mL and 0.25-0.5 µg/mL, respectively, and a larger subpopulation<br />

(n=72) exhibiting elevated enr<strong>of</strong>loxacin and nalidixic acid MICs <strong>of</strong> 0.06-0.25 µg/mL and 8-64 µg/mL,<br />

respectively. All isolates that showed low MICs for enr<strong>of</strong>loxacin also showed low MICs for nalidixic<br />

acid. Only a single isolate, no. 1521, which was isolated from rainbow trout with clinical signs <strong>of</strong> ERM<br />

in 2011 in Hessen, exhibited elevated MICs for sulfamethoxazole ( 1024 µg/mL) and for the<br />

combination sulfamethoxazole/trimethoprim ( 32/608 µg/mL). Three isolates showed increased MICs<br />

<strong>of</strong> ≥ 32 µg/mL for colistin.<br />

117


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

3.2 Analysis <strong>of</strong> the quinolone resistance-determining region (QRDR) <strong>of</strong> the gyrA gene<br />

Thirteen Y. ruckeri isolates were chosen according to motility, fish and geographic origins, year <strong>of</strong><br />

isolation and PFGE patterns. Sequences <strong>of</strong> the QRDR <strong>of</strong> the gyrA gene <strong>of</strong> these 13 isolates revealed that<br />

the subpopulation (n=3) with lower (fluoro)quinolone MICs showed the non-mutated gyrA wildtype<br />

QRDR sequence. In contrast, four different single bp mutations were identified among the ten<br />

representative isolates <strong>of</strong> the subpopulation with the elevated (fluoro)quinolone MICs (Table 2). Three<br />

single bp mutations resulted in single amino acid substitutions at position 83 <strong>of</strong> the GyrA protein<br />

(according to E. coli numbering). Two <strong>of</strong> them resulted in the same amino acid substitutions: AGC <br />

CGC or AGC AGA (Ser83 Arg83). The third mutation, AGCATC, resulted in a Ser83 Ile83<br />

exchange. A fourth mutation (GAC TAC) led to a substitution at position 87 (Asp87 Tyr87).<br />

Depending on the mutation, the MICs increased 8- to 32-fold for enr<strong>of</strong>loxacin or 32- to 128-fold for<br />

nalidixic acid (Table 2).<br />

3.3 Detection <strong>of</strong> plasmid-borne antimicrobial resistance genes and sequence analysis<br />

The resistance genes sul2, strB and dfrA14 and a larger amplicon for the strA gene were detected<br />

in Y. ruckeri isolate 1521. Linkage PCR assays suggested the insertion <strong>of</strong> the dfrA14 gene into the strA<br />

gene <strong>of</strong> a sul2-strA-strB gene cluster. The same PCR results and sulfamethoxazole MIC 1024 µg/mL<br />

and sulfamethoxazole/trimethoprim MIC 32/608 µg/mL were obtained for both the Y. ruckeri 1521<br />

isolate and the E. coli JM107 transformant that carried plasmid pYR1521. The size <strong>of</strong> this plasmid was<br />

estimated to be ~8.9 kb. Sequence analysis <strong>of</strong> the 5,021-bp segment confirmed the insertion <strong>of</strong> a 568 bp<br />

dfrA14 gene cassette into the strA gene, which accounted for the larger amplicons detected by the PCR<br />

118


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

assays for the strA gene and the linkage <strong>of</strong> the genes sul2-strA. The strA-strB genes encode<br />

aminoglycoside phosphotransferases which confer resistance to streptomycin. This has also been<br />

confirmed by disc diffusion with a 10 µg streptomycin disc (data not shown). However, in this case due<br />

to the inactivation <strong>of</strong> the strA gene by the insertion <strong>of</strong> dfrA14 gene cassette this resistance phenotype<br />

was not expressed. The sul2 gene encodes a sulfonamide-resistant dihydropteroate synthase and the<br />

dfrA14 gene a trimethoprim-resistant dihydr<strong>of</strong>olate reductase. Further sequence analysis identified,<br />

upstream <strong>of</strong> sul2, the partial sequence <strong>of</strong> a repC gene for a replication initiation protein. Downstream <strong>of</strong><br />

the strB gene, an inverted repeat (IR R ) <strong>of</strong> transposon Tn5393 and the rcr2 gene for a remnant <strong>of</strong> a<br />

truncated replication protein <strong>of</strong> the common region element ISCR2 (formerly designated CR2 element)<br />

were found (Yau et al. 2010; Anantham and Hall, 2012). The partial sequence <strong>of</strong> pYR1521 (from<br />

position 1 to 3,810) including the genes sul2, strA, dfrA14, strB and the IR R <strong>of</strong> Tn5393 showed<br />

99-100 % identity to sequences previously found in plasmids from E. coli obtained in Nigeria<br />

(pSTOJO1, accession number AJ313522.1), Kenya (accession number AJ884725.1), Australia<br />

(pCERC1, accession number JN012467.1), and in a plasmid <strong>of</strong> an uncultured bacterium from a<br />

wastewater treatment plant in Germany (pRSB206, accession number JN102344) (Fig. 1). The 100 %<br />

identity was seen with the sequence <strong>of</strong> E. coli plasmid pCERC1, but according to the sequences flanking<br />

this 3810-bp segment, the plasmids pYR1521 and pCERC1 do not seem to be related otherwise.<br />

Plasmid pYR1521 seems to be more related to the Salmonella Typhimurium plasmid pSCR15<br />

(accession number GQ379901) which, however, lacks the insertion <strong>of</strong> the dfrA14 gene cassette into the<br />

strA gene.<br />

119


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

4. Discussion<br />

The CLSI guideline VET04-A (CLSI, 2006b) states that in general antimicrobial susceptibility<br />

testing <strong>of</strong> bacteria from aquatic animals should be performed either at 22 2 °C with a reading <strong>of</strong> the<br />

results after 24-28 h or after 44-48 h. In case an isolate shows poor growth at 22 2 °C, incubation at 28<br />

2 °C with a reading <strong>of</strong> the results after 24-28 h is also an approved alternative. In the case <strong>of</strong> Y. ruckeri,<br />

we observed good growth after incubation <strong>of</strong> the inoculated microtitre plates for 24 h at 22 2 °C. The<br />

MIC values changed only marginally – if at all – when the incubation time was extended by 4 h whereas<br />

more pronounced changes in the MICs were seen for most antimicrobial agents tested after incubation<br />

for 48 h. As such, a suggestion is made that Y. ruckeri should be tested at 22 2 °C – which also more<br />

closely resembles the real life situation than 28 2 °C – and that incubation times longer than 24 h are<br />

not necessary.<br />

Although the Y. ruckeri isolates included in this study have been collected during several years<br />

from a large number <strong>of</strong> geographically different locations in North West Germany and their molecular<br />

and phenotypic analysis identified distinct differences (Huang et al., 2013), they showed unimodal MIC<br />

distributions and low MICs for most <strong>of</strong> antimicrobial agents tested. However, due the lack <strong>of</strong> clinical<br />

breakpoints applicable to Y. ruckeri (CLSI, 2010), the isolates cannot be assigned to any <strong>of</strong> the three<br />

categories „susceptible‟, „intermediate‟ or „resistant‟. Nevertheless, the data presented in this study may<br />

be used in the future to establish epidemiological cut-<strong>of</strong>f values for Y. ruckeri. For this, a larger number<br />

<strong>of</strong> isolates obtained from different countries, and tested following the same standardized methodology<br />

will be necessary.<br />

120


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Three isolates showing elevated colistin MICs belonged to the typing groups 25 and 26, as they<br />

clustered into PFGE pattern B and showed different band patterns in repetitive sequence-based PCR and<br />

API 20E pr<strong>of</strong>iles, compared other isolates in previous study. (Huang et al., 2013). A clear bimodal<br />

distribution <strong>of</strong> the MICs was observed for (fluoro)quinolones, as well as sulfamethoxazole and<br />

sulfamethoxazole/trimethoprim. The isolates <strong>of</strong> the subpopulation with the higher MICs were likely to<br />

harbor acquired or mutational resistance mechanisms to the aforementioned antimicrobial agents. This<br />

assumption was confirmed by the molecular analysis <strong>of</strong> representative isolates for resistance-mediating<br />

mutations or resistance genes. Resistances are known to develop quickly under the selective pressure<br />

imposed by the application <strong>of</strong> the respective antimicrobial agents. This can occur either by mutational<br />

changes in the target genes or by the acquisition <strong>of</strong> mobile genetic elements that carry resistance genes.<br />

It has been reported that quinolones and sulfonamides have been used for the treatment <strong>of</strong> ERM<br />

outbreaks in aquaculture (Bosse and Post 1983; Bullock et al., 1983; Horsberg et al., 1997;<br />

Treves-Brown, 2000).<br />

Gibello and colleagues (2004) described the first substitution in the QRDR region <strong>of</strong> the GyrA<br />

protein in Y. ruckeri isolates. It was a single amino acid substitution, Ser-83 Arg-83 (E. coli<br />

numbering) and the isolates exhibited reduced susceptibility to nalidixic acid and oxolinic acid.<br />

Recently, the same mutation was observed with <strong>of</strong> Y. ruckeri in Norway. However, no mutations were<br />

found in QRDR <strong>of</strong> other target genes, such as gyrB, parC and parE (Shah et al., 2012). Substitutions in<br />

QRDR at position 83 and 87 <strong>of</strong> the GyrA protein have been commonly reported in other Gram-negative<br />

bacteria. Such substitutions are usually related to reduced susceptibility to fluoroquinolones (Eaves et al.,<br />

2004; Sihvonen et al., 2011). The substitution Ser-83 Arg-83, found in this study, has already been<br />

121


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

reported in Y. ruckeri (Shah et al., 2012) and Y. enterocolitica (Sihvonen et al., 2011). However, this is,<br />

to the best <strong>of</strong> our knowledge, the first description <strong>of</strong> the substitution Asp87 Tyr87 in Y. ruckeri.<br />

Resistance to both trimethoprim and sulfonamides is widely disseminated among all major species<br />

<strong>of</strong> bacteria. So far, many resistant bacteria have been reported worldwide, including various<br />

fish-pathogenic Aeromonas spp. (Kadlec et al., 2011) and Y. ruckeri isolates (De Grandis and Stevenson,<br />

1985; Welch et al., 2007). It has been reported that the sul2-strA-strB gene cluster arose from the<br />

transposition <strong>of</strong> transposon Tn5393 (harboring strA-strB gene cluster) in the ISCR2-sul2 region<br />

followed by a partial deletion <strong>of</strong> both Tn5393 and ISCR2 (Yau et al. 2010). Interestingly, the multidrug<br />

resistance plasmid pYR1 (accession number NC_009139.1) <strong>of</strong> Y. ruckeri YR71 harbors the strA-strB<br />

gene cluster (location 136,344 - 137,147 bp) and the ISCR2-sul2 configuration (32,085-33,635 bp)<br />

located about 102 kb apart from each other (Welch et al., 2007). So far, few reports described the<br />

dfrA14 gene cassette integrated into the strA gene and all these reports referred to Enterobacteriaceae or<br />

an uncultured bacterium (Ojo et al., 2002; Anantham and Hall, 2012; Eikmeyer et al., 2012). This is to<br />

the best <strong>of</strong> our knowledge, the first description <strong>of</strong> the presence <strong>of</strong> a dfrA14 gene cassette at this<br />

particular secondary site in Y. ruckeri. According to the sequence analysis, all <strong>of</strong> the<br />

non-integron-associated dfrA14 gene cassettes were integrated at a secondary site (GATAT) in the strA<br />

gene (Ojo et al., 2002; Kikuvi et al., 2007; Anantham and Hall, 2012; Eikmeyer et al., 2012). The highly<br />

conserved sequence <strong>of</strong> the dfrA14 gene, its 59-base element and the whole segment from sul2 gene to<br />

the IR R <strong>of</strong> Tn5393 (Fig. 1a-b) on unrelated plasmids from diverse bacteria and sources (clinical or<br />

non-clinical, human or animals) in different continents suggest a common origin and/or en bloc<br />

dissemination <strong>of</strong> the sul2-strA-dfrA14-strA-strB gene cluster. Noteworthy, in North West Germany<br />

122


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

this cluster was found on the small transferable plasmid pYR1521 from Y. ruckeri obtained in 2011<br />

from a rainbow trout and has been reported in 2012 to be located on the larger conjugative<br />

multi-resistance plasmid pRSB206 <strong>of</strong> an unculturable bacterium obtained from a wastewater treatment<br />

plant (Eikmeyer et al., 2012).<br />

5. Conclusion<br />

Incubation for 24 h at 22 ± 2 °C appears to be sufficient for susceptibility testing <strong>of</strong> Y. ruckeri.<br />

The largely unimodal MIC distribution suggests that most <strong>of</strong> the isolates tested represent the wildtype<br />

population that has not acquired resistance mechanisms. Exceptions were the isolates with elevated<br />

(fluoro)quinolone MICs that also had resistance-mediating mutations in the QRDR region <strong>of</strong> gyrA gene<br />

and a single isolate that carried plasmid-borne sul2, dfrA14, strA and strB genes. These observations<br />

confirmed that Y. ruckeri is able to develop mutations for reduced susceptibility to (fluoro)quinolone,<br />

but can also acquire plasmid-borne resistance genes.<br />

123


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Acknowledgement<br />

The authors acknowledge the generous support <strong>of</strong> Dr. W. Schäfer and D. Mock (Fish Health<br />

Service, North Rhine-Westphalia) during the sample collection in aquaculture ponds. Moreover, the<br />

authors thank Dr. S. Braune (LAVES, Lower Saxony), Dr. C. Gould (MSD Animal Heath, UK) and Dr.<br />

A. Nilz (LHL, Hessen) for providing Yersinia ruckeri isolates.<br />

Funding<br />

This study was financially supported by the Landesamt für Natur, Umwelt und Verbraucherschutz<br />

Nordrhein-Westfalen (LANUV, North Rhine-Westphalia). Y. Huang received a scholarship by the<br />

China Scholarship Council.<br />

Transparency declarations<br />

None to declare<br />

References<br />

Anantham, S., Hall, R.M., 2012. pCERC1, a small, globally disseminated plasmid carrying the dfrA 14<br />

cassette in the strA gene <strong>of</strong> the sul2-strA-strB gene cluster. Microb. Drug Resist. 18, 364-371.<br />

Angulo, F., 2000. Antimicrobial agents in aquaculture: protential impact on public health. APUA Newsl.<br />

18, 1-5.<br />

Barnard, F.M., Maxwell, A., 2001. Interaction between DNA gyrase and quinolones: effects <strong>of</strong> alanine<br />

mutations at GyrA subunit residues Ser83 and Asp87. Antimicrob. Agents Chemother. 45,<br />

1994-2000.<br />

124


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Bosse, M.P., Post, G., 1983. Tribrissen and tiamulin for control <strong>of</strong> enteric redmouth disease. J. Fish Dis. 6,<br />

27-32.<br />

Bullock, G.L., Maestrone, G., Starliper, C., Schill, B., 1983. Potentiated sulfonamide therapy <strong>of</strong> enteric<br />

redmouth disease. Can. J. Fish. Aquat. Sci. 40, 101-102.<br />

Bywater, R., Silley, P., Simjee, S., 2006. Antimicrobial breakpoints – definitions and conflicting<br />

requirements. Vet. Microbiol. 118, 158-159.<br />

CLSI. 2006a. Methods for antimicrobial disk susceptibility testing <strong>of</strong> bacteria isolated from aquatic<br />

animals; approved guideline. CLSI document VET03-A. Clinical and Laboratory Standards<br />

<strong>Institute</strong>; Wayne, PA, USA.<br />

CLSI. 2006b. Methods for broth dilution susceptibility testing <strong>of</strong> bacteria isolated from aquatic animals;<br />

approved guideline. CLSI document VET04-A. Clinical and Laboratory Standards <strong>Institute</strong>;<br />

Wayne, PA, USA.<br />

CLSI. 2010. Performance standards for antimicrobial susceptibility testing <strong>of</strong> bacteria isolated from<br />

aquatic animals; first informational supplement. CLSI document VET03/VET04-S1. Clinical and<br />

Laboratory Standards <strong>Institute</strong>; Wayne, PA, USA.<br />

De Grandis, S.A., Stevenson, R.M.W., 1985. Antimicrobial susceptibility patterns and R<br />

plasmid-mediated resistance <strong>of</strong> the fish pathogen Yersinia ruckeri. Antimicrob. Agents Chemother.<br />

27, 938-942.<br />

Eaves, D.J., Randall, L., Gray, D.T., Buckley, A., Woodward, M.J., White, A.P., Piddock, L.J.V., 2004.<br />

Prevalence <strong>of</strong> mutations within the Quinolone Resistance-Determining Region <strong>of</strong> gyrA, gyrB, parC,<br />

and parE and association with antibiotic resistance in quinolone-resistant Salmonella enterica.<br />

Antimicrob. Agents Chemother. 48, 4012-4015.<br />

Eikmeyer, F., Hadiati, A., Szczepanowski, R., Wibberg, D., Schneiker-Bekel, S., Rogers, L.M., Brown,<br />

C.J., Top, E.M., Puhler, A., Schluter, A., 2012. The complete genome sequences <strong>of</strong> four new IncN<br />

plasmids from wastewater treatment plant effluent provide new insights into IncN plasmid<br />

diversity and evolution. Plasmid 68, 13-24.<br />

125


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Feßler, A.T., Olde Riekerink, R.G., Rothkamp, A., Kadlec, K., Sampimon, O.C., Lam, T.J., Schwarz, S.,<br />

2012. Characterization <strong>of</strong> methicillin-resistant Staphylococcus aureus CC398 obtained from<br />

humans and animals on dairy farms. Vet. Microbiol. 160, 77-84.<br />

Gebreyes, W.A., Altier, C., 2002. Molecular characterization <strong>of</strong> multidrug-resistant Salmonella enterica<br />

subsp. enterica serovar Typhimurium isolates from swine. J. Clin. Microbiol. 40, 2813-2822.<br />

Gibello, A., Porrero, M.C., Blanco, M.M., Vela, A.I., Liébana, P., Moreno, M.A., Fernández-Garayzábal,<br />

J.F., Domínguez, L., 2004. Analysis <strong>of</strong> the gyrA gene <strong>of</strong> clinical Yersinia ruckeri isolates with<br />

reduced susceptibility to quinolones. Appl. Environ. Microbiol. 70, 599-602.<br />

Grape, M., Sundström, L., Kronva, G., 2003. Sulphonamide resistance gene sul3 found in Escherichia<br />

coli isolates from human sources. J. Antimicrob. Chemother. 52, 1022-1024.<br />

Horsberg, T.E., Martinsen, B., Sandersen, K., Zernichow, L., 1997. Potentiated sulfonamides: In vitro<br />

inhibitory effect and pharmacokinetic properties in atlantic swalmon in seawater. J. Aqua. Ani.<br />

Health. 9, 203-210.<br />

Huang, Y., Runge, M., Michael, G.B., Schwarz, S., Jung, A., Steinhagen, D., 2013. Biochemical and<br />

molecular heterogeneity among isolates <strong>of</strong> Yersinia ruckeri from rainbow trout (Oncorhynchus<br />

mykiss, Walbaum) in North West Germany. BMC Vet. Res. 9,<br />

215; doi:10.1186/1746-6148-9-215 (published: 21 October 2013).<br />

Kadlec, K., Kehrenberg, C., Schwarz, S., 2005. Molecular basis <strong>of</strong> resistance to trimethoprim,<br />

chloramphenicol and sulphonamides in Bordetella bronchiseptica. J. Antimicrob. Chemother. 56,<br />

485-490.<br />

Kadlec, K., Czapiewski, E.v., Kaspar, H., Wallmann, J., Michael, G.B., Steinacker, U., Schwarz, S., 2011.<br />

Molecular basis <strong>of</strong> sulfonamide and trimethoprim resistance in fish-pathogenic Aeromonas isolates.<br />

Appl. Environ. Microbiol. 77, 7147-7150.<br />

126


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Kehrenberg, C., Schwarz, S., 2001. Occurence and linkage <strong>of</strong> genes coding for resistance to<br />

sulphonamides, streptomycin and chloramphenicol in bacteria <strong>of</strong> the genera Pasteurella and<br />

Mannheimia. FEMS Microbiol. Lett. 205, 283-290.<br />

Kikuvi, G.M., Schwarz, S., Ombui, J.N., Mitema, E.S., Kehrenberg, C., 2007. Streptomycin and<br />

chloramphenicol resistance genes in Escherichia coli isolates from cattle, pigs, and chicken in<br />

Kenya. Microb. Drug Resist. Spring 13, 62-68.<br />

Macrina, F.L., Kopecko, D.J., Jones, D.J., Ayers, D.J., McCowen, S.M., 1978. A multiple<br />

plasmid-containing Escherichia coli strain: convenient source <strong>of</strong> size reference plasmid molecules.<br />

Plasmid 1, 417-420.<br />

Miller, R.A., Walker, R.D., Baya, A., Clemens, K., Coles, M., Hawke, J.P., Henricson, B.E., Hsu, H.M.,<br />

Mathers, J.J., Oaks, J.L., Papapetropoulou, M., Reimschuessel, R., 2003. Antimicrobial<br />

susceptibility testing <strong>of</strong> aquatic bacteria: quality control disk diffusion ranges for Escherichia coli<br />

ATCC 25922 and Aeromonas salmonicida subsp. salmonicida ATCC 33658 at 22 and 28 C. J.<br />

Clin. Microbiol. 41, 4318-4323.<br />

Miller, R.A., Walker, R.D., Carson, J., Coles, M., Coyne, R., Dalsgaad, I., Gieseker, C., Hsu, H.M.,<br />

Mathers, J.J., Papapetropoulou, M., Petty, B., Teitzel, C., Reimschuessel, R., 2005. Standardization<br />

<strong>of</strong> a broth microdilution susceptibility testing method to determine minimum inhibitory<br />

concentrations <strong>of</strong> aquatic bacteria. Dis. Aquat. Org. 64, 211-222.<br />

Noga, E.J., 2010. Fish Disease: Diagnosis and Treatment. 2. Edition. Wiley-Blackwell, Iowa, pp.<br />

197-199.<br />

Ojo, K.K., Kehrenberg, C., Schwarz, S., Odelola, H.A., 2002. Identification <strong>of</strong> a complete dfrA14 gene<br />

cassette integrated at a secondary site in a resistance plasmid <strong>of</strong> uropathogenic Escherichia coli<br />

from Nigeria. Antimicrob. Agents Chemother. 46, 2054-2055.<br />

Rigos, G., Troisi, G.M., 2005. Antibacterial agents in Mediterranean finfish farming: A synopsis <strong>of</strong> drug<br />

pharmacokinetics in important euryhaline fish species and possible environmental implications.<br />

Rev. Fish. Biol. Fisher. 15, 53-73.<br />

127


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Rodgers, C.J., 2001. Resistance <strong>of</strong> Yersinia ruckeri to antimicrobial agents in vitro. Aquaculture. 196,<br />

325-345.<br />

Schwarz, S., Silley, P., Simjee, S., Woodford, N., van Duijkeren, E., Johnson, AP., Gaastra, W., 2010.<br />

Assessing the antimicrobial susceptibility <strong>of</strong> bacteria obtained from animals. Vet. Microbiol. 141,<br />

1-4.<br />

Shah, S.Q.A., Karatas, S., Nilsen, H., Steinum, T.M., 2012. Characterization and expression <strong>of</strong> the gyrA<br />

gene from quinolone resistant Yersinia ruckeri strains isolated from Atlantic salmon (Slamo salar<br />

L.) in Norway. Aquaculture 350, 37-41.<br />

Sihvonen, L.M., Toivonen, S., Haukka, K., Kuusi, M., Skurnik, M., Siitonen, A., 2011. Multilocus<br />

variable-number tandem-repeat analysis, pulsed-field gel electrophoresis, and antimicrobial<br />

susceptibility patterns in discrimination <strong>of</strong> sporadic and outbreak-related strains <strong>of</strong> Yersinia<br />

enterocolitica. BMC Microbiol. 11, 1-10.<br />

Treves-Brown, K.M., 2000. Applied Fish Pharmacology. Kluwer Academic publishers, London, pp.118.<br />

Welch, T.J., Fricke, W.F., McDermott, P.F., White, D.G., Rosso, M.L., Rasko, D.A., Mammel, M.K.,<br />

Eppinger, M., Rosovitz, M.J., Wagner, D., Rahalison, L., Leclerc, J.E., Hinshaw, J.M., Lindler,<br />

L.E., Cebula, T.A., Carniel, E., Ravel, J., 2007. Multiple antimicrobial resistance in plague: an<br />

emerging public health risk. PLoS ONE 2, e309.<br />

Yau, S., Liu, X., Djordjevic, S.P., Hall, R.M., 2010. RSF1010-like plasmids in Australian Salmonella<br />

enterica serovar Typhimurium and origin <strong>of</strong> their sul2-strA-strB antibiotic resistance gene cluster.<br />

Microb. Drug Resist. 16, 249-252.<br />

128


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Table 1:<br />

MIC distribution <strong>of</strong> antimicrobial agents, MIC 50 and MIC 90 values <strong>of</strong> Yersinia ruckeri isolates from North West Germany and the type<br />

strain Y. ruckeri DSM 18506 (n=83) after incubation times <strong>of</strong> 24 and 48 h at 22 ± 2 °C. The white area between the two grey-shaded<br />

areas indicates the range <strong>of</strong> test concentrations. The numbers in the right-hand grey-shaded area indicate the numbers <strong>of</strong> isolates that<br />

grew in the highest test concentration.<br />

Antimicrobial Incubation Number <strong>of</strong> isolates with MIC (µg/mL) <strong>of</strong>:<br />

MIC (µg/mL)<br />

Agent a time (h) ≤0.008 0.015 0.03 0.06 0.12 0.25 0.5 1 2 4 8 16 32 64 128 256 512 ≥1024 MIC 50 MIC 90<br />

NAL 24 9 2 3 65 4 16 16<br />

48 7 4 1 31 31 1 16 32<br />

ENR 24 8 3 18 52 2 0.12 0.12<br />

48 7 4 3 66 3 0.12 0.12<br />

SMX 24 82 1 16 16<br />

48 82 1 16 16<br />

SXT 24 29 49 4 1 0.06 0.06<br />

48 8 69 5 1 0.06 0.06<br />

APR 24 19 62 2 4 4<br />

48 7 48 27 1 4 8<br />

GEN 24 26 57 1 1<br />

48 14 50 20 1 2<br />

PEN 24 1 61 21 1 16 32<br />

48 9 71 3 32 32<br />

AMP 24 70 12 1 1 2<br />

48 22 59 2 2 2<br />

AUG 24 2 81 2 2<br />

129


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

48 62 21 2 4<br />

Antimicrobial Incubation Number <strong>of</strong> isolates with MIC (µg/mL) <strong>of</strong>:<br />

MIC (µg/mL)<br />

Agent a time (h) ≤0.008 0.015 0.03 0.06 0.12 0.25 0.5 1 2 4 8 16 32 64 128 256 512 ≥1024 MIC 50 MIC 90<br />

CEP 24 83 32 32<br />

48 73 10 32 32<br />

XNL 24 70 13 0.25 0.25<br />

48 11 70 2 0.5 0.5<br />

FOT 24 77 6 0.06 0.06<br />

48 27 55 1 0.12 0.12<br />

FOP 24 2 70 11 0.25 0.25<br />

48 1 31 49 2 0.5 0.5<br />

CEQ 24 3 76 4 0.06 0.06<br />

48 2 58 23 0.06 0.12<br />

IMP 24 83 0.12 0.12<br />

48 81 2 0.12 0.12<br />

TIA 24 53 25 5 16 32<br />

48 4 65 12 2 32 64<br />

SPI 24 48 29 6 64 128<br />

48 15 46 22 128 256<br />

TIL 24 36 42 4 1 64 64<br />

48 19 46 16 2 64 128<br />

TUL 24 6 62 10 5 16 32<br />

48 39 36 7 1 32 32<br />

CHL 24 11 72 4 4<br />

48 80 3 4 4<br />

FFN 24 71 11 1 2 2<br />

48 20 62 1 4 4<br />

130


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

COL 24 59 19 2 3 0.25 0.5<br />

48 58 20 2 3 0.25 0.5<br />

TET 24 21 62 1 1<br />

48 67 16 1 2<br />

DOX 24 26 57 0.5 0.5<br />

48 82 1 0.5 0.5<br />

a NAL: nalidixic acid, ENR: enr<strong>of</strong>loxacin, SMX: sulfamethoxazole, SXT: sulfamethoxazole/trimethoprim (19: 1), APR: apramycin, GEN: gentamicin, PEN: penicillin,<br />

AMP: ampicillin, AUG: amoxicillin/clavulanic acid (2:1), CEP: cefalotin, XNL: cefti<strong>of</strong>ur, FOT: cefotaxime, FOP: cefoperazone, CEQ: cefquinome, IMP: imipenem,<br />

TIA: tiamulin, SPI: spiramycin, TIL: tilmicosin, TUL: tulathromycin, CHL: chloramphenicol, FFN: florfenicol, COL: colistin, TET: tetracycline, DOX: doxycycline.<br />

131


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

Table 2:<br />

Characteristics <strong>of</strong> Y. ruckeri isolates, their (fluoro)quinolone<br />

MICs and their GyrA protein substitutions in the QRDR<br />

Isolate Fish species Isolation Origin <strong>of</strong> MIC (µg/mL) GyrA substitution e<br />

no.<br />

a<br />

(year) isolates b ENR c NAL d (gyrA mutations)<br />

111-1 NS 2005 LS ≤0.008 0.25 -<br />

L-1 RT 2011 NRW ≤0.008 0.25 -<br />

285-1 Pike 2005 LS 0.015 0.5 -<br />

115-1 RT 2005 LS 0.06 8 Asp87Tyr<br />

201-1 RT 2008 LS 0.06 32 Ser83Arg (CGC)<br />

72-2 RT 2005 LS 0.12 16 Ser83Arg (CGC)<br />

SRT2-2 RT 2012 NRW 0.12 16 Ser83Arg (CGC)<br />

3279 RT 2010 HS 0.12 32 Ser83Arg (AGA)<br />

339-1 RT 2005 LS 0.12 16 Ser83Arg (AGA)<br />

B1 RT 2011 NRW 0.12 32 Ser83Arg (AGA)<br />

lay3 RT 2011 NRW 0.12 16 Ser83Arg (AGA)<br />

204-1 Brown trout 2004 LS 0.25 32 Ser83Ile<br />

70-1 Koi 2005 LS 0.25 32 Ser83Ile<br />

a<br />

NS: not specified, RT: rainbow trout; b LS: Lower Saxony, NRW: North Rhine-Westphalia,<br />

HS: Hessen; c ENR: enr<strong>of</strong>loxacin; d NAL: nalidixic acid; e Numbering system according to<br />

Escherichia coli.<br />

132


Chapter 6 Pheno- and genotypic analysis <strong>of</strong> antimicrobial resistance properties <strong>of</strong> Y. ruckeri<br />

(a)<br />

Integration site <strong>of</strong> dfrA14 gene cassette<br />

into the strA gene<br />

-G ATAT-<br />

pSCR15<br />

pYR1521<br />

sul2 strA<br />

strB rcr2 (5„ end)<br />

sul2<br />

strB<br />

rcr2 (5„ end)<br />

pCERC1<br />

sul2<br />

strB<br />

pSTOJO1<br />

rcr2<br />

(patial sequence)<br />

pRSB206<br />

sul2<br />

strB<br />

500<br />

(b)<br />

strA<br />

(5„ end)<br />

dfrA14<br />

strA<br />

(3„ end)<br />

2R<br />

GCAGCGATA-GTTAGC-GCC-TTTTTCC--GCATCACGCGC<br />

1R<br />

stop | ||||| |||||| ||| |||||| ||| | ||| T<br />

TTAACCCAGGATGAGAACCTTGAAA- -GGTTAACAAAGCTATGCAATCGACGGCAAAAAGCTTCGT--T-CGCT<br />

RBS start<br />

dfrA14<br />

1L<br />

2L<br />

59-base element<br />

dfrA14 gene cassette<br />

Fig 1. (a) Structural organization and alignment <strong>of</strong> segments <strong>of</strong> plasmids pSCR15<br />

(accession number GQ379901), pYR1521 (accession number HG423538), pCERC1<br />

(accession number JN012467.1), pSTOJO1 (accession number AJ313522.1) and pRSB206<br />

(accession number JN102344). The arrows represent the coding reading frames <strong>of</strong> genes<br />

sul2, strA, dfrA14, strB and rcr2. The arrowheads indicate the direction <strong>of</strong> transcription.<br />

The gray areas show matches between the segments, 99 or 100% identity was found in these<br />

areas. The vertical black bars represent the IR R <strong>of</strong> Tn5393. Sequence <strong>of</strong> strA gene (GATAT)<br />

involved in the integration <strong>of</strong> dfrA14 gene cassette. (b) Structure <strong>of</strong> the dfrA14 gene cassette<br />

inserted into the strA gene. Coding reading frames <strong>of</strong> dfrA14 gene is represented by an<br />

arrow. The ribosome biding site (RBS), start and stop codons are underlined. The 59-base<br />

element is shown in boldface and the integrase biding domains (1R, 2R, 2L and 1L) are<br />

indicated by arrows.<br />

133


134


Chapter 7. General Discussion<br />

135


Chapter 7 General discussion<br />

General Discussion<br />

This project investigated Yersinia ruckeri collected from trout hatcheries in North West<br />

Germany from 2004 to 2012 with regard to (1) their genetic and phenotypic diversity<br />

correlated with epidemiological factors, and (2) their resistance to antimicrobial substances.<br />

Since the first occurrence <strong>of</strong> non-motile Y. ruckeri isolates was reported (AUSTIN et al. 2003),<br />

isolates with this biological trait have already spread over most European countries (AUSTIN<br />

et al. 2003; FOUZ et al. 2006; BASTARDO et al. 2012), as well as in American countries,<br />

such as USA (ARIAS et al. 2007) and Chile(BASTARDO et al. 2011). In recent years,<br />

non-motile strains caused outbreaks <strong>of</strong> enteric redmouth disease in the winter period in North<br />

West Germany and attracted the attention <strong>of</strong> the local fish health department.<br />

7.1 Non-motile Yersinia ruckeri<br />

Y. ruckeri infection were usually caused by the motile „Hagerman‟ strain (DAVIES 1991).<br />

This „Hagerman‟ strain is belonged to serotype O1 (DAVIES 1991), and provided the basis<br />

for the development <strong>of</strong> successful vaccines since the 1970ies (STEVENSON 1997). These<br />

vaccines were used effectively to control outbreaks caused by motile Y. ruckeri for three<br />

decades, until the non-motile Y. ruckeri variant emerged and was reported have a reduced<br />

sensitivity to immersion vaccination (AUSTIN et al. 2003). The new non-motile variant,<br />

belonging to serotype O1 as well, established a new biotype for Y. ruckeri, biotype 2, and<br />

caused outbreaks in previously vaccinated ponds (AUSTIN et al. 2003). It was believed that<br />

the non-motile strain might have evolved from a related motile one (BASTARDO et al. 2012).<br />

Although the Hagerman strain and the non-motile variant belong to the same serotype, they<br />

made up different clonal groups, clone 2 for the non-motile variant and clone 5 for the<br />

Hagerman strain, on the basis <strong>of</strong> biotype and outer membrane protein (OMP) typing<br />

(DAVIES 1991). There was no report in the 1980s about non-motile isolates from other<br />

serotypes (DAVIES and FRERICHS 1989). However, recently, biotype 2 isolates were also<br />

identified within other seroptypes besides O1 (J W TINSLEY et al. 2011). Non-motile strains<br />

were dominant in North West Germany. In field study <strong>of</strong> the present investigation, motile<br />

strains were found present only in fish farm L.<br />

Bacteria from the non-motile strain, as the name indicates, are lacking motility due to the<br />

absence <strong>of</strong> flagella (AUSTIN et al. 2003). Unlike biotype 1, these biotype 2 isolates cannot<br />

hydrolyze tween 80 and tween 20 (DAVIES and FRERICHS 1989). Thus, tween 80<br />

136


Chapter 7 General discussion<br />

hydrolysis is used as one <strong>of</strong> the methods to identify non-motile isolates. Moreover, the<br />

lipopolysaccharide (O antigen) type <strong>of</strong> non-motile variants was found different from other<br />

motile isolates (J W TINSLEY et al. 2011).<br />

7.2 Analysis <strong>of</strong> typing methods<br />

Different typing methods were applied in this study in order to analyse intraspecies diversity<br />

<strong>of</strong> Y. ruckeri isolates from North West Germany.<br />

The API system is a simplified biochemical identification kit for bacteria. Three API kits are<br />

available for identifying enterobacteria, namely, a screening kit <strong>of</strong> 10 tests (10S), a basic set<br />

<strong>of</strong> 20 tests (20E), and <strong>of</strong> 50 tests (50E)(HOLMES et al. 1978), among which the API 20E is<br />

used <strong>of</strong>ten for the identification <strong>of</strong> Y. ruckeri. Although the non-motile strain found in Spain<br />

shared the same API 20E pr<strong>of</strong>ile <strong>of</strong> 5307100 (FOUZ et al. 2006) with some non-motile<br />

isolates in the study, it is difficult to separate non-motile strains from motile only according to<br />

API 20E pr<strong>of</strong>iles, because non-motile strains do not exhibit specific API 20E pr<strong>of</strong>iles (J W<br />

TINSLEY et al. 2011). In this study during biochemical identification, the API pr<strong>of</strong>ile<br />

5307100 was exhibited by some motile isolates as well. Eight different numeric pr<strong>of</strong>iles were<br />

obtained from the North West German isolates with a discriminatory power <strong>of</strong> 0.763.<br />

PCR-based DNA fingerprinting methods for a differentiation <strong>of</strong> microorganisms have been<br />

established by using different primers. Repetitive-sequence-based PCRs (rep-PCR) are based<br />

on the binding <strong>of</strong> primers to specific regions <strong>of</strong> the DNA and when this binding happens in<br />

the right orientation with an optimum distance, species- or strain- specific amplicons could be<br />

generated and observed (MALATHUM et al. 1998). So far, different rep-PCRs were used in<br />

the evaluation <strong>of</strong> the intraspecific diversity <strong>of</strong> particular bacteria and were considered as a<br />

useful molecular technique (VERSALOVIC et al. 1991; MOHAPATRA and MAZUMDER<br />

2008). For instance, the repetitive extragenic palindromic-PCR coupled with (GTG) 5 primers<br />

[(GTG) 5 -PCR] was reported as a complementary molecular tool for the rapid determination <strong>of</strong><br />

Escherichia coli (MOHAPATRA et al. 2008) and Enterococcus spp. (ŠVEC et al. 2005).<br />

Enterobacterial repetitive intergenic consensus sequences (ERIC-PCR) generate multiple<br />

distinct amplification products ranging from around 50bp to 3000bp (DI-GIOVANNI et al.<br />

1999) and the ERIC elements are distributed throughout the extragenic regions <strong>of</strong> many<br />

Gram-negative bacterial genomes (HULTON et al. 1991). Other two rep-PCRs used in this<br />

study to differentiate Y. ruckeri isolates were repetitive extragenic palindromic PCR<br />

137


Chapter 7 General discussion<br />

(REP-PCR) and BOX-PCR, which was based on a 154-bp BOX element (KOETH et al.<br />

1995). REP and BOX elements may play an important role in the organization <strong>of</strong> the bacterial<br />

genome (KRAWIEC and RILEY 1990; LUPSKI and WEINSTOCK 1992). In this study,<br />

ERIC-PCR turned out to be the most effective rep-PCR to investigate Y. ruckeri isolates with<br />

several distinct bands. However, the discriminatory index <strong>of</strong> rep-PCRs was lower than 0.2,<br />

indicating the intraspecies diversity was not clearly shown via rep-PCRs.<br />

Although the sensitivity and the discriminatory power <strong>of</strong> pulsed-field electrophoresis (PFGE)<br />

rely on the bacterial species and the restriction enzyme used, it could permit PFGE to be a<br />

useful tool for clustering and differentiating many bacterial pathogens (SKOV et al. 1995;<br />

GARCIA et al. 2000; ARAL et al. 2007). Compared with rep-PCRs, pulsed-field<br />

electrophoresis with the restriction enzyme NotⅠ is a powerful tool to differentiate Y. ruckeri<br />

isolates. When all the typing methods combined together, a more systematic category was<br />

formed, including 27 different groups, providing the basis for the selection <strong>of</strong> isolates to be<br />

analysed for cytotoxicity in vitro.<br />

7.3 Relationship among phenotypic and genetic traits and epidemiological<br />

characteristics<br />

Biochemically, a fatty acid is an either saturated or unsaturated carboxylic acid with a long<br />

aliphatic chain. As one phenotypic trait, fatty acids <strong>of</strong> microorganisms are chemically<br />

complex and metabolically active (O'LEARY 1962). Although the fatty acid composition<br />

differs markedly between lab-cultured bacteria and those growing in the natural environment<br />

(O'LEARY 1962), it doesn‟t hinder the analysis <strong>of</strong> the fatty acid composition to become a tool<br />

to identify specific bacteria. There were several reports about bacterial identification on the<br />

basis <strong>of</strong> cellular fatty acid composition (CANONICA and PISANO 1985; OSTERHOUT et al.<br />

1991; SHOEMAKER et al. 2013). Leclercq et al (LECLERCQ et al. 2000) found that the<br />

fatty acid composition <strong>of</strong> Yersinia pestis is very homogeneous. Furthermore, the ratio for<br />

12:0/16:0 and 14:0/16:0 fatty acids could be used for differentiating pathogenic Yersinia<br />

enterocolitica isolates, and for separating Yersinia pseudotuberculosis and Y. pestis strains<br />

(LECLERCQ et al. 2000). It is interesting to study the fatty acid pr<strong>of</strong>iles <strong>of</strong> North West<br />

German Y. ruckeri isolates and its connection with epidemiological parameters. From<br />

obtained in the present study, Y. ruckeri isolates shared also homogeneous fatty acid pr<strong>of</strong>iles.<br />

Considering the biochemical traits and biotypes together with the fatty acid pr<strong>of</strong>iles, no<br />

138


Chapter 7 General discussion<br />

specific relationship was found. Isolates positive for sorbitol fermentation were mainly<br />

obtained from other fish species rather than rainbow trout and all <strong>of</strong> these isolates were motile.<br />

There were only two isolates from rainbow trout found positive for sorbitol fermentation. One<br />

<strong>of</strong> these isolates was obtained during the field study and it shared the same biochemical and<br />

genetic traits as the one isolate obtained from a brown trout. Interestingly, brown trouts were<br />

also cultivated in the same farm. It was proven that Y. ruckeri isolates were not species<br />

specific for particular fish species.<br />

From previous results published on pulsotyping <strong>of</strong> Y. ruckeri, it is suggested that serotype O1<br />

isolates in UK and in mainland Europe were from different and non-overlapping<br />

subpopulations (WHEELER et al. 2009). Recently, one isolate from Germany was found<br />

dispersed from other two main clusters by using multilocus sequence typing (MLST)<br />

(BASTARDO et al. 2012), indicating that this German isolate might not be derived from the<br />

same clonal complexes as other isolates from different countries. In the present study, all<br />

isolates were local North West German ones, and variations among different isolates were<br />

still observed. There was no difference in pulsotype between biotype 1 and 2 strains. Isolates<br />

from rainbow trout clustered together in the PFGE dendrogram. Great differences were shown<br />

among isolates from different fish hosts. Most isolates in the same pulsotype distributed<br />

equally in North West Germany. When all the typing methods were combined together, two<br />

groups <strong>of</strong> isolates could be found in Lower Saxony, Hessen and North Rhine-Westphalia.<br />

7.4 Cytotoxicity and Virulence genes<br />

In vitro culture <strong>of</strong> fish cells can be traced back to the early 1960s, and cultivated cells were<br />

derived from rainbow trout, fathead minnow and blue-striped grunt (WOLF and QUIMBY<br />

1969). More fish cell lines were added to the list later (HIGHTOWER and RENFRO 1988;<br />

FRYER and LANNAN 1994; OTT 2004). Rachlin and Perlmutter (RACHLIN and<br />

PERLMUTTER 1968) investigated for the first time the use <strong>of</strong> cultured fathead minnow<br />

epithelial cells (FHM) in a cytotoxicity assay. Nonetheless, in vitro cytotoxicity is not<br />

comparable with in vivo data (BABICH and BORENFREUND 1991). It was reported in<br />

previous studies that in cytotoxicity tests Y. ruckeri isolates have a preference for particular<br />

cell lines, for instance a strong host cell preference for Fathead minnow epithelial (FHM)<br />

cells, but not for human derived Hep-2 cells (KAWULA et al. 1996). Y. ruckeri was reported<br />

to be able to adhere to and invade into cells from the chinook salmon embryo (CHSE-214)<br />

and rainbow trout liver cell lines (R1)(TOBBACK et al. 2010). In the present study, four<br />

139


Chapter 7 General discussion<br />

common fish cell lines: common carp brain (CCB), epithelioma papulosum cyprini (EPC),<br />

FHM and rainbow trout gonad (RTG-2) cells, were used to perform cytotoxicity tests in vitro.<br />

Isolates from all 27 typing groups showed a strong preference for FHM and EPC cells.<br />

However, they caused a lower cytotoxicity to rainbow trout gonad cells. Since one trait <strong>of</strong> this<br />

non-motile strain is that it can cause ERM outbreaks even in winter, two incubation<br />

temperatures were tested in this study, and the results indicated that non-motile Y. ruckeri<br />

isolates were more active in the first 24 hours at lower temperature. The data also showed that<br />

non motile isolates had the same ability to cause cytotoxicity in vitro as motile isolates. The<br />

data also underlined that the flagellum is not necessary for Y. ruckeri virulence.<br />

Although it was reported earlier that flagellin may not be required for proinflammatory innate<br />

immune response in rainbow trout by Y. ruckeri (JOHN TINSLEY 2010), it is still interesting<br />

to study flagellar genes and their expression in non-motile isolates. Non-motile knock-out<br />

mutants <strong>of</strong> Y. ruckeri with the absence <strong>of</strong> specific flagellar genes were usually constructed to<br />

study gene function (YOUNG et al. 1999; EVENHUIS et al. 2009). However, in real<br />

non-motile strain, the flagellar genes were still present but stay unfunctional, and there was no<br />

great difference <strong>of</strong> flagellar gene expression between motile and non-motile strains. It was<br />

suggested that the lack <strong>of</strong> motility in biotype 2 strains may not just rely on the absence <strong>of</strong><br />

certain flagellar genes, nor a lower expression <strong>of</strong> flagellar genes, but probably due to complex<br />

mechanisms.<br />

According to Secades and Guijarro, the extracellular serralysin metalloprotease Yrp1 was<br />

identified in some Y. ruckeri isolates, leading to the classification into two different groups:<br />

Azo + (presence <strong>of</strong> the Yrp1 proteolytic activity) and Azo - ( absence <strong>of</strong> the Yrp1 proteolytic<br />

activity)(SECADES and GUIJARRO 1999). All isolates from the present study were found<br />

positive for the yrp1 gene by means <strong>of</strong> PCR detection, indicating that they belong to the Azo +<br />

group.<br />

7.5 Antimicrobial resistance <strong>of</strong> Y. ruckeri<br />

In previous studies by Rucker (RUCKER 1966), sulphamethazine, chloramphenicol and<br />

oxytetracycline, followed by oxolinic acid (RODGERS and AUSTIN 1983) were<br />

recommended for the treatment <strong>of</strong> enteric redmouth disease, caused by Y. ruckeri. However,<br />

the potential risk <strong>of</strong> misuse <strong>of</strong> antimicrobial agents in treatment <strong>of</strong> ERM and the presence <strong>of</strong><br />

resistant isolates could not be overlooked. As early as 1987, Post (POST 1987) already<br />

mentioned Y. ruckeri isolates which were resistant to the therapeutic level <strong>of</strong> both<br />

140


Chapter 7 General discussion<br />

sulphamethazine and oxytetracycline. Rodgers (RODGERS 2001) found a potential ability <strong>of</strong><br />

Y. ruckeri to be resistant to oxolinic acid, oxytetracycline and the potentiated sulphonamide<br />

after 15 subcultures. In the present study, for enr<strong>of</strong>loxacin and nalidixic acid, a bimodal MIC<br />

distribution was observed with one population showing MICs <strong>of</strong> 0.008-0.015 mg/L <strong>of</strong><br />

enr<strong>of</strong>loxacin and 0.25-0.5 mg/L <strong>of</strong> nalidixic acid, respectively, and the other subpopulation<br />

exhibiting MICs <strong>of</strong> 0.06-0.25 mg/L <strong>of</strong> enr<strong>of</strong>loxacin and 8-64 mg/L <strong>of</strong> nalidixic acid,<br />

respectively. And a single isolate showed high MIC values for sulfamethoxazole and<br />

sulfamethoxazole/trimethoprim.<br />

7.5.1 Quinolone resistance<br />

For nalidixic acid, as a quinolone derivative, a bactericidal activity was first found in 1962<br />

(LESCHER et al. 1962). Since then, as a result <strong>of</strong> its wide spectrum <strong>of</strong> activity, several<br />

generations <strong>of</strong> quinolones were developed and used widely. In bacteria, there are two main<br />

chromosomal-mediated mechanisms <strong>of</strong> quinolone resistance: alterations in the targets <strong>of</strong><br />

quinolones, and decreased accumulation inside the bacterial cell because <strong>of</strong> an impermeability<br />

<strong>of</strong> the membrane and/or an overexpression <strong>of</strong> afflux pump systems (RUIZ 2003). Generally,<br />

the quinolone target in Gram-negative bacteria is the DNA gyrase, encoded by gyrA and gyrB<br />

genes, and is topoisomerase Ⅳ (encoded by the parC and parE genes) in Gram-positive<br />

bacteria (RUIZ 2003). Nucleotide substitutions varied from different positions in the sequence<br />

<strong>of</strong> GyrA, and this section is recognized as quinolone resistance-determining region (QRDR)<br />

since 1990 (YOSHIDA et al. 1990). It is also reported that the presence <strong>of</strong> a single mutation<br />

<strong>of</strong> QRDR usually leads to a high-level resistance to nalidixic acid, but additional mutation(s)<br />

in the gyrA sequence is needed for a high level <strong>of</strong> resistance to fluoroquinolones (VILA et al.<br />

1994; RUIZ et al. 2002). In this study, three different substitutions were found in QRDR at<br />

different positions, contributing to different quinolone susceptibility levels. It is suggested<br />

that the substitution <strong>of</strong> different amino acids may influence in various ways the affinity <strong>of</strong> the<br />

enzyme for the quinolone molecule, or even the interaction <strong>of</strong> this enzyme with quilonones by<br />

affecting the whole protein structure (RUIZ 2003).<br />

7.6.2 Trimethoprim/Sulfonamides resistance<br />

Sulfonamides are synthetic chemicals and were used effectively for the treatment <strong>of</strong> bacterial<br />

disease for a long period <strong>of</strong> time. Sulfamethoxazole, combined with trimethoprim is an<br />

example for the current generation <strong>of</strong> sulfa drugs. In contrary to eukaryotes, bacteria need to<br />

synthetise the folate skeleton de novo. Sulfonamides and trimethoprim play the role in<br />

141


Chapter 7 General discussion<br />

blocking a step in folic acid metabolism by either acting as a competitive inhibitor <strong>of</strong><br />

para-aminobenzoic acid (PABA) in dihydropteroate synthesis, or by blocking the enzymatic<br />

recycling <strong>of</strong> the folate coenzymes (NEU and GOOTZ 1996; BERMINGHAM and DERRICK<br />

2002). The resistance to both sulfonamides and trimethoprim is due to a horizontal and<br />

resistant gene flow among bacteria (SKOELD 2001). There are two types <strong>of</strong> resistance to<br />

sulfonamides and trimethoprim: plasmid-borne and a chromosomal mutation (SKOELD<br />

2001). The former was observed <strong>of</strong>ten in different cases (GALIMAND et al. 1997;<br />

PERRETEN and BOERLIN 2003). The plasmid-borne genes sul1 and sul2 are associated<br />

with the long known and very common sulfonamide resistance in Gram-negative bacteria<br />

(SKOELD 2001). In the present study, a ~8.9 kb plasmid was found in a single<br />

sulfonamides/trimethoprim resistant Y. ruckeri isolate, carrying the genes sul2, strB and a<br />

dfrA14 gene cassette integrated into the strA gene. The plasmid was transformable and the<br />

resistant traits could be inherited by recipient bacteria, indicating the potential risk <strong>of</strong> a wider<br />

spreading <strong>of</strong> sulfonamide/thrimethoprim resistant Y. ruckeri isolates.<br />

7.6 Conclusions<br />

In fishes from aquaculture ponds in North West Germany, non-motile strains <strong>of</strong> Y. ruckeri<br />

were dominant, especially in winter. Two or more different Y. ruckeri isolates were present in<br />

most farms visited in field study. Among the 27 typing groups <strong>of</strong> Y. ruckeri, only two groups<br />

were found distributed in the three federal states. The fatty acid composition <strong>of</strong> Y. ruckeri<br />

isolates was homogenous. Y. ruckeri isolates were more cytotoxic to EPC and FHM cells.<br />

There were no significant differences between non-motile and motile strains <strong>of</strong> Y. ruckeri on<br />

either the presence <strong>of</strong> virulence genes detected in this study or by the expression <strong>of</strong> these<br />

genes. Most <strong>of</strong> the isolates from North West Germany have not acquired resistance<br />

mechanisms. However, Y. ruckeri could still develop mutations for (fluoro) quinolone<br />

resistance, but is also able to acquire plasmid-borne resistance genes.<br />

142


Chapter 7 General discussion<br />

References<br />

ARAL, H., Y. MORITA, S. IZUMI, T. KATAGIRI and H. KIMURA (2007):<br />

Molecular typing by pulsed-field gel electrophoresis <strong>of</strong> Flavobacterium psychrophilum isolates derived from<br />

Japanese fish.<br />

J Fish Dis 30, 345-355<br />

ARIAS, C. R., O. OLIVARES-FUSTER, K. HAYDEN, C. A. SHOEMAKER, J. M. GRIZZLE and P. H.<br />

KLESIUS (2007):<br />

First report <strong>of</strong> Yersinia ruckeri biotype 2 in the USA.<br />

J Aquat Anim Health 19, 35-40<br />

AUSTIN, D. A., P. A. W. ROBERTSON and B. AUSTIN (2003):<br />

Recovery <strong>of</strong> a New Biogroup <strong>of</strong> Yersinia ruckeri from Diseased Rainbow Trout(Oncorhynchus mykiss,<br />

Walbaum).<br />

Syst Appl Microbiol 26, 127-131<br />

BABICH, H. and E. BORENFREUND (1991):<br />

Cytotoxicity and genotoxicity assays with cultured fish cells: A review.<br />

Toxicol in Vitro 5, 91-100<br />

BASTARDO, A., H. BOHLE, C. RAVELO, A. E. TORANZO and J. L. ROMALDE (2011):<br />

Serological and molecular heterogenetity among Yersinia ruckeri strains isolated from farmed Atlantic salmon<br />

Salmo salar in Chile.<br />

Dis Aqua Organ 93, 207-214<br />

BASTARDO, A., C. RAVELO and J. L. ROMALDE (2012):<br />

Multilocus sequence typing reveals high genetic diversity and epidemic population structure for the fish<br />

pathogen Yersinia ruckeri.<br />

Environ microbiol 14, 1888-1897<br />

BERMINGHAM, A. and J. P. DERRICK (2002):<br />

The folic acid biosynthesis pathway in bacteria: evaluation <strong>of</strong> potential for antibacterial drug discovery.<br />

BioEssays 24, 637-648<br />

CANONICA, F. P. and M. A. PISANO (1985):<br />

Identification <strong>of</strong> Hydroxy Fatty Acids in Aeromonas hydrophila, Aeromonas sobria and Aeromonas caviae.<br />

J clin Microbiol 22, 1061-1062<br />

DAVIES, R. L. (1991):<br />

Clonal analysis <strong>of</strong> Yersinia ruckeri based on biotypes, serotypes and outer membrane protein-types.<br />

J Fish Dis 14, 221-228<br />

DAVIES, R. L. and G. N. FRERICHS (1989):<br />

Morphological and biochemical differences among isolates <strong>of</strong> Yersinia ruckeri obtained from wide geographical<br />

areas.<br />

J Fish Dis 12, 357-365<br />

DI-GIOVANNI, G. D., L. S. WATRUD, R. J. SEIDLER and F. WIDMER (1999):<br />

Comparsion <strong>of</strong> parental and Transgenic alfalfa rhizosphere bacterial communities using biolog GN metabolic<br />

fingerprinting and enterobacterial repetitive intergenic consensus sequence-PCR (ERIC-PCR).<br />

Microbial Ecol 37, 129-139<br />

EVENHUIS, J. P., S. E. LAPATRA, D. W. VERNER-JEFFREYS, I. DALSGAARD and T. J. WELCH (2009):<br />

Identification <strong>of</strong> Flagellar Motility Genes in Yersinia ruckeri by Transposon Mutagenesis.<br />

Antimicrob Agents Ch 75, 6630-6633<br />

FOUZ, B., C. ZARZA and C. AMARO (2006):<br />

First description <strong>of</strong> non-motile Yerinia ruckeri serovar I strains causing disease in rainbow trout, Oncorhynchus<br />

mykiss (Walbaum), cultured in Spain.<br />

J Fish Dis 29, 339-346<br />

143


Chapter 7 General discussion<br />

FRYER, J. L. and C. N. LANNAN (1994):<br />

Three decades <strong>of</strong> fish cell culture: A current listing <strong>of</strong> cell lines derived from fishes.<br />

Journal <strong>of</strong> Tissue Culture Methods 16, 87-94<br />

GALIMAND, C., A. GUIYOULE, G. GERBAUD, B. RASOAMANANA, S. CHANTEAU, E. CARNIEL and<br />

P. COURVALIN (1997):<br />

Multidrug resistance in Yersinia pestis mediated by a transferable plasmid.<br />

the New England Journal <strong>of</strong> <strong>Medicine</strong> 337, 677-681<br />

GARCIA, J. A., J. L. LARSEN, I. DALSGAARD and K. PEDERSEN (2000):<br />

Pulsed-field gel electrophoresis analysis <strong>of</strong> Aeromonas salmonicida ssp. salmonicida.<br />

FEMS Microbiol Lett 190, 163–166<br />

HIGHTOWER, L. and E. RENFRO (1988):<br />

Recent applications <strong>of</strong> fish cell culture to biomedical research.<br />

Journal <strong>of</strong> Exp Zool 248, 290-302<br />

HOLMES, B., W. R. WILLCOX and S. P. LAPAGE (1978):<br />

Identification <strong>of</strong> Enterobacteriaceae by the API 20E system.<br />

J Clin Pathol 31, 22-30<br />

HULTON, C. S. J., C. F. HIGGINS and P. M. SHARP (1991):<br />

ERIC sequences: a novel family <strong>of</strong> repetitive elements in the genomes <strong>of</strong> Escherichia coli, Salmonella<br />

typhimurium and other enterobacteria.<br />

Mol Microbiol 5, 825-834<br />

KAWULA, T. H., M. J. LELIVELT and P. E. ORNDORFF (1996):<br />

Using a new inbred fish model and cultured fish tissue cells to study Aeromonas hydrophila and Yersinia ruckeri<br />

pathogenesis.<br />

Microb Pathogenesis 20, 119-125<br />

KOETH, T., J. VERSALOVIC and J. R. LUPSKI (1995):<br />

Differential subsequence conservation <strong>of</strong> interspersed repetitive Streptococcus pneumoniae BOX elements in<br />

diverse bacteria.<br />

Genome Res 5, 408-418<br />

KRAWIEC, S. and M. RILEY (1990):<br />

Organization <strong>of</strong> the bacterial chromosome.<br />

Microbiol Rev 54, 502-539<br />

LECLERCQ, A., A. GUIYOULE, M. E. LIOUI, E. CARNIEL and J. DECALLONNE (2000):<br />

High homogeneity <strong>of</strong> the Yersinia pestis fatty acid composition.<br />

J clin Microbiol 38, 1545-1551<br />

LESCHER, G. Y., E. J. FROELICH, M. D. GRUETT, J. H. BAILEY and R. P. BRUNDAGE (1962):<br />

1,8-Naphthyridine derivatives: a new class <strong>of</strong> chemotherapy agents.<br />

Journal <strong>of</strong> Medicinal and Pharmaceutical Chemistry 5, 1063-1068<br />

LUPSKI, J. R. and G. M. WEINSTOCK (1992):<br />

Short, interspersed repetitive DNA sequence in prokaryotic genomes.<br />

J Bacteriol 174, 4525-4529<br />

MALATHUM, K., K. V. SINGH, G. M. WEINSTOCK and B. E. MURRAY (1998):<br />

Repetitive sequence-based PCR versus pulsed-field gel electrophoresis for typing <strong>of</strong> Enterococcus faecalis at the<br />

subspeces level.<br />

J clin Microbiol 36, 211-215<br />

MOHAPATRA, B. R., K. BROERSMA and A. MAZUMDER (2008):<br />

144


Chapter 7 General discussion<br />

Differentiation <strong>of</strong> fecal Escherichia coli form poultry and free-living birds by (GTG) 5 -PCR genomic<br />

fingerprinting.<br />

Int J Med Microbiol 298, 245-252<br />

MOHAPATRA, B. R. and A. MAZUMDER (2008):<br />

Comparative efficacy <strong>of</strong> five different rep-PCR methods to discriminate Escherichia coli populations in aquatic<br />

enviroments.<br />

Water Sci Technol 58, 537-547<br />

NEU, H. C. and T. D. GOOTZ (1996):<br />

Antimicrobial Chemotherapy.<br />

In: Medical Microbiology<br />

4, <strong>University</strong> <strong>of</strong> Texas Medical Branch at Galveston, Galveston, USA, S.<br />

O'LEARY, W. M. (1962):<br />

The fatty acids <strong>of</strong> bacteria.<br />

Bacteriol Rev 26, 421-447<br />

OSTERHOUT, C. J., V. H. SHULL and J. D. DICK (1991):<br />

Identification <strong>of</strong> clinical isolates <strong>of</strong> gram-negative nonfermentative bacteria by an automated cellular fatty acid<br />

identification system.<br />

J clin Microbiol 29, 1822-1830<br />

PERRETEN, V. and P. BOERLIN (2003):<br />

A New Sulfonamide Resistance Gene (sul3) in Escherichia coli is widespread in the pig population <strong>of</strong><br />

switzerland.<br />

Antimicrob Agents Ch 47, 1169-1172<br />

POST, G. (1987):<br />

Enteric redmouth disease (Yersiniosis).<br />

TFH Publication, Neptune city, New Jersey, USA<br />

RACHLIN, J. W. and A. PERLMUTTER (1968):<br />

Fish cells in culture for study <strong>of</strong> aquatic toxicants.<br />

Water Res 2, 409-414<br />

RODGERS, C. J. (2001):<br />

Resistance <strong>of</strong> Yerisinia ruckeri to antimicrobial agents in vitro.<br />

Aquaculture 196, 325-345<br />

RODGERS, C. J. and B. AUSTIN (1983):<br />

Oxolinic acid for control <strong>of</strong> enteric redmouth disease in rainbow trout.<br />

Vet Rec 112, 83<br />

RUCKER, R. R. (1966):<br />

Redmouth disease <strong>of</strong> rainbow trout (Salmo gairdner).<br />

Bull <strong>of</strong>f int Epizoot. 65, 825-830<br />

RUIZ, J. (2003):<br />

Mechanisms <strong>of</strong> resistance to quinolones: target alterations, decreased accumulation and DNA gyrase protection.<br />

J Antimicrob Ch 51, 1109-1117<br />

RUIZ, J., J. G MEZ, M. M. NAVIA, A. RIBERA, J. M. SIERRA, F. MARCO, J. MENSA and J. VILA (2002):<br />

High prevalence <strong>of</strong> nalidixic acid resistant, cipr<strong>of</strong>loxacin susceptible phenotype among clinical isolates <strong>of</strong><br />

Escherichia coli and other Enterobaceriaceae.<br />

Diagn Micr Infec Dis 42, 257-261<br />

SECADES, P. and J. A. GUIJARRO (1999):<br />

Purification and characterization <strong>of</strong> an extracellular protease from the fish pathogen Yersinia ruckeri and effect<br />

<strong>of</strong> culture conditions on production.<br />

145


Chapter 7 General discussion<br />

Appl Environ Microbiol 65, 3969-3975<br />

SHOEMAKER, C. A., C. R. ARIAS, P. H. KLESIUS and T. L. WELKER (2013):<br />

Technique for identifying Flavobacterium columnare using whole-cell fatty acid pr<strong>of</strong>iles.<br />

J Aqua Anim Health 17, 267-274<br />

SKOELD, O. (2001):<br />

Resistance to trimethoprim and sulfonamides.<br />

Vet Res 32, 261-273<br />

SKOV, M. N., K. PEDERSEN and J. L. LARSEN (1995):<br />

Comparison <strong>of</strong> pulsed-field gel electrophoresis, ribotyping, and plasmid pr<strong>of</strong>iling for typing <strong>of</strong> Vibrio<br />

anguillarum serovar O1.<br />

Appl Environ Microbiol 61, 1540–1545<br />

STEVENSON, R. M. W. (1997):<br />

Immuniztion with bacterial antigens: yersiniosis.<br />

Dev Biol Stand 90, 117-124<br />

ŠVEC, P., M. VANCANNEYT, M. SEMAN, C. SNAUWAERT, K. LEFEBVRE, I. SEDL CEK and J.<br />

SWINGS (2005):<br />

Evaluation <strong>of</strong> (GTG) 5 -PCR for identification <strong>of</strong> Enterococcus spp.<br />

FEMS Microbiol Lett 247, 59-63<br />

TINSLEY, J. (2010)<br />

Studies on the pathogenicity <strong>of</strong> Yersinia ruckeri biotype 2 to rainbow trout (Oncorhynchus mykiss, Walbaum)<br />

Edinburgh, UK School <strong>of</strong> Life Sciences<br />

TINSLEY, J. W., D. A. AUSTIN, A. R. LYNDON and B. AUSTIN (2011):<br />

Novel non-motile phenotypes <strong>of</strong> Yersinia ruckeri suggest expansion <strong>of</strong> the current clonal complex theory.<br />

J Fish Dis 34, 311-317<br />

TOBBACK, E., A. DECOSTERE, K. HERMANS, W. VAN DEN BROECK, F. HAESEBROUCK and K.<br />

CHIERS (2010):<br />

In vitro markers for virulence in Yersinia ruckeri.<br />

J Fish Dis 33, 197-209<br />

VERSALOVIC, J., T. KOEUTH and J. R. LUPSKI (1991):<br />

Distribution <strong>of</strong> repetitive DNA sequences in eubacteria and application to fingerprinting <strong>of</strong> bacterial genomes.<br />

Nucleic Acids Res. 19, 6823-6831<br />

VILA, J., J. RUIZ, F. MARCO, A. BARCELO, P. GO I, E. GIRALT and T. JIMENEZ-DE-ANTA (1994):<br />

Association between double mutation in gyrA gene <strong>of</strong> cipr<strong>of</strong>loxacin-resistant clinical isolates <strong>of</strong> Escherichia coli<br />

and MICs.<br />

Antimicrob Agents Ch 38, 2477-2479<br />

WHEELER, R. W., R. L. DAVIES, I. DALSGAARD, J. GARCIA, T. J. WELCH, S. WAGLEY, K. S.<br />

BATEMAN and D. W. VERNER-JEFFREYS (2009):<br />

Yersinia ruckeri biotype 2 isolates from mainland Europe and the UK represent different clonal groups.<br />

Dis Aqua Organ 84, 25-33<br />

WOLF, K. and M. C. QUIMBY (1969):<br />

Fish cell and tissue culture.<br />

In: Fish Physiology<br />

Academic Press, New York, S. 253-305<br />

YOSHIDA, H., M. BOGAKI, M. NAKAMURA and S. NAKAMURA (1990):<br />

Quinolone resistance-determining region in the DNA gyrase gene <strong>of</strong> Escherichia coli.<br />

Antimicrob Agents Ch 34, 1271-1272<br />

146


Chapter 7 General discussion<br />

YOUNG, G. M., M. J. SMITH, S. A. MINNICH and V. L. MILLER (1999):<br />

The Yersinia enterocolitica motility master regulatory operon, flhDC, is required for flagellin production,<br />

swimming motility and swarming motility.<br />

J Bacteriol 181, 2823-2833<br />

147


148


Chapter 8. Summary<br />

149


Chapter 8 Summary<br />

Summary<br />

Epidemiological Study on Yersinia ruckeri isolates from Rainbow trout<br />

(Oncorhynchus mykiss, Walbaum) in North West Germany. Yidan Huang<br />

Yersinia ruckeri is the causative agent <strong>of</strong> Enteric Redmouth Disease, one <strong>of</strong> the most<br />

important infectious diseases in rainbow trout hatcheries in Europe. Although generally Y.<br />

ruckeri was well controlled by vaccination and antibiotic treatment, outbreaks <strong>of</strong> this disease<br />

were still periodically observed, especially in endemic areas. Moreover, some Y. ruckeri<br />

strains seem not to be affected by several <strong>of</strong> the commercial vaccines. All these strains were<br />

lacking motility, while strains from previous outbreaks were all motile. It was hypothesized<br />

that vaccination exerted a selective pressure that enabled the emergence <strong>of</strong> non-motile strains<br />

that are resistant to the commercial vaccines. Therefore, there is a high risk that these<br />

non-motile vaccine-resistant strains spread and originate severe outbreaks <strong>of</strong> disease in trout<br />

farms. The aim <strong>of</strong> this project was to investigate the genotypic and phenotypic diversity <strong>of</strong> Y.<br />

ruckeri isolates collected in North West Germany to figure out the relationship according to<br />

the colony development, compare intraspecies differences with in vitro cytotoxicity to<br />

different fish cell lines, and the expression <strong>of</strong> pathogenic factors, especially between motile<br />

and non-motile isolates, and study their susceptibility to different antimicrobial agents and the<br />

genetic determents <strong>of</strong> antimicrobial resistance to understand the situation trout farming in<br />

North Western Germany now is facing.<br />

Since 2011, 48 Y. ruckeri isolates were collected from 12 trout farms in<br />

North-Rheine-Westphalia during different seasons. Additional 33 isolates, <strong>of</strong>fered by LAVES<br />

<strong>Hannover</strong> and the fish disease service <strong>of</strong> Hessen, were from clinical cases in Lower Saxony<br />

and Hessen. One Reference strain was provided by the Clinic for Poultry at the <strong>University</strong> <strong>of</strong><br />

<strong>Veterinary</strong> <strong>Medicine</strong>, <strong>Hannover</strong>, and one typical non-motile strain was <strong>of</strong>fered by Dr. Gould<br />

from MSD Animal Health. The isolates were characterized by biochemical tests, pulsed-field<br />

gel electrophoresis (PFGE) and repetitive sequence-based PCR assays, including<br />

(GTG)5-PCR, BOX-PCR, ERIC-PCR and REP-PCR. Whole cell fatty acid composition was<br />

analyzed by gas chromatography. Cytotoxicity assays were performed using Common carp<br />

brain (CCB), epithelioma papulosum cyprini (EPC), fathead minnow epithelial cell (FHM)<br />

and rainbow trout gonad-2 (RTG-2) cells at different incubation times and temperatures. A<br />

multiplex PCR was established to detect 10 virulence factor genes (hemolysin genes yhlAB,<br />

ruckerbactin genes rucC and rupG, ABC exporter protein system genes yrp1 and yrpDEF ,<br />

150


Chapter 8 Summary<br />

flagellar secretion chaperones gene flgA, flagellar secretion apparatus gene flhA) found in Y.<br />

ruckeri. The expression <strong>of</strong> these genes was analyzed and compared by semi-quantitative PCR.<br />

Susceptibility tests to antimicrobial substances and genetic determents for a resistance to these<br />

substances were detected by microdilution test and PCR.<br />

Eight different API 20E pr<strong>of</strong>iles were obtained, which were different from that <strong>of</strong> the Y.<br />

ruckeri reference strain DSM18506. Five isolates were confirmed as Y. ruckeri by 16S rDNA<br />

sequencing. 17 isolates hydrolyzed Tween 80/20. All Y. ruckeri isolates exhibited 15 major<br />

fatty acids, including 12:0, 13:0, 13.957 (equivalent chain length, ECL unknown), 14:0,<br />

14.502 (ECL unknown), 15:0, 16:1ω5c, 16:0, 17:1ω8c, 17:0 CYCLO, 17:0, 16:1 2OH,<br />

18:1ω9c, 18:1ω7c and 18:0. In addition to 17 PFGE patterns, two different REP-PCR patterns,<br />

five ERIC-PCR patterns, four (GTG) 5 -PCR patterns and three BOX-PCR patterns were<br />

obtained. According to the results <strong>of</strong> API 20E, PFGE and the various PCR assays, the isolates<br />

could be divided into 27 different groups. In rainbow trout hatcheries, isolates from more than<br />

two different typing groups were present in the same fish farm. Two groups <strong>of</strong> isolates were<br />

found to be present in all three federal states in North West Germany. Non-motile strains were<br />

more active to cause in vitro cytotoxicity to fish cell lines than motile strain in the first 24 h,<br />

but without significant difference after longer incubation at the lower temperature. All studied<br />

virulence genes were found present in both motile and non-motile strains <strong>of</strong> Y. ruckeri. In<br />

addition, no significant differences were found in the expression <strong>of</strong> these genes. For most <strong>of</strong><br />

the antimicrobial agents tested, a unimodal distribution <strong>of</strong> MIC values was observed. In<br />

general, the MIC values increased with the incubation time. For enr<strong>of</strong>loxacin and nalidixic<br />

acid, a bimodal MIC distribution was observed with one population showing MICs <strong>of</strong><br />

0.008-0.015 mg/L <strong>of</strong> enr<strong>of</strong>loxacin and 0.25-0.5 mg/L <strong>of</strong> nalidixic acid, respectively, and the<br />

other subpopulation exhibiting MICs <strong>of</strong> 0.06-0.25 mg/L <strong>of</strong> enr<strong>of</strong>loxacin and 8-64 mg/L <strong>of</strong><br />

nalidixic acid, respectively. While isolates (n=3) from the subpopulation with the lower<br />

(fluoro)quinolone MICs showed the non-mutated gyrA wildtype QRDR sequence, another ten<br />

isolates from the subpopulation with the higher (fluoro)quinolone MICs exhibited different<br />

single bp mutations that resulted in single amino acid substitutions in the GyrA protein: Ser<br />

Arg or Ser Ile (at position 83) or Asn Tyr (at position 87) [Escherichia coli<br />

numbering]. A single isolate showed high MIC values for sulfamethoxazole and<br />

sulfamethoxazole/trimethoprim. A ~8.9 kb plasmid was found in this isolate, which carried<br />

the genes sul2, strB and a dfrA14 gene cassette integrated into the strA gene. Neither class 1<br />

nor class 2 integrons were found in this isolate.<br />

151


Chapter 8 Summary<br />

In conclusion, there was a variety <strong>of</strong> Y. ruckeri isolates in North West German aquaculture,<br />

dominated by non-motile strain. In most farms from the field study, two or more different Y.<br />

ruckeri isolates were present. It was easier for non-motile strains to cause outbreaks in winter,<br />

since they were more active than motile strain at lower temperature. The fatty acid<br />

composition <strong>of</strong> Y. ruckeri isolates was homogenous and there was no obvious link between<br />

the fatty acid pr<strong>of</strong>iles and the epidemiological parameters. Non-motile strains <strong>of</strong> Y. ruckeri<br />

could not be differentiated from motile strains by either the presence <strong>of</strong> virulence genes<br />

detected in this study or by the expression <strong>of</strong> these genes. Incubation for 24h-28h at 22±2 °C<br />

appears to be sufficient for the testing <strong>of</strong> Y. ruckeri for susceptibility to antimicrobial<br />

substances. The largely unimodal MIC distribution suggests that most <strong>of</strong> the isolates tested<br />

represent the wildtype population that has not acquired resistance mechanisms. Exceptions<br />

were (i) the finding <strong>of</strong> isolates with elevated (fluoro)quinolone MICs that also had<br />

resistance-mediating mutations in the QRDR region <strong>of</strong> gyrA, and (ii) the single isolate that<br />

carried plasmid-borne sul2, dfrA14, ΔstrA and strB genes. This observation confirmed that Y.<br />

ruckeri is able to develop mutations for (fluoro)quinolone resistance, but can also acquire<br />

plasmid-borne resistance genes.<br />

152


Chapter 9. Zusammenfassung<br />

153


Chapter 9 Zusammenfassung<br />

Zusammenfassung<br />

Eine epidemiologische Studie an Yersinia ruckeri-Isolaten aus Regenbogenforellen<br />

(Oncorhynchus mykiss, Walbaum) in Nord-West Deutschland. Yidan Huang<br />

Yersinia ruckeri, der Erreger der Rotmaulseuche (Enteric Redmouth Disease,ERM), ist eine<br />

der bedeutendsten Infektionskrankheiten in der Regenbogenforellenzucht in Europa. Obwohl<br />

im allgemeinen Y. ruckeri gut durch Impfungen und Antibiotikabehandlungen kontrolliert<br />

werden konnte, waren Ausbrüche dieser Krankheit noch regelmäßig zu beobachten, vor allem<br />

in endemischen Gebieten. Darüber hinaus scheinen einige Y. ruckeri Stämme nicht auf<br />

bestimmte kommerzielle Impfst<strong>of</strong>fe zu reagieren. Alle diese Stämme waren unbeweglich,<br />

während in früheren Ausbrüchen isolierte Stämme von beweglich waren. Es wurde vermutet,<br />

dass die Impfung einen selektiven Druck auf die Bakterien ausübt, der das Entstehen von<br />

nicht beweglichen Stämmen ermöglichte, die gegen die kommerziellen Vakzine<br />

unempfindlich sind. Daher besteht ein hohes Risiko, dass sich diese nicht beweglichen<br />

Vakzine-resistenten Stämme ausbreiten und schwere Krankheitsausbrüche in Forellenfarmen<br />

verursachen können. Das Ziel dieses Projektes war, genotypische und phenotypische<br />

Unterschiede von Y. ruckeri Isolaten aus Nord-West Deutschland zu erfassen und zu<br />

analysiere. Es wurden unterschiedliches Koloniewachstum ermittelt, Intraspezies-Vergleiche<br />

in Hinblick auf in vitro Toxizität gegenüber verschiedenen Fischzelllinien, der Expression von<br />

Pathogenitätsfaktoren insbesondere zwischen beweglichen und unbeweglichen Isolaten<br />

angestellt und die Anfälligkeit von Isolaten gegenüber verschiedenen antimikrobiellen<br />

Substanzen sowie die genetischen Faktoren einer antimikrobiellen Resistenz untersucht. Ziel<br />

war es die Situation zu verstehen, mit der die Forellenzucht in Nord-West Deutschland jetzt<br />

konfrontiert ist.<br />

Seit 2011 wurden 48 Y ruckeri Isolate von 12 Forellenteichwirtschaften in<br />

Nordrhein-Westfalen während verschiedener Jahreszeiten gesammelt. Zusätzliche 33 Isolate<br />

wurden vom LAVES <strong>Hannover</strong> und dem Fischgesundheitsdienst Hessen aus klinischen<br />

Ausbrüchen von ERM in Niedersachsen und Hessen zur Verfügung gestellt. Ein<br />

Referenzstamm wurde von der Klinik für Geflügel der Tierärztlichen Hochschule <strong>Hannover</strong><br />

und ein typischer nicht beweglicher Stamm wurde von Dr. Gould vom MSD Animal Health<br />

bereitgestellt. Alle Isolate wurden durch biochemische Tests, Puls-Feld-Gelelektrophorese<br />

(PFGE) und Repetitive Sequenz-basierte PCR Analysen, inklusive (GTG)5-PCR, BOX-PCR,<br />

ERIC-PCR und REP-PCR, charakterisiert. Die Fettsäuremuster der Isolate wurden mittels<br />

Gaschromatographie analysiert. Die Cytotoxicität wurde an „Common carp brain“ (CCB),<br />

„Epithelioma papulosum cyprini (EPC), „fathead minnow“ Epithelzellen (FHM) und<br />

154


Chapter 9 Zusammenfassung<br />

„Rainbow trout gonad-2“ (RTG-2) Zellinien bei verschiedenen Inkubationszeiten und<br />

Temperaturen untersucht. Eine Multiplex PCR wurde etabliert, um 10 verschiedene<br />

Virulenzfaktor-Gene (Hämolysingene yhlAB, Ruckerbactingene rucC and rupG, Gene des<br />

ABC Proteinexportersystems yrp1 and yrpDEF, Flagellare Secretionschaperon- Gene flgA,<br />

Gene des Flagellaren Sekretionsapparates flhA) von Y. ruckeri zu detektieren. Die Expression<br />

dieser Gene wurde mit einer semi-quantitativen PCR analysiert und verglichen.<br />

Empfindlichkeitstests gegenüber antimikrobiellen Substanzen und genetischen Eigenschaften<br />

für eine Resistenz gegenüber diesen Substanzen wurde mittels eines Mikrodillutionstest und<br />

einer PCR untersucht.<br />

Aus den Isolaten konnten acht verschiedenen API 20E Pr<strong>of</strong>ile erstellt werden, die sich vom Y.<br />

ruckeri Referenzstamm DSM 18506 unterschieden. Fünf Isolate wurden mittels 16s<br />

rDNA-Sequenzierung als Y. ruckeri bestätigt. 17 Isolate konnten Tween 80/20 hydrolisieren.<br />

Alle Y. ruckeri Isolate zeigten 15 wesentliche Fettsäuren, darunter 12:0, 13:0, 13.957 (Gleiche<br />

Kettenlänge, ECL unbekannt), 14:0, 14.502 (ECL unbekannt) 15:0 16:1ω5c, 16:0, 17:1ω8c,<br />

17:0 CYCLO, 17:0, 16:1 2OH, 18:1ω9c, 18:1ω7c und 18:0. Zusätzlich zu 17 verschiedenen<br />

Mustern der PFGE wurden 2 verschiedene REP-PCR, 5 verschiedene ERIC-PCR, vier<br />

(GTG) 5 –PCR und drei Muster in der BOX-PCR Analyse entdeckt. Entsprechend der<br />

Resultate aus API 20, PFGE und den verschiedenen PCR Analysen konnten die Isolate in 27<br />

verschiedene Gruppen unterteilt werden. In den Regenbogenforellen der meisten untersuchten<br />

Fischfarmen wurden Isolate aus mehr als zwei verschiedenen Typisierungs Gruppen isoliert.<br />

Es stellte sich heraus, dass zwei der Isolat-Gruppen in allen drei Bundesländern Nord-West<br />

Deutschlands nachgewiesen werden konnten. Nicht bewegliche Stämme haben auf<br />

Fischzelllinien in den ersten 24 Stunden eine höhere in vitro Zytotoxizität als bewegliche<br />

Stämme, aber nach längerer Inkubationszeit bei niedrigen Temperaturen gibt es keine<br />

signifikanten Unterschiede mehr. Alle untersuchten Virulenz-Gene wurden sowohl in den<br />

beweglichen sowie auch in den beweglichen Y. ruckeri Stämmen gefunden. Weiterhin<br />

konnten in der Expression dieser Gene keine signifikanten Unterschiede nachgewiesen<br />

werden. Für die meisten Antibiotika konnte eine unimodale Verteilung der MHK beobachtet<br />

werden. Im Allgemeinen stiegen die MHK mit der Inkubationszeit an. Für Enr<strong>of</strong>loxacin und<br />

Nalidixinsäure wurde eine bimodale Verteilung der MHK Werte festgestellt. Die eine<br />

Population zeigte MHK Werte von 0,008-0,015 mg/l bei Enr<strong>of</strong>loxacin und 0,25-0,5 mg/l bei<br />

Nalidixinsäure, während die andere Subpopulation MHK Werte von0,06-0,25 mg/l für<br />

Enr<strong>of</strong>loxacin und 8-64 mg/l für Nalidixinsäure zeigte. Während Isolate (n=3) aus der<br />

Subpopulation mit der niedrigeren MHK für (Fluor)chinolon eine nicht mutierte gyrA<br />

155


Chapter 9 Zusammenfassung<br />

Wildtyp QRDR Gensequenz besaßen, zeigten (n=10) Isolate der Subpopulation mit der<br />

höheren MHK für (Fluor)chinolon eine Punktmutation eines Basenpaares in der gyrA<br />

Gensequenz. Dies resultiert in einer Änderung einer Aminosäure im GyrA Protein: Ser zu Arg<br />

oder Ser zu Ile (an Position 83) oder Asn zu Tyr (an Position 87) (Escherichia coli<br />

Nummerierung). Ein einzelnes Isolat zeigte hohe MHK für Sulfamethoxazol und<br />

Sulfamthoxazol/Trimethoprim. Ein etwa 8.9kb großes Plasmid wurde in diesem Isolat<br />

gefunden, dass die Gene sul2 strB und eine dfrA14 Genkasette die in das strA Gen integriert<br />

war aufwies. Weder Klasse 1 noch Klasse 2 Integrine konnten in diesen Isolaten gefunden<br />

werden.<br />

Schlussfolgernd hieraus kann gesagt werden, das es eine Vielzahl von Y. ruckeri Stämmen in<br />

Nordwest Deutschland gibt, die zu einem Großteil aus nicht motilen Stämmen bestehen. In<br />

den meisten Farmen aus der Feldstudie waren zwei oder mehr verschiedene Y. ruckeri Isolate<br />

vorhanden. Im Winter war es einfacher für nicht bewegliche Stämme Krankheitsausbrüche<br />

hervorzurufen, da sie bei niedrigeren Temperaturen aktiver als die beweglichen Stämme<br />

waren. Die Fettsäure Zusammensetzung von Y. ruckeri Isolaten war homogen und es gab<br />

keine <strong>of</strong>fensichtliche Verbindung zwischen den Fettsäuremustern und den epidemiologischen<br />

Parametern. Bewegliche Stämme von Y. ruckeri konnten in dieser Studie von nicht<br />

beweglichen Stämmen weder durch die Präsenz von Virulenzantigenen noch durch die<br />

Expression dieser Gene unterschieden werden. Die Inkubation von 24-28 Stunden bei<br />

22+/-2°C scheint ausreichend zu sein, um die Empfindlichkeit von Y. ruckeri gegenüber<br />

antimikrobiellen Substanzen zu testen. Die große unimodale MHK Verteilung lässt vermuten,<br />

dass die meisten getesteten Isolate den Wildtyp repräsentieren, der keine<br />

Resistenzmechanismen erworben hat. Erwartungen waren (i) das Auffinden von Isolaten mit<br />

erhöhten MHK für (Fluor)chinolone die außerdem eine Resistenz vermittelnde Mutation in<br />

der QRDR Region des gyrA Gens, und (ii) Isolate, die Plasmidvermittelte sul2, dfrA14, ΔstrA<br />

und strB Gene tragen. Diese Untersuchungen bestätigten, dass Y. ruckeri fähig ist, Mutationen<br />

für eine Fluor)chinolon Resistenz zu entwickeln, aber ebenso Plasmid vermittelte Resistenzen<br />

zu erwerben.<br />

156


Acknowledgement<br />

Acknowledgement<br />

Firstly, I would like to thank my supervisor Pr<strong>of</strong>. Dr. Dieter Steinhagen for his help and<br />

support during my three-year PhD work. I think I am lucky enough to meet him and to be<br />

supervised by him. Pr<strong>of</strong>. Steinhagen is always nice to guide my work, patient to answer and<br />

explain all my questions, and gave me lots <strong>of</strong> great ideas. He encouraged me to be<br />

self-confident in my presentations, scientific work and daily life.<br />

Secondly, I want to thank other two members in my supervisor group, Pr<strong>of</strong>. Dr. Martin Runge<br />

and Pr<strong>of</strong>. Dr. Lothar Kreienbrock. It is helpful to have your guidance during my work. Thanks<br />

for the patience and comment for all my meetings, presentations and manuscripts.<br />

I am particularly grateful to Pr<strong>of</strong>. Dr. Friedhelm Jaeger in the Ministry <strong>of</strong> the Climate<br />

Protection, Environment, Agriculture, Conservation and Consumer Protection <strong>of</strong> the State <strong>of</strong><br />

North-Rhine Westphalia. He <strong>of</strong>fered me the chance to study in Germany and gave me all<br />

kinds <strong>of</strong> help, concern and support in my project and my daily life.<br />

I really appreciate the help from Pr<strong>of</strong>. Dr. Stefan Schwarz and his team in FLI Mariensee.<br />

Pr<strong>of</strong>. Schwarz, Dr. Geovana Brenner Michael and Ms. Roswitha Becker helped me a lot with<br />

my work. With their expertise, my experience went well. Travelling to Mariensee and<br />

working with this group are great experience for my life in Germany. I have learnt a lot. I<br />

thank Dr. Arne Jung, Dr. Martin Ryll and all technicians in clinic for poultry in <strong>University</strong> <strong>of</strong><br />

<strong>Veterinary</strong> <strong>Medicine</strong>, <strong>Hannover</strong> for the help <strong>of</strong> identification and fatty acid analysis. I also<br />

want to thank the people in LAVES, <strong>Hannover</strong>, especially Sabina, Sandra and Marie for all<br />

help and caring in my life. I thank everyone in our “fish group”: Patricia, Verena, Hamdan,<br />

John, Karina, Agnes, Marc, Birgit, Mikolaj, Graham, and Charlie for the technical support<br />

and valuable ideas for my experiment. It is really nice to work together as a big „fish family‟.<br />

I thank Mr. Dieter Mock and the whole group in LANUV, Albaum for assisting my field<br />

study. For the whole sampling year, Mr. Mock drove me to different fish farms in<br />

North-Rhine Westphalia and helped a lot with the sampling and isolation work. Special<br />

Thanks are given to Dr. Johannes-Werner Schaefer and his family: Berta, Katrin, Andrea and<br />

Sebastian. Thanks for the accommodation during my sampling time and the kind invitation<br />

157


Acknowledgement<br />

for Christmas. It made me feel like home to stay with them. It is my sweet memories in my<br />

life.<br />

Many thanks to Dr. Silke Braune (LAVES <strong>Hannover</strong>), Dr. Chris Gould (MSD Animal Health)<br />

and Dr. Joachim Nilz (LHL Hessen) for providing me some Yersinia ruckeri isolates for my<br />

project.<br />

I would like to thank all my friends in Germany and China, especially Rindra, Natasha and<br />

Xin for the nice suggestions and help with my work, and all the encouragement and supports.<br />

It is great to spend time with you together and it is my precious memory.<br />

I owe my deepest gratitude to my parents. Thanks a lot for letting me pursue my dream freely,<br />

for the caring and supporting in my whole life.<br />

Finally, I want to thank China Scholarship Council for financial support during my stay in<br />

Germany.<br />

158

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!